Next Article in Journal
Plasma-Treated Water Effect on Sporulating Bacillus cereus vs. Non-Sporulating Listeria monocytogenes Biofilm Cell Vitality
Previous Article in Journal
Applied Microbiology for Sustainable Agricultural Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanistic Role of Heavy Metals in Driving Antimicrobial Resistance: From Rhizosphere to Phyllosphere

by
Rahul Kumar
1,
Tanja P. Vasić
2,
Sanja P. Živković
2,
Periyasamy Panneerselvam
3,
Gustavo Santoyo
4,
Sergio de los Santos Villalobos
5,
Adeyemi Nurudeen Olatunbosun
6,
Aditi Pandit
7,
Leonard Koolman
8,
Debasis Mitra
1,* and
Pankaj Gautam
1,*
1
Department of Microbiology, Graphic Era (Deemed to be University), Dehradun 248002, Uttarakhand, India
2
Faculty of Agriculture, University of Niš, Kruševac 4, 37000 Niš, Serbia
3
Crop Production Division, ICAR-Central Rice Research Institute, Cuttack 753006, Odisha, India
4
Instituto de Investigaciones Químico-Biológicas, Universidad Michoacana de San Nicolás de Hidalgo, Morelia 58030, Michoacán, Mexico
5
Instituto Tecnológico de Sonora, 5 de Febrero 818 sur, Ciudad Obregón 85000, Sonora, Mexico
6
Department of Plant Physiology and Crop Production, Federal University of Agriculture Abeokuta, Abeokuta P.M.B 2240, Nigeria
7
Genetics Department, University of Georgia, Athens, GA 30602, USA
8
Centre for AMR & One Health Research, School of Biological, Health & Sports Sciences, Technological University Dublin, D07 XN77 Dublin, Ireland
*
Authors to whom correspondence should be addressed.
Appl. Microbiol. 2025, 5(3), 79; https://doi.org/10.3390/applmicrobiol5030079
Submission received: 25 May 2025 / Revised: 17 July 2025 / Accepted: 30 July 2025 / Published: 4 August 2025

Abstract

Heavy metal pollution represents a pervasive environmental challenge that significantly exacerbates the ever-increasing crisis of antimicrobial resistance and the capacity of microorganisms to endure and proliferate despite antibiotic interventions. This review examines the intricate relationship between heavy metals and AMR, with an emphasis on the underlying molecular mechanisms and ecological ramifications. Common environmental metals, including arsenic, mercury, cadmium, and lead, exert substantial selective pressures on microbial communities. These induce oxidative stress and DNA damage, potentially leading to mutations that enhance antibiotic resistance. Key microbial responses include the overexpression of efflux pumps that expel both metals and antibiotics, production of detoxifying enzymes, and formation of protective biofilms, all of which contribute to the emergence of multidrug-resistant strains. In the soil environment, particularly the rhizosphere, heavy metals disrupt plant–microbe interactions by inhibiting beneficial organisms, such as rhizobacteria, mycorrhizal fungi, and actinomycetes, thereby impairing nutrient cycling and plant health. Nonetheless, certain microbial consortia can tolerate and detoxify heavy metals through sequestration and biotransformation, rendering them valuable for bioremediation. Advances in biotechnology, including gene editing and the development of engineered metal-resistant microbes, offer promising solutions for mitigating the spread of metal-driven AMR and restoring ecological balance. By understanding the interplay between metal pollution and microbial resistance, we can more effectively devise strategies for environmental protection and public health.

Graphical Abstract

1. Introduction

The continuing increase in antimicrobial resistance (AMR) is an international health crisis that threatens to destroy the pillars of contemporary medicine [1]. The World Health Organization (WHO) has ranked AMR as a global health threat that threatens a post-antibiotic era in which routine infections become lethal [2]. As estimated recently AMR accounted fully for 1.27 million deaths in 2019, and antimicrobial resistance remains an increasingly looming global health and economic danger [3]. In 2019, AMR was attributable to 1.27 million deaths and linked to 4.95 million deaths worldwide [4]. The most recent exhaustive estimate in 2021 indicates a decline with 1.14 million directly attributable deaths due to AMR and 4.71 million linked deaths [5]. AMR will be responsible for a maximum of 169 million deaths between 2025 and 2050 [6]. The mortality rates for individuals above the age of 70 are anticipated to increase to 1.26 million by 2050 [7]. Antimicrobial resistance refers to the capacity of microorganisms, including bacteria, fungi, parasites, and viruses, to respond to medications previously effective in treating infections. When microbes develop resistance, standard treatments become ineffective, infections persist, and the risk of transmission increases. While AMR occurs naturally through genetic mutations, its progression is accelerated by misuse and overuse of medicines for human and veterinary pesticides and biofertilizers for agriculture [8,9,10], including the use of inappropriate prescribing, lack of diagnostic tools, incomplete treatment courses, and extensive antibiotic use in animal feed. Resistant organisms can endure multiple antibiotics, causing common infections, such as pneumonia, urinary tract infections, and bloodstream infections, which are increasingly difficult to treat [11]. This resistance results in prolonged illness, elevated medical costs, increased mortality, and jeopardizes modern medical procedures, including surgery, chemotherapy, and organ transplants, which depend on effective antimicrobial [12].
The WHO recognizes AMR as one of the top ten global public health threats requiring urgent coordinated efforts across sectors to monitor, prevent, and control its spread [13]. This mounting crisis requires an in-depth understanding of the complex factors that fuel AMR, such as the frequently neglected contribution of heavy metal pollution [14]. Heavy metals (HMs) such as mercury (Hg), lead (Pb), cadmium (Cd), arsenic (As), copper (Cu), and zinc (Zn) are widespread environmental contaminants released from industrial effluents, mining, agricultural activities, and inadequate waste management [15,16]. Repeated exposure to soil, water, and sediments imposes constant selective pressure on microbial populations, enabling the development and growth of metal-resistant bacteria [17]. A surprisingly large percentage of these metal-resistant bacteria are cross-resistant to antibiotics, constituting a formidable public health threat [18]. Nonetheless, resource mobilization is unbalanced because sharp cuts in aid for low- and middle-income countries (LMICs) have threatened AMR programs [19].
Surveillance and diagnostic capacity are also at risk, with more than 700 global laboratories facing closure [20], whereas areas such as Africa strengthen their pathogen surveillance systems using a One Health approach [21]. Innovation is encouraged by international platforms such as the GAMRIF Summit 2025 [22], complemented by regional research such as India’s Symposium Conference, which focuses on AI-assisted diagnostics [23]. Policymaking remains a process, although 10% of the 178 nations with National Action Plans (NAPs) currently allocate AMR budgets [24]. The pathways through which HMs drive AMR are both multifaceted and interlinked. HMs trigger oxidative stress in bacterial cells through the production of reactive oxygen species (ROS) that harm DNA, proteins, and lipids [25]. They also directly interact with cellular structures, interfering with important metabolic processes and cellular functions [26]. Recent research has reaffirmed previous reservations regarding trace heavy metals, especially Cd, which are key to the development of bacterial resistance. A study conducted by Ali et al. (2022) revealed that concentrations of Cd as low as 0.1 mg/L were able to suppress microbial growth and initiate adaptive resistance processes [27]. Following this, Shaw and Dussan (2015) indicated that above 0.5 mg/L concentrations triggered the induction of metal efflux pumps in Pseudomonas and Escherichia coli [28].
Recent research (2021–2024) has identified even more complex mechanisms. For example, regulatory factors such as CadR and CzcR3 control resistance genes [29] and emphasize microbial mechanisms such as sequestration and enzymatic detoxification of Cd ions. Synthetic biology strategies have also facilitated increased resistance by designing efflux systems and biosensors [30]. Similarly, exposure to lead (Pb) at concentrations of 0.5 and 1.0 mg/L has been proven to increase biofilm development and horizontal gene transfer (HGT), both determinants of AMR dissemination [31]. Mercury (Hg), although at very low levels (>0.01 mg/L), particularly in its methylated state (MeHg), interferes with DNA repair and enhances mutagenesis within bacterial populations [32]. Arsenic (As) tolerance and resistance gene expressions occur at levels greater than 0.5 mg/L, and Bacillus and Pseudomonas exhibit arsenate reductase ability [33]. Copper between 1.5 and 2.5 mg/L causes oxidative stress and enhances multidrug resistance (MDR) gene expression [34]. Zinc levels of 5–10 mg/L have been correlated with enhanced antibiotic resistance through co-selection [35].
These mechanisms tend to overlap with those that confer antibiotic resistance, resulting in the co-selection and spread of MDR [36]. Existing evidence suggests that co-selection of metal and antibiotic resistance genes is ubiquitous, with integrons, transposons, and plasmids acting as vectors for the transmission of resistance [37]. Co-selection is the most concerning in heavy-metal- and antibiotic-contaminated environments, including hospital wastewater, industrial effluents, and agricultural soils [38]. Recent studies have revealed highly evolved mechanisms through which microbes interact with and alter plant cell walls, a dynamic interaction that is at the heart of both plant defense and pathogen invasion [39]. Pathogenic microbes release a versatile repertoire of cell wall-degrading enzymes (CWDEs), including cellulases, pectinases, xylanases, and ligninases, that break down structural polysaccharides and allow penetration into plant tissues [40]. Several bacteria and fungi also secrete effector proteins that regulate host cell wall remodeling pathways, inhibit immune responses, or take over host metabolism to facilitate colonization [41]. Conversely, valuable microbes, such as rhizobacteria and mycorrhizal fungi, can activate plant genes that mediate cell wall strengthening, including callose deposition and lignification, within systemic mechanisms of resistance [42]. Recent developments in imaging, transcriptomics, and metabolomics have shown that the microbial impact on plant cell wall structure is more suitable than had previously been appreciated, including not just degradation but also structural remodeling and signaling crosstalk that redefine host-microbe interactions.
A study in India found a strong correlation between heavy metal pollution in hospital effluents and the occurrence of multidrug-resistant bacteria, emphasizing the clinical significance of this phenomenon [43]. Heavy metals have implications beyond the clinical environment, affecting soil microbial populations and plant–microbe interactions [44]. The rhizosphere, an important region for plant growth, is highly susceptible to metal toxicity. HMs such as Pb, Cd, and Hg inhibit nitrogen fixation, causing a decline in crop yield and soil fertility [45]. For example, research has indicated that Cd pollution can decrease nitrogen fixation in legumes by as much as 50% [46]. Nevertheless, some rhizospheric bacteria such as Pseudomonas and Bacillus have impressive metal tolerance through detoxification enzymes, biosorption, and bioaccumulation. These microorganisms are important for reducing metal toxicity and for soil health. This review provides mechanistic information regarding the effects of HMs on antimicrobial resistance.

1.1. Molecular Mechanisms

Heavy metal contamination imposes a severe selective pressure on microorganisms, inducing molecular reactions that unintentionally foster AMR. Major environmental metals such as arsenic (As), mercury (Hg), cadmium (Cd), and lead (Pb) impose oxidative stress, destabilize the cellular redox balance, and cause DNA damage, all of which have the potential to contribute to mutations [47]. These stressors induce global regulatory systems, such as the SOS response, where RecA and LexA proteins control error-prone DNA repair, leading to the accumulation of resistance mutations [48]. Moreover, heavy metals induce biofilm development by activating quorum sensing and stress-response regulators such as rpoS to ensure survival in adverse environments [49]. Detoxifying enzymes, such as arsenate reductases, superoxide dismutase, and catalase, are also expressed to neutralize the toxicity of metals. Antibiotic and metal resistance genes tend to co-locate on plasmids, integrons, or transposons to facilitate their transfer through HGT [50]. Consequently, without antibiotic pressure, heavy metal exposure can select for antibiotic resistance traits through co-selection mechanisms. The molecular co-overlap between antibiotic resistance and metal resistance emphasizes the importance of controlling AMR in metal-contaminated environments and indicates that control measures must be integrated and account for both selective agents at the same time.

1.2. Ecological Impacts

Heavy metal pollution has significant ecological implications, particularly in terms of microbial diversity and ecosystem processes [51]. On land, particularly in the rhizosphere, the region around plant roots that is affected by metal buildup interferes with symbiotic plant–microbe relationships. Heavy metals decrease the populations of Plant Growth-Promoting Rhizobacteria (PGPR), arbuscular mycorrhizal fungi, and actinomycetes, which are involved in nutrient cycling, nitrogen fixation, and soil fertility [52]. The resulting microbial imbalance inhibits plant growth, diminishes crop yield, and degrades the long-term soil health. Freshwater ecosystems are also targeted; heavy metal inputs in industrial effluents, mining, and agriculture reshape microbial community composition, repressing indigenous, non-resistant species in favor of metal- and drug-resistant bacteria. These resistant populations are potential reservoirs of AMR genes, enabling their dissemination through horizontal gene transfer to environmental and pathogenic microbes [53]. Furthermore, metals such as Cu and Zn are commonly applied as agricultural fertilizers and feed additives, inadvertently favoring AMR in animals and nearby soil or water environments [54]. Bioaccumulation of metals in microbial food webs also amplifies environmental risks, with ripples extending to soil fauna, aquatic life, and human health via tainted produce or water. Metal-induced microbial stress also increases the generation of reactive oxygen species (ROS), triggering inflammation and DNA damage in higher organisms [55]. Hence, heavy metal pollution not only endangers microbial ecosystem services but also intensifies the global AMR burden. Unraveling these interactive processes is critical for ecosystem restoration and formulating sustainable agriculture and waste management practices in metal-affected areas [56,57].

1.3. Bioremediation Strategies

Bioremediation, or the application of microorganisms to decontaminate toxic environments, has immense potential for reducing heavy metal pollution and its role in AMR. Pseudomonas, Bacillus, Rhizobium, and Streptomyces are among the microbial groups that have developed resistance to toxic metals and can transform them into less toxic forms using mechanisms such as biosorption, biotransformation, bioaccumulation, and biomineralization [58]. These microbes sequester metal ions from their cell walls, extracellular polymeric substances (EPS), or metal-chelating proteins such as metallothioneins. Others form enzymes such as arsenate reductase, mercuric reductase, and peroxidases to enzymatically convert metals into insoluble or less toxic forms [59]. When integrated with phytoremediation, these microbes increase the plant uptake of metals and prevent oxidative damage, facilitating the recovery of contaminated soils. The use of engineered microbial communities composed of strains with specially designed detoxification abilities can also enhance remediation in situations where contamination is complex. Advancements in synthetic biology and CRISPR-mediated gene editing enable the creation of super-tolerant strains that are more resistant and effective at metal transformation [60]. Microbial bioreactors and engineered wetlands inoculated with these microbes have been found to effectively remediate wastewater and industrial effluents. In addition, horizontal gene transfer suppression and quorum-sensing inhibition research has the potential to minimize the risk of resistance spreading during remediation [61]. To achieve optimal efficiency, bioremediation needs to be complemented by environmental regulations that control the release of metals and limit metal-based antimicrobial use in agriculture [62]. Finally, microbial bioremediation presents a sustainable, low-cost means of tackling both environmental pollution and AMR risks, conserving soil fertility and water quality.

2. Effect of Metals on Plant Rhizospheric Flora

HMs such as Cd, Pb, Hg, and As are key environmental pollutants that harm rhizospheric microbiota and are very important for plant health, nutrient cycling, and soil fertility [63]. Contaminants can cause toxicity, alter microbial populations, and impede physiological functions. They bind with cellular proteins, DNA, and enzymes, causing structural and functional injuries. The primary damage is the suppression of microbial variety and growth, especially in beneficial actinomycetes, mycorrhizal fungi, and rhizobacteria [64]. Rhizobacteria provide nitrogen fixation and phosphorus solubilization, but heavy metal pollution diminishes their populations, decreases nutrient content, and damages plant growth. Mycorrhizal fungi, such as Glomus spp., enhance nutrient acquisition and resistance to soil pathogens; however, heavy metal toxicity inhibits fungal spore germination, hyphal growth, and mycorrhizal development, thereby increasing plant susceptibility to environmental stress [65]. Actinomycetes that decompose organic matter and produce antibiotics are adversely affected by heavy metal pollution (HMP) [66]. Bioremediation, organic amendments, and phytoremediation-based remediation are helpful for achieving microbial equilibrium, ensuring soil ecosystem sustainability, and enhancing agricultural productivity [67].

2.1. Nitrogen-Fixing and Phosphorus-Solubilizing Rhizobacteria

Rhizobacteria play a critical role in promoting plant growth through nitrogen fixation and phosphorus solubilization, which are essential for plant nutrition and soil fertility [68]. These bacteria, including Rhizobium, Azospirillum, and Bacillus species, convert atmospheric nitrogen into bioavailable forms and release organic acids that solubilize insoluble phosphorus compounds. However, heavy metal contamination severely disrupts the microbial communities. Cadmium, lead, arsenic, and mercury exert toxic effects by damaging cell membranes, interfering with enzyme functions, and generating oxidative stress [69]. This leads to a decline in bacterial population density and enzymatic activity, which, in turn, reduces nutrient availability to plants. Consequently, plants grown in contaminated soils often show stunted growth, chlorosis, and reduced yields [70]. Moreover, the loss of rhizobacterial functionality weakens the soil structure and resilience, further compounding the ecological impact. Restoring these microbial populations through organic amendments or microbial inoculants is crucial for sustaining soil health and ensuring efficient nutrient cycling in agroecosystems affected by heavy-metal pollution [71,72].

2.2. Mycorrhizal Fungi

Mycorrhizal fungi, particularly arbuscular mycorrhizal fungi (AMF), such as Glomus spp., form symbiotic associations with plant roots, enhance nutrient uptake, especially phosphorus, and improve plant resistance to biotic and abiotic stressors [73]. These fungi extend their hyphal networks into the soil, thereby increasing the root surface area and facilitating nutrient and water absorption. However, heavy metals such as cadmium and arsenic adversely affect mycorrhizal fungi by inhibiting spore germination, disrupting hyphal elongation, and impairing the colonization of plant roots [74]. This diminishes the ability of fungi to protect plants against droughts, pathogens, and other environmental stressors. Additionally, heavy metal stress can alter the community structure and biodiversity of mycorrhizal fungi, weaken their ecological functions, and compromise their mutualistic relationship with host plants [75]. The resulting nutrient deficiency and stress sensitivity in plants negatively affect growth and productivity. To mitigate these effects, practices such as applying mycorrhizal bioinoculants and organic soil amendments are recommended to promote mycorrhizal recovery and support plant–microbe symbiosis in contaminated environments [76].

3. Impact of Metal on Soil and Plant Health

3.1. Antioxidant Enzymes in Microbial Ros Detoxification

Microbes synthesize antioxidant enzymes to counteract oxidative stress, which neutralizes ROS and restores cellular homeostasis [77]. The three most important enzymes involved in ROS detoxification are superoxide dismutase (SOD), catalase (CAT), and peroxidases [78]. SOD is a protective mechanism against ROS, catalyzing the dismutation of superoxide radicals into hydrogen peroxide (H2O2) and molecular oxygen [79]. Several forms of SOD in bacteria increase SOD expression in metal-contaminated environments to avoid oxidative damage [80]. Catalase, in addition to SOD, destroys hydrogen peroxide, a stable ROS, and generates hydroxyl radicals in the presence of transition metals such as iron [81]. Most metal-resistant bacteria exhibit high catalase activity to prevent oxidative damage. Peroxidases, including glutathione peroxidase and alkyl hydroperoxide reductase, also break down hydrogen peroxide and organic peroxides using reducing agents such as glutathione or NADPH [82]. Peroxidases are essential in metal-tolerant bacteria to avoid lipid peroxidation, which is the primary effect of oxidative stress that causes membrane damage [83].

3.2. Metal-Reducing Enzymes: Detoxification Through Biotransformation

Metal-reducing enzymes in microbes, such as arsenate reductase (ArsC) in Pseudomonas and other metal-resistant bacteria, help transform toxic metal ions into less toxic or less bioavailable forms [84]. Arsenic contamination is a serious environmental problem, particularly in groundwater and agricultural soil. Pseudomonas, along with ArsC, can thrive in arsenic-polluted environments and has bioremediation applications. Other metal-reducing enzymes include Mercury Reductase (MerA), Chromate Reductase, and Ferric Reductase, which oxidize toxic metals into less toxic or bioavailable forms [85]. Heavy metal pollution, including Cd, Cu, Pb, Hg, and Zn, is toxic to bacteria, even at low doses, interfering with cell functions and forming reactive oxygen species. Metal efflux pumps are membrane-bound proteins that are transported across the bacterial cell membrane and play a critical role in bacterial survival in metal-contaminated niches and bioremediation [86]. Efflux pumps enable bacteria to endure and remove heavy metals from the ecosystem, thereby maintaining metal homeostasis and minimizing intracellular toxicity [87].

3.3. Relationships of Symbiotic Alteration

Mycorrhizal fungi, such as Glomus spp., form symbiotic associations with plant roots to enhance water and nutrient acquisition. However, heavy metals, such as Cd, Pb, and Zn, disrupt the germination of fungal spores, elongation of hyphae, and root colonization. This leads to reduced mycorrhizal benefits and phosphorus availability, nutrient starvation, reduced growth, and increased exposure to environmental stresses [88]. Legume–rhizobial symbiosis, an important process in biological nitrogen fixation, is negatively affected by heavy metals. Rhizobia develop nodules on leguminous plant roots, which fix atmospheric nitrogen to ammonia, which can be absorbed by the plant, suppress rhizobial adherence to host roots, decrease nodule formation, and disrupt nitrogenase enzyme action, leaving less nitrogen available to support plant development [89]. In addition, heavy metals restructure the synthesis of chemical signaling molecules, including flavonoids and lipochitooligosaccharides, which allows communication between plants and symbiotic microorganisms [90]. This modification compromises the mutualistic relationships required by plants to achieve improved nutrient uptake and stress resistance. To counteract these effects, interventions such as soil amendments, bioremediation using metal-tolerant microbial strains, and phytoremediation with metal-accumulating plants can assist in the recovery of microbial symbiosis and enhance plant resistance in polluted environments [91].

3.4. Increased Pathogen Susceptibility

HM contamination in the soil disturbs microbial equilibrium, weakening beneficial microbes that are pivotal to plant defense and disease susceptibility. Beneficial bacteria and fungi suppress plant diseases by competing for nutrients, synthesizing antimicrobial substances, and boosting the plant immune system [92]. This imbalance makes plants more vulnerable to disease, resulting in tremendous agricultural losses. One of the major effects of HM stress is the reduction of PGPR, such as Pseudomonas and Bacillus, which are responsible for the production of secondary metabolites that inhibit the growth of soil-borne pathogens [93]. Heavy metals disrupt bacterial metabolism and decrease their capacity to produce protective agents; hence, pathogens such as Fusarium and Pythium have a competitive edge [94]. Mycorrhizal fungi, which create symbiotic relationships with plant roots and act as protectors against soil pathogens, are negatively affected by heavy metals. The collective effects of heavy metal stress on beneficial microorganisms and plant immunity lead to increased disease prevalence and reduced agricultural yields [95]. To reduce the impact of heavy-metal-induced pathogen susceptibility, bioremediation processes involving metal-tolerant microbial strains, organic amendments, and phytoremediation with metal-hyperaccumulating plants can be used to reverse the loss of microbial diversity and increase plant resistance.

4. Metal Efflux Pumps in Bacteria Mechanisms

Microorganisms employ diverse and specialized metal efflux systems to maintain metal homeostasis and prevent toxicity. Among them, P-type ATPases like CopA actively export toxic metal ions such as copper (Cu2+) from the cytoplasm using ATP hydrolysis, thus protecting bacterial enzymes and DNA from metal-induced damage [96]. Resistance-Nodulation-Division (RND) transporters, common in Gram-negative bacteria, utilize proton motive force to export heavy metals and antibiotics, exemplified by the CzcABC system in Cupriavidus metallidurans, which facilitates the efflux of cobalt, zinc, and cadmium [97]. Cation Diffusion Facilitators (CDFs), such as ZntA in Escherichia coli, operate via passive transport or ion exchange to export divalent metals like Zn2+ and Cd2+, preventing metabolic disruption [98] (Figure 1). ATP-Binding Cassette (ABC) transporters, including the ArsAB system, use ATP hydrolysis to actively expel metals like arsenite (As3+) and are involved in both import and efflux, impacting stress response, antibiotic resistance, and virulence [99] (Figure 2). These systems are vital for microbial survival in metal-rich environments and hold therapeutic and agricultural relevance.

5. Positive Effects: Role of Metal-Tolerant Microbes

5.1. Metal Immobilization and Detoxification

Metal-resistant fungi and bacteria have developed mechanisms to immobilize and clean toxic metals via biosorption, biotransformation, and biomineralization [100]. Such strategies reduce the access of metals to plants for root uptake and limit toxicity. Bioremediation is a non-polluting green technology, in which microbial processes are employed to detoxify HMs [101]. Biosorption immobilizes heavy metals by the biotransformation of cell wall components biotransformation converting metals into less toxic or insoluble states, and biomineralization precipitates metals into stable mineral phases (Table 1).

5.2. Metal Stress Enhancement of Plant Growth

Plant Growth-Promoting Rhizobacteria (PGPR) are very important in plant adaptation to metal stress through physio-biochemical mechanisms to grow resistance [102]. These environmentally beneficial microbes neutralize metal toxicity by producing stress-relieving enzymes, enhancing nutrient availability, and activating plant defense systems, ultimately optimizing the health and productivity of plants under contaminated conditions [103,104]. Another important method for facilitating plant growth under metal stress is the production of 1-aminocyclopropane-1-carboxylate (ACC) deaminase [105]. The enzyme breaks down ACC into α-ketobutyrate and ammonia, which decreases the ethylene content in plants, promoting root elongation, nutrient uptake, and overall plant growth [106]. PGPR also increase metal stress tolerance in plants by triggering system resistance to oxidative stress. Exposure to heavy metals, such as cadmium, lead, and arsenic, produces reactive oxygen species (ROS), leading to lipid peroxidation, DNA damage, and protein oxidation (Table 1).
Table 1. Some examples of PGPR with demonstrated efficacy in mitigating heavy metal stress.
Table 1. Some examples of PGPR with demonstrated efficacy in mitigating heavy metal stress.
PGPR StrainTarget Heavy Metal (s)Host PlantMechanism of ActionReference
Pseudomonas fluorescensZinc (Zn), Cadmium (Cd)Brassica junceaACC deaminase activity, Zn/Cd biosorption, antioxidant enzyme induction[107]
Bacillus subtilisLead (Pb)Zea maysCell wall binding, phytohormone production, Pb detoxification[108]
Enterobacter cloacaeArsenic (As)Oryza sativa (rice)Arsenite oxidation, metal exclusion, enhanced root architecture[109]
Azospirillum brasilenseNickel (Ni)Triticum aestivumIndole-3-acetic acid (IAA) production, Ni stress alleviation[110]
Rhizobium leguminosarumChromium (Cr)Lens culinarisSiderophore production, improved nutrient uptake, Cr sequestration[111]
Serratia marcescensCopper (Cu)Solanum lycopersicumBiofilm formation, Cu efflux, antioxidant enzyme upregulation[112]

5.3. Reactive Oxygen Species (Ros) Generation

Management of bacterial ROS under metal stress is crucial for their survival, affecting microbial diversity and ecological processes in contaminated environments. One of the major sources of ROS in bacteria exposed to heavy metal stress is the Fenton reaction between Fe2+ and hydrogen peroxide, which forms hydroxyl radicals (•OH) [113]. Such harmful ROS may interact with cellular macromolecules and induce mutations, enzyme inactivation, and membrane disruptions. Redox cycling metals, such as iron and copper, also help produce ROS by being involved in electron transfer reactions, shuttling between their oxidized and reduced forms, and unintentionally causing persistent generation of ROS [114]. In the absence of functional SOD and catalase, bacterial cells accumulate toxic ROS, resulting in damage to the cell membrane, protein misfolding, and apoptosis-like cell death. Heavy metals also interfere with the bacterial electron transport chain (ETC) of cellular respiration, enhancing ROS leakage [115]. Arsenic and chromium interfere with electron transfer in the ETC, resulting in electron leakage and suboptimal oxygen reduction, with the formation of excessive superoxide and hydrogen peroxide [116]. Hydrogen-mediated ROS formation has significant ecological impacts, modifying the microbial community composition and function. Metal-resistant organisms, such as Pseudomonas and Bacillus strains, develop adaptive processes to endure oxidative stress and strengthen antioxidant systems, metal efflux pumps, and biofilm formation capabilities [117]. Knowledge of these processes is essential for creating bioremediation applications that utilize metal-resistant bacteria to recover from contaminated environments and avoid heavy metal toxicity.

5.4. Plant-Related Oxidative Stress Impacts

Oxidative stress in plants occurs when reactive oxygen species (ROS), such as superoxide (O2), hydrogen peroxide (H2O2), and hydroxyl radicals (•OH), accumulate beyond the detoxifying ability of plant defense mechanisms. This accumulation is generally due to environmental stresses caused by heavy metal toxicity, salinity, drought, and pathogen infection [118]. The effects of ROS overproduction include lipid peroxidation, which disrupts cellular membrane integrity, DNA, and protein oxidation, which disrupts genetic stability and enzyme activity and inhibits photosynthesis and cellular signaling [119]. To offset these effects, plants trigger a set of enzymatic antioxidants such as superoxide dismutase (SOD), catalase (CAT), and ascorbate peroxidase (APX), which act in combination to neutralize ROS [120]. Plant Growth-Promoting Rhizobacteria (PGPR) further develop plant resistance through the synthesis of antioxidant enzymes, nutrient mobilization, and root growth promotion [121,122]. PGPR also aid metal detoxification by biosorption and ACC deaminase activity, thus reducing oxidative stress and enhancing plant health under stressful conditions.

5.5. Bioremediation Applications and Link to AMR

Bioremediation uses metal-tolerant microorganisms to remove heavy metals by detoxification via processes such as biosorption, biotransformation, and biomineralization, which enhance the quality of soil and facilitate plant growth in polluted environments. Microbial activities not only reduce metal toxicity but also enhance sustainable ecosystem recovery [123]. However, exposure to pollutants causes oxidative stress in microbes, resulting in the buildup of reactive oxygen species (ROS) that damage DNA, initiate genetic mutations, and induce biofilm growth as a defense mechanism [124]. Oxidative stress also initiates stress response regulons, which increase microbial survival in unfavorable environments. Such adverse environments are highly associated with the development and dissemination of AMR [125]. Chronic oxidative stress enhances horizontal gene transfer, whereas selective pressure from metal exposure enhances the survival of resistant populations [126]. Heavy metal pollution also tends to co-select antibiotic resistance genes in addition to the AMR threat. Thus, microbial bioremediation not only rehabilitates environmental well-being, but is also critical in truncating the spread of AMR, providing both ecological and public health advantages [127].

5.6. Oxidative Damage to Cellular Components

Oxidative stress compromises bacterial viability and modulates microbial communities in metal-polluted habitats [128]. DNA damage is a serious outcome, because ROS target nucleotides, resulting in base alterations, strand breaks, and mutations. Hydroxyl radicals form double-strand breaks (DSBs), which are extremely toxic if not repaired accurately. Additionally, oxidative stress enhances the potential for mutagenesis, possibly resulting in antibiotic-resistant mutations [129]. Bacteria subjected to heavy metal stress can acquire adaptive mutations in DNA repair genes, efflux pump regulators, or ribosomal proteins, which increases their survival in metal-contaminated and antibiotic-exposed environments [130]. Such interactions between oxidative stress and genetic adaptation are of concern regarding the evolution of multidrug-resistant bacterial strains in contaminated environments [131]. Certain metal-resistant bacteria have developed adaptive strategies to survive in polluted soils and water bodies, including increased production of antioxidant enzymes, biofilm formation, and activation of DNA repair processes [132]. Uncovering the molecular effects of oxidative damage is important for designing efficient bioremediation approaches and alleviating the ecological hazards caused by heavy metal contamination.

6. Link Between Oxidative Stress and Antimicrobial Resistance

Oxidative stress plays a critical role in the rapid development of AMR. Reactive oxygen species, such as superoxide, hydroxyl radicals, and hydrogen peroxide, damage bacterial DNA, proteins, and membranes, thereby inducing genetic instability [133]. Damage enhances the mutation frequency and facilitates HGT, resulting in the development of multidrug-resistant (MDR) bacterial strains. DNA damage may cause single-strand breaks, double-strand breaks, and base modifications, especially if they take place in antibiotic target site-containing genes, efflux pump control genes, and DNA repair pathway genes [134]. Alterations in these genes may lead to resistance to fluoroquinolone antibiotics and aminoglycosides such as streptomycin and kanamycin [135]. The SOS response is a bacterial global regulatory network that responds to DNA damage, especially under stress conditions such as oxidative exposure [136]. This response allows cells to trigger DNA repair processes in addition to ensuring survival under genotoxic stress. RecA, the central regulator of the SOS response, is a highly significant DNA-binding protein [137]. When detecting single-stranded DNA resulting from damage, RecA is activated and promotes autocleavage of LexA, a repressor of transcription that normally suppresses SOS genes [138,139]. Cleavage derepresses DNA repair, cell cycle control, and error-prone replication genes. Therefore, mutagenic DNA polymerases are produced, elevating the mutation frequency and allowing rapid adaptation of bacterial populations to adverse environments [140]. In addition to beneficial mutations, lethal mutations confer adaptive advantages such as resistance to several antibiotics. Hybrid gene transfer is another major determinant of AMR evolution [141]. Metal-resistant bacteria have metal and antibiotic resistance genes co-located in the same MGEs, resulting in co-selection for AMR [142]. Oxidative stress also affects efflux pump expression and drug resistance. Oxidative stress promotes the development of multidrug-resistant bacteria by increasing the rate of mutation, SOS induction, horizontal gene transfer, and efflux pump regulation [143].

6.1. Fluoroquinolone Resistance Results from Mutations in DNA Gyrase (gyrA and gyrB) Caused by ROS Damage

Reactive oxygen species play a role in bacterial mutagenesis by inducing mutations in the DNA gyrase genes gyrA and gyrB which are critical for DNA replication and supercoiling [144]. Oxidative damage results in fluoroquinolone resistance, which is a considerable problem in both clinical and environmental settings. Fluoroquinolone antibiotics such as ciprofloxacin and levofloxacin inhibit the enzymatic activity of DNA gyrase, resulting in double-strand breaks and bacterial cell death [145]. However, oxidative stress due to heavy metals or antimicrobial compounds can trigger mutations in gyrA and gyrB changing the conformation of the DNA gyrase enzyme and decreasing the fluoroquinolone-binding affinity [146,147]. ROS-mediated DNA damage triggers the SOS response, a RecA- and LexA-regulated bacterial stress response [148]. The error-prone DNA polymerases Pol IV and Pol V are induced by inserting random mutations into the bacterial genome, including gyrA and gyrB [149]. This accelerates bacterial adaptation and favors the evolution of fluoroquinolone resistance. The heavy metals Cd, Pb, and As induce ROS production, enhancing mutation rates in bacteria [150]. Such bacteria co-evolve resistance mechanisms and tend to develop cross-resistance to both fluoroquinolones and metals owing to genetic mutations in DNA gyrase [151]. It is important to understand these mechanisms to develop a new therapeutic approach to combat antibiotic resistance and restrict the transfer of drug-resistant bacteria into contaminated ecosystems [152].

6.2. Mutations in Ribosomal Proteins (rpsL and rrl) Confer Aminoglycoside Resistance

Aminoglycoside antibiotics such as streptomycin, gentamicin, kanamycin, and amikacin bind to bacterial ribosomes, inhibit protein synthesis, and cause cell death [153]. However, mutations in ribosomal protein genes (rpsL and rrl) modify the structure of ribosomes, decreasing the binding affinity of aminoglycosides and giving rise to resistance [154]. Oxidative stress increases these mutations, which, in turn, drives the emergence of aminoglycoside-resistant bacterial strains. Aminoglycosides interact with the 30S ribosomal subunit, where they bind to 16S rRNA and ribosomal protein S12, inducing misreading of mRNA and inclusion of incorrect amino acids, resulting in defective protein synthesis and cell death [155,156]. Such rpsL resistance mutations are K43R and K88R in Escherichia coli and Mycobacterium tuberculosis, respectively, which compromise the ability of the antibiotic to cause translational errors and allow bacteria to survive despite treatment [157]. Universal resistance mutations in rrl involve alterations in nucleotide interactions and reduced aminoglycoside affinities [158]. General resistance mutations in rrl are A1408G and C1409U in Pseudomonas aeruginosa and Klebsiella pneumoniae, respectively, which disrupt hydrogen bonding with aminoglycosides, thereby reducing their effectiveness [159]. HMs and antibiotic-induced oxidative stress cause ROS-mediated DNA damage, activation of the SOS response, and selective pressure, favoring the survival of aminoglycoside-resistant mutants [160,161]. Aminoglycoside-resistant microorganisms pose a significant challenge for the treatment of E. coli, Mycobacterium tuberculosis, and Staphylococcus aureus infections [162]. The horizontal transfer of genes in resistant bacteria occurs at a high rate, increasing the possibility of drug resistance in hospitals and environmental settings. Identification of such mechanisms is critical for designing alternative therapies and mitigating the development of drug-resistant bacterial infections.

6.3. Metal Stress-Induced SOS Response Increasing Resistance Mutations

Heavy metal (HM) stress can cause direct damage to bacterial DNA, thus inducing the SOS response, a RecA- and LexA-regulated conserved bacterial survival strategy [163]. Under unstressed, normal conditions, LexA is a transcriptional repressor that maintains the repressed status of DNA repair genes [164]. However, when HMs create DNA lesions, RecA, which detects single-stranded DNA, is activated and induces autocleavage of LexA [165]. This depression in SOS genes triggers a DNA repair response. Although the SOS response is crucial for the survival of bacteria under genotoxic stress, it occurs at the expense of genomic stability [166]. These mutations may result in clinically relevant resistance, such as changes in DNA gyrase genes, causing resistance to fluoroquinolones, or altering target sites for aminoglycosides. Furthermore, SOS induction increases the mobilization of mobile genetic elements, promoting increased horizontal transfer of resistance genes. As a result, metal stress not only causes mutations in individual cells, but also selects and enables the transmission of multidrug-resistant (MDR) bacterial strains in the environment [167]. It is important to identify this causal mechanism to design novel therapeutic interventions and policies to reduce the co-selection pressure exerted by environmental metal pollution on antimicrobial resistance.

6.4. Horizontal Gene Transfer (HGT) and Co-Selection of AMR Genes

HGT is a system by which bacteria share genetic information regardless of cell division, enabling the rapid establishment of new traits, including AMR. HGT occurs through three principal mechanisms: transformation (free DNA uptake), transduction (phage-mediated gene transfer), and conjugation (plasmid transfer among cells). Stress factors such as heavy metals, antibiotics, and oxidative stress may boost HGT through enhanced genetic mobility and stress-induced mutagenesis [168]. This provides a context in which the resistance genes, although not related to immediate selective pressure, are maintained because of co-selection. Co-selection occurs when antibiotic and heavy metal resistance genes exist on the same mobile genetic elements as plasmids, integrons, or transposons and are exchanged during HGT. Exposure to a single selective agent such as heavy metals can induce persistence and transmission of antibiotic resistance genes, exacerbating the AMR crisis in both clinical and environmental contexts [169].
HM and antibiotics exert selective pressures on bacterial populations that lead to the co-selection of resistance genes as well as their accumulation (Table 1). Bacteria subjected to HM have evolved means of resistance, which often overlaps with those used for antibiotic resistance. This cross-protection is primarily brought about by efflux pumps, biofilms, enzymatic detoxification, and HGT, which enable bacteria to tolerate metal toxicity and antibiotic treatment, including CzcABC, P-type ATPases, and AcrAB-TolC, which actively export toxic metal ions and antibiotics [170]. Upregulation of biofilm production in response to HM stress indirectly enhances antibiotic resistance. Enzymatic detoxification pathways, such as arsenate reductase (arsC), catalase (CAT), and superoxide dismutase (SOD), are essential for preventing metal- and antibiotic-induced oxidative stress [171]. Heavy metals generate ROS, which oxidize bacterial lipids, proteins, and DNA, thereby triggering bacteria to activate detoxification enzymes. Hybrid gene transfer contributes to increased co-selection of metal and antibiotic resistance genes. Resistance genes are most often found in plasmids, transposons, and integrons and may be transferred between bacteria via conjugation, transformation, and transduction [172]. HM in the environment selects heavy-metal-resistant bacteria that may or may not carry resistance genes on plasmids, promoting the dissemination of antibiotic resistance, even in the absence of antibiotics (Table 2).

6.5. Plasmid-Mediated Resistance Metal Stress

Plasmid-mediated resistance is facilitated by heavy metal stress, particularly with respect to the increased transfer of resistance plasmids carrying antibiotic resistance genes (ARGs) [181]. Under metal-polluted conditions, bacteria are subjected to selective pressure, which selects for the survival and growth of strains possessing mobile genetic elements (MGEs), such as plasmids, transposons, and integrons. This facilitates rapid adaptation and dissemination of resistance features by HGT, mostly through conjugation. Under heavy metal stress, bacteria induce stress response mechanisms such as the SOS response, which increases the expression of conjugative transfer genes [182]. Metal ions cause DNA damage and oxidative stress, leading to RecA-dependent induction of the SOS system. This enhances the mobility of R-plasmids and facilitates the exchange of resistance genes among bacterial populations [183]. Metal exposure upregulates type IV secretion systems (T4SS) necessary for plasmid transfer and promotes the propagation of AMR [184]. Numerous R-plasmids have both heavy metal resistance (HMR) and antibiotic resistance determinants; as a result, co-selection ensues where metal stress inadvertently results in antibiotic-resistant bacteria, even without exposure to antibiotics [185].

7. Biofilm Formation and Antibiotic Tolerance

Bacterial survival and resilience are facilitated by reactive oxygen species (ROS) and cellular ROS, which are produced as metabolic end products or triggered by extracellular stress agents, such as heavy metals, antibiotics, and host immunity, and affect the survival and tolerance of bacteria to extreme environmental conditions [186]. In accordance with oxidative stress, biofilm-associated regulatory mechanisms in bacteria induce biofilm formation as a mode of protective survival [187]. One major mechanism is the stimulation of quorum sensing (QS), a cell-to-cell communication system that controls gene expression as a function of cell population density [188]. Oxidative stress can activate QS gene expression by modifying redox-sensitive regulatory networks, thus increasing the synthesis of signaling molecules such as N-acyl homoserine lactones (AHLs) in gram-negative bacteria or autoinducing peptides in gram-positive bacteria [189]. These signaling molecules, in turn, orchestrate the collective behavior required for strong biofilm establishment and stress tolerance [190]. One key participant in this stress-adaptive mechanism is the rpoS gene, which produces the alternative sigma factor σ^S, a general stress response master regulator in most gram-negative bacteria such as Escherichia coli and Pseudomonas aeruginosa [191]. During oxidative stress, higher ROS levels activate rpoS transcription and stabilization, which in turn activates a set of genes for antioxidant defense, DNA repair, and notably, biofilm formation [192]. rpoS-regulated responses favor the production of extracellular polymeric compounds (EPS) and increase the expression of QS-regulated biofilm genes [193], thus connecting oxidative stress perception to quorum-sensing-mediated communal behavior [194]. In the biofilm mode, the bacteria are ensconced in a protective matrix of EPS that is replete with antioxidant enzymes such as catalase and superoxide dismutase, which neutralize ROS and protect cells against antibiotics and host immune responses. The role of rpoS here is emphasized by the evidence that rpoS-defective mutants produce substantially reduced numbers of biofilms, with a crucial role in oxidative stress-regulated biofilm control [195]. Knowledge of the molecular crosstalk between oxidative stress, quorum sensing, and rpoS-controlled biofilm formation provides exciting prospects for anti-biofilm therapy development [196]. These involve quorum-sensing inhibitors, agents modulating ROS, and rpoS-targeting approaches to counteract multidrug-resistant (MDR) bacterial infections and environmental biofilm-related challenges [197].

Metals and Antibiotics Are Biofilms That Increase Multidrug Tolerance

Biofilms, which are colonies of microorganisms, are protected from environmental stressors such as heavy metals and antibiotics by an EPS matrix. Biofilms provide bacteria with the means to survive hostile conditions, which results in chronic infections and multidrug tolerance (MDT) [198]. The EPS matrix, consisting of proteins, lipids, polysaccharides, and extracellular DNA, creates a barrier that prevents the entry of metals and antibiotics and lowers the effective concentration of toxic chemicals [199]. Some bacteria produce negatively charged compounds that chelate metal ions, thereby lowering toxicity and bioavailability. Antibiotics experience limited diffusion and decreased efficacy in biofilms, allowing bacteria to trigger stress responses before reaching lethal concentrations [200]. β-lactamases and aminoglycoside-modifying enzymes are typically integrated into the matrix, degrading antibiotics before they reach bacterial cells. Biofilms create heterogeneous microenvironments, resulting in various metabolic states. Latent or slow-growing bacteria in the lower layers become tolerant to antibiotics through loss of sensitivity, and consequently, reiterative attacks produce chronic infections. Multidrug-resistant infections caused by biofilms are major contributors to hospital and environmental infections [201]. Determining how biofilms protect bacteria against metals and antibiotics is important for the development of new antibiofilm approaches. Exposure to high Cu concentrations triggers oxidative and envelope stress responses that stimulate biofilm production, a primary survival trait that increases bacterial resistance to environmental stress and antibiotics [202]. This stress also causes overproduction of efflux pumps, such as the MexAB-OprM system, which contributes to intrinsic and acquired resistance to several antibiotics, including carbapenems, a vital class of β-lactam antibiotics for the treatment of multidrug-resistant infections [203]. The MexAB-OprM efflux pump expels toxic metals and antibiotics that directly connect metal stress to antimicrobial resistance in a direct manner [204]. Moreover, exposure to copper has been found to induce the expression of quorum sensing and rpoS-associated genes, further contributing to biofilm maturation and persistence [205]. These are examples of how exposure to heavy metals indirectly promotes the evolution and persistence of AMR.

8. Environmental and Clinical Implications

Heavy metal contamination is a major factor in the development of antibiotic-resistant bacteria, making AMR a global health concern. Pb, Cd, Cu, and Ag are among the heavy metals released into the environment from mining, agricultural, and industrial effluents [206]. In addition, to directly poison microbes, these metals also function as selection pressures to promote the growth and survival of antibiotic-resistant bacteria. As most bacterial metal resistance mechanisms, such as efflux pumps, biofilm formation, and oxidative stress responses, are similar to those of antibiotic resistance, this choice was selected [207]. AMR hotspots have emerged in heavy-metal-contaminated environments, and resistant bacteria are likely to spread in clinical settings. Long-term pollution is caused by the discharge of hazardous metal ions into soil and water bodies through mining operations, metal processing businesses, and factory effluents. Selective pressure that supports resistant microbial communities is introduced by agricultural runoff from pesticides, fertilizers, and animal manure that contains trace metals [208]. Once bacteria acquire metal resistance, they often develop cross-resistance to antibiotics, owing to the shared resistance mechanisms of metal efflux pumps and stress response pathways. A specific source of AMR hotspots is hospital wastewater, which contains a cocktail of antibiotics, disinfectants, heavy metals from medical devices, plumbing agents, and sterilizing agents. Combined exposure to these stressors has a synergistic effect that promotes co-selection, favoring bacterial populations resistant to both metals and antibiotics [209]. In contrast to antibiotics, the long-term environmental persistence of HMs places a continuous selective pressure on microbial populations, enabling resistant bacteria to propagate and spread within polluted ecosystems, thus increasing the risk of human transmission via food, water, and direct exposure to the environment. Antimicrobials based on heavy metals, including silver nanoparticles and copper antimicrobials, have been used in medical devices, wound dressings, disinfectants, and agricultural products because of their high-spectrum antimicrobial effects and correlation metal and antibiotic resistance in clinical and environmental implications (Table 3).

9. Bioremediation Using Metal-Resistant Bacteria

Bioremediation is a suggested approach to mitigate environmental metal pollution. This involves the use of metal-resistant fungi and bacteria to adsorb, convert, or precipitate poisonous metals, thereby decreasing their bioavailability [225]. Microorganisms such as Pseudomonas putida, Bacillus subtilis, and Rhodococcus spp. immobilize heavy metals via biosorption and bioaccumulation, producing exopolysaccharides and metallothioneins are produced [226]. This can reverse the ecological balance and minimize selective pressure on AMR bacteria. Control of HMP from industry and agriculture is necessary to curb pollution and affect bacterial resistance. The implementation of wastewater treatment procedures and sustainable farming practices can mitigate the accumulation of metals in agricultural fields and products [227]. The combination of drugs acting against efflux pumps and antibiotic resistance mechanisms is promising for the reversal of cross-resistance. Scientists are currently investigating efflux pump inhibitors (EPIs) to inhibit metal and antibiotic efflux in bacteria, thereby rendering pathogens more accessible for treatment. Combination treatments target bacterial vulnerabilities when stressed with metals. For example, antibiotics combined with metal chelators act synergistically [228]. EDTA and chelators have been established to increase β-lactam antibiotic activity against resistant bacteria through destabilization of biofilms and inhibition of metalloenzymes. Table 4 and Table 5 provide a comprehensive summary of the bioremediation of heavy metals, highlighting metal-accumulating bacteria in soil and their remediation potential as well as the bacterial species in water and the corresponding heavy metal concentrations that they mitigate.

10. Conclusions

Heavy metal contamination markedly promotes AMR by imposing selective pressure on microbial populations in environmental niches ranging from the rhizosphere to the phyllosphere. Cadmium, lead, mercury, and arsenic cause oxidative stress, DNA damage, and mutations that increase antibiotic resistance through mechanisms such as overexpression of efflux pumps, detoxification enzyme expression, and biofilm formation. These processes, combined with horizontal gene transfer and the co-selection of resistance genes, accelerate the development of multidrug-resistant strains, which pose an extreme danger to global health. The rhizosphere, which is essential for plant–microbe interactions, is affected by heavy metal toxicity, which hampers beneficial microbes, such as rhizobacteria and mycorrhizal fungi, affecting nutrient cycling and plant health. Nevertheless, metal-tolerant microorganisms provide promising solutions through bioremediation strategies involving biosorption, biotransformation, and biomineralization, which counteract metal toxicity and promote the recovery of ecosystems. Advances in synthetic biology and gene editing continue to boost the chances of engineered microbes tackling metal-induced AMR. Notwithstanding, challenges exist, such as resource inequalities in low- and middle-income nations and disparities in global surveillance systems. Mitigating the intersection of heavy metal pollution and AMR necessitates interconnected strategies, including the amalgamation of bioremediation, policy measures, and optimal resource utilization. By understanding and addressing the molecular and ecological processes connecting heavy metals and AMR, we can establish sustainable approaches to stem resistance, ensure environmental health, and preserve public health for generations.

Author Contributions

Conceptualization, D.M., P.G. and R.K.; software, R.K.; investigation, R.K., P.G. and D.M.; resources, D.M. and P.G.; data curation, R.K., T.P.V., S.P.Ž., A.N.O., A.P. and D.M.; writing—original draft preparation, R.K., T.P.V., S.P.Ž., P.P., G.S., S.d.l.S.V., L.K., A.N.O., P.G. and D.M.; writing—review and editing, R.K., T.P.V., S.P.Ž., P.P., G.S., S.d.l.S.V., L.K., A.N.O., D.M. and P.G.; visualization, R.K., D.M. and P.G.; supervision, P.G.; project administration, P.G.; funding acquisition, D.M. 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.

Acknowledgments

All authors are thankful to Graphic Era (Deemed to be University), India; University of Niš, Serbia; ICAR-Central Rice Research Institute, India; Universidad Michoacana de San Nicolás de Hidalgo, Mexico; Instituto Tecnológico de Sonora, Mexico; Federal University of Agriculture Abeokuta, Nigeria; The University of Georgia, USA; Technological University Dublin, Ireland for support. R.K., D.M. and P.G. special thanks to Graphic Era (Deemed to be University), Dehradun, India, for their support. We would like to thanks Paperpal (https://paperpal.com/), Cactus Communications Services Pte Ltd., Singapore for their invaluable online assistance in refining the language and grammar of our manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Puri, B.; Vaishya, R.; Vaish, A. Antimicrobial resistance: Current challenges and future directions. Med. J. Armed Forces India 2025, 81, 247–258. [Google Scholar] [CrossRef]
  2. Saeed, U.; Insaf, R.A.; Piracha, Z.Z.; Tariq, M.N.; Sohail, A.; Abbasi, U.A.; Fazal, I. Crisis averted: A world united against the menace of multiple drug-resistant superbugs-pioneering anti-AMR vaccines, RNA interference, nanomedicine, CRISPR-based antimicrobials, bacteriophage therapies, and clinical artificial intelligence strategies to safeguard global antimicrobial arsenal. Front. Microbiol. 2023, 14, 1270018. [Google Scholar]
  3. Ferraz, M.P. Antimicrobial resistance: The impact from and on society according to one health approach. Societies 2024, 14, 187. [Google Scholar] [CrossRef]
  4. Kim, C.; Holm, M.; Frost, I.; Hasso-Agopsowicz, M.; Abbas, K. Global and regional burden of attributable and associated bacterial antimicrobial resistance avertable by vaccination: Modelling study. BMJ Glob. Health 2023, 8, e011341. [Google Scholar] [CrossRef]
  5. Javadi, H.; Lehnen, A.C.; Hartlieb, M. Bioinspired Cationic Antimicrobial Polymers. Angew. Chem. Int. Ed. 2025, 64, e202503738. [Google Scholar] [CrossRef] [PubMed]
  6. Samtiya, M.; Matthews, K.R.; Dhewa, T.; Puniya, A.K. Antimicrobial resistance in the food chain: Trends, mechanisms, pathways, and possible regulation strategies. Foods 2022, 11, 2966. [Google Scholar] [CrossRef] [PubMed]
  7. Li, L.; Liu, X.; Fang, Y.; Guo, K.; Li, L.; Cai, S.; Hu, B. Global patterns of change in the burden of malnutrition in older adults from 1990 to 2021 and the forecast for the next 25 years. Front. Nutr. 2025, 12, 1562536. [Google Scholar] [CrossRef] [PubMed]
  8. Irfan, M.; Almotiri, A.; AlZeyadi, Z.A. Antimicrobial resistance and its drivers—A review. Antibiotics 2022, 11, 1362. [Google Scholar] [CrossRef]
  9. Kumar, R.; Kamboj, N.; Kumar, V.; Kumar, S.; Kumar, N.; Gautam, P. Assessing the Contamination Risk of Salmonella enterica subsp. in Green Leafy Vegetables. J. Pure Appl. Microbiol. 2025, 19, 459. [Google Scholar] [CrossRef]
  10. Kumar, R.; Adeyemi, N.O.; Chattaraj, S.; Alloun, W.; Thamarsha, A.K.A.N.W.M.R.K.; Anđelković, S.; Mitra, D.; Gautam, P. Antimicrobial resistance in Salmonella: One Health perspective on global food safety challenges. Sci. One Health 2025, 4, 100117. [Google Scholar] [CrossRef] [PubMed]
  11. Bassetti, M.; Vena, A.; Sepulcri, C.; Giacobbe, D.R.; Peghin, M. Treatment of bloodstream infections due to Gram-negative bacteria with difficult-to-treat resistance. Antibiotics 2020, 9, 632. [Google Scholar] [CrossRef] [PubMed]
  12. Salam, M.A.; Al-Amin, M.Y.; Salam, M.T.; Pawar, J.S.; Akhter, N.; Rabaan, A.A.; Alqumber, M.A. Antimicrobial resistance: A growing serious threat for global public health. Healthcare 2023, 11, 1946. [Google Scholar] [CrossRef]
  13. Coque, T.M.; Cantón, R.; Pérez-Cobas, A.E.; Fernández-de-Bobadilla, M.D.; Baquero, F. Antimicrobial resistance in the global health network: Known unknowns and challenges for efficient responses in the 21st century. Microorganisms 2023, 11, 1050. [Google Scholar] [CrossRef]
  14. Ye, Z.; Li, M.; Jing, Y.; Liu, K.; Wu, Y.; Peng, Z. What Are the Drivers Triggering Antimicrobial Resistance Emergence and Spread? Outlook from a One Health Perspective. Antibiotics 2025, 14, 543. [Google Scholar] [CrossRef] [PubMed]
  15. Rahman, Z.; Singh, V.P. The relative impact of toxic heavy metals (THMs) (arsenic (As), cadmium (Cd), chromium (Cr)(VI), mercury (Hg), and lead (Pb) on the total environment: An overview. Environ. Monit. Assess. 2019, 191, 419. [Google Scholar] [CrossRef]
  16. Kumar, R.; Gupta, G.; Hussain, A.; Rani, A.; Thapliyal, A.; Gunsola, D.; Mitra, D. Pioneering zero-waste technologies utilization and its framework on sustainable management: International, national and state level. Discov. Appl. Sci. 2025, 7, 224. [Google Scholar] [CrossRef]
  17. Gillieatt, B.F.; Coleman, N.V. Unravelling the mechanisms of antibiotic and heavy metal resistance co-selection in environmental bacteria. FEMS Microbiol. Rev. 2024, 48, fuae017. [Google Scholar] [CrossRef]
  18. Muteeb, G.; Rehman, M.T.; Shahwan, M.; Aatif, M. Origin of antibiotics and antibiotic resistance, and their impacts on drug development: A narrative review. Pharmaceuticals 2023, 16, 1615. [Google Scholar] [CrossRef]
  19. Sulis, G.; Sayood, S.; Gandra, S. Antimicrobial resistance in low-and middle-income countries: Current status and future directions. Expert Rev. Anti-Infect. Ther. 2022, 20, 147–160. [Google Scholar] [CrossRef]
  20. Kirchhelle, C. Giants on Clay Feet—COVID-19, infection control and public health laboratory networks in England, the USA and (West-) Germany (1945–2020). Soc. Hist. Med. 2022, 35, 703–748. [Google Scholar] [CrossRef] [PubMed]
  21. Dunga, K.E.; Okoro, C.I.; Onyenama, A.C.; Ekuma, U.O.; Ohanusi, I.N.; Izah, S.C. Implementing One Health approach to emerging zoonotic diseases: Bridging surveillance, sustainability and global governance. Exon 2025, 2, 200–223. [Google Scholar] [CrossRef]
  22. Wasan, H.; Singh, D.; Reeta, K.H.; Gupta, Y.K. Landscape of push funding in antibiotic research: Current status and way forward. Biology 2023, 12, 101. [Google Scholar] [CrossRef] [PubMed]
  23. Gou, F.; Liu, J.; Xiao, C.; Wu, J. Research on artificial-intelligence-assisted medicine: A survey on medical artificial intelligence. Diagnostics 2024, 14, 1472. [Google Scholar] [CrossRef]
  24. Ahmed, S.M.; Naher, N.; Tune, S.N.B.K.; Islam, B.Z. The Implementation of National Action Plan (NAP) on Antimicrobial Resistance (AMR) in Bangladesh: Challenges and lessons learned from a cross-sectional qualitative study. Antibiotics 2022, 11, 690. [Google Scholar] [CrossRef]
  25. Seixas, A.F.; Quendera, A.P.; Sousa, J.P.; Silva, A.F.; Arraiano, C.M.; Andrade, J.M. Bacterial response to oxidative stress and RNA oxidation. Front. Genet. 2022, 12, 821535. [Google Scholar] [CrossRef]
  26. Zhu, J.; Thompson, C.B. Metabolic regulation of cell growth and proliferation. Nat. Rev. Mol. Cell Biol. 2019, 20, 436–450. [Google Scholar] [CrossRef]
  27. 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]
  28. Shaw, D.R.; Dussan, J. Mathematical modelling of toxic metal uptake and efflux pump in metal-resistant bacterium Bacillus cereus isolated from heavy crude oil. Water Air Soil Pollut. 2015, 226, 112. [Google Scholar] [CrossRef]
  29. Liu, H.; Zhang, Y.; Wang, Y.; Xie, X.; Shi, Q. The connection between Czc and Cad systems involved in cadmium resistance in Pseudomonas putida. Int. J. Mol. Sci. 2021, 22, 9697. [Google Scholar] [CrossRef]
  30. Ashwath, P.; Sannejal, A.D. A quest to the therapeutic arsenal: Novel strategies to combat multidrug-resistant bacteria. Curr. Gene Ther. 2022, 22, 79–88. [Google Scholar] [CrossRef] [PubMed]
  31. Sarti, G.C.; Paz-González, A.; Cristóbal-Miguez, J.A.E.; García, A.R.; Galelli, M.E. Production of a Microbial Biofilm and Its Application on Tomato Seeds to Improve Crop Development in a Lead-Contaminated Substrate. Processes 2025, 13, 767. [Google Scholar] [CrossRef]
  32. Basu, M. Impact of mercury and its toxicity on health and environment: A general perspective. In Mercury Toxicity: Challenges and Solutions; Springer Nature: Singapore, 2023; pp. 95–139. [Google Scholar]
  33. Anand, V.; Kaur, J.; Srivastava, S.; Bist, V.; Dharmesh, V.; Kriti, K.; Srivastava, S. Potential of methyltransferase containing Pseudomonas oleovorans for abatement of arsenic toxicity in rice. Sci. Total Environ. 2023, 856, 158944. [Google Scholar] [CrossRef]
  34. Raro, O.H.F.; Poirel, L.; Nordmann, P. Effect of zinc oxide and copper sulfate on antibiotic resistance plasmid transfer in Escherichia coli. Microorganisms 2023, 11, 2880. [Google Scholar] [CrossRef]
  35. Dinesh, R.; Sreena, C.P.; Sheeja, T.E.; Srinivasan, V.; Subila, K.P.; Sona, C.; Sajith, V. Co-resistance is the dominant mechanism of co-selection and dissemination of antibiotic resistome in nano zinc oxide polluted soil. J. Hazard. Mater. 2025, 485, 136885. [Google Scholar] [CrossRef]
  36. Murray, L.M.; Hayes, A.; Snape, J.; Kasprzyk-Hordern, B.; Gaze, W.H.; Murray, A.K. Co-selection for antibiotic resistance by environmental contaminants. NPJ Antimicrob. Resist. 2024, 2, 9. [Google Scholar] [CrossRef]
  37. Engin, A.B.; Engin, E.D.; Engin, A. Effects of co-selection of antibiotic-resistance and metal-resistance genes on antibiotic-resistance potency of environmental bacteria and related ecological risk factors. Environ. Toxicol. Pharmacol. 2023, 98, 104081. [Google Scholar] [CrossRef]
  38. de Araújo, L.C.A.; Lima, A.V.A.; Barbosa, A.V. Impacted Aquatic Environment Microorganisms. Emerg. Contam. 2021, 1, 191. [Google Scholar]
  39. Kim, I.S.; Yang, W.S.; Kim, C.H. Physiological properties, functions, and trends in the matrix metalloproteinase inhibitors in inflammation-mediated human diseases. Curr. Med. Chem. 2023, 30, 2075–2112. [Google Scholar] [CrossRef]
  40. Lorrai, R.; Ferrari, S. Host cell wall damage during pathogen infection: Mechanisms of perception and role in plant-pathogen interactions. Plants 2021, 10, 399. [Google Scholar] [CrossRef]
  41. Gaudet, R.G.; Bradfield, C.J.; MacMicking, J.D. Evolution of cell-autonomous effector mechanisms in macrophages versus non-immune cells. Myeloid Cells Health Dis. A Synth. 2017, 1, 615–635. [Google Scholar]
  42. Pontiggia, D.; Giulietti, S.; Gramegna, G.; Lionetti, V.; Lorrai, R.; Marti, L.; Cervone, F. The Ancient Battle between Plants and Pathogens: Resilience of the Plant Cell Wall and Damage-Associated Molecular Patterns (DAMPs) Drive Plant Immunity. In Plant Cell Walls; CRC Press: Boca Raton, FL, USA, 2023; pp. 393–411. [Google Scholar]
  43. Kothari, A.; Kumar, P.; Gaurav, A.; Kaushal, K.; Pandey, A.; Yadav, S.R.M.; Omar, B.J. Association of antibiotics and heavy metal arsenic to horizontal gene transfer from multidrug-resistant clinical strains to antibiotic-sensitive environmental strains. J. Hazard. Mater. 2023, 443, 130260. [Google Scholar] [CrossRef] [PubMed]
  44. Makar, O.; Kavulych, Y.; Terek, O.; Romanyuk, N. Plant-microbe interaction: Mechanisms and applications for improving crop yield and quality. Stud. Biol. 2023, 17, 225–242. [Google Scholar] [CrossRef]
  45. Khatun, J.; Intekhab, A.; Dhak, D. Effect of uncontrolled fertilization and heavy metal toxicity associated with arsenic (As), lead (Pb) and cadmium (Cd), and possible remediation. Toxicology 2022, 477, 153274. [Google Scholar] [CrossRef]
  46. Jach, M.E.; Sajnaga, E.; Ziaja, M. Utilization of legume-nodule bacterial symbiosis in phytoremediation of heavy metal-contaminated soils. Biology 2022, 11, 676. [Google Scholar] [CrossRef] [PubMed]
  47. Balali-Mood, M.; Naseri, K.; Tahergorabi, Z.; Khazdair, M.R.; Sadeghi, M. Toxic mechanisms of five heavy metals: Mercury, lead, chromium, cadmium, and arsenic. Front. Pharmacol. 2021, 12, 643972. [Google Scholar] [CrossRef]
  48. Pasic, L.; Avgustin, J.A.; Erjavec, M.S.; Herzog-Velikonja, B.; Podlesek, Z.; Zgur-Bertok, D. Two tales of prokaryotic genomic diversity: Escherichia coli and Halophiles. Food Technol. Biotechnol. 2014, 52, 158. [Google Scholar]
  49. Mahto, K.U.; Kumari, S.; Das, S. Unraveling the complex regulatory networks in biofilm formation in bacteria and relevance of biofilms in environmental remediation. Crit. Rev. Biochem. Mol. Biol. 2022, 57, 305–332. [Google Scholar] [CrossRef]
  50. Zhang, J.; Lei, H.; Huang, J.; Wong, J.W.; Li, B. Co-occurrence and co-expression of antibiotic, biocide, and metal resistance genes with mobile genetic elements in microbial communities subjected to long-term antibiotic pressure: Novel insights from metagenomics and metatranscriptomics. J. Hazard. Mater. 2025, 489, 137559. [Google Scholar] [CrossRef]
  51. Jia, T.; Wang, R.; Chai, B. Effects of heavy metal pollution on soil physicochemical properties and microbial diversity over different reclamation years in a copper tailings dam. J. Soil Water Conserv. 2019, 74, 439–448. [Google Scholar] [CrossRef]
  52. Williams, A.; Sinanaj, B.; Hoysted, G.A. Plant–microbe interactions through a lens: Tales from the mycorrhizosphere. Ann. Bot. 2024, 133, 399–412. [Google Scholar] [CrossRef]
  53. Sharma, M.; Bharti, M.; Modeel, S.; Khurana, H.; Negi, T.; Negi, R.K. Challenges of Multidrug-Resistant Microbes on Public Health. In Biomedical Research, Medicine, and Disease; CRC Press: Boca Raton, FL, USA, 2023; pp. 363–376. [Google Scholar]
  54. James, C.; James, S.J.; Onarinde, B.A.; Dixon, R.A.; Williams, N. A critical review of AMR risks arising as a consequence of using biocides and certain metals in food animal production. Antibiotics 2023, 12, 1569. [Google Scholar] [CrossRef]
  55. Yoon, H.J.; Shin, S.H.; Yoon, J.H. The Role of Reactive Oxygen Species from Heavy Metal: Effect on reactivity of Fish and Defensive Mechanism of Antibiotic Resistant Bacteria in Aquatic Environment. Water Air Soil Pollut. 2024, 235, 768. [Google Scholar] [CrossRef]
  56. Zaib, M.; Aryan, M.; Khaliq, A.; Haider, K.; Ahmad, S.; Zeeshan, A.; Zubair, H. Essential Insights for Effective Environmental Management and Human Well-being: Strategies for Remediation in Soil-Plant-Environment Systems. J. Asian Dev. Stud. 2023, 12, 1453–1469. [Google Scholar]
  57. Kumar, R.; Kamboj, N.; Mitra, D. Microbial innovations in agriculture: Interdisciplinary approaches to leveraging microbes for food sustainability and security. Indian. J. Microbiol. Res. 2024, 11, 129–139. [Google Scholar]
  58. Fakhar, A.; Gul, B.; Gurmani, A.R.; Khan, S.M.; Ali, S.; Sultan, T.; Rizwan, M. Heavy metal remediation and resistance mechanism of Aeromonas, Bacillus, and Pseudomonas: A review. Crit. Rev. Environ. Sci. Technol. 2022, 52, 1868–1914. [Google Scholar] [CrossRef]
  59. Sher, S.; Rehman, A. Use of heavy metals resistant bacteria—A strategy for arsenic bioremediation. Appl. Microbiol. Biotechnol. 2019, 103, 6007–6021. [Google Scholar] [CrossRef]
  60. Erdoğan, İ.; Cevher-Keskin, B.; Bilir, Ö.; Hong, Y.; Tör, M. Recent developments in CRISPR/Cas9 genome-editing technology related to plant disease resistance and abiotic stress tolerance. Biology 2023, 12, 1037. [Google Scholar] [CrossRef]
  61. Qiu, X.; Wang, B.; Ren, S.; Liu, X.; Wang, Y. Regulation of quorum sensing for the manipulation of conjugative transfer of antibiotic resistance genes in wastewater treatment system. Water Res. 2024, 253, 121222. [Google Scholar] [CrossRef] [PubMed]
  62. Tarfeen, N.; Nisa, K.U.; Hamid, B.; Bashir, Z.; Yatoo, A.M.; Dar, M.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]
  63. Guerra Sierra, B.E.; Muñoz Guerrero, J.; Sokolski, S. Phytoremediation of Heavy Metals in Tropical Soils an Overview. Sustainability 2021, 13, 2574. [Google Scholar] [CrossRef]
  64. Yadav, S.; Singh, K.; Chandra, R. Plant growth–promoting rhizobacteria (PGPR) and bioremediation of industrial waste. In Microbes for Sustainable Development and Bioremediation; CRC Press: Boca Raton, FL, USA, 2019; pp. 207–241. [Google Scholar]
  65. Madouh, T.A.; Davidson, M.K. Management of Environmental Pollution: Hyperaccumulator Plants, Arbuscular Mycorrhizal Fungi (AMF), and Biochar in Heavy Metal Remediation. In Ecosystem Management: Climate Change and Sustainability; John Wiley & Sons: Hoboken, NJ, USA, 2024; pp. 115–169. [Google Scholar]
  66. Zhang, H.; Liu, W.; Xiong, Y.; Li, G.; Cui, J.; Zhao, C.; Zhang, L. Effects of dissolved organic matter on distribution characteristics of heavy metals and their interactions with microorganisms in soil under long-term exogenous effects. Sci. Total Environ. 2024, 947, 174565. [Google Scholar] [CrossRef]
  67. Arora, D.; Arora, A.; Panghal, V.; Singh, A.; Bala, R.; Kumari, S.; Kumar, S. Unleashing the feasibility of nanotechnology in phytoremediation of heavy metal–contaminated soil: A critical review towards sustainable approach. Water Air Soil Pollut. 2024, 235, 57. [Google Scholar] [CrossRef]
  68. Timofeeva, A.M.; Galyamova, M.R.; Sedykh, S.E. Plant growth-promoting soil bacteria: Nitrogen fixation, phosphate solubilization, siderophore production, and other biological activities. Plants 2023, 12, 4074. [Google Scholar] [CrossRef]
  69. Sharma, M.; Sharma, S.; Paavan Gupta, M.; Goyal, S.; Talukder, D.; Baskoutas, S. Mechanisms of microbial resistance against cadmium–a review. J. Environ. Health Sci. Eng. 2024, 22, 13–30. [Google Scholar] [CrossRef]
  70. Vasilachi, I.C.; Stoleru, V.; Gavrilescu, M. Analysis of heavy metal impacts on cereal crop growth and development in contaminated soils. Agriculture 2023, 13, 1983. [Google Scholar] [CrossRef]
  71. Wong, C.K.F.; Teh, C.Y. Impact of biofertilizers on horticultural crops. In Biofertilizers: Study and Impact; Wiley: Hoboken, NJ, USA, 2021; pp. 39–103. [Google Scholar]
  72. Kumar, R.; Kamboj, N.; Mitra, D.; Verma, D.; Choudhury, T. Bacterial Strategies: Novel Pathways for Sustainable Development Through Wastewater. In AI Technologies for Enhancing Recycling Processes; IGI Global Scientific Publishing: Hershey, PA, USA, 2025; Volume 1, pp. 421–440. [Google Scholar]
  73. Wahab, A.; Muhammad, M.; Munir, A.; Abdi, G.; Zaman, W.; Ayaz, A.; Reddy, S.P.P. Role of arbuscular mycorrhizal fungi in regulating growth, enhancing productivity, and potentially influencing ecosystems under abiotic and biotic stresses. Plants 2023, 12, 3102. [Google Scholar] [CrossRef]
  74. Delano-Frier, J.P.; Castro-Guillén, J.L.; Blanco-Labra, A. Recent findings on the multifaceted functionality of enzyme inhibition by natural compounds: A review. Curr. Enzym. Inhib. 2008, 4, 121–152. [Google Scholar] [CrossRef]
  75. Lin, L.; Chen, Y.; Qu, L.; Zhang, Y.; Ma, K. Cd heavy metal and plants, rather than soil nutrient conditions, affect soil arbuscular mycorrhizal fungal diversity in green spaces during urbanization. Sci. Total Environ. 2020, 726, 138594. [Google Scholar] [CrossRef]
  76. Dubey, R.K.; Vishal Tripathi, V.T.; Edrisi, S.A.; Mansi Bakshi, M.B.; Dubey, P.K.; Ajeet Singh, A.S.; Abhilash, P.C. Role of plant growth-promoting microorganisms in sustainable agriculture and environmental remediation. In Advances in PGPR Research; CABI: Wallingford, UK, 2017; Volume 1, pp. 75–125. [Google Scholar]
  77. Zandi, P.; Schnug, E. Reactive oxygen species, antioxidant responses and implications from a microbial modulation perspective. Biology 2022, 11, 155. [Google Scholar] [CrossRef]
  78. Ighodaro, O.M.; Akinloye, O.A. First line defence antioxidants-superoxide dismutase (SOD), catalase (CAT) and glutathione peroxidase (GPX): Their fundamental role in the entire antioxidant defence grid. Alex. J. Med. 2018, 54, 287–293. [Google Scholar] [CrossRef]
  79. Batool, R.; Umer, M.J.; Hussain, B.; Anees, M.; Wang, Z. Molecular mechanisms of superoxide dismutase (SODs)-mediated defense in controlling oxidative stress in plants. In Antioxidant Defense in Plants: Molecular Basis of Regulation; Springer Nature: Singapore, 2022; pp. 157–179. [Google Scholar]
  80. Bhattacharya, P.; Chakraborty, N.; Pal, R. Bioremediation of toxic metals using algae. In Algal Biorefinery: An Integrated Approach; Springer: Berlin/Heidelberg, Germany, 2015; pp. 439–462. [Google Scholar]
  81. Parcheta, M.; Świsłocka, R.; Orzechowska, S.; Akimowicz, M.; Choińska, R.; Lewandowski, W. Recent developments in effective antioxidants: The structure and antioxidant properties. Materials 2021, 14, 1984. [Google Scholar] [CrossRef]
  82. Kühn, H.; Borchert, A. Regulation of enzymatic lipid peroxidation: The interplay of peroxidizing and peroxide reducing enzymes. Free Radic. Biol. Med. 2002, 33, 154–172. [Google Scholar] [CrossRef]
  83. Aragaw, T.A.; Bogale, F.M.; Gessesse, A. Adaptive response of thermophiles to redox stress and their role in the process of dye degradation from textile industry wastewater. Front. Physiol. 2022, 13, 908370. [Google Scholar] [CrossRef]
  84. 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]
  85. Ding, C.; Ding, Z.; Liu, Q.; Liu, W.; Chai, L. Advances in mechanism for the microbial transformation of heavy metals: Implications for bioremediation strategies. Chem. Commun. 2024, 60, 12315–12332. [Google Scholar] [CrossRef]
  86. Adhikary, S.; Saha, J.; Dutta, P.; Pal, A. Bacterial homeostasis and tolerance to potentially toxic metals and metalloids through diverse transporters: Metal-specific insights. Geomicrobiol. J. 2024, 41, 496–518. [Google Scholar] [CrossRef]
  87. Nnaji, N.D.; Anyanwu, C.U.; Miri, T.; Onyeaka, H. Mechanisms of heavy metal tolerance in bacteria: A review. Sustainability 2024, 16, 11124. [Google Scholar] [CrossRef]
  88. Branco, S.; Schauster, A.; Liao, H.L.; Ruytinx, J. Mechanisms of stress tolerance and their effects on the ecology and evolution of mycorrhizal fungi. New Phytol. 2022, 235, 2158–2175. [Google Scholar] [CrossRef]
  89. Stambulska, U.Y.; Bayliak, M.M. Legume-rhizobium symbiosis: Secondary metabolites, free radical processes, and effects of heavy metals. In Co-Evolution of Secondary Metabolites; Springer International Publishing: Cham, Switzerland, 2020; pp. 291–322. [Google Scholar]
  90. Antar, M.; Gopal, P.; Msimbira, L.A.; Naamala, J.; Nazari, M.; Overbeek, W.; Smith, D.L. Inter-organismal signaling in the rhizosphere. In Rhizosphere Biology: Interactions Between Microbes and Plants; Springer: Singapore, 2020; pp. 255–293. [Google Scholar]
  91. Karnwal, A.; Martolia, S.; Dohroo, A.; Al-Tawaha, A.R.M.S.; Malik, T. Exploring bioremediation strategies for heavy metals and POPs pollution: The role of microbes, plants, and nanotechnology. Front. Environ. Sci. 2024, 12, 1397850. [Google Scholar] [CrossRef]
  92. Ali, S.; Waqas, W.; Bakky, M.A.H.; Zada, S.; Saif, U.M.; Hasan, M.T.; Hui, W. Implications of Microalgal–Bacterial Interactions in Modern Aquaculture Practices: A Review of the Current Knowledge. Rev. Aquac. 2025, 17, e12980. [Google Scholar] [CrossRef]
  93. Gupta, R.; Khan, F.; Alqahtani, F.M.; Hashem, M.; Ahmad, F. Plant growth–promoting Rhizobacteria (PGPR) assisted bioremediation of Heavy Metal Toxicity. Appl. Biochem. Biotechnol. 2024, 196, 2928–2956. [Google Scholar] [CrossRef]
  94. Rai, M.; Ingle, A.P.; Paralikar, P.; Anasane, N.; Gade, R.; Ingle, P. Effective management of soft rot of ginger caused by Pythium spp. and Fusarium spp.: Emerging role of nanotechnology. Appl. Microbiol. Biotechnol. 2018, 102, 6827–6839. [Google Scholar] [CrossRef]
  95. Xu, W.; Chen, L.; Hu, X.; Zhang, L.; Huang, D.; Li, J.; Zhu, M. Botanic signal monitor: Advanced wearable sensor for plant health analysis. Adv. Funct. Mater. 2024, 34, 2410544. [Google Scholar] [CrossRef]
  96. Rensing, C.; Moodley, A.; Cavaco, L.M.; McDevitt, S.F. Resistance to metals used in agricultural production. In Antimicrobial Resistance in Bacteria from Livestock and Companion Animals; John Wiley & Sons: Hoboken, NJ, USA, 2018; pp. 83–107. [Google Scholar]
  97. Schulz, V.; Schmidt-Vogler, C.; Strohmeyer, P.; Weber, S.; Kleemann, D.; Nies, D.H.; Herzberg, M. Behind the shield of Czc: ZntR controls expression of the gene for the zinc-exporting P-type ATPase ZntA in Cupriavidus metallidurans. J. Bacteriol. 2021, 203, 10–1128. [Google Scholar] [CrossRef]
  98. Wilcox, D.E. Isothermal titration calorimetry of metal ions binding to proteins: An overview of recent studies. Inorg. Chim. Acta 2008, 361, 857–867. [Google Scholar] [CrossRef]
  99. Klein, J.S.; Lewinson, O. Bacterial ATP-driven transporters of transition metals: Physiological roles, mechanisms of action, and roles in bacterial virulence. Metallomics 2011, 3, 1098–1108. [Google Scholar] [CrossRef]
  100. Alotaibi, B.S.; Khan, M.; Shamim, S. Unraveling the underlying heavy metal detoxification mechanisms of Bacillus species. Microorganisms 2021, 9, 1628. [Google Scholar] [CrossRef]
  101. Priya, A.K.; Gnanasekaran, L.; Dutta, K.; Rajendran, S.; Balakrishnan, D.; Soto-Moscoso, M. Biosorption of heavy metals by microorganisms: Evaluation of different underlying mechanisms. Chemosphere 2022, 307, 135957. [Google Scholar] [CrossRef] [PubMed]
  102. Vocciante, M.; Grifoni, M.; Fusini, D.; Petruzzelli, G.; Franchi, E. The role of plant growth-promoting rhizobacteria (PGPR) in mitigating plant’s environmental stresses. Appl. Sci. 2022, 12, 1231. [Google Scholar] [CrossRef]
  103. Rohmanna, N.A.; Agustina, L.; Saputra, R.A.; Sari, N.; Noor, I.; Majid, Z.A.N.M. Enhancing phytoremediation of heavy metals: A comprehensive review of performance microorganism-assisted Cyperus rotundus L. J. Ecol. Eng. 2025, 26, 391–401. [Google Scholar] [CrossRef]
  104. Kumar, R.; Farda, B.; Mignini, A.; Djebaili, R.; Koolman, L.; Paul, A.; Mitra, D. Microbial Solutions in Agriculture: Enhancing Soil Health and Resilience Through Bio-Inoculants and Bioremediation. Bacteria 2025, 4, 28. [Google Scholar] [CrossRef]
  105. Chandwani, S.; Wahab, R.; Ahmad, N.; Manoharadas, S.; Natarajan, A. Siderophore and 1-Aminocyclopropane-1-Carboxylic-Acid Deaminase (ACCD) Producing Lead-Tolerant Bacillus spp. Alleviate Lead Stress and Improve N and P Dynamics in Black Gram (Vigna mungo L.) Plants. Geomicrobiol. J. 2025, 42, 496–506. [Google Scholar] [CrossRef]
  106. Shahid, M.; Singh, U.B.; Khan, M.S.; Singh, P.; Kumar, R.; Singh, R.N.; Singh, H.V. Bacterial ACC deaminase: Insights into enzymology, biochemistry, genetics, and potential role in amelioration of environmental stress in crop plants. Front. Microbiol. 2023, 14, 1132770. [Google Scholar] [CrossRef]
  107. Nazli, F.; Mustafa, A.; Ahmad, M.; Hussain, A.; Jamil, M.; Wang, X.; El-Esawi, M.A. A review on practical application and potentials of phytohormone-producing plant growth-promoting rhizobacteria for inducing heavy metal tolerance in crops. Sustainability 2020, 12, 9056. [Google Scholar] [CrossRef]
  108. Menhas, S.; Hayat, K.; Niazi, N.K.; Zhou, P.; Amna Bundschuh, J.; Chaudhary, H.J. Microbe-EDTA mediated approach in the phytoremediation of lead-contaminated soils using maize (Zea mays L.) plants. Int. J. Phytoremediation 2021, 23, 585–596. [Google Scholar] [CrossRef]
  109. Sultana, R.; Islam, S.M.N.; Sultana, T. Arsenic and other heavy metals resistant bacteria in rice ecosystem: Potential role in promoting plant growth and tolerance to heavy metal stress. Environ. Technol. Innov. 2023, 31, 103160. [Google Scholar] [CrossRef]
  110. Wu, R.; Sun, X.; Zhu, M.; Wang, Y.; Zhu, Y.; Fang, Z.; Du, S. Abscisic acid-producing bacterium Azospirillum brasilense effectively reduces heavy metals (cadmium, nickel, lead, and zinc) accumulation in pak choi across various soil types. Ecotoxicol. Environ. Saf. 2025, 298, 118277. [Google Scholar] [CrossRef]
  111. Naz, H.; Sayyed, R.Z.; Khan, R.U.; Naz, A.; Wani, O.A.; Maqsood, A.; Show, P.L. Mesorhizobium improves chickpea growth under chromium stress and alleviates chromium contamination of soil. J. Environ. Manag. 2023, 338, 117779. [Google Scholar] [CrossRef]
  112. Mandal, S.; Saha, K.K.; Mandal, N.C. Molecular insight into key eco-physiological process in bioremediating and plant-growth-promoting bacteria. Front. Agron. 2021, 3, 664126. [Google Scholar] [CrossRef]
  113. Janani, B.; Sre, V.V.; Syed, A.; Elgorban, A.M.; Abid, I.; Wong, L.S.; Khan, S.S. Engineering defects and lattice disorientation in layered double hydroxides by coupling 2D-Co (OH) 2 platelets via pn heterojunction for enhanced photocatalytic degradation chloramphenicol. Colloids Surf. A Physicochem. Eng. Asp. 2025, 705, 135674. [Google Scholar] [CrossRef]
  114. Marcon, L.; Oliveras, J.; Puntes, V.F. In situ nanoremediation of soils and groundwaters from the nanoparticle’s standpoint: A review. Sci. Total Environ. 2021, 791, 148324. [Google Scholar] [CrossRef]
  115. Salam, A.; Afridi, M.S.; Khan, A.R.; Azhar, W.; Shuaiqi, Y.; Ulhassan, Z.; Gan, Y. Cobalt induced toxicity and tolerance in plants: Insights from omics approaches. In Heavy Metal Toxicity and Tolerance in Plants: A Biological, Omics, and Genetic Engineering Approach; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2023; pp. 207–229. [Google Scholar]
  116. Keunen, E.; Remans, T.; Bohler, S.; Vangronsveld, J.; Cuypers, A. Metal-induced oxidative stress and plant mitochondria. Int. J. Mol. Sci. 2011, 12, 6894–6918. [Google Scholar] [CrossRef]
  117. Prabhu, C.; Satyaprasad, A.U.; Deekshit, V.K. Understanding bacterial resistance to heavy metals and nanoparticles: Mechanisms, implications, and challenges. J. Basic Microbiol. 2025, 65, e2400596. [Google Scholar] [CrossRef]
  118. Agarwal, P.; Giri, B.S.; Rani, R. Unravelling the role of rhizospheric plant-microbe synergy in phytoremediation: A genomic perspective. Curr. Genom. 2020, 21, 334–342. [Google Scholar] [CrossRef]
  119. Sachdev, S.; Ansari, S.A.; Ansari, M.I.; Fujita, M.; Hasanuzzaman, M. Abiotic stress and reactive oxygen species: Generation, signaling, and defense mechanisms. Antioxid 2021, 10, 277. [Google Scholar] [CrossRef]
  120. Malik, Z.; Afzal, S.; Danish, M.; Abbasi, G.H.; Bukhari, S.A.H.; Khan, M.I.; Ali, S. Role of nitric oxide and calcium signaling in abiotic stress tolerance in plants. In Protective Chemical Agents in the Amelioration of Plant Abiotic Stress: Biochemical and Molecular Perspectives; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2020; pp. 563–581. [Google Scholar]
  121. Ha-Tran, D.M.; Nguyen, T.T.M.; Hung, S.H.; Huang, E.; Huang, C.C. Roles of plant growth-promoting rhizobacteria (PGPR) in stimulating salinity stress defense in plants: A review. Int. J. Mol. Sci. 2021, 22, 3154. [Google Scholar] [CrossRef]
  122. Kamboj, N.; Chugh, P.; Wijenayake, W.P.T.; Mitra, D.; Panneerselvam, P.; Kumar, R. Multifunctionality and Diversity of Antagonistic Potential Fungi as Biocontrol Agent. In Bio-Control Agents for Sustainable Agriculture: Diversity, Mechanisms and Applications; Springer Nature: Singapore, 2025; Volume 1, pp. 167–208. [Google Scholar]
  123. Nastro, R.A.; Leccisi, E.; Toscanesi, M.; Liu, G.; Trifuoggi, M.; Ulgiati, S. Exploring avoided environmental impacts as well as energy and resource recovery from microbial desalination cell treatment of brine. Energies 2021, 14, 4453. [Google Scholar] [CrossRef]
  124. da Cruz Nizer, W.S.; Inkovskiy, V.; Versey, Z.; Strempel, N.; Cassol, E.; Overhage, J. Oxidative stress response in Pseudomonas aeruginosa. Pathogens 2021, 10, 1187. [Google Scholar] [CrossRef]
  125. Ballash, G.A.; Parker, E.M.; Mollenkopf, D.F.; Wittum, T.E. The One Health dissemination of antimicrobial resistance occurs in both natural and clinical environments. J. Am. Vet. Med. Assoc. 2024, 262, 451–458. [Google Scholar] [CrossRef]
  126. Chen, X.; Yin, H.; Li, G.; Wang, W.; Wong, P.K.; Zhao, H.; An, T. Antibiotic-resistance gene transfer in antibiotic-resistance bacteria under different light irradiation: Implications from oxidative stress and gene expression. Water Res. 2019, 149, 282–291. [Google Scholar] [CrossRef]
  127. Fischer, F. The Beneficial Microbiota: Exploring the Roles, Mechanisms, and Potential of Good Bacteria in Environmental, Human, and Industrial Contexts. Insight Into Epidemiol. 2024, 1, e1113. [Google Scholar]
  128. Rizvi, A.; Ahmed, B.; Khan, M.S.; Rajput, V.D.; Umar, S.; Minkina, T.; Lee, J. Maize associated bacterial microbiome linked mitigation of heavy metal stress: A multidimensional detoxification approach. Environ. Exp. Bot. 2022, 200, 104911. [Google Scholar] [CrossRef]
  129. Yang, Y.L.; Amnuaykanjanasin, A. Natural Product Reports. Nat. Prod. Rep. 2022, 39, 991–1014. [Google Scholar]
  130. Vats, P.; Kaur, U.J.; Rishi, P. Heavy metal-induced selection and proliferation of antibiotic resistance: A review. J. Appl. Microbiol. 2022, 132, 4058–4076. [Google Scholar] [CrossRef] [PubMed]
  131. Alnsour, A.R.; Daghmash, R.M.; Masadeh, M.M.; Alzoubi, K.H.; Masadeh, M.M.; Bataineh, N.H.; Al-Ogaidi, M.S. The pharmaceutical role of silver nanoparticles in treating multidrug-resistant bacteria and biofilms. Curr. Nanosci. 2024, 20, 471–494. [Google Scholar] [CrossRef]
  132. Pal, A.; Bhattacharjee, S.; Saha, J.; Sarkar, M.; Mandal, P. Bacterial survival strategies and responses under heavy metal stress: A comprehensive overview. Crit. Rev. Microbiol. 2022, 48, 327–355. [Google Scholar] [CrossRef]
  133. Andrés Juan, C.; Pérez de Lastra, J.M.; Plou Gasca, F.J.; Pérez-Lebeña, E. The chemistry of reactive oxygen species (ROS) revisited: Outlining their role in biological macromolecules (DNA, lipids and proteins) and induced pathologies. Int. J. Mol. Sci. 2021, 22, 4642. [Google Scholar] [CrossRef]
  134. Klitgaard, R.N.; Jana, B.; Guardabassi, L.; Nielsen, K.L.; Løbner-Olesen, A. DNA damage repair and drug efflux as potential targets for reversing low or intermediate ciprofloxacin resistance in E. coli K-12. Front. Microbiol. 2018, 9, 1438. [Google Scholar] [CrossRef]
  135. Bush, N.G.; Diez-Santos, I.; Abbott, L.R.; Maxwell, A. Quinolones: Mechanism, lethality and their contributions to antibiotic resistance. Molecules 2020, 25, 5662. [Google Scholar] [CrossRef]
  136. Podlesek, Z.; Žgur Bertok, D. The DNA damage inducible SOS response is a key player in the generation of bacterial persister cells and population wide tolerance. Front. Microbiol. 2020, 11, 1785. [Google Scholar] [CrossRef] [PubMed]
  137. Gao, B.; Liang, L.; Su, L.; Wen, A.; Zhou, C.; Feng, Y. Structural basis for regulation of SOS response in bacteria. Proc. Natl. Acad. Sci. USA 2023, 120, e2217493120. [Google Scholar] [CrossRef]
  138. Skutel, M.; Yanovskaya, D.; Demkina, A.; Shenfeld, A.; Musharova, O.; Severinov, K.; Isaev, A. RecA-dependent or independent recombination of plasmid DNA generates a conflict with the host EcoKI immunity by launching restriction alleviation. Nucleic Acids Res. 2024, 52, 5195–5208. [Google Scholar] [CrossRef]
  139. Liu, S.K.; Eisen, J.A.; Hanawalt, P.C.; Tessman, I. recA mutations that reduce the constitutive coprotease activity of the RecA1202 (Prtc) protein: Possible involvement of interfilament association in proteolytic and recombination activities. J. Bacteriol. 1993, 175, 6518–6529. [Google Scholar] [CrossRef]
  140. Tran, Q.G.; Le, T.T.; Choi, D.Y.; Cho, D.H.; Yun, J.H.; Choi, H.I.; Lee, Y.J. Progress and challenges in CRISPR/Cas applications in microalgae. J. Microbiol. 2025, 63, e2501028. [Google Scholar] [CrossRef]
  141. Andersson, D.I.; Balaban, N.Q.; Baquero, F.; Courvalin, P.; Glaser, P.; Gophna, U.; Tønjum, T. Antibiotic resistance: Turning evolutionary principles into clinical reality. FEMS Microbiol. Rev. 2020, 44, 171–188. [Google Scholar] [CrossRef]
  142. Singh, C.K.; Sodhi, K.K.; Shree, P.; Nitin, V. Heavy metals as catalysts in the evolution of antimicrobial resistance and the mechanisms underpinning co-selection. Curr. Microbiol. 2024, 81, 148. [Google Scholar] [CrossRef]
  143. Henrici De Angelis, L.; D’Andrea, M.; Di Pilato, V.; Demattè, E.; Brovarone, F.; Nardini, P.; Rossolini, G. First report of multi-drug resistant, VEB-6 producing Proteus mirabilis isolates from Italy. In Abstract Book Congresso; Italian society of microbiology; SIM: Genoa, Italy, 2017; Available online: https://shorturl.at/QLHqC (accessed on 15 July 2025).
  144. Briguglio, I.; Piras, S.; Corona, P.; Antonietta Pirisi, M.; Jabes, D.; Carta, A. SAR and anti-mycobacterial activity of quinolones and triazoloquinolones: An update. Anti-Infect. Agents 2013, 11, 75–89. [Google Scholar] [CrossRef]
  145. Drlica, K.; Zhao, X. Bacterial death from treatment with fluoroquinolones and other lethal stressors. Expert Rev. Anti-Infect. Ther. 2021, 19, 601–618. [Google Scholar] [CrossRef]
  146. Saleh, N.M.; Moemen, Y.S.; Mohamed, S.H.; Fathy, G.; Ahmed, A.A.; Al-Ghamdi, A.A.; El Sayed, I.E.T. Experimental and molecular docking studies of cyclic diphenyl phosphonates as DNA gyrase inhibitors for fluoroquinolone-resistant pathogens. Antibiotics 2022, 11, 53. [Google Scholar] [CrossRef]
  147. Suwanthada, P.; Kongsoi, S.; Miura, N.; Belotindos, L.P.; Piantham, C.; Toyting, J.; Suzuki, Y. The Impact of Substitutions at Positions 1 and 8 of Fluoroquinolones on the Activity Against Mutant DNA Gyrases of Salmonella Typhimurium. Microb. Drug Resist. 2023, 29, 552–560. [Google Scholar] [CrossRef]
  148. Blanchard, L.; de Groot, A. Coexistence of SOS-dependent and SOS-independent regulation of DNA repair genes in radiation-resistant Deinococcus bacteria. Cells 2021, 10, 924. [Google Scholar] [CrossRef]
  149. Mercolino, J.; Lo Sciuto, A.; Spinnato, M.C.; Rampioni, G.; Imperi, F. RecA and specialized error-prone DNA polymerases are not required for mutagenesis and antibiotic resistance induced by fluoroquinolones in Pseudomonas aeruginosa. Antibiotics 2022, 11, 325. [Google Scholar] [CrossRef]
  150. Xu, Y.; Tan, L.; Li, Q.; Zheng, X.; Liu, W. Sublethal concentrations of heavy metals Cu2+ and Zn2+ can induce the emergence of bacterial multidrug resistance. Environ. Technol. Innov. 2022, 27, 102379. [Google Scholar] [CrossRef]
  151. Muttiah, B.; Hanafiah, A. Bioinspired Nanoplatforms: Polydopamine and Exosomes for Targeted Antimicrobial Therapy. Polymers 2025, 17, 1670. [Google Scholar] [CrossRef]
  152. Elshobary, M.E.; Badawy, N.K.; Ashraf, Y.; Zatioun, A.A.; Masriya, H.H.; Ammar, M.M.; Assy, A.M. Combating antibiotic resistance: Mechanisms, multidrug-resistant pathogens, and novel therapeutic approaches: An updated review. Pharmaceuticals 2025, 18, 402. [Google Scholar] [CrossRef]
  153. Webster, C.M.; Shepherd, M. A mini-review: Environmental and metabolic factors affecting aminoglycoside efficacy. World J. Microbiol. Biotechnol. 2023, 39, 7. [Google Scholar] [CrossRef]
  154. Hou, K.; Jabeen, R.; Sun, L.; Wei, J. How do Mutations of Mycobacterium Genes Cause Drug Resistance in Tuberculosis? Curr. Pharm. Biotechnol. 2024, 25, 724–736. [Google Scholar] [CrossRef]
  155. Baran, A.; Kwiatkowska, A.; Potocki, L. Antibiotics and bacterial resistance—A short story of an endless arms race. Int. J. Mol. Sci. 2023, 24, 5777. [Google Scholar] [CrossRef]
  156. Mikami, M.; Shimizu, H.; Iwama, N.; Yajima, M.; Kuwasako, K.; Ogura, Y.; Nameki, N. Stalled ribosome rescue factors exert different roles depending on types of antibiotics in Escherichia coli. NPJ Antimicrob. Resist. 2024, 2, 22. [Google Scholar] [CrossRef]
  157. Sun, H.; Zeng, J.; Li, S.; Liang, P.; Zheng, C.; Liu, Y.; Sun, Q. Interaction between rpsL and gyrA mutations affects the fitness and dual resistance of Mycobacterium tuberculosis clinical isolates against streptomycin and fluoroquinolones. Infect. Drug Resist. 2018, 11, 431–440. [Google Scholar] [CrossRef]
  158. Springer, B.; Kidan, Y.G.; Prammananan, T.; Ellrott, K.; Böttger, E.C.; Sander, P. Mechanisms of streptomycin resistance: Selection of mutations in the 16S rRNA gene conferring resistance. Antimicrob. Agents Chemother. 2001, 45, 2877–2884. [Google Scholar] [CrossRef]
  159. Mwangi, Z.; Naeku, G.; Mureithi, M.; Onyambu, F.; Bulimo, W. Mutation patterns of resistance genes for macrolides, aminoglycosides, and rifampicin in non-tuberculous mycobacteria isolates from Kenya. F1000Research 2023, 11, 962. [Google Scholar] [CrossRef] [PubMed]
  160. Händel, N.; Hoeksema, M.; Freijo Mata, M.; Brul, S.; ter Kuile, B.H. Effects of stress, reactive oxygen species, and the SOS response on de novo acquisition of antibiotic resistance in Escherichia coli. Antimicrob. Agents Chemother. 2016, 60, 1319–1327. [Google Scholar] [CrossRef]
  161. Crane, J.K.; Alvarado, C.L.; Sutton, M.D. Role of the SOS response in the generation of antibiotic resistance in vivo. Antimicrob. Agents Chemother. 2021, 65, 10–1128. [Google Scholar] [CrossRef]
  162. Borgio, J.F.; Rasdan, A.S.; Sonbol, B.; Alhamid, G.; Almandil, N.B.; AbdulAzeez, S. Emerging status of multidrug-resistant bacteria and fungi in the arabian peninsula. Biology 2021, 10, 1144. [Google Scholar] [CrossRef]
  163. Blázquez, J.; Rodríguez-Beltrán, J.; Matic, I. Antibiotic-induced genetic variation: How it arises and how it can be prevented. Annu. Rev. Microbiol. 2018, 72, 209–230. [Google Scholar] [CrossRef]
  164. Fornelos, N.; Browning, D.F.; Butala, M. The use and abuse of LexA by mobile genetic elements. Trends Microbiol. 2016, 24, 391–401. [Google Scholar] [CrossRef]
  165. Sass, T.H.; Ferrazzoli, A.E.; Lovett, S.T. DnaA and SspA regulation of the iraD gene of Escherichia coli: An alternative DNA damage response independent of LexA/RecA. Genetics 2022, 221, iyac062. [Google Scholar] [CrossRef]
  166. Heininger, K. The deprivation syndrome is the driving force of phylogeny, ontogeny and oncogeny. Rev. Neurosci. 2001, 12, 217–288. [Google Scholar] [CrossRef]
  167. da Silva, P.B.; Araújo, V.H.; Fonseca-Santos, B.; Solcia, M.C.; Ribeiro, C.M.; da Silva, I.C.; Chorilli, M. Highlights regarding the use of metallic nanoparticles against pathogens considered a priority by the world health organization. Curr. Med. Chem. 2021, 28, 1906–1956. [Google Scholar] [CrossRef]
  168. Sazykin, I.S.; Sazykina, M.A.; Litsevich, A.R. Distribution of Antibiotic Resistance Genes in Microbial Communities: The Impact of Anthropogenic Pollution. Mol. Biol. 2024, 58, 1049–1062. [Google Scholar] [CrossRef]
  169. Bazzi, W.; Abou Fayad, A.G.; Nasser, A.; Haraoui, L.P.; Dewachi, O.; Abou-Sitta, G.; Matar, G.M. Heavy metal toxicity in armed conflicts potentiates AMR in A. baumannii by selecting for antibiotic and heavy metal co-resistance mechanisms. Front. Microbiol. 2020, 11, 68. [Google Scholar] [CrossRef]
  170. Delmar, J.A.; Su, C.C.; Yu, E.W. Bacterial multidrug efflux transporters. Annu. Rev. Biophys. 2014, 43, 93–117. [Google Scholar] [CrossRef] [PubMed]
  171. Sarangi, B.K.; Kalve, S.; Pandey, R.A.; Chakrabarti, T. Transgenic plants for phytoremediation of arsenic and chromium to enhance tolerance and hyperaccumulation. Transgenic Plant J. 2009, 3, 57–86. [Google Scholar]
  172. Geleta, D.; Abebe, G.; Alemu, B.; Workneh, N.; Beyene, G. Mechanisms of bacterial drug resistance with special emphasis on phenotypic and molecular characterization of extended spectrum beta-lactamase. New Microbiol. 2024, 47, 1–14. [Google Scholar]
  173. Priyadarshanee, M.; Chatterjee, S.; Rath, S.; Dash, H.R.; Das, S. Cellular and genetic mechanism of bacterial mercury resistance and their role in biogeochemistry and bioremediation. J. Hazard. Mater. 2022, 423, 126985. [Google Scholar] [CrossRef]
  174. Hikal, A.F.; Hasan, S.; Gudeta, D.; Zhao, S.; Foley, S.; Khan, A.A. The acquired pco gene cluster in Salmonella enterica mediates resistance to copper. Front. Microbiol. 2024, 15, 1454763. [Google Scholar] [CrossRef]
  175. Islam, M.R.; Mondol, S.M.; Hossen, M.A.; Khatun, M.P.; Selim, S.; Amiruzzaman; Rahaman, M.M. First report on comprehensive genomic analysis of a multidrug-resistant Enterobacter asburiae isolated from diabetic foot infection from Bangladesh. Sci. Rep. 2025, 15, 424. [Google Scholar] [CrossRef]
  176. Yan, G.; Chen, X.; Du, S.; Deng, Z.; Wang, L.; Chen, S. Genetic mechanisms of arsenic detoxification and metabolism in bacteria. Curr. Genet. 2019, 65, 329–338. [Google Scholar] [CrossRef]
  177. Chatterjee, S.; Barman, P.; Barman, C.; Majumdar, S.; Chakraborty, R. Multimodal cadmium resistance and its regulatory networking in Pseudomonas aeruginosa strain CD3. Sci. Rep. 2024, 14, 31689. [Google Scholar] [CrossRef]
  178. Taghavi, S.; Lesaulnier, C.; Monchy, S.; Wattiez, R.; Mergeay, M.; van der Lelie, D. Lead (II) resistance in Cupriavidus metallidurans CH34: Interplay between plasmid and chromosomally-located functions. Antonie Van. Leeuwenhoek 2009, 96, 171–182. [Google Scholar] [CrossRef]
  179. Zhu, T.; Tian, J.; Zhang, S.; Wu, N.; Fan, Y. Identification of the transcriptional regulator NcrB in the nickel resistance determinant of Leptospirillum ferriphilum UBK03. PLoS ONE 2011, 6, e17367. [Google Scholar] [CrossRef]
  180. Gu, R.; Gao, J.; Dong, L.; Liu, Y.; Li, X.; Bai, Q.; Xiao, H. Chromium metabolism characteristics of coexpression of ChrA and ChrT gene. Ecotoxicol. Environ. Saf. 2020, 204, 111060. [Google Scholar] [CrossRef]
  181. Wang, Q.; Liu, L.; Hou, Z.; Wang, L.; Ma, D.; Yang, G.; Luo, Y. Heavy metal copper accelerates the conjugative transfer of antibiotic resistance genes in freshwater microcosms. Sci. Total Environ. 2020, 717, 137055. [Google Scholar] [CrossRef]
  182. Zhang, Y.; Gu, A.Z.; Cen, T.; Li, X.; He, M.; Li, D.; Chen, J. Sub-inhibitory concentrations of heavy metals facilitate the horizontal transfer of plasmid-mediated antibiotic resistance genes in water environment. Environ. Pollut. 2018, 237, 74–82. [Google Scholar] [CrossRef]
  183. Vrancianu, C.O.; Popa, L.I.; Bleotu, C.; Chifiriuc, M.C. Targeting plasmids to limit acquisition and transmission of antimicrobial resistance. Front. Microbiol. 2020, 11, 761. [Google Scholar] [CrossRef]
  184. Sánchez-Corona, C.G.; Gonzalez-Avila, L.U.; Hernández-Cortez, C.; Rojas-Vargas, J.; Castro-Escarpulli, G.; Castelán-Sánchez, H.G. Impact of Heavy Metal and Resistance Genes on Antimicrobial Resistance: Ecological and Public Health Implications. Genes 2025, 16, 625. [Google Scholar] [CrossRef] [PubMed]
  185. Vermeire, M.L.; Thiour-Mauprivez, C.; De Clerck, C. Agroecological transition: Towards a better understanding of the impact of ecology-based farming practices on soil microbial ecotoxicology. FEMS Microbiol. Ecol. 2024, 100, fiae031. [Google Scholar] [CrossRef] [PubMed]
  186. Obruca, S.; Sedlacek, P.; Pernicova, I.; Kovalcik, A.; Novackova, I.; Slaninova, E.; Marova, I. Interconnection between PHA and stress robustness of bacteria. In The Handbook of Polyhydroxyalkanoates; CRC Press: Boca Raton, FL, USA, 2020; pp. 107–132. [Google Scholar]
  187. Francolini, I.; Piozzi, A. Role of antioxidant molecules and polymers in prevention of bacterial growth and biofilm formation. Curr. Med. Chem. 2020, 27, 4882–4904. [Google Scholar] [CrossRef] [PubMed]
  188. Holm, A.; Vikström, E. Quorum sensing communication between bacteria and human cells: Signals, targets, and functions. Front. Plant Sci. 2014, 5, 309. [Google Scholar] [CrossRef]
  189. Ramanathan, J. Microbiome in Defence Against Pathogens. In Pathogens and Environmental Impact on Life Forms: Understanding Pathogens and Host Defence Mechanisms; Springer: Cham, Switzerland, 2024; pp. 343–422. [Google Scholar]
  190. Sionov, R.V.; Steinberg, D. Targeting the holy triangle of quorum sensing, biofilm formation, and antibiotic resistance in pathogenic bacteria. Microorganisms 2022, 10, 1239. [Google Scholar] [CrossRef]
  191. Bouillet, S.; Bauer, T.S.; Gottesman, S. RpoS and the bacterial general stress response. Microbiol. Mol. Biol. Rev. 2024, 88, e00151-22. [Google Scholar] [CrossRef]
  192. Akter, S.; Rahman, M.A.; Ashrafudoulla, M.; Mahamud, A.S.U.; Chowdhury, M.A.H.; Ha, S.D. Mechanistic and bibliometric insights into RpoS-mediated biofilm regulation and its strategic role in food safety applications. Crit. Rev. Food Sci. Nutr. 2025, 1–15. [Google Scholar] [CrossRef] [PubMed]
  193. Scott, R.A. Transcriptional Responses of Pseudomonas Syringae to Factors Inherent During Plant Host Associations, with a Focus on Quorum Sensing and its Regulation. PhD Thesis, University of California, Berkeley, CA, USA, 2013. [Google Scholar]
  194. Gao, J.; Liu, H.; Zhang, Z.; Liang, Z. Quorum sensing-mediated lipid oxidation further regulating the environmental adaptability of Aspergillus ochraceus. Metabolites 2023, 13, 491. [Google Scholar] [CrossRef]
  195. Duan, X.; Pan, Y.; Cai, Z.; Liu, Y.; Zhang, Y.; Liu, M.; Yang, L. rpoS-mutation variants are selected in Pseudomonas aeruginosa biofilms under imipenem pressure. Cell Biosci. 2021, 11, 138. [Google Scholar] [CrossRef]
  196. Nadar, S.; Khan, T.; Patching, S.G.; Omri, A. Development of antibiofilm therapeutics strategies to overcome antimicrobial drug resistance. Microorganisms 2022, 10, 303. [Google Scholar] [CrossRef]
  197. Hetta, H.F.; Ramadan, Y.N.; Al-Harbi, A.I.A.; Ahmed, E.; Battah, B.; Abd Ellah, N.H.; Donadu, M.G. Nanotechnology as a promising approach to combat multidrug resistant bacteria: A comprehensive review and future perspectives. Biomedicines 2023, 11, 413. [Google Scholar] [CrossRef] [PubMed]
  198. François, P.; Schrenzel, J.; Götz, F. Biology and regulation of staphylococcal biofilm. Int. J. Mol. Sci. 2023, 24, 5218. [Google Scholar] [CrossRef] [PubMed]
  199. Eissa, M. Genus burkholderia: A double-edged sword with widespread implications for human health, agriculture, and the environment. J. Biol. Res. Rev. 2024, 1, 79–96. [Google Scholar] [CrossRef]
  200. Roy, R.; Tiwari, M.; Donelli, G.; Tiwari, V. Strategies for combating bacterial biofilms: A focus on anti-biofilm agents and their mechanisms of action. Virulence 2018, 9, 522–554. [Google Scholar] [CrossRef]
  201. Huang, L.; Ahmed, S.; Gu, Y.; Huang, J.; An, B.; Wu, C.; Cheng, G. The effects of natural products and environmental conditions on antimicrobial resistance. Molecules 2021, 26, 4277. [Google Scholar] [CrossRef]
  202. Mosselhy, D.A.; Assad, M.; Sironen, T.; Elbahri, M. Nanotheranostics: A Possible Solution for Drug-Resistant Staphylococcus aureus and their Biofilms? Nanomaterials 2021, 11, 82. [Google Scholar] [CrossRef]
  203. Zhao, Y.; Xu, H.; Wang, H.; Wang, P.; Chen, S. Multidrug resistance in Pseudomonas aeruginosa: Genetic control mechanisms and therapeutic advances. Mol. Biomed. 2024, 5, 62. [Google Scholar] [CrossRef] [PubMed]
  204. Huang, L.; Wu, C.; Gao, H.; Xu, C.; Dai, M.; Huang, L.; Cheng, G. Bacterial multidrug efflux pumps at the frontline of antimicrobial resistance: An overview. Antibiotics 2022, 11, 520. [Google Scholar] [CrossRef]
  205. Dong, T.; Schellhorn, H.E. Role of RpoS in virulence of pathogens. Infect. Immun. 2010, 78, 887–897. [Google Scholar] [CrossRef]
  206. Mishra, S.; Bharagava, R.N.; More, N.; Yadav, A.; Zainith, S.; Mani, S.; Chowdhary, P. Heavy metal contamination: An alarming threat to environment and human health. In Environmental Biotechnology: For Sustainable Future; Springer: Singapore, 2018; pp. 103–125. [Google Scholar]
  207. Nguyen, T.H.T.; Nguyen, H.D.; Le, M.H.; Nguyen, T.T.H.; Nguyen, T.D.; Nguyen, D.L.; Pham, H.N. Efflux pump inhibitors in controlling antibiotic resistance: Outlook under a heavy metal contamination context. Molecules 2023, 28, 2912. [Google Scholar] [CrossRef]
  208. Zhang, S.; Lu, J.; Wang, Y.; Verstraete, W.; Yuan, Z.; Guo, J. Insights of metallic nanoparticles and ions in accelerating the bacterial uptake of antibiotic resistance genes. J. Hazard. Mater. 2022, 421, 126728. [Google Scholar] [CrossRef]
  209. Xing, Y.; Kang, X.; Zhang, S.; Men, Y. Specific phenotypic, genomic, and fitness evolutionary trajectories toward streptomycin resistance induced by pesticide co-stressors in Escherichia coli. ISME Commun. 2021, 1, 39. [Google Scholar] [CrossRef] [PubMed]
  210. Kaur, U.J.; Preet, S.; Rishi, P. Augmented antibiotic resistance associated with cadmium induced alterations in Salmonella enterica serovar Typhi. Sci. Rep. 2018, 8, 12818. [Google Scholar] [CrossRef]
  211. Zhai, Y.; He, Z.; Kang, Y.; Yu, H.; Wang, J.; Du, P.; Gao, Z. Complete nucleotide sequence of pH11, an IncHI2 plasmid conferring multi-antibiotic resistance and multi-heavy metal resistance genes in a clinical Klebsiella pneumoniae isolate. Plasmid 2016, 86, 26–31. [Google Scholar] [CrossRef] [PubMed]
  212. Rojo-Bezares, B.; Azcona-Gutiérrez, J.M.; Martin, C.; Jareño, M.S.; Torres, C.; Sáenz, Y. Streptococcus agalactiae from pregnant women: Antibiotic and heavy-metal resistance mechanisms and molecular typing. Epidemiol. Infect. 2016, 144, 3205–3214. [Google Scholar] [CrossRef]
  213. Garhwal, D.; Vaghela, G.; Panwala, T.; Revdiwala, S.; Shah, A.; Mulla, S. Lead tolerance capacity of clinical bacterial isolates and change in their antibiotic susceptibility pattern after exposure to a heavy metal. Int. J. Med. Public. Health 2014, 4, 253. [Google Scholar]
  214. Nair, R.; Thapaliya, D.; Su, Y.; Smith, T.C. Resistance to zinc and cadmium in Staphylococcus aureus of human and animal origin. Infect. Control Hosp. Epidemiol. 2014, 35 (Suppl. S3), S32–S39. [Google Scholar] [CrossRef]
  215. Sandegren, L.; Linkevicius, M.; Lytsy, B.; Melhus, Å.; Andersson, D.I. Transfer of an Escherichia coli ST131 multiresistance cassette has created a Klebsiella pneumoniae-specific plasmid associated with a major nosocomial outbreak. J. Antimicrob. Chemother. 2012, 67, 74–83. [Google Scholar] [CrossRef]
  216. Hofmann, L.; Hirsch, M.; Ruthstein, S. Advances in understanding of the copper homeostasis in Pseudomonas aeruginosa. Int. J. Mol. Sci. 2021, 22, 2050. [Google Scholar] [CrossRef]
  217. Giacometti, F.; Shirzad-Aski, H.; Ferreira, S. Antimicrobials and food-related stresses as selective factors for antibiotic resistance along the farm to fork continuum. Antibiotics 2021, 10, 671. [Google Scholar] [CrossRef] [PubMed]
  218. Islam, T.N.; Meem, F.S.; Yasmin, R.; Amin, M.B.; Rahman, T.; Mohasin, M. Co-exposure of chromium or cadmium and a low concentration of amoxicillin are responsible to emerge amoxicillin resistant Staphylococcus aureus. J. Glob. Antimicrob. Resist. 2023, 35, 279–288. [Google Scholar] [CrossRef] [PubMed]
  219. Sun, S.; Wang, M.; Xiang, J.; Shao, Y.; Li, L.; Sedjoah, R.C.A.A.; Xin, Z. BON domain-containing protein-mediated co-selection of antibiotic and heavy metal resistance in bacteria. Int. J. Biol. Macromol. 2023, 238, 124062. [Google Scholar] [CrossRef] [PubMed]
  220. Robas, M.; Probanza, A.; González, D.; Jiménez, P.A. Mercury and antibiotic resistance co-selection in Bacillus sp. isolates from the Almadén mining district. Int. J. Environ. Res. Public Health 2021, 18, 8304. [Google Scholar] [CrossRef]
  221. Longhi, C.; Maurizi, L.; Conte, A.L.; Marazzato, M.; Comanducci, A.; Nicoletti, M.; Zagaglia, C. Extraintestinal pathogenic Escherichia coli: Beta-lactam antibiotic and heavy metal resistance. Antibiotics 2022, 11, 328. [Google Scholar] [CrossRef]
  222. KassimGhaima, K.; Mohamed, A.I.; Al Meshhdany, W.Y.; Abdulhassan, A.A. Resistance and bioadsorption of cadmium by Pseudomonas aeruginosa isolated from agricultural soil. Int. J. Appl. Environ. Sci. 2017, 12, 1649–1660. [Google Scholar]
  223. Figueiredo, R.; Card, R.M.; Nunez-Garcia, J.; Mendonca, N.; da Silva, G.J.; Anjum, M.F. Multidrug-resistant Salmonella enterica isolated from food animal and foodstuff may also be less susceptible to heavy metals. Foodborne Pathog. Dis. 2019, 16, 166–172. [Google Scholar] [CrossRef]
  224. 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]
  225. Dinakarkumar, Y.; Ramakrishnan, G.; Gujjula, K.R.; Vasu, V.; Balamurugan, P.; Murali, G. Fungal bioremediation: An overview of the mechanisms, applications, and future perspectives. Environ. Chem. Ecotoxicol. 2024, 6, 293–302. [Google Scholar] [CrossRef]
  226. Ghosh, A.; Sah, D.; Chakraborty, M.; Rai, J.P.N. Mechanism and Application of Bacterial Exopolysaccharides: An Advanced Approach for Sustainable Heavy Metal Absorption from Soil. Carbohydr. Res. 2024, 544, 109247. [Google Scholar] [CrossRef]
  227. Khan, M.M.; Siddiqi, S.A.; Farooque, A.A.; Iqbal, Q.; Shahid, S.A.; Akram, M.T.; Khan, I. Towards Sustainable Application of Wastewater in Agriculture: A Review on Reusability and Risk Assessment. Agronomy 2022, 12, 1397. [Google Scholar] [CrossRef]
  228. Xu, Q.; Xiong, X.; Shi, Y.; Qian, L.; Zhou, X.; Tian, X.; Fang, L. Antagonism or synergism? Contrasting toxicity mechanisms of combined antibiotic and metal pollution in Eisenia fetida. Environ. Pollut. 2025, 374, 126166. [Google Scholar] [CrossRef]
  229. Jally, B.; Laubie, B.; Chour, Z.; Muhr, L.; Qiu, R.; Morel, J.L.; Simonnot, M.O. A new method for recovering rare earth elements from the hyperaccumulating fern Dicranopteris linearis in China. Miner. Eng. 2021, 166, 106879. [Google Scholar] [CrossRef]
  230. Khan, A.M.; Yusoff, I.; Abu Bakar, N.K.; Abu Bakar, A.F.; Alias, Y.; Mispan, M.S. Accumulation, uptake, and bioavailability of rare earth elements (rees) in soil-grown plants from ex-mining area in Perak, Malaysia. Appl. Ecol. Environ. Res. 2017, 15, 117–133. [Google Scholar] [CrossRef]
  231. Qin, B.; Liu, W.; He, E.; Li, Y.; Liu, C.; Ruan, J.; Tang, Y. Vacuum pyrolysis method for reclamation of rare earth elements from the hyperaccumulator Dicranopteris dichotoma grown in contaminated soil. J. Clean. Prod. 2019, 229, 480–488. [Google Scholar] [CrossRef]
  232. Dalla Vecchia, F.; Nardi, S.; Santoro, V.; Pilon-Smits, E.; Schiavon, M. Brassica juncea and the Se-hyperaccumulator Stanleya pinnata exhibit different patterns of chromium and selenium accumulation and distribution while activating distinct oxidative stress-response signatures. Environ. Pollut. 2023, 320, 121048. [Google Scholar] [CrossRef]
  233. Bouslimi, H.; Ferreira, R.; Dridi, N.; Brito, P.; Martins-Dias, S.; Caçador, I.; Sleimi, N.; Laboratory of Resources; Materials and Ecosystems; Faculty of Sciences of Bizerte; et al. Effects of barium stress in Brassica juncea and Cakile maritima: The indicator role of some antioxidant enzymes and secondary metabolites. Phyton 2021, 90, 145. [Google Scholar] [CrossRef]
  234. Kouki, R.; Dridi, N.; Vives-Peris, V.; Gómez-Cadenas, A.; Caçador, I.; Pérez-Clemente, R.M.; Sleimi, N. Appraisal of Abelmoschus esculentus L. Response to Al and Ba Stress. Plants 2023, 12, 179. [Google Scholar] [CrossRef] [PubMed]
  235. Redondo-Gómez, S.; Mateos-Naranjo, E.; Vecino-Bueno, I.; Feldman, S.R. Accumulation and tolerance characteristics of chromium in a cordgrass Cr-hyperaccumulator, Spartina argentinensis. J. Hazard. Mater. 2011, 185, 862–869. [Google Scholar] [CrossRef]
  236. Dridi, N.; Bouslimi, H.; Duarte, B.; Caçador, I.; Sleimi, N. Evaluation of physiological and biochemical parameters and some bioindicators of barium tolerance in Limbarda crithmoides and Helianthus annuus. Int. J. Plant Biol. 2022, 13, 115–131. [Google Scholar] [CrossRef]
  237. Andrade, A.F.M.D.; Amaral Sobrinho, N.M.B.D.; Santos, F.S.D.; Magalhães, M.O.L.; Tolón-Becerra, A.; Lima, L.D.S. EDTA-induced Phytoextraction of Pb and Ba by Brachiaria (B. decumbens cv. Basilisk) in soil contaminated with oil exploration drilling waste. Acta Scientiarum. Agron. 2014, 36, 495–500. [Google Scholar]
  238. Smillie, C. Salicornia spp. as a biomonitor of Cu and Zn in salt marsh sediments. Ecol. Indic. 2015, 56, 70–78. [Google Scholar] [CrossRef]
  239. Mellem, J.J.; Baijnath, H.; Odhav, B. Translocation and accumulation of Cr, Hg, As, Pb, Cu and Ni by Amaranthus dubius (Amaranthaceae) from contaminated sites. J. Environ. Sci. Health Part A 2009, 44, 568–575. [Google Scholar] [CrossRef] [PubMed]
  240. Fellet, G.; Pošćić, F.; Casolo, V.; Marchiol, L. Metallophytes and thallium hyperaccumulation at the former Raibl lead–zinc mining site (Julian Alps, Italy). Plant Biosyst. Int. J. Deal. All Asp. Plant Biol. 2012, 146, 1023–1036. [Google Scholar] [CrossRef]
  241. Lutts, S.; Lefèvre, I.; Delpérée, C.; Kivits, S.; Dechamps, C.; Robledo, A.; Correal, E. Heavy Metal Accumulation by the Halophyte Species Mediterranean Saltbush. J. Environ. Qual. 2004, 33, 1271–1279. [Google Scholar] [CrossRef]
  242. Joshi, A.; Kanthaliya, B.; Rajput, V.; Minkina, T.; Arora, J. Assessment of Phytoremediation Capacity of Three Halophytes: Suaeda monoica, Tamarix indica, and Cressa critica. Biol. Futur. 2020, 71, 301–312. [Google Scholar] [CrossRef] [PubMed]
  243. Kachout, S.S.; Mansoura, A.B.; Mechergui, R.; Leclerc, J.C.; Rejeb, M.N.; Ouerghi, Z. Accumulation of Cu, Pb, Ni, and Zn in the halophyte plant Atriplex grown on polluted soil. J. Sci. Food Agric. 2012, 92, 336–342. [Google Scholar] [CrossRef]
  244. Ayyappan, D.; Sathiyaraj, G.; Ravindran, K.C. Phytoextraction of heavy metals by Sesuvium portulacastrum L., a salt marsh halophyte from tannery effluent. Int. J. Phytoremediation 2016, 18, 453–459. [Google Scholar] [CrossRef]
  245. Grosjean, N.; Le Jean, M.; Berthelot, C.; Chalot, M.; Gross, E.M.; Blaudez, D. Accumulation and fractionation of rare earth elements are conserved traits in the Phytolacca genus. Sci. Rep. 2019, 9, 18458. [Google Scholar] [CrossRef]
  246. Nezhadasad-Aghbash, B.; Radjabian, T.; Hajiboland, R. Tolerance to Zn toxicity in the halophyte Lepidium latifolium L. and the effect of salt on Zn tolerance and accumulation, respectively. Acta Agric. Slov. 2023, 119, 1–17. [Google Scholar] [CrossRef]
  247. Liu, J.; Duan, C.Q.; Zhang, X.H.; Zhu, Y.N.; Hu, C. Characteristics of chromium (III) uptake in the hyperaccumulator Leersia hexandra Swartz. Environ. Exp. Bot. 2011, 74, 122–126. [Google Scholar] [CrossRef]
  248. Kalve, S.; Sarangi, B.K.; Pandey, R.A.; Chakrabarti, T. Arsenic and chromium hyperaccumulation by an ecotype of Pteris vittata, prospective for phytoextraction from contaminated water and soil. Curr. Sci. 2011, 100, 888–894. [Google Scholar]
  249. Lokhande, V.H.; Patade, V.Y.; Srivastava, S.; Suprasanna, P.; Shrivastava, M.; Awasthi, G. Copper accumulation and biochemical responses of Sesuvium portulacastrum (L.). Mater. Today Proc. 2020, 31, 679–684. [Google Scholar] [CrossRef]
  250. Hajri, A.K.; Hamdi, N.; Alharbi, A.A.; Alsherari, S.A.; Albalawi, D.A.; Kelabi, E.; Ghnaya, T. Evaluation of the potential of two halophytes to extract Cd and Zn from contaminated saltwater. Environ. Sci. Pollut. Res. 2023, 30, 114525–114534. [Google Scholar] [CrossRef] [PubMed]
  251. Feng, J.; Lin, Y.; Yang, Y.; Shen, Q.; Huang, J.S.; Wang, S.; Zhu, X.; Li, Z. Tolerance and bioaccumulation of Cd and Cu in Sesuvium portulacastrum. Ecotoxicol. Environ. Saf. 2018, 147, 306–312. [Google Scholar] [CrossRef]
  252. Fatnani, D.; Patel, M.; Parida, A.K. Regulation of chromium translocation to shoot and physiological, metabolomic, and ionomic adjustments confer chromium stress tolerance in the halophyte Suaeda maritima. Environ. Pollut. 2023, 320, 121046. [Google Scholar] [CrossRef]
  253. Eslava-Silva, F.D.J.; Muñíz-Díaz de León, M.E.; Jiménez-Estrada, M. Pteridium aquilinum (Dennstaedtiaceae), a novel hyperaccumulator species of hexavalent chromium. Appl. Sci. 2023, 13, 5621. [Google Scholar] [CrossRef]
  254. Chai, M.; Shi, F.; Li, R.; Qiu, G.; Liu, F.; Liu, L. Growth and physiological responses to copper stress in a halophyte Spartina alterniflora (Poaceae). Acta Physiol. Plant. 2014, 36, 745–754. [Google Scholar] [CrossRef]
  255. Buendía-González, L.; Orozco-Villafuerte, J.; Cruz-Sosa, F.; Barrera-Díaz, C.E.; Vernon-Carter, E.J. Prosopis laevigata a potential chromium (VI) and cadmium (II) hyperaccumulator desert plant. Bioresour. Technol. 2010, 101, 5862–5867. [Google Scholar] [CrossRef] [PubMed]
  256. Adki, V.S.; Jadhav, J.P.; Bapat, V.A. Nopalea cochenillifera, a potential chromium (VI) hyperaccumulator plant. Environ. Sci. Pollut. Res. 2013, 20, 1173–1180. [Google Scholar] [CrossRef] [PubMed]
  257. Tangahu, B.V.; Abdullah, S.R.S.; Basri, H.; Idris, M.; Anuar, N.; Mukhlisin, M. Phytoremediation of wastewater containing lead (Pb) in pilot reed bed using Scirpus grossus. Int. J. Phytoremediation 2013, 15, 663–676. [Google Scholar] [CrossRef] [PubMed]
  258. Tangahu, B.V.; Abdullah, S.R.S.; Basri, H.; Idris, M.; Anuar, N.; Mukhlisin, M. Lead (Pb) removal from contaminated water using constructed wetland planted with Scirpus grossus: Optimization using response surface methodology (RSM) and assessment of rhizobacterial addition. Chemosphere 2022, 291, 132952. [Google Scholar] [CrossRef]
Figure 1. Cation Diffusion Facilitator (CDF) family transporters: illustrates the structure and passive efflux mechanism of divalent metal ions (e.g., Zn2+ and Cd2+) via CDF proteins across cellular compartments for metal detoxification.
Figure 1. Cation Diffusion Facilitator (CDF) family transporters: illustrates the structure and passive efflux mechanism of divalent metal ions (e.g., Zn2+ and Cd2+) via CDF proteins across cellular compartments for metal detoxification.
Applmicrobiol 05 00079 g001
Figure 2. ABC transporters in metal homeostasis: depicts the three-component model (SBP, TMD, and NBD) of ABC transporters mediating ATP-driven translocation of various metal ions across bacterial and eukaryotic membranes.
Figure 2. ABC transporters in metal homeostasis: depicts the three-component model (SBP, TMD, and NBD) of ABC transporters mediating ATP-driven translocation of various metal ions across bacterial and eukaryotic membranes.
Applmicrobiol 05 00079 g002
Table 2. Heavy metal resistance genes and plasmid resistance sites.
Table 2. Heavy metal resistance genes and plasmid resistance sites.
Resistance MechanismGenes InvolvedBinding SiteResistance SiteReferences
MercurymerA, merB, merC, merD, merE, merR, merTOperonMercury detoxification and efflux[173]
CoppercopA, copB, copC, copD, and pcoA, pcoC, pcoDOperonInvolved in copper efflux[174]
ZinczntA, zntB, zntCOperonInvolved in zinc efflux[175]
ArsenicarsA, arsB, arsC, arsD, arsR, arsHOperonArsenic detoxification and efflux[176]
Cadmium czcA, czcB, czcC, czcDOperonCadmium involved in efflux[177]
LeadpbrA, pbrB, pbrC, pbrTLocusLead involved in efflux and uptake[178]
NickelncrA, ncrB, ncrCOperonInvolved in nickel efflux[179]
ChromiumchrA, chrB, chrCOperonInvolved in chromium efflux[180]
Table 3. According to some studies, there are clinical and environmental repercussions when metals and antibiotic resistance are linked.
Table 3. According to some studies, there are clinical and environmental repercussions when metals and antibiotic resistance are linked.
Heavy MetalsClinically Relevant Microorganism(s) Exhibiting
Co-Selection
AntibioticsReference
CadmiumSalmonella TyphiCiprofloxacin, ceftizoxime, ampicillin, chloramphenicol[210]
Lead, cadmium, mercury, tellurite, cobalt, copperKlebsiella pneumoniaeAminoglycosides, Streptomycin, Rifampicin, Beta-lactams, Gentamicin, Trimethoprim, Chloramphenicol[211]
Cadmium, copperStreptococcus agalactiaeLincosamide, Macrolide[212]
LeadPseudomonas spp.Cefepime, Colistin, Levofloxacin[213]
LeadSalmonella TyphiAmpicillin-Sulbactam, Nalidixic Acid, Levofloxacin[213]
LeadMembers of EnterobacteriaceaeTetracycline, Co-Trimoxazole, Third and Fourth generation Cephalosporins, Ciprofloxacin, Gentamycin[213]
LeadStaphylococcus aureusAmoxicillin/Clavulanic Acid, Cefoxitin, Co-Trimoxazole, Amoxicillin/Clavulanic Acid[213]
Cadmium, ZincStaphylococcus aureusMethicillin[214]
Arsenic, Silver, CopperKlebsiella pneumoniae, Escherichia coliMacrolides, Trimethoprim, Beta-lactams, Tetracyclines, Aminoglycosides[215]
Copper, ZincPseudomonas aeruginosaCarbapenem[216]
CadmiumKlebsiella pneumoniae, Pseudomonas aeruginosaSulphonamides, Kanamycin, Oxacillin, Nalidixic Acid[217]
Chromium, LeadStaphylococcus aureusOxacillin, Penicillin G, Methicillin[218]
CadmiumStenotrophomonas maltophiliaMacrolide[219]
MercuryMembers of Enterobacteriaceae, Oral StreptococciTetracycline, Ampicillin, Streptomycin[220]
MercuryEscherichia coliKanamycin, Chloramphenicol, Streptomycin, Tetracycline[221]
Arsenic, Mercury, CadmiumPseudomonas aeruginosaChloramphenicol, Streptomycin, Kanamycin[222]
Mercury, SilverSalmonella TyphimuriumAmpicillin, Chloramphenicol[223]
MercuryStaphylococcus aureusPenicillin[224]
Table 4. Heavy metal accumulator species in soil.
Table 4. Heavy metal accumulator species in soil.
Heavy MetalsAccumulator SpeciesClimate ConditionExperimental ConditionConcentration in PlantsReferences
Group of rare earth elements (REE)Dicranopteris linearisAnnual Rainfall Temperature (ART)Chinese plants grown from mining tailingsDry Weight (DW) Biomass milligram/g Total REE; 2.70 billion[229]
ScCyperus Rotundus L.ARTPlant grown on pH 4.4–7.5 media soil from a former mining region in Lahat, MalasiaFlower (8.21 µg/g), Leaves (204.60 µg/g), Stem (18.84 microgram/g), Root (280.83 µg/g) DW[230]
ScDicranopteris linearis (Burn) (B)ARTSoil (pH 4.4–7.5); plant from a former mining region near Lahat, MalasiaLamina (478.50 µg/g), Petiole (26.57 µg/g), Shoot (24.53 µg/g), Root (312.41 microgram/g) DW[230]
Sc, REEDicranopteris dichotomaARTPlants from a former mining region in Lahat, Malasia Media; soil (pH 4.4–7.5); plants gathered from ChinaBiomass (mg/kg)
Se 121
Cr 155
Nd 511
La 929
Roots (33.16 µg/g), Shoots (177.83 µg/g), Leaves (33.40 µg/g) DW
[230,231]
Se, Cr, BaBrassica juncea (glycophyte)ART 26/21 °CMedia: perlite/gravel combination (2:1, v/v)Mas. fresh weight; Roots (4017 mg), Shoots (9471mg), Whole plant (13,488 mg); (Ba concentration 100µM)
Tolerance index; 112–183% of the whole plant leaf.
Se 2600 mgCr/kg DW
Cr 1000 mgSe/killogram DW
Se 35 mgCr/kg/dl root
Cr 1150 mg Se/kg DW
[232,233]
BaAbelmoshush esculentus L.ART 12/25 °C (Night/Day)Plant: Tunisia Media, Baddar Company; blend gravel and perlite/gravel (1:2, v/v)Ba in Fruits (150–450 mg/kg), Shoots (200–900 microgram/killogram), Roots (450–1100 mg/kg) DW[234]
BaCakile maritima (Halophyes)ART 25 ± 5 °CMedia: perlite/gravel combination (2:1, v/v)Mas. Fresh weight in roots (1339 mg), Shoots (13,019 mg), Whole plants (14,358 mg), (Ba concentration: 500µM)
71–88% of whole plants
[233]
CrSpartina argentinensisART 21–25 °CPlants; Argentina
Media; inert substrate, pearlite in glass
Cr absorption in root and cultivator dry masses; 4–15 milligram/gram[235]
BaHelianthus annussART 25 ± 5 °CSunflower-seed land farm located north of Tunisia
Media; mixture of perlite and gravel substrate (2:1, v/v)
Shoots 1350 µg/g DW[236]
BaLimbarda crithmoidesARTNorth seeds shore of the Bizerte lagon in Menzel emil, Bizerte, TunisiaIn each organ; be around 3000 µg/g DW[236]
Ba, PbBrachiaria decumbensARTMedia; soil tainted by waste from exploration and oil well drilling (pH soil 8.3) Cv. Basilisk plants from Brazil
Including EDTA in the therapy
In Leaf (Pb 0.5–9.02 mg/kg), (Ba; 24.3–27.3 mg/kg) DW
In Root (Pb; 1.41–5.3 mg/kg), (Ba; 17.6–20.9 mg/kg) DW
[237]
Zn, CuSalicornia spp.ARTSediment samples from Restronguet CreekCu in Sediment (3420), Aerial water (551), Roots (1954) mg/kg DW
Zn in Sediment (1991), Aerial water (4831), Roots (720) mg/kg DW
[238]
Ni, Cu, Pb, As, Hg, CrAmaranthus dubiusARTSoil pH 7–8
Plants dispersed in media throughout Africa, Asia, and South America
In Leaves Cr, Hg, As, Pb, Cu, Ni (17–118, 3–12, 4–188, 3–9, 20–45, 9–36) ppm
In Stems Cr, Hg, As, Pb, Cu, Ni (17–126, 17–61, 13–201, 6–21, 27–67, 9–38) ppm
In Roots Cr, Hg, As, Pb, Cu, Ni (30–308, 24–66-, 7–127, 10–138, 118–173, 38–106) ppm
[239]
Zn, Ti, Pb, CdMinuartia verna subsp. VernaTotal annual rainfall: 2100 mm
The temperature was 2.4 °C in January and 16.4 °C in August.
Soil Raibl Lead/Zinc mining site, Cave del Predial, Italy; Media and Plants (pH 7.84–8.44)M. verna subsp.
In shoot Zn, Ti, Pb, Cd (24.1–18,372, 18.3–3632, 40–5998, 0.55–24.7) mg/kg
In Root Cd, Pb, Ti, Zn (0.60–29.2, 151–9329, 0–1286, 342–27,521) mg/kg
[240]
Zn, Ti, Pb, CdThlaspo rotundifolium subsp. CepaeifoliumAnnual rainfall 2100 mm
Temperature 2.4 °C January and 16.4 °C August
Plant and Media; Cave del Predial, Soil Raibl lead, zinc mining site, Italy (pH 784–8.44)Root and Shoot (mg/kg)
Cd; 1.30–47.9 and 1.35–78
Pb; 282–14,435 and 29.2–2817
Ti; 35.8–2275 and 0–874
Zn; 71.7–11,548 and 28.1–11,573
[240]
Cd, Pb, Ti, ZnBiscutella laevigata subsp.Annual rainfall 2100 mm
Temperature 2.4 °C January and 16.4 °C August
Media and Plants; Zinc mining site, Soil Raibl lead, Cave del Predial, Italy (pH 7.84–8.44)Shoot and Root (mg/kg)
Cd; 0–4.75 and 0–8.45
Pb; 0–874 and 0–1939
Ti; 33.4–32,661 and 0–9984
Zn; 20.2–20,980 and 34.7–5692
[240]
Cd, Pb, Ti, ZnAlyssum wulfenianumAnnual rainfall 2100 mm Temperature
2.4 °C January and 16.4 °C August
Media and Plants; Cave del Predial, Soil Raibl lead, Zinc mining site, Italy (pH 7.84–8.44)Root and Shoot (mg/kg)
Cd; 2.20–6.23 and 0.59–5.95
Pb; 253–1504 and 28–826
Ti; 0–4347 and 31.5–1946
Zn; 354–5670 and 194–1934
[240]
Zn, CdAtriplex halimus L. (Halophyte species Mediterranean saltbush)ART 25 ± 2 °CMedia; soil
Plant; Spain
Leaves, Roots, Stems (mg/kg DW)
Zn; 460, 1423, 431
Cd; 618, 3174, 1151
[241]
Cd, Cr, Cu, Mn, Fe, ZnCressa critica (Halophytes)Rainfall range: 650–700 mm annuallyGujrat, India’s Gulf of Bhabha soil, pH 8.93, EC 112 dS/mStems and Leaves (mg/kg DW)
Cd; 2 and 2.1
Cu; 11 and 12
Mn; 30 and 35
Fe; 160 nm and 150
Zn; 15 and 14
[242]
Zn, Fe, Mn, Cu, Cr, CdTamarix indica (Halophytes)Annual rainfall; 650–700 mm per yearSoil from Gulf of Bhabha, Gujrat, India, pH 8.93, EC 112 dS/mIn Stems Zn, Fe, Mn, Cu, Cd (19,150, 25, 10, 3.5) mg/kg DW
In Leaves Zn, Fe, Mn, Cu, Cd (17, 160, 39, 9, 3.6) mg/kg DW
[242]
Zn, Pb, CuHalophyte Atriplex species; A. roseaARTPlants; CN Seed Ltd., Ely, UK
Media; Contaminated soil in France
mg/kg DW
In Root Pb (725)
In Roots Cu (385)
In Shoots Cu (48)
In Root Zn (1179)
[243]
Cu, PbHalophytes Atriplex species; A. hortensis var. purpureaARTMedia; French plants with contaminated soil; CN Seed Ltd., Ely, UKIn Root Cu (265 mg/kg DW) and Pb (1064 mg/kg DW)[243]
Cu, Ni, Zn, PbHalophyte Atriplex species: Atriplex hortensis var. rubraARTMedia; Contaminated soil in France
Plants; CN Seed Ltd., Ely, UK
In Root Ni, Pb, Zn (1593, 982, 2512) mg/kg DW
In Shoot Ni, Cu (61, 30 min) mg/kg DW
[243]
Zn, Cu, Cr, CdSesuvium portulacastrum (Halophytes) Rainfall total during the year: 135.6 cm Temperature range: 33.26 °C in the summer to 29.68 °C in the winterMedia: soil treated with tanning effluent (pH 8.3, EC 5.25 dSm-1) and salt (pH 7., EC 4.48 dSm-1)
Plants from the Cuddalore District’s Pichavaram Mangrove Forest in Tail Nasa, India
Zn (70.10), Cu (35.10), Cd (22.10), Cr (49.82) mg/g DW[244]
Zn, Cu, Cr, CdSuaeda monoica (Salt marsh halophyte)Rainfall totals: 135.6 cm annually Maximum summer temperature: 33.26 °C; minimum winter temperature: 29.68 °CMedia; soil (pH 8.40, EC 4.75 dS/m) treated with paper mill effluent; plants; Tamil Nadu, India.
Media: Salt-treated soil (pH 7.7, EC 4.48 dSm-1) and tannery wastewater samples (pH 8.3, EC 5.25 dSm-1). Gujrat, India’s Gulf of Bhabha soil, pH 8.93, EC 112 dS/m
Plants from the Cuddalore District’s Pichavaram Mangrove Forest in Tail Nasa, India
Mg/kg DW
Zn (60.20)
Cu (29.2)
Cd (17.50)
Cr (40.89)
Stems and Leaves (mg/kg DW)
Cr (2.5 and 2)
Cd (4.0 and 4.5)
Cu (10 °C and 12)
Mn (40 and 45)
Fr (160 and 170)
Zn (21 and 20)
Mg/kg DW
Zn 22.33
Cu 23.89
Cr 18.66
Cd 15
[244]
Table 5. Heavy metal accumulator species in water.
Table 5. Heavy metal accumulator species in water.
Heavy MetalsAccumulator SpeciesClimate ConditionExperimental ConditionConcentration in PlantsReferences
Rare earth elements (REEs)Atriplex rosea
Phytolacca bogotensis
P. acinosa Roxb.
P. icosandra L.
P. clavigera W.W. Smith
P. americana L.
Annual rainfall Temperature (ART) 23/18 °CMedia: water
Plants: UK, Canada, France
Dry Weight (DW)
Shoot
Ni; 90–460 microgram/gram DW
Cu; 160–390 µg/g DW
Cd; 210–340 µg/g DW
Zn; 126–1341 microgram/g DW
Pb; 157–1184 µg/g DW
Leaves and Roots; 180–500 and 1500–8000 mg/kg
Leaves and Roots; 150–250 and 1600–10,000 mg/kg
Leaves and Roots; 50–400 and 1600–13,000 mg/kg
Leaves and Roots; 250–500 and 1500–6000 mg/kg
Leaves and Roots; 300–900 and 1500–5000 mg/kg
[245]
Cd, Cu, Zn, PbAtriplex hortensisARTMedia; hydroponic
Plants from Denmark
Shoot
Zn; 351–1257 microgram/gram DW
Pb; 773–2276 µg/g DW
Cd; 210–340 microgram/g DW
Ni; 210–770 µg/g DW
Cu; 340–400 µg/gram DW
[243]
ZnHalophytes Lepidium latifolium L.ART 25/17 °CMedia; Water
Plants; Meghan Playa, Iran
Old Leaves, Roots, Young Leaves (0.7, 2.9, 1 mg/g DW)[246]
Cr IIILeersia hexandra (Swartz)ART 25/18 °C Day/NightMedia; Water
The Taohua River in Guilin
Uptake Cr(III) 48 h; 3300 mg/kg DW[247]
Cr, AsPterus vittata (Ecotype)ART 30–35 °CMedia; Hoagland medium (water)
Plants in India
Root, Leaf, Stem AS; (1173–10,170, 2307–3474, 1299–6400) microgram/kg DW
Cr; 9200–17,427, 611–2609, 1556–20,675 mg/kg DW
[248]
Cu, Cd, ZnSesuvium portulacastrum (Halophytes) ARTMS Medium; water for Indian plantsRoots and Shoots
Cd; 300–800 and 340–620 µg/g DW
Zn; 910–1200 and 420–650 µg/g DW
[249,250]
Zn, CdCarpobrotus edulis L.ARTMedia; waterRoots and Shoots
Zn; 660–1450 and 560–700 µg/g DW
Cd; 670–1250 and 330–670 µg/g DW
[250]
Cd, CuSesuvium portulacastrum (Halophyte) ARTMedia; water
Plant from Halling Levee, Shatoulong Village, Yangjiang, China
Leaf, Root, Stem (mg/kg DW)
Cu; 29.3–1032, 22.9–2090, 33.4–106.1
Cd; 16.26–64.7, 1707–3815, 77.8–251
[251]
CrSuaeda maritima HalophyteARTMedia: Central Salt and Marine Chemical Research Institute (CSMCRI), Bhavnagar, Gujarat, India; hydroponic plant from a salt farm.
400 µM Cr was applied to the plant.
Cr. In Shoot (90.83 µg/g DW)
Cr. In Root (1874.97 µg/g DW)
[252]
Cr (VI)Pteridium aquilinum (Dennstaedtiaceae)ARTMedia; in vitro and via hydroponic treatment
Spores in La Cantera, Maxico Use of the gametophyte and sporophytes phase
Sporophytes 11,854 mg Cr/kg DW in underground part.
Gametophytes 915 mg Cr/kg DW
[253]
Se, CrStanleya pinnataART
26/21 °C
Media; WaterIn leaf; Se (12 microgram Cr/kg DW), Cr (1500 microgram Se/killogram DW)
In Root; Se (500 mg Cr/kg DW), Cr (950 mg Se/kg DW)
[232]
CuSpartina alterniflora (Poaceae) HalophyteART
22–28 °C (Day) and 16–22 °C (Night)
Media: Hoagland’s nutrition solution pH 6.5, clean, dry vermiculite
Seeds, China’s Bohai Bay
Rhozomes; 49 µg/g DW
Steam; 49 µg/g DW
Fine roots; 197.12 µg/g DW
Leaves; 47 µg/g DW
[254]
Cr, CdProsopis laevigataART 25 ± 2 °CMedia; liquid culture media, often known as Murashige and Skoog (MS) medium
Humb and Bonel Seed. Ex. Mexican State’s Willd. M. C. Johnston
3.4 mM Cr (VI) and 0.65 mM Cd (II) were used to cultivate the plants.
In Root; (Cr 8090 mg/kg DW), (Cd; 21,437 mg/kg DW)
In Shoot; (Cr 5461 mg/kg DW), (Cd; 8176 mg/kg DW)
[255]
Cr (VI)Nopalea cochenilliferaARTMedia; Shivaji University, Kolhapur, India seeds; Murashige and Skoog’s (MS) basal medium (pH 5.80 corrected)Shoots; 705.714 microgram Cr/kg DW
Roots; 25,263.396 microgram Cr/kg DW
[256]
PbScirpus grossusARTMedium sand and water with spikesPb Treatment
3236 mg/kg for 50 mg/L
4909 mg/kg for 30 mg/L
1343 milligram/killogram for 10 milligram/L
[257,258]
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

Kumar, R.; Vasić, T.P.; Živković, S.P.; Panneerselvam, P.; Santoyo, G.; de los Santos Villalobos, S.; Olatunbosun, A.N.; Pandit, A.; Koolman, L.; Mitra, D.; et al. Mechanistic Role of Heavy Metals in Driving Antimicrobial Resistance: From Rhizosphere to Phyllosphere. Appl. Microbiol. 2025, 5, 79. https://doi.org/10.3390/applmicrobiol5030079

AMA Style

Kumar R, Vasić TP, Živković SP, Panneerselvam P, Santoyo G, de los Santos Villalobos S, Olatunbosun AN, Pandit A, Koolman L, Mitra D, et al. Mechanistic Role of Heavy Metals in Driving Antimicrobial Resistance: From Rhizosphere to Phyllosphere. Applied Microbiology. 2025; 5(3):79. https://doi.org/10.3390/applmicrobiol5030079

Chicago/Turabian Style

Kumar, Rahul, Tanja P. Vasić, Sanja P. Živković, Periyasamy Panneerselvam, Gustavo Santoyo, Sergio de los Santos Villalobos, Adeyemi Nurudeen Olatunbosun, Aditi Pandit, Leonard Koolman, Debasis Mitra, and et al. 2025. "Mechanistic Role of Heavy Metals in Driving Antimicrobial Resistance: From Rhizosphere to Phyllosphere" Applied Microbiology 5, no. 3: 79. https://doi.org/10.3390/applmicrobiol5030079

APA Style

Kumar, R., Vasić, T. P., Živković, S. P., Panneerselvam, P., Santoyo, G., de los Santos Villalobos, S., Olatunbosun, A. N., Pandit, A., Koolman, L., Mitra, D., & Gautam, P. (2025). Mechanistic Role of Heavy Metals in Driving Antimicrobial Resistance: From Rhizosphere to Phyllosphere. Applied Microbiology, 5(3), 79. https://doi.org/10.3390/applmicrobiol5030079

Article Metrics

Back to TopTop