Next Article in Journal
Effect of Tilapia Parvovirus (TiPV) on Fish Health: An In Vitro Approach
Previous Article in Journal
Prevention of Recurrent Urinary Tract Infection in Women: An Update
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microbial Metallophores in the Productivity of Agroecosystems

by
Lily X. Zelaya-Molina
1,*,
Ismael F. Chávez-Díaz
1,*,
José A. Urrieta-Velázquez
2,
Marco A. Aragón-Magadan
1,
Cristo O. Puente-Valenzuela
2,
Mario Blanco-Camarillo
3,
Sergio de los Santos-Villalobos
4 and
Juan Ramos-Garza
5
1
Centro Nacional de Recursos Genéticos-INIFAP, Boulevard de la Biodiversidad No. 400, Tepatitlán de Morelos 47600, Jalisco, Mexico
2
CENID-RASPA-INIFAP, Km. 6.5 margen derecha, Canal de Sacramento, Gómez Palacio 35079, Durango, Mexico
3
Colegio de Postgraduados, Campus Montecillo, Carretera México-Texcoco Km. 36.5, Texcoco 56230, Estado de México, Mexico
4
Departamento de Ciencias Agronómicas y Veterinarias, Instituto Tecnológico de Sonora, 5 de febrero 818 sur, Ciudad Obregón 85000, Sonora, Mexico
5
Escuela de Ciencias de la Salud, Universidad del Valle de México, Campus Coyoacán, Calzada de Tlalpan 3016/3058, Coyoacán 04910, Ciudad de México, Mexico
*
Authors to whom correspondence should be addressed.
Microbiol. Res. 2025, 16(3), 67; https://doi.org/10.3390/microbiolres16030067
Submission received: 20 February 2025 / Revised: 9 March 2025 / Accepted: 12 March 2025 / Published: 14 March 2025

Abstract

:
Microbial metallophores are low-molecular-weight chelating agents produced by microorganisms to acquire essential metal ions. Their biosynthesis, transport, and regulation involve complex processes, specialized enzymatic machinery, and intricate regulatory networks. This review examines the multifaceted roles of metallophores in microbial ecology and their potential applications in sustainable agriculture, emphasizing their key role in trace metal acquisition, nutrient cycling, and plant–microbe interactions. Furthermore, it explores the potential applications of metallophores in agriculture, bioremediation, and biotechnology, connecting their potential to the development of novel strategies for sustainable agriculture.

1. Introduction

In recent years, among the green strategies implemented to mitigate the negative effects of agrochemicals on soil, water quality, and human health, the development of bioproducts based on beneficial microorganisms has gained significant attention in the academic, industrial, and governmental sectors [1]. Plant-growth-promoting bacteria (PGPB) and plant-growth-promoting fungi (PGPF) establish symbiotic relationships in the rhizosphere, phyllosphere or plant-endosphere, resulting in mutual benefits [2,3,4]. PGPB and PGPF promote crop growth and yield through both direct and indirect mechanisms, mediated by different types of symbiotic interactions with host plants. Direct mechanisms involve improving the plant’s nutritional status by enhancing root exploration, water uptake, nutrient availability and absorption, and overall plant physiology [5,6,7]. Indirect mechanisms include protection against abiotic and biotic stresses, such as biological control of phytopathogens through the activation of induced or acquired systemic resistance, as well as tolerance to extreme environmental conditions, heavy metal toxicity, and nutrient imbalances [6,8,9]. In both types of mechanisms, the ability to produce siderophores is a key characteristic of many effective PGPB and PGPF strains, that can be used for the development of biofertilizers or biofungicides [10,11].
Metal ions such as copper, iron, manganese, molybdenum, nickel, vanadium, and zinc are essential trace elements required for the enzymatic functioning of microorganisms and plants [12]. However, these metals are often present in low concentrations in the soil and may be insoluble or biologically unavailable. Archaea, bacteria, fungi, and plants have evolved mechanisms to scavenge these metals by secreting chelating agents called metallophores, which sequester metal ions from other chemical species generally from the soil [13,14]. Once transported into cells, metal ions are released from metallophores, becoming bioavailable. Metallophores are low-molecular-mass organic molecules that facilitate the complexation and mobilization of metal ions through ligand reactions [15,16]. Their biosynthesis and exudation depend on the organism’s internal metal ion concentrations [17]. Siderophores, the best-known and most extensively studied metallophores, are primarily involved in iron acquisition. These compounds, consisting of approximately 500 different chelators ranging from 400 to 1500 Da, exhibit a high affinity for Fe3+ but can also chelate other metallic micronutrients or toxic ions, forming stable complexes with elements such as aluminum, cadmium, copper, gallium, indium, lead, and zinc [18,19,20,21]. The best-known siderophores are pyoverdine, pyochelin, and pyocyanin from Pseudomonas strains, as well as enterobactin from Escherichia coli and Salmonella typhimurium [22,23,24]. Siderophores reported from fungi include dimerum acid, fusigen, coprogen, and ferricrocin from Trichoderma spp. [25]; ferrirubin from Paecilomyces variotii [26]; ferrichrome from Ustilago sphaerogena [27]; and fusarinine C from Fusarium roseum [27]. Other metallophores include staphylopine, which uptakes a broad spectrum of metals [28], pseudopaline, specific for nickel and zinc [29], and yersiniabactin, which chelates iron and nickel [30]. Recent research has also begun to explore metallophores such as chalkophores (copper metallophores), molybdophores, vanadhophores, and zincophores, although these remain less studied [31,32,33]. Additionally, metallophores specific to cobalt, gold, and lanthanides have been reported [34,35,36].
Microbial metallophores play a crucial role in microbe–microbe and plant–microbe interactions, directly and indirectly contributing to plant-growth-promoting mechanisms. Consequently, these molecules hold great potential for sustainable agriculture. Despite extensive studies on microbial siderophores involved in Fe3+ chelation, relatively little is known about the role of metallophores in chelating additional micronutrients or heavy metals [37,38,39]. To explore the full potential of metallophores in sustainable agricultural practices, further research is needed to investigate their diversity, mechanisms of action, and practical applications. This review provides a comprehensive overview of metallophore types, their roles in plant–microbe interactions, and their potential contributions to sustainable agriculture.

2. Biosynthesis and Transcription of Metallophores

Under conditions of metal scarcity, metallophore production is upregulated to facilitate the scavenging of essential metal ions; but environmental factors such as pH, temperature, oxygen availability, and the presence of competing ligands can influence metallophore synthesis. These factors regulate transcription factors, which bind to specific DNA sequences to activate metallophore biosynthesis genes. Complex regulatory networks involving transcription factors, small RNAs, and signaling molecules further fine-tune metallophore production in response to cellular needs. For instance, siderophore production is often controlled by the ferric uptake regulator (Fur), which represses iron acquisition genes in iron-replete conditions [40,41]. Several well-characterized gene clusters are involved in metallophore biosynthesis. For example, the ent genes in Escherichia coli encode enzymes for enterobactin production [42], while the pvd genes are responsible for pyoverdine biosynthesis [43]. Similarly, the dhb genes are involved in bacillibactin production [44]. These gene clusters contain conserved domains, such as 2,3-dihydro-2,3-dihydroxybenzoate dehydrogenase and the TonB-dependent siderophore receptor domain. Moreover, in fungi, the siderophore operon includes sidA, sidC, sidD, sidG, sidF, and sidL gene, but the locations of these genes are different on the fungal chromosomes [45].
Metallophores are synthesized by an intricate enzymatic machinery, primarily involving nonribosomal peptide synthetases (NRPSs) and polyketide synthases (PKSs). Siderophores synthesized via NPRS/PKS include enterobactin [46], and vibriobactin [47]. Additionally, some siderophores are assembled from dicarboxylic acid subunits and dinucleophile linkers by the action of a conserved family of adenylation/condensation enzymes distinct from NRPS and PKS [48,49,50]. These NRPS-independent synthases (NISs) catalyze the biosynthesis of achromobactin [51,52], and staphyloferrin A and B [53,54]. All these modular enzymes catalyze a series of reactions, including substrate activation, peptide bond formation, and side-chain modification, to assemble amino acids and other building blocks into diverse metallophore structures. Non-standard amino acids and post-translational modifications, such as methylation, hydroxylation, and cyclization, further diversify the metallophore chemistry. Additionally, even though chalkophore biosynthesis has been linked to several NRPS gene clusters [55,56], methanobactins are produced as ribosomally synthesized and post-translationally modified peptides (RiPPs) [32,57].
Metallophore biosynthetic pathways have evolved through a gradual and complex process [58]. These systems are thought to have originated from pre-existing enzymes involved in peptide and polyketide synthesis, and gene duplication, followed by mutations and recombination events, has facilitated the diversification of these pathways and the emergence of metallophore synthesis. Horizontal gene transfer has also contributed significantly to the evolution of metallophores [58,59,60].

3. Mechanisms of Excretion, Action, and Transport of Microbial Metallophores

Microorganisms employ a variety of strategies to transport metallophores out of the cell (Figure 1). These compounds can be actively effluxed across the cell membrane via ATP-binding cassette (ABC) transporters or secreted through specialized secretion systems such as the T1SS, T2SS, T3SS, and T4SS [61,62]. Alternatively, metallophores may be passively released through porins or channels; some microorganisms package metallophores into membrane-bound vesicles to facilitate targeted release and protect them from degradation, while others release them through cell lysis [63]. Recently, a multidrug and toxic compound extrusion (MATE) transporter, MbnM, has been associated to be in charge of the excretion of some methanobactins [64].
Once secreted, metallophores bind tightly to metal ions in the environment, forming highly stable complexes that significantly enhance the solubility and bioavailability of the metal ion [65]. This chelation process is mediated by various functional groups, including hydroxamates, catecholates, carboxylates, amines, imidazoles, thiols, phosphonates, and others [66,67]. Metal–metallophore complexes, in combination with dissolved organic matter, contribute to a ligand-sphere that enriches the local environment with accessible metal complexes, promoting microbial growth and ecological functions [68,69]. After the metal–metallophore complex is formed, it is transported back into the cell (Figure 1).
The metal–metallophore complex can enter the cell through various mechanisms. The complex can directly cross the cell membrane via specialized transporters, such as those in the major facilitator superfamily (MFS) or the ATP-binding cassette (ABC) transporter family [61,67]. Moreover, the metal–metallophore complex may bind to specific receptors on the cell surface, triggering endocytosis [45]. Other systems, including those from the Resistance–Nodulation–Division (RND) superfamily and the Tripartite ATP-Independent Periplasmic (TRAP) transporter family, may also facilitate the uptake of metal–metallophore complexes, depending on the microorganism and the metal ion being transported [61,70,71]. Once inside the cell, the metal ion is released from the metallophore and becomes available for essential cellular processes, including enzymatic catalysis, electron transfer, and biosynthesis (Figure 1).
Siderophores exemplify the sophisticated strategies microorganisms use to acquire essential metals. They are typically secreted by efflux pumps of the major facilitator superfamily (MFS), the resistance, nodulation, and cell division (RND) superfamily, and the ATP-binding cassette (ABC) superfamily [70,72]. Siderophores contain functional groups such as hydroxamates, phenol-catecholates, carboxylates, and mixed groups, which form high-affinity complexes with ferric iron (Fe3+). The coordination of iron to these functional groups varies based on the siderophore structure: hydroxamates generally form hexadentate complexes, while catecholates often form octahedral complexes [73]. This structural diversity allows siderophores to bind iron effectively under a range of environmental conditions.
Once formed, the siderophore–iron complex is recognized by specific receptors on the bacterial outer membrane and internalized via active transport systems. These mechanisms involve proteins like TonB- or TonB-dependent transporters (TBDT), periplasmic siderophore-binding proteins (SBP), ABC transporters, or siderophore-mediated endocytosis [74]. While the specific genes fepA, fepB, fepC, fepD, and fepG are primarily found in Escherichia coli, other bacterial genera like Pseudomonas, Yersinia, Salmonella, and Klebsiella possess analogous genes that encode for proteins involved in the uptake of siderophores [75,76,77]. In addition, in Gram-positive bacteria, siderophore-dependent ferrichelatases have been recently proposed to act as iron shuttles [78]. Within the cell, the iron is released from the siderophore and utilized for various metabolic processes, while the siderophore may either be destroyed or recycled [79,80]. Although iron is the primary target of siderophores, they can also chelate other metal ions, including divalent cations (e.g., Zn2+, Ni2+, Cd2+, Cu2+, and Pb2+), trivalent cations (e.g., Al3+, In3+, Mn3+, V3+, and Co3+), actinides (e.g., Th⁴+, U⁴+, and Pu⁴+) [74]. This broad chelation capacity highlights the versatility of siderophores in adapting to diverse environmental metal availability [81,82,83], or maybe they correspond to metallophores that that have been misidentified and named. In the case of fungi, they produce hydroxamate-type siderophores that are grouped into the families: rhodothorulic acid, fusarinines, coprogens, and ferrichromes [20]. Little is known about the mechanism of fungal siderophore secretion [37]; however, some mechanisms proposed for the iron-loaded siderophore uptake depend on the specific recognition of various siderophores, and Fe3+ reduction occurs at the membrane and Fe2+ is then absorbed into the cell [84]. For example, in fungi such as Rhodotorula spp., Fe3+ transfer across the membrane from the siderophore rhodotorulic acid to intracellular ligands occurs without the entry of rhodotorulic acid into the cell [85]. Meanwhile, in Ustilago sphaerogena, the iron-loaded siderophore ferrichrome A is not transported into the cell; instead, Fe3+ reduction occurs at the membrane, and Fe2+ is then absorbed into the cell [84].

4. Metallophore Detection, Extraction, Purification, and Quantification

Metallophore detection is crucial for understanding their biological roles in agriculture, mainly in complex matrices such as soil. Chromogenic assays, such as the CAS assay, apply the ability of metallophores to form colored complexes with specific metal ions, providing a direct method for their detection. The CAS assay has undergone various modifications, including changes to the metal ion, surfactant, and dye, as well as the addition of adsorbent resins to facilitate the growth of fungi and yeast [86,87,88,89]. Additionally, the assay has been adapted to use a concentrated CAS reagent and different growth media for broader applications [90,91]. Furthermore, the technique has been implemented in microplates to enable both a qualitative and quantitative estimation of siderophore production [92,93]. Other colorimetric methods can be used to detect different types of siderophores, such as the Arnow’s and FeCl3 tests for catecholates; the ferric perchlorate assay, and tetrazolium Csaky’s tests for hydroxamates, and the Vogel’s chemical test for carboxylates [82], whereas fluorescent siderophores, like pyoverdines, can be detected using fluorescence-based methods [94]. Complementary to these assays, spectroscopic techniques like ultraviolet–visible spectroscopy (UV–Vis) and nuclear magnetic resonance (NMR) offer detailed structural information about metallophores. Moreover, bioassays using mutant strains that are deficient in metal uptake are used to identify the function, biosynthesis, and regulation of metallophores [93]; for instance, the mutant strains ΔpvdA and ΔpvdE of P. fluorescens PF08 indicated the participation of pyoverdine in biofilm formation and motility, and in the regulation of flagellum and stress resistance of the strains [95].
The in-depth characterization of metallophores and methallophore–metal complexes requires their isolation and purification. For example, siderophore extraction into organic solvents, such as ethyl acetate for the catechol type or either benzyl alcohol or chloroform:phenol (1:1) for the hydroxamate types, have been used as an effective purification step for many types of siderophores since both salts and macromolecules are removed effectively [96]. Solid-phase extraction (SPE) offers a more selective approach, using solid sorbents to retain metallophores. Polystyrene resins like XAD-2 or Amberlite XAD-7 can be used to adsorb metallophores from complex mixtures. For instance, the resin amberlite XAD has been widely used for the purification of the neutral, ferrichrome type of siderophores and polyamide resin for the chromatographic separation of the catechol type [97]. Siderophore-rich cell-free concentrated broth from Alcaligenes faecalis BCCM ID 2374 was purified on the Amberlite XAD-4 column; it resulted in the separation of two fractions having absorption maxima at 264 and 224 nm, that belonged to the hydroxamate and catecholate type of siderophores, respectively [96]. With the cultivation of Gordonia rubripertincta CWB2 on succinate and extraction using an adsorbing mixture of XAD4 and XAD16 resins, 28 desferrioxamine-like compounds were purified; five of these were identified as desferrioxamines B, E, and G1, and bisucaberin and bisucaberin B, and the siderophore function of the remaining compounds remains to be elucidated [98]. Moreover, Fourier-transform infrared analysis (FT-IR) has been used to determine the functional groups in a siderophore and the functional groups involved in the interaction of the siderophore with the metal [99]; for instance, by the FT-IR spectra, Mazari et al. [100] characterized the pyoverdine of the PGPR strains Pseudomonas atacamensis PO22, Pseudomonas lactis R2P30, and Pseudomonas sp.PS11, isolated from roots of grapevine. Recently, ionization and analytical techniques, such as electrospray ionization mass spectrometry, matrix-assisted laser desorption/ionization mass spectrometry, inductively coupled plasma mass spectrometry, and mass spectrometry imaging have been used to identify and characterize siderophores [82].
Subsequent purification steps, such as high-performance liquid chromatography (HPLC), electrophoresis, and mass spectrometry and associated technologies such as liquid chromatography–mass spectrometry and gas chromatography–mass spectrometry, further separate and characterize metallophores based on their physicochemical properties. A reversed-phase HPLC method can screen for siderophore-producing strains, and can be used to control their purification and immobilization, as demonstrated in Gordonia rubripertincta CWB2, which produces desferrioxamine-like siderophores [101]. Moreover, crystallization yields pure crystals suitable for X-ray crystallography, providing a high-resolution structure of the metallophore. The zincophores isoavenaciol and 7-hydroxy-isoavenaciol produced by the root-endophyte Pezicula ericae w12-25 of the zinc-accumulating plant Aucuba japonica were determined to be (3aS,4S,6aR)-3ahydroxy-3-methylene-4-octyldihydrofuro[3,4-b]furan-2,6(3H,4H)-dione and (3S,3aS,4S,6aR)-3a-hydroxy-3-(hydroxymethyl)-4-octyldihydrofuro[3,4-b]furan-2,6(3H,4H)-dione using HRMS, 1H and 13C NMR, and 2D NMR spectroscopic methods and X-ray crystallography [102]. Recently, immobilized metal affinity chromatography, metal oxide affinity chromatography, and titanium dioxide affinity chromatography are new and improved methodologies that have been used for the purification of siderophores [82]. Additionally, ultra-high-performance liquid chromatography coupled to a time-of-flight mass spectrometer (UHPLC-ToF-MS) has been developed to quantify and monitor the production of metallophores. Using this methodology, Deicke et al. [103] reported that, under a low Mo concentration, Azotobacter vinelandii OP produces peaks of Mo-protochelin concentration in the middle lag phase, and the Fe-protochelin concentration rises twice, at the beginning of the exponential phase and during the stationary phase, indicating that, in A. vinelandii OP, the production of metallophores is highly dynamic throughout growth [103].
Recent advancements in high-throughput sequencing and bioinformatic tools, such as sequence homology searches, domain architecture analyses, and comparative genomics, have significantly accelerated the discovery and prediction of metallophore biosynthesis gene clusters, and revealed that some of them are conserved across diverse microbial taxa, suggesting a widespread role of metallophores in diverse environments. For instance, the dhbACEBF cluster has been predicted in Bacillus subtilis EA-CB0575 to encode the enzymes responsible for bacillibactin production [44]; this siderophore was first reported in B. subtilis ATCC 21332 [104]. Moreover, genome mining focusing on the core operon Mbn has revealed the presence of methanobactins in a broader spectrum of bacteria beyond methanotrophs; bacterial groups such as Proteobacteria (e.g., Pseudomonas, Burkholderia, Gluconacetobacter, and Azospirillum), Actinobacteria (e.g., Streptomyces), and Firmicutes (e.g., Bacillus) have been identified as producers of these chalkophores [105,106]. Furthermore, Calderón-Celis et al. [107] developed an approach to detect metallophores using a two-dimensional solid-phase extraction–liquid chromatography–mass spectrometry (SPE-LC-MlS) coupled with the taxonomic profiling of the prokaryotic community and mining of the antiSMASH database to link metallophores with potential producers; this approach enables the direct detection and identification of metallophores in native environments, including agricultural settings. Additionally, molecular modeling techniques, including homology modeling, molecular docking, and molecular dynamics simulations, could provide atomic-level insights into metallophore structure and function [108]. For instance, docking and molecular dynamics simulations reveal that, in rice, the two yellow stripe-like (YSL) transporters responsible for carrying siderophore–Fe3+ complexes into roots show a higher binding affinity for the bacterial siderophore bacillibactin–Fe3+ complex compared to the plant siderophore 2′-deoxymugineic acid (DMA)–Fe3+ complex. This suggests that the use of siderophores produced by PGPR can enhance the iron uptake in rice, as this crop possesses receptors that also bind for bacterial siderophores [109].

5. Microbial Metallophores in Microbial Communities

Microbiomes form intricate interactions within their habitats that contribute to ecosystem balance. In this way, microbial diversity is essential for maintaining ecosystem equilibrium [110,111]. These microbial interactions are largely governed by secondary metabolites, including metallophores, which mediate cooperation, exploitation, and competition among microbial populations and communities. For example, siderophores can be acquired by cells of the same or different species possessing the matching receptors, and distributing the costs and benefits of siderophore production among individuals. However, siderophores can also induce iron starvation in cells lacking specific receptors [39,112,113]. Thus, in microbial communities, siderophores can be shared cooperatively or exploited by non-producing “cheater” cells [114,115]. The production of multiple siderophores with different binding affinities can limit the success of cheaters, leading to the stable coexistence of producers and non-producers. When metal availability is limited, the production of metallophores by microorganisms can shape the microbial community structure by promoting both cooperative and competitive interactions [107,116]. For example, the majority of fungi possess the ability to take up iron carried by xenosiderophores (siderophores synthesized by other microbial communities), and Aspergillus nidulans exhibits affinity for ferrirubin released from Aspergillus ochraceus [117] and for bacterially synthesized enterobactin [118]; and Saccharomyces cerevisiae also depends on siderophores produced by other species [119]. Furthermore, siderophores influence quorum sensing, nutrient availability, and stress responses within and between species [18,120]. For example, siderophores participate in sporulation stimulation, the induction of siderophore production in other species, protection against reactive oxygen species and heavy metals, and the promotion of growth in cohabitating organisms [18,120]. Microorganisms are social organisms that employ quorum sensing to coordinate their behavior [121]. In this context, they can produce redundant metallophores for different metal ions, such as zinc, molybdenum, and vanadium, depending on the environmental conditions [62,68]. Moreover, metallophores can contribute to biofilm formation by facilitating cell adhesion and promoting intercellular communication. In the strains P. fluorescens PF08, pyoverdine is involved in the production of exopolysaccharides, formation of biofilm and motility, and regulation of the transcription levels of htpX, phoA, flip, flgA, and rpoS, influencing the strain’s flagellar biosynthesis and resistance to stress [95].
The phyllosphere, rhizosphere, and plant-endosphere are environments shaped by complex interactions between plants and microorganisms. The structural and functional diversity of microbial communities significantly impacts plant health and ecosystem sustainability [122,123,124,125]. The rhizosphere and root-endosphere maintain a dynamic, nutrient-rich environment around and inside roots, sustaining specific bacterial populations involved in activities that ensure the stability and productivity of agroecosystems, where metallophores must play a fundamental role in the microbiome and plant–microbiome interactions [126,127,128,129,130]. Some of these microorganisms are known as PGPB and PGPF, that, due to their effective properties in promoting plant growth, specifically, increasing yield and controlling pests and diseases in agricultural crops, have been selected for the development of bioproducts. These bioproducts are considered suitable methods for introducing probiotics to agricultural soil or increasing soil microbial biodiversity by breaking the dormancy of soil microbial seed banks’ dormancy and reinforcing the biodiversity–ecosystem functioning relationship [131,132,133].

6. The Role of Microbial Metallophores in Promoting Plant Growth

Metallophores play a crucial role in the cycling of essential metals in the environment; several of these metal ions are indispensable for plant growth and development. For instance, iron is critical for chlorophyll synthesis and maintaining chloroplast structure and function [134]. Manganese plays a key role in plant metabolism and development, participating in various enzymatic processes [135]. Molybdenum is a vital component of proteins involved in nitrogen assimilation, sulfur metabolism, phytohormone biosynthesis, and stress responses [136]. Nickel is an essential component of plant ureases, which protect against phytopathogens [137]. Zinc plays an important role in seed development and chloroplast function, with numerous zinc-dependent enzymes performing essential tasks [138]. Copper is essential for photosynthesis, carbon and nitrogen metabolism, oxidative stress protection, and cell wall synthesis [139]. Recent studies have demonstrated a hormetic effect of lanthanides in plants; they stimulate chlorophyll synthesis, enhance photochemical reactions in thylakoids, and improve stress tolerance; but the precise mechanisms underlying these hormetic effects remain largely unknown, and gaps in this area will be filled in future studies [140]. In environments with limited metal ion availability or under stress conditions, metallophores become crucial for bacterial communities to acquire and supply essential metals to plants [18]. For instance, siderophore-producing strains such as Pseudomonas sp. GRP3A and PRS have been shown to increase germination rates and promote plant growth in maize [141]. Similarly, the siderophore-producing strain Pseudomonas aerugionsa RSP5 enhances iron transport to maize plants, leading to increased shoot and root length, larger cobs, higher grain yields, and a greater iron content in plant tissues [142]. Siderophores also influence plant hormone levels, with siderophore-producing strains like Trichoderma asperellum Q1 supporting optimal root system development in Arabidopsis thaliana by increasing auxin concentrations [143]. Siderophores of Streptomyces sp. CoT10, an endophyte of Camellia oleifera, directly mediate Fe–P mobilization; this enhances P and Fe acquisition in C. oleifera and improves microbial interactions within the plant rhizosphere [144] (Figure 2). It has been proposed that plants acquire xenometallophores produced by PGPR through the transporters employed for the acquisition of phytometallophores; for example, bacillibactin produced by the PGPR Bacillus subtilis is possibly transported by the YSL15 transporter and siderophore import proteins in Oryza sativa plants, and YSL15 transports PS–Fe3+ complexes from the rhizosphere to the root and from the root to flowering regions [145].
Furthermore, bacterial siderophores can significantly impact plant gene expression. In A. thaliana, the apo-pyoverdine produced by Pseudomonas fluorescens C7R12 upregulates genes related to development and iron acquisition while downregulating defense-related genes [79]. Siderophores also play a crucial role in the fitness of phyllospheric bacteria, as demonstrated by their role in promoting the population growth of biocontrol agents like Pseudomonas syringae pv. syringae 22d/93 against the plant pathogen Pseudomonas syringae pv. glycinea [146]. Additionally, lanthanophores produced by the bacterial community of crop phyllosphere can enhance growth, crop yields, and drought resistance [34]. For instance, genes involved in metallophore–metal uptake support an increased AMF colonization of the arbuscular mycorrhizal fungi Rhizophagus irregularis [147]. Furthermore, as bacterial communities provide nitrogen to plants, they have high nutritional requirements for molybdenum, iron, and vanadium [19,148]. The free-living diazotroph Azotobacter vinelandii, a well-studied nitrogen-fixing bacteria, produces metallophores like protochelin, azotochelin, azotobactin, aminochelin, and vibrioferrin efficiently to the three metal ions under low-concentration conditions [31,148,149,150,151,152]. Additionally, siderophores produced by the fungi Aspergillus niger, Penicillium citrinum, and Trichoderma harzianum increase the shoot and root lengths of chickpeas [153]; and ectomycorrhizal fungi, such as Aspergillus niger, Penicillium citrinum, and Trichoderma harzianum, supply iron to the host roots of plants [154].
Microbial metallophores also play a critical role in suppressing plant diseases. In the rhizosphere, microbial communities compete with plant pathogens for space and access to metal ion pools. Moreover, siderophores contribute to disease suppression through various mechanisms, including the production of antibiotic compounds, the synthesis of fungal cell wall-lytic enzymes, and the induction of systemic acquired resistance [18,155,156,157,158,159,160]. For example, the siderophore enantio-pyochelin produced by Pseudomonas protegens CS1 exhibits antagonistic activity against Xanthomonas citri subsp. citri by generating reactive oxygen species (ROS) from pyochelin. Similarly, the enantio-pyochelin produced by Pseudomonas fluorescens CHA0 provides protection against Tobacco necrosis virus infection in tobacco, and the pseudobactin produced by P. fluorescens WCS374 provides resistance to Fusarium wilt in radish [161,162]. Pyoverdine and enantio-pyochelin produced by P. protegens are also responsible for its resistance to fusaric acid [163]. Siderophore-producing strains, such as Bacillus subtilis MF497, Pseudomonas koreensis MG20938, and Pseudomonas taiwanensis MTCC11631, positively affect maize (Zea mays L.) and anthurium (Anthurium andreanum L.) by inducing a priming effect. They activate plant defense enzymes, such as polyphenol oxidase (PPO), phenylalanine ammonia-lyase (PAL), and peroxidases (POX), which are associated with lignification processes that strengthen cell walls and reduce the incidence and severity of diseases caused by Cephalosporium maydis and Xanthomonas axonopodis pv. dieffenbachiae, respectively [164,165].
Furthermore, certain siderophore-producing strains improve plant growth and development while also playing a crucial role in biological control. For example, Pseudomonas aeruginosa P4 promotes the growth, nutrition, and health of peanut plants (Arachis hypogaea L.) through siderophore production, organic acid secretion, iron and phosphate supply, and the production of secondary metabolites [166]. The siderophores of Rhizobium meliloti function as biocontrol agents against Macrophomina phaseolina in Mucuna pruriens while simultaneously increasing the seedling biomass and fresh nodule formation [167]. Additionally, siderophores produced by Streptomyces sp. AzR-051 and Bacillus velezensis NKG-2 promote plant growth and antagonizes major fungal pathogens in tomato plants (Solanum lycopersicum L.) [168,169]. Additionally, the plant-growth-promoting fungus Aureobasidium pullulans L1 produces siderophores to improve the iron availability in soils with a low iron content and acts as an antagonist to the soil-borne pathogen Rhizoctonia solani through iron competition [170].
Additionally, phytometallophores are also involved in the uptake of Fe, Zn, Cu, Mn, Ni, Cd, and Co. Specifically, phytosiderophores, including 2′-deoxymugineic acid (DMA) and eight different compounds from the mugineic acid (MA) family, are secreted by the root cells of cereals via MA transporters. This constitutes the chelation strategy (Strategy II) for Fe uptake in gramineous plants [171,172,173].

7. Applications of Microbial Metallophores in Agroecosystems

In the context of agriculture, microbial metallophores play a crucial role in promoting plant health and productivity (Figure 2). Several biofertilizers, biostimulants, and soil inoculants incorporate metallophore-producing microbial strains that improve crop yields. Microbial metallophores enhance the availability of essential nutrients to plants; these molecules reduce the need for synthetic fertilizers and improve nutrient availability. At the same time, microbial metallophores stimulate the growth of beneficial soil microorganisms that improve soil structure and fertility, and contribute to enhanced soil health. The use of metallophores in nutrient cycling can develop targeted fertilization strategies, reduce fertilizer inputs, and minimize environmental impacts. For instance, siderophore-producing strains show promise as auxiliaries in specific production systems where iron availability is limited, such as hydroponics. For example, in hydroponically grown strawberry (Fragaria ananassa cv. Pajaro), a significant increase in root area, plant growth, and iron content in its tissues was observed. This effect was greater in plants inoculated with the hydroximate-siderophore-producing Glauconoacetobacter diazotrophicus PAL5 than in those inoculated with the catechol-siderophore-producing Azospirillum brasilense REC3, likely due to the hydrophilic nature of hydroximate siderophore [174]. Moreover, nocardicin A has the ability to improve nitrogen fixation in legumes like alfalfa by increasing the availability of cobalt for nitrogen-fixing bacteria [175]. In this way, metallophores enhance metal homeostasis mechanisms, making metal ions available to plants and favoring their growth [107,176].
Microbial metallophores also offer a sustainable and environmentally friendly approach to mitigating heavy metal contamination in agricultural soils (Figure 3). For instance, the production of siderophores with diverse chemical properties enables microorganisms to sequester iron under different environmental conditions or perform complementary functions, making them ideal for ecosystem bioremediation; moreover, by reducing metal stress, metallophores indirectly protect crops from both biotic and abiotic stresses [82,177,178,179]. Thus, metallophores contribute to bioremediation efforts by cleaning up contaminated soils and waterways, and enhancing phytoremediation processes. Microbial metallophores are effective because, generally, they are stable in soil environments and resistant to degradation; they form stable complexes with heavy metal ions that remain in the soil, or bind to the cell surfaces of microorganisms or plant roots [73]. In this way, microbial metallophores immobilize heavy metals and prevent their uptake, protecting plants from oxidative stress, growth inhibition, and other harmful effects [67]. The mechanisms through which metallophores mitigate metal toxicity induce chelation (forming stable metal–metallophore complexes), biosorption (binding to cell surfaces), and bioaccumulation (intracellular accumulation of metal–metallophore complexes) [82]. For instance, siderophores produced from Pseudomonas putida KNP9 in the presence of Cd increase the biomass of Phaseolus vulgaris without any toxicity symptoms [180]. Moreover, siderophore-containing culture filtrate from Streptomyces acidiscabies E13 supplied cowpea plants with Fe in the presence of high levels of Al, Cu, Mn, Ni, and U, metals that inhibit Fe acquisition; this led to an increased chlorophyll content and reduced the formation of free radicals, thereby protecting microbial auxins from degradation and enabling them to enhance plant growth [181]. As well, piochelin remediate soils contaminated with copper and nickel, demonstrating a reduction in the bioaccumulation of these metals in plants such as mustard [73], and acinetoferrin enhance the tolerance of crops like rice to copper toxicity in saline soils [182].
Additionally, microbial metallophores produced by the microbial community associated with hyperaccumulator plants enable those plants to tolerate higher concentrations of metals in their tissues. For example, zincophores produced by Pseudomonas monteilii, Microbacterium saperdae, and Enterobacter cancerogenes enhanced four-fold the accumulation of Zn in the Zn-hyperaccumulate plant Thlaspi caerulescens [183]. However, chalkophores such as methanobactins can negatively impact agricultural production by affecting denitrification, specifically at the step of nitrous oxide (N2O) reduction and possibly also at nitrite reduction [184,185]. The enzyme N2O reductase (NosZ) is copper-dependent [186], and, in soil, methanobactin-producing bacteria can inhibit nosZ expression in the denitrifying microbial community [187,188], leading to increased N2O emissions, as reported in rice paddies and forest soils. This process can intensify the impacts of climate change on agriculture, contribute to nitrogen loss from the soil, and promote the use of nitrogen fertilizers [189]. Furthermore, the use of siderophore-producing microbial strains can reduce Fe deficiency and plant stress under saline soil conditions. For instance, Bacillus aryabhattai MS3 showed the maximum siderophore production at 200 mM NaCl, due to the activation of the entD gene (gene in charge of siderophore biosynthesis) by salinity [190]; and the strain Streptomyces sp. C increased siderophore production at the high concentration of 300 mM of NaCl, increasing the iron content in the shoot of wheat plants grown in saline soil [191].
In addition to the rise in the use of microbial strains that produce metallophores in agricultural production, the potential use of metallophores in the industry, medicine, and environmental field should be reflected in the agroecosystem stability (Figure 3). For example, in the medical sector, siderophores have been used as a “Trojan horse” to simplify the management of multi-resistant strains, prevent the emergence of new antibiotic resistance, and reduce ecological and environmental impacts [192,193,194]. The Trojan horse strategy consists of metalloantibiotics, metal complexes with antimicrobial activity that depend on deceiving the target bacteria into actively internalizing a metallophore–antibiotic conjugate, but more advanced antibiotics synthesis methods based in the use of metallophores are needed; these methods could involve transporting the metallophore outside the cell, capturing and blocking the free metallophore or the metallophore–metal complex outside the bacterial cell, or taking up metal ions loaded onto metallophores [195]. In the industrial sector, advancements in metallophore research can significantly impact agricultural practices. Bioremediation, utilizing metallophores to remove heavy metals from industrial waste and contaminated sites, is crucial for agriculture as it prevents the contamination of agricultural lands and water sources, ensuring the long-term health and productivity of agricultural ecosystems [73]. The chelation efficacy of Pseudomonas lactis and Pseudomonas atacamensis pyoverdine is comparable to that of EDTA and citric acid, emphasizing its viability as an eco-friendly alternative for the remediation of metal-contaminated soils, which can be used in environmentally sustainable soil decontamination strategies [100]. In the remotion of harmful heavy metals from water or wastewater, metallophores can be used as an alternative treatment technique; polymeric adsorption resins can be modified with chemically or naturally synthesized metallophores such as staphylopine, pseudopaline, acinetobactin, or yersiniabactin to remove iron, copper, nickel, or a broader range of different metal ions [196]. Furthermore, resource recovery processes facilitated by metallophores can efficiently extract valuable metals from industrial waste streams. This reduces the need for mining and minimizes the environmental impact associated with resource extraction, thereby promoting sustainable agricultural practices [67]. Additionally, siderophores from fungi such as Gloeophyllum trabeum and Coriolus versicolor have been used during the bleaching process in paper and pulp industries as biobleaching agents to depolymerize the cellulose, lignocellulose, and hemicellulose in wood pulp [197]. In this way, industrial applications of metallophore research can significantly contribute to the stability and long-term sustainability of agroecosystems.

8. The Future of Metallophores

The interest in finding molecules capable of processing metals for use in various scientific and technological areas is not recent; the first siderophores were described in the 1950s [198]. Currently, research directed toward the study of metallophores is focused mainly on the medical, environmental, and industrial fields [199,200,201]; even so, the use of siderophores in human health is limited to few cases, such as desferrioxamine B (Desferal) applied in the treatment of chronic iron overload [202] and the antibiotic cefiderocol [203]. In agriculture, it is necessary that we encourage the evaluation of plant-growth-promoting microbial strains on their ability to produce metallophores, not only of iron, but also of metallophores of the different metals important for plants, as well as include in labels or the promotion of microbial strains their ability to produce these different metallophores.
Several challenges need to be addressed to enable the widespread application of microbial metallophores (Figure 3). Advanced omics techniques, such as genomics, transcriptomics, and proteomics, can be used to identify novel metallophores and elucidate their regulatory mechanisms. These efforts include finding metallophores with a high specificity for target metals, and scaling up their production to meet agricultural demands. Furthermore, the characterization of the membrane receptors involved in metal ion uptake opens a wide opportunity to improve current applications and develop new ones. Examples of emergent areas include their use as biosensors, diagnostic molecules, and vaccines, and in the production of safer and healthier food, thus contributing to a more just and sustainable society [204].

9. Conclusions

Metallophores represent a promising class of molecules with significant potential for addressing global challenges in agriculture sustainability. Their ability to chelate metals and influence microbial interactions offers a wide range of applications, including nutrient management, plant disease control, and soil remediation. Future research should focus on searching new metallophores, and elucidating the molecular mechanisms underlying metallophore biosynthesis, transport, and regulation, expanding our knowledge of their diversity and distribution in microbial communities; as well as developing novel biotechnological applications to support a more sustainable and resilient agricultural systems, while mitigating the negative impacts of environmental pollution.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Anjaria, P.; Vaghela, S. Toxicity of agrochemicals: Impact on environment and human health. J. Toxicol. Stud. 2024, 2, 250. [Google Scholar] [CrossRef]
  2. Pedraza, R.O.; Teixeira, K.R.S.; Fernández-Scavino, A.; García-de Salamone, I.; Baca, B.E.; Azcón, R.; Baldani, V.L.D.; Bonilla, R. Microorganisms that enhance plant growth and soil quality. Review. Corpoica Cienc. Tecnol. Agropecu. 2010, 11, 155–164. [Google Scholar] [CrossRef]
  3. Turner, T.R.; James, E.K.; Poole, P.S. The plant microbiome. Genome Biol. 2013, 14, 209. [Google Scholar] [CrossRef]
  4. Woźniak, M.; Gałązka, A. The rhizosphere microbiome and its beneficial effects on plants–current knowledge and perspectives. Post. Microbiol. 2019, 58, 59–69. [Google Scholar] [CrossRef]
  5. Kumar, A.; Bahadur, I.; Maurya, B.R.; Raghuwanshi, R.; Meena, V.S.; Singh, D.K.; Dixit, J. Does a plant growth-promoting rhizobacteria enhance agricultural sustainability? J. Pure Appl. Microbiol. 2015, 9, 715–724. [Google Scholar]
  6. Gouda, S.; Kerry, R.G.; Das, G.; Paramithiotis, S.; Shin, H.S.; Patra, J.K. Revitalization of plant growth promoting rhizobacteria for sustainable development in agriculture. Microbiol. Res. 2018, 206, 131–140. [Google Scholar] [CrossRef]
  7. Moreno-Reséndez, A.; Carda-Mendoza, V.; Reyes Carrillo, J.L.; Vásquez Arroyo, J.; Cano Ríos, P. Plant growth promoting rhizobacteria: A biofertilization alternative for sustainable agriculture. Rev. Colomb. Biotecnol. 2018, 20, 68–83. [Google Scholar] [CrossRef]
  8. Shameer, S.; Prasad, T.N.V.K.V. Plant growth promoting rhizobacteria for sustainable agricultural practices with special reference to biotic and abiotic stresses. Plant Growth Regul. 2018, 84, 603–615. [Google Scholar] [CrossRef]
  9. Enebe, M.C.; Babalola, O.O. The influence of plant growth-promoting rhizobacteria in plant tolerance to abiotic stress: A survival strategy. Appl. Microbiol. Biotechnol. 2018, 102, 7821–7835. [Google Scholar] [CrossRef]
  10. Verma, P.P.; Shelake, R.M.; Das, S.; Sharma, P.; Kim, J.Y. Plant growth-promoting rhizobacteria (PGPR) and fungi (PGPF): Potential biological control agents of diseases and pests. In Microbial Interventions in Agriculture and Environment; Singh, D.P., Kumar-Gupta, V., Prabha, R., Eds.; Springer Nature: Singapore, 2019; pp. 281–311. [Google Scholar]
  11. El-Maraghy, S.S.; Tohamy, A.T.; Hussein, K.A. Plant protection properties of the Plant Growth-Promoting Fungi (PGPF): Mechanisms and potentiality. Curr. Res. Environ. Appl. Mycol. 2021, 11, 391–415. [Google Scholar] [CrossRef]
  12. Johnston, A.E. Trace elements in soil: Status and management. In Essential Trace Elements for Plants, Animals and Humans; Thorvaldsson, G., Jónsdóttir, R.S., Eds.; Nordic Association of Agricultural Scientists: Reykjavík, Iceland, 2005; pp. 7–14. [Google Scholar]
  13. Callahan, D.L.; Baker, A.J.; Kolev, S.D.; Wedd, A.G. Metal ion ligands in hyperaccumulating plants. J. Biol. Inorg. Chem. 2006, 11, 2–12. [Google Scholar] [CrossRef] [PubMed]
  14. Hänsch, R.; Mendel, R.R. Physiological functions of mineral micronutrients (Cu, Zn, Mn, Fe, Ni, Mo, B, Cl). Curr. Opin. Plant Biol. 2009, 12, 259–266. [Google Scholar] [CrossRef] [PubMed]
  15. Pesch, M.L.; Hoffmann, M.; Christl, I.; Kraemer, S.M.; Kretzschmar, R. Competitive ligand exchange between Cu-humic acid complexes and methanobactin. Geobiology 2013, 11, 44–54. [Google Scholar] [CrossRef]
  16. Kraemer, S.M. Iron oxide dissolution and solubility in the presence of siderophores. Aquat. Sci. 2004, 66, 3–18. [Google Scholar] [CrossRef]
  17. Ghssein, G.; Matar, S.F. Chelating mechanisms of transition metals by bacterial metallophores “pseudopaline and staphylopine”: A quantum chemical assessment. Computation 2018, 6, 56. [Google Scholar] [CrossRef]
  18. Kramer, J.; Özkaya, Ö.; Kümmerli, R. Bacterial siderophores in community and host interactions. Nat. Rev. Microbiol. 2020, 18, 152–163. [Google Scholar] [CrossRef]
  19. Kraepiel, A.M.L.; Bellenger, J.P.; Wichard, T.; Morel, F.M.M. Multiples roles of siderophores in free-living nitrogen-fixing bacteria. Biometals 2009, 22, 573–581. [Google Scholar] [CrossRef]
  20. Schalk, I.J.; Hannauer, M.; Braud, A. New roles for bacterial siderophores in metal transport and tolerance. Environ. Microbiol. 2011, 13, 2844–2854. [Google Scholar] [CrossRef]
  21. Yu, S.; Teng, C.; Bai, X.; Liang, J.; Song, T.; Dong, L.; Jin, Y.; Qu, J. Optimization of siderophore production by Bacillus sp. PZ-1 and its potential enhancement of phytoextration of Pb from soil. J. Microbiol. Biotechnol. 2017, 27, 1500–1512. [Google Scholar] [CrossRef]
  22. Ghssein, G.; Ezzeddine, Z. A review of Pseudomonas aeruginosa metallophores: Pyoverdine, pyochelin and pseudopaline. Biology 2022, 11, 1711. [Google Scholar] [CrossRef]
  23. Gehring, A.M.; Bradley, K.A.; Walsh, C.T. Enterobactin biosynthesis in Escherichia coli: Isochorismate lyase (EntB) is a bifunctional enzyme that is phosphopantetheinylated by EntD and then acylated by EntE using ATP and 2, 3-dihydroxybenzoate. Biochemistry 1997, 36, 8495–8503. [Google Scholar] [CrossRef] [PubMed]
  24. Mohsen, Y.; Tarchichi, N.; Barakat, R.; Kawtharani, I.; Ghandour, R.; Ezzeddine, Z.; Ghssein, G. The different types of metallophores produced by Salmonella enterica: A review. Microbiol. Res. 2023, 14, 1457–1469. [Google Scholar] [CrossRef]
  25. Lehner, S.M.; Atanasova, L.; Neumann, N.K.; Krska, R.; Lemmens, M.; Druzhinina, I.S.; Schuhmacher, R. Isotope-assisted screening for iron-containing metabolites reveals a high degree of diversity among known and unknown siderophores produced by Trichoderma spp. Appl. Environ. Microbiol. 2013, 79, 18–31. [Google Scholar] [CrossRef]
  26. Holinsworth, B.; Martin, J.D. Siderophore production by marine derived fungi. BioMetals 2009, 22, 625–632. [Google Scholar] [CrossRef]
  27. Saharan, B.S.; Nehra, V. Plant growth promoting rhizobacteria: A critical review. Life Sci. Med. Res. 2011, 21, 1–30. [Google Scholar]
  28. Ghssein, G.; Brutesco, C.; Ouerdane, L.; Fojcik, C.; Izaute, A.; Wang, S.; Hajjar, C.; Lobinski, R.; Lemaire, D.; Richaud, P.; et al. Biosynthesis of a broad-spectrum nicotianamine-like metallophore in Staphylococcus aureus. Science 2016, 352, 1105–1109. [Google Scholar] [CrossRef]
  29. Lhospice, S.; Gomez, N.O.; Ouerdane, L.; Brutesco, C.; Ghssein, G.; Hajjar, C.; Liratni, A.; Wang, S.; Richaud, P.; Bleves, S.; et al. Pseudomonas aeruginosa zinc uptake in chelating environment is primarily mediated by the metallophore pseudopaline. Sci. Rep. 2017, 7, 17132. [Google Scholar] [CrossRef]
  30. Robinson, A.E.; Lowe, J.E.; Koh, E.I.; Henderson, J.P. Uropathogenic enterobacteria use the yersiniabactin metallophore system to acquire nickel. J. Biol. Chem. 2018, 293, 14953–14961. [Google Scholar] [CrossRef]
  31. Bellenger, J.P.; Wichard, T.; Kraepiel, A.M.L. Vanadium requirements and uptake kinetics in the dinitrogen-fixing bacterium Azotobacter vinelandii. Appl. Environ. Microbiol. 2008, 74, 1478–1484. [Google Scholar] [CrossRef]
  32. Kenney, G.E.; Rosenzweig, A.C. Chalkophores. Annu. Rev. Biochem. 2018, 87, 645–676. [Google Scholar] [CrossRef]
  33. Morey, J.R.; Kehl-Fie, T.E. Bioinformatic mapping of opine-like zincophore biosynthesis in bacteria. Msystems 2020, 5, e00554-20. [Google Scholar] [CrossRef]
  34. Cotruvo, J.A., Jr. The chemistry of lanthanides in biology: Recent discoveries, emerging principles, and technological applications. ACS Cent. Sci. 2019, 5, 1496–1506. [Google Scholar] [CrossRef]
  35. Johnston, C.W.; Wyatt, M.A.; Li, X.; Ibrahim, A.; Shuster, J.; Southam, G.; Magarvey, N.A. Gold biomineralization by a metallophore from a gold-associated microbe. Nat. Chem. Biol. 2013, 9, 241–243. [Google Scholar] [CrossRef]
  36. Saito, M.A.; Goepfert, T.J.; Ritt, J.T. Some thoughts on the concept of colimitation: Three definitions and the importance of bioavailability. Limnol. Oceanogr. 2008, 53, 276–290. [Google Scholar] [CrossRef]
  37. Haas, H.; Eisendle, M.; Turgeon, B.G. Siderophores in fungal physiology and virulence. Annu. Rev. Phytopathol. 2008, 46, 149–187. [Google Scholar] [CrossRef]
  38. Swinburne, T.R. Iron, Siderophores, and Plant Diseases; Springer: New York, NY, USA, 2012; p. 350. [Google Scholar]
  39. Butaitė, E.; Baumgartner, M.; Wyder, S.; Kümmerli, R. Siderophore cheating and cheating resistance shape competition for iron in soil and freshwater Pseudomonas communities. Nat. Commun. 2017, 8, 414. [Google Scholar] [CrossRef]
  40. Xie, P.; Xu, Y.; Tang, J.; Wu, S.; Gao, H. Multifaceted regulation of siderophore synthesis by multiple regulatory systems in Shewanella oneidensis. Commun. Biol. 2024, 7, 498. [Google Scholar] [CrossRef]
  41. Rodrigues-de Souza Santos, R.E.; Bontempi-Batista, B.; da Silva-Neto, J.F. Ferric uptake regulator Fur coordinates siderophore production and defense against iron toxicity and oxidative stress and contributes to virulence in Chromobacterium violaceum. Appl. Environ. Microbiol. 2020, 86, e01620-20. [Google Scholar]
  42. Coderre, P.E.; Earhart, C.F. The entD gene of the Escherichia coli K12 enterobactin gene cluster. Microbiology 1989, 135, 3043–3055. [Google Scholar] [CrossRef]
  43. Dell’Anno, F.; Vitale, G.A.; Buonocore, C.; Vitale, L.; Palma Esposito, F.; Coppola, D.; Della Sala, G.; Tedesco, P.; de Pascale, D. Novel insights on pyoverdine: From biosynthesis to biotechnological application. Int. J. Mol. Sci. 2022, 23, 11507. [Google Scholar] [CrossRef]
  44. May, J.J.; Wendrich, T.M.; Marahiel, M.A. The dhb operon of Bacillus subtilis encodes the biosynthetic template for the catecholic siderophore 2, 3-dihydroxybenzoate-glycine-threonine trimeric ester bacillibactin. J. Biol. Chem. 2001, 276, 7209–7217. [Google Scholar] [CrossRef]
  45. El-Maraghy, S.S.; Tohamy, T.A.; Hussein, K.A. Expression of SidD gene and physiological characterization of the rhizosphere plant growth promoting yeasts. Heliyon 2020, 6, e04384. [Google Scholar] [CrossRef]
  46. McRose, D.L.; Seyedsayamdost, M.R.; Morel, F.M.M. Multiple siderophores: Bug or feature? J. Biol. Inorg. Chem. 2018, 23, 983–993. [Google Scholar] [CrossRef]
  47. Keating, T.A.; Marshall, C.G.; Walsh, C.T. Vibriobactin biosynthesis in Vibrio cholerae: VibH is an amide synthase homologous to nonribosomal peptide synthetase condensation domains. Biochemistry 2000, 39, 15513–15521. [Google Scholar] [CrossRef]
  48. Barry, S.M.; Challis, G.L. Recent advances in siderophore biosynthesis. Curr. Opin. Chem. Biol. 2009, 13, 205–215. [Google Scholar] [CrossRef]
  49. Crosa, J.H.; Walsh, C.T. Genetics and Assembly line enzymology of siderophore biosynthesis in bacteria. Microbiol. Mol. Biol. Rev. 2002, 66, 223–249. [Google Scholar] [CrossRef]
  50. Khan, A.; Singh, P.; Srivastava, A. Synthesis, nature and utility of universal iron chelator—Siderophore: A review. Microbiol. Res. 2018, 212–213, 103–111. [Google Scholar] [CrossRef]
  51. Berti, A.D.; Thomas, M.G. Analysis of achromobactin biosynthesis by Pseudomonas syringae pv. syringae B728a. J. Bacteriol. 2009, 191, 4594–4604. [Google Scholar] [CrossRef]
  52. Schalk, I.J. Siderophore-antibiotic conjugates: Exploiting iron uptake to deliver drugs into bacteria. Clin. Microbiol. Infect. 2019, 24, 801–802. [Google Scholar] [CrossRef]
  53. Haag, H.; Fiedler, H.P.; Meiwes, J.; Drechsel, H.; Jung, G.; Zähner, H. Isolation and biological characterization of staphyloferrin B, a compound with siderophore activity from staphylococci. FEMS Microbiol. Lett. 1994, 115, 125–130. [Google Scholar] [CrossRef]
  54. Ghssein, G.; Ezzeddine, Z. The key element role of metallophores in the pathogenicity and virulence of Staphylococcus aureus: A Review. Biology 2022, 11, 1525. [Google Scholar] [CrossRef]
  55. Buettner, H.; Hoerl, J.; Krabbe, J.; Hertweck, C. Discovery and biosynthesis of anthrochelin, a growth-promoting metallophore of the human pathogen Luteibacter anthropi. ChemBioChem 2023, 24, e202300322. [Google Scholar]
  56. Buglino, J.A.; Ozakman, Y.; Xu, Y.; Chowdhury, F.; Tan, D.S.; Glickman, M.S. Diisonitrile lipopeptides mediate resistance to copper starvation in pathogenic mycobacteria. mBio 2022, 13, e0251322. [Google Scholar] [CrossRef]
  57. Semrau, J.D.; DiSpirito, A.A.; Obulisamy, P.K.; Kang-Yun, C.S. Methanobactin from methanotrophs: Genetics, structure, function and potential applications. FEMS Microbiol. Lett. 2020, 367, fnaa045. [Google Scholar] [CrossRef]
  58. Reitz, Z.L.; Medema, M.H. Genome mining strategies for metallophore discovery. Curr. Opin. Biotechnol. 2022, 77, 102757. [Google Scholar] [CrossRef]
  59. Leprevost, L.; Jünger, S.; Lippens, G.; Guillaume, C.; Sicoli, G.; Oliveira, L.; Falcone, E.; de-Santis, E.; Rivera-Millot, A.; Billon, G.; et al. A widespread family of ribosomal peptide metallophores involved in bacterial adaptation to metal stress. Proc. Natl. Acad. Sci. USA 2024, 121, e2408304121. [Google Scholar] [CrossRef]
  60. Antelo, G.T.; Vila, A.J.; Giedroc, D.P.; Capdevila, D.A. Molecular evolution of transition metal bioavailability at the host–pathogen interface. Trends Microbiol. 2021, 29, 441–457. [Google Scholar] [CrossRef]
  61. Hider, R.C.; Kong, X. Chemistry and biology of siderophores. Nat. Prod. Rep. 2010, 27, 637–657. [Google Scholar] [CrossRef]
  62. Johnstone, T.C.; Nolan, E.M. Beyond iron: Non-classical biological functions of bacterial siderophores. Dalton Trans. 2015, 44, 6320–6339. [Google Scholar] [CrossRef]
  63. Bellotti, D.; Rowińska-Żyrek, M.; Remelli, M. How zinc-binding systems, expressed by human pathogens, acquire zinc from the colonized host environment: A critical review on zincophores. Curr. Med. Chem. 2021, 28, 7312. [Google Scholar]
  64. Kusakizako, T.; Miyauchi, H.; Ishitani, R.; Nureki, O. Structural biology of the multidrug and toxic compound extrusion superfamily transporters. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183154. [Google Scholar] [CrossRef]
  65. Kraemer, S.M.; Duckworth, O.W.; Harrington, J.M.; Schenkeveld, W.D. Metallophores and trace metal biogeochemistry. Aquat. Geochem. 2015, 21, 159–195. [Google Scholar] [CrossRef]
  66. Laschat, S.; Bilitewski, U.; Blodgett, J.; Duhme-Klair, A.K.; Dallavalle, S.; Routledge, A.; Schobert, R. Chemical and biological aspects of nutritional immunity-perspectives for new antiinfectives targeting iron uptake systems. Angew. Chem. Int. Ed. 2017, 56, 14360. [Google Scholar]
  67. Hofmann, M.; Retamal-Morales, G.; Tischler, D. Metal binding ability of microbial natural metal chelators and potential applications. Nat. Prod. Rep. 2020, 37, 1262. [Google Scholar] [CrossRef]
  68. Mohr, J.F.; Baldeweg, F.; Deicke, M.; Morales-Reyes, C.F.; Hoffmeister, D.; Wichard, T. Frankobactin metallophores produced by nitrogen-fixing Frankia actinobacteria function in toxic metal sequestration. J. Nat. Prod. 2021, 84, 1216–1225. [Google Scholar] [CrossRef]
  69. Maknun, L.; Kińska, K.; González-Álvarez, I.; Ouerdane, L.; Lauga, B.; Siripinyanond, A.; Szpunar, J.; Lobinski, R. Quantitative determination of iron–siderophore complexes in peat by isotope-exchange size-exclusion UPLC–electrospray ionization high-resolution accurate mass (HRAM) mass spectrometry. Anal. Chem. 2023, 95, 9182–9190. [Google Scholar] [CrossRef]
  70. Ramesh, B.; Kameswaran, S.; Venkatrayulu, C.; Silpa, S.; Chandra, M.S.; Reddy, G.V.S.; Kumar, K.N. Microbial interaction with metals and metalloids. In Innovations in Biotechnology for a Sustainable Future; Maddela, N.R., García, L.C., Eds.; Elsevier: Cham, Switzerland, 2021; pp. 243–272. [Google Scholar]
  71. Ramke, R.; Jeyaraman, A. Microbial interaction with metals and metalloids. In Microbes and Microbial Biotechnology for Green Remediation; Malik, J.A., Ed.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 29–46. [Google Scholar]
  72. Wilson, B.R.; Bogdan, A.R.; Miyazawa, M.; Hashimoto, K.; Tsuji, Y. Siderophores in Iron metabolism: From mechanism to therapy potential. Trends Mol. Med. 2016, 22, 1077–1090. [Google Scholar] [CrossRef]
  73. Timofeeva, A.M.; Galyamova, M.R.; Sedykh, S.E. Bacterial siderophores: Classification, biosynthesis, perspectives of use in agriculture. Plants 2022, 11, 3065. [Google Scholar] [CrossRef]
  74. Puja, H.; Mislin, G.L.A.; Rigouin, C. Engineering siderophore biosynthesis and regulation pathways to increase diversity and availability. Biomolecules 2023, 13, 959. [Google Scholar] [CrossRef]
  75. Furrer, J.L.; Sanders, D.N.; Hook-Barnard, I.G.; McIntosh, M.A. Export of the siderophore enterobactin in Escherichia coli: Involvement of a 43 kDa membrane exporter. Mol. Microbiol. 2002, 44, 1225–1234. [Google Scholar] [CrossRef]
  76. Salvail, H.; Lanthier-Bourbonnais, P.; Sobota, J.M.; Caza, M.; Benjamin, J.A.M.; Mendieta, M.E.S.; Lépine, F.; Dozois, C.M.; Imlay, J.; Massé, E. A small RNA promotes siderophore production through transcriptional and metabolic remodeling. Proc. Natl. Acad. Sci. USA 2010, 107, 15223–15228. [Google Scholar] [CrossRef]
  77. Schubert, S.; Fischer, D.; Heesemann, J. Ferric enterochelin transport in Yersinia enterocolitica: Molecular and evolutionary aspects. J. Bacteriol. 1999, 181, 6387–6395. [Google Scholar] [CrossRef]
  78. Endicott, N.P.; Rivera, G.S.M.; Yang, J.; Wencewicz, T.A. Emergence of ferrichelatase activity in a siderophore-binding protein supports an iron shuttle in bacteria. ACS Cent. Sci. 2020, 6, 493–506. [Google Scholar] [CrossRef]
  79. Trapet, P.; Avoscan, L.; Klinguer, A.; Pateyron, S.; Citerne, S.; Chervin, C.; Mazurier, S.; Lemanceau, P.; Wendehenne, D.; Besson-Bard, A. The Pseudomonas fluorescens siderophore pyoverdine weakens Arabidopsis thaliana defense in favor of growth in iron-deficient conditions. Plant Physiol. 2016, 171, 675–693. [Google Scholar] [CrossRef]
  80. Rajkumar, M.; Ae, N.; Prasad, M.N.V.; Freitas, H. Potential of siderophore-producing bacteria for improving heavy metal phytoextraction. Trends Biotechnol. 2010, 28, 142–149. [Google Scholar] [CrossRef]
  81. Albelda-Berenguer, M.; Monachon, M.; Joseph, E. Siderophores: From natural roles to potential applications. Adv. Appl. Microbiol. 2019, 106, 193–225. [Google Scholar]
  82. Gomes, A.F.; Almeida, M.C.; Sousa, E.; Resende, D.I. Siderophores and metallophores: Metal complexation weapons to fight environmental pollution. Sci. Total Environ. 2024, 932, 173044. [Google Scholar] [CrossRef]
  83. Francis, M.S.; Thomas, C.J. Mutants in the CtpA copper transporting P-type ATPase reduce virulence of Listeria monocytogenes. Microb. Pathog. 1997, 22, 67–78. [Google Scholar] [CrossRef]
  84. Renshaw, J.C.; Robson, G.D.; Trinci, A.P.J.; Wiebe, M.G.; Livens, F.R.; Collison, D.; Taylor, R.J. Fungal siderophores: Structures, functions and applications. Mycol. Res. 2002, 106, 1123–1142. [Google Scholar] [CrossRef]
  85. Müller, G.; Isowa, Y.; Raymond, K.N. Stereospecificity of siderophore-mediated iron uptake in Rhodotorula pilimanae as probed by enantiorhodotorulic acid and isomers of chromic rhodotorulate. J. Biol. Chem. 1985, 260, 13921–13926. [Google Scholar] [CrossRef]
  86. Fekete, F.A.; Chandhoke, V.; Jellison, J. Iron-binding compounds produced by wood-decaying basidiomycetes. Appl. Environ. Microbiol. 1989, 55, 2720–2722. [Google Scholar] [CrossRef] [PubMed]
  87. Sushma, J.; Neeti, K.; Sarrah, R.; Pratibha, S.; Hajra, G.A.S. Optimization of universal chrome Azurol S (CAS) siderophore detecting media with an economical dye substitute using Pseudomonas fluorescens as a test organism. Res. J. Chem. Environ. 2021, 25, 27–34. [Google Scholar]
  88. Dimkpa, C. Microbial siderophores: Production, detection and application in agriculture and environment. Endocytobiosis Cell Res. 2016, 27, 7–16. [Google Scholar]
  89. Andrews, M.Y.; Santelli, C.M.; Duckworth, O.W. Layer plate CAS assay for the quantitation of siderophore production and determination of exudation patterns for fungi. J. Microbiol. Methods 2016, 121, 41–43. [Google Scholar] [CrossRef]
  90. Murakami, C.; Tanaka, A.R.; Sato, Y.; Kimura, Y.; Morimoto, K. Easy detection of siderophore production in diluted growth media using an improved CAS reagent. J. Microbiol. Methods 2021, 189, 106310. [Google Scholar] [CrossRef]
  91. Bhattacharya, S.; John, P.J.; Ledwani, L. Microbial siderophores an envisaged tool for asbestos bioremediation—A microcosm approach. Mater. Today Proc. 2021, 43, 3110–3116. [Google Scholar] [CrossRef]
  92. Arora, N.K.; Verma, M. Modified microplate method for rapid and efficient estimation of siderophore produced by bacteria. 3 Biotech 2017, 7, 381. [Google Scholar] [CrossRef]
  93. Gu, S.; Wan, W.; Shao, Z.; Zhong, W. High-throughput method for detecting siderophore production by rhizosphere bacteria. Bio Protoc. 2021, 11, e4001. [Google Scholar] [CrossRef]
  94. Yoder, M.F.; Kisaalita, W.S. Fluorescence of pyoverdin in response to iron and other common well water metals. J. Environ. Sci. Health A 2006, 41, 369–380. [Google Scholar] [CrossRef]
  95. Chen, J.; Sun, Z.; Jin, J.; Wang, F.; Yang, Q.; Yu, H.; Yu, J.; Wang, Y. Role of siderophore in Pseudomonas fluorescens biofilm formation and spoilage potential function. Food Microbiol. 2023, 109, 104151. [Google Scholar] [CrossRef]
  96. Sayyed, R.Z.; Chincholkar, S.B. Purification of siderophores of Alcaligenes feacalis on Amberlite XAD. Bioresour. Technol. 2006, 97, 1026–1029. [Google Scholar] [CrossRef] [PubMed]
  97. Sayyed, R.Z.; Chincholkar, S.B.; Reddy, M.S.; Gangurde, N.S.; Patel, P.R. Siderophore producing PGPR for crop nutrition and phytopathogen suppression. In Bacteria in Agrobiology: Disease Management; Sayyed, R.Z., Chincholkar, S.B., Reddy, M.S., Gangurde, N.S., Patel, P.R., Eds.; Springer: Berlin, Germany, 2012; pp. 449–471. [Google Scholar]
  98. Schwabe, R.; Senges, C.H.R.; Bandow, J.E.; Heine, T.; Lehmann, H.; Wiche, O.; Schlömanna, M.; Levicáne, G.; Tischler, D. Cultivation dependent formation of siderophores by Gordonia rubripertincta CWB2. Microbiol. Res. 2020, 238, 126481. [Google Scholar] [CrossRef] [PubMed]
  99. Meesungnoen, O.; Chantiratikul, P.; Thumanu, K.; Nuengchamnong, N.; Hokura, A.; Nakbanpote, W. Elucidation of crude siderophore extracts from supernatants of Pseudomonas sp. ZnCd2003 cultivated in nutrient broth supplemented with Zn, Cd, and Zn plus Cd. Arch. Microbiol. 2021, 203, 2863–2874. [Google Scholar] [CrossRef]
  100. Mazari, H.E.; Meliani, A.; Berkat, S.; Aliane, S.; Djibaoui, R.; Bouderoua, K. Washing of heavy metal-contaminated soils using pyoverdine extracted from plant growth-promoting bacteria Pseudomonas lactis and P. atacamensis. Carpathian J. Earth Environ. Sci. 2024, 19, 169–178. [Google Scholar] [CrossRef]
  101. Schwabe, R.; Anke, M.K.; Szymańska, K.; Wiche, O.; Tischler, D. Analysis of desferrioxamine-like siderophores and their capability to selectively bind metals and metalloids: Development of a robust analytical RP-HPLC method. Res. Microbiol. 2018, 169, 598–607. [Google Scholar] [CrossRef]
  102. Doyama, K.; Haruma, T.; Hishiyama, S.; Kato, A.; Masuya, H.; Yamaji, K. Isoavenaciol and 7-hydroxy-isoavenaciol: Zn-chelating metallophores produced by root-endophytic Pezicula ericae in a Zn-accumulating plant, Aucuba japonica. Phytochemistry 2023, 206, 113547. [Google Scholar] [CrossRef]
  103. Deicke, M.; Bellenger, J.P.; Wichard, T. Direct quantification of bacterial molybdenum and iron metallophores with ultra-high-performance liquid chromatography coupled to time-of-flight mass spectrometry. J. Chromatogr. A 2013, 1298, 50–60. [Google Scholar] [CrossRef]
  104. Franco-Sierra, N.D.; Posada, L.F.; Santa-María, G.; Romero-Tabarez, M.; Villegas-Escobar, V.; Álvarez, J.C. Bacillus subtilis EA-CB0575 genome reveals clues for plant growth promotion and potential for sustainable agriculture. Funct. Integr. Genomics 2020, 20, 575–589. [Google Scholar] [CrossRef]
  105. Kenney, G.E.; Rosenzweig, A.C. Methanobactins: Maintaining copper homeostasis in methanotrophs and beyond. J. Biol. Chem. 2018, 293, 4606–4615. [Google Scholar] [CrossRef]
  106. Dassama, L.M.; Kenney, G.E.; Rosenzweig, A.C. Methanobactins: From genome to function. Metallomics 2017, 9, 7–20. [Google Scholar] [CrossRef]
  107. Calderón-Celis, F.; González-Álvarez, I.; Fabjanowicz, M.; Godin, S.; Ouerdane, L.; Lauga, B.; Łobiński, R. Unveiling the pool of metallophores in native environments and correlation with their potential producers. Environ. Sci. Technol. 2023, 57, 17302–17311. [Google Scholar] [CrossRef] [PubMed]
  108. Garber, A.I.; Nealson, K.H.; Okamoto, A.; McAllister, S.M.; Chan, C.S.; Barco, R.A.; Merino, N. FeGenie: A comprehensive tool for the identification of iron genes and iron gene neighborhoods in genome and metagenome assemblies. Front. Microbiol. 2020, 11, 37. [Google Scholar] [CrossRef] [PubMed]
  109. Singh, P.; Khan, A.; Kumar, R.; Ojha, K.K.; Singh, V.K.; Srivastava, A. In silico analysis of comparative affinity of phytosiderophore and bacillibactin for iron uptake by YSL15 and YSL18 receptors of Oryza sativa. J. Biomol. Struct. Dyn. 2023, 41, 2733–2746. [Google Scholar] [CrossRef] [PubMed]
  110. Berg, G.; Rybakova, D.; Fischer, D.; Cernava, T.; Verges, M.C.C.; Charles, T.; Chen, X.; Cocolin, L.; Eversole, K.; Corral, G.H.; et al. Microbiome definition re-visited: Old concepts and new challenges. Microbiome 2020, 8, 103. [Google Scholar]
  111. Hassani, M.A.; Durán, P.; Hacquard, S. Microbial interactions within the plant holobiont. Microbiome 2018, 6, 58. [Google Scholar] [CrossRef]
  112. Cordero, O.X.; Ventouras, L.A.; DeLong, E.F.; Polz, M.F. Public good dynamics drive evolution of iron acquisition strategies in natural bacterioplankton populations. Proc. Natl Acad. Sci. USA 2012, 109, 20059–20064. [Google Scholar] [CrossRef]
  113. Leinweber, A.; Inglis, R.F.; Kümmerli, R. Cheating fosters species co-existence in well-mixed bacterial communities. ISME J. 2017, 11, 1179–1188. [Google Scholar] [CrossRef]
  114. MacLean, R.C. The tragedy of the commons in microbial populations: Insights from theoretical, comparative and experimental studies. Heredity 2008, 100, 471–477. [Google Scholar] [CrossRef]
  115. Özkaya, Ö.; Xavier, K.B.; Dionisio, F.; Balbontin, R. Maintenance of microbial cooperation mediated by public goods in single and multiple traits scenarios. J. Bacteriol. 2017, 199, e00297-17. [Google Scholar] [CrossRef]
  116. Lin, H.; Liu, C.; Li, B.; Dong, Y. Trifolium repens L. regulated phytoremediation of heavy metal contaminated soil by promoting soil enzyme activities and beneficial rhizosphere associated microorganisms. J. Hazard. Mater. 2021, 402, 123829. [Google Scholar] [CrossRef]
  117. Oberegger, H.; Schoeser, M.; Zadra, I.; Abt, B.; Haas, H. SREA is involved in regulation of siderophore biosynthesis, utilisation and uptake in Aspergillus nidulans. Mol. Microbiol. 2001, 41, 1077–1089. [Google Scholar] [CrossRef]
  118. Fiedler, H.P.; Krastel, P.; Muller, J.; Gebhardt, K.; Zeeck, A. Enterobactin: The characteristic catecholate siderophore of Enterobacteriaceae is produced by Streptomyces species. FEMS Microbiol. Lett. 2001, 115, 125–130. [Google Scholar] [CrossRef]
  119. Haas, D.; Keel, C. Regulation of antibiotic production in root-colonizing Pseudomonas spp and relevance for biological control of plant disease. Annu. Rev. Phytopathol. 2003, 41, 117–153. [Google Scholar] [CrossRef]
  120. O’Brien, S.; Hodgson, D.J.; Buckling, A. Social evolution of toxic metal bioremediation in Pseudomonas aeruginosa. Proc. Biol. Sci. 2014, 281, 20140858. [Google Scholar] [CrossRef]
  121. Gonzales, D.T.; Suraritdechachai, S.; Zechner, C.; Tang, T.Y.D. Bidirectional communication between droplet interface bilayers driven by cell-free quorum sensing gene circuits. ChemSystemsChem 2023, 5, e202300029. [Google Scholar] [CrossRef]
  122. Lynch, J.M. Beneficial interactions between micro-organisms and roots. Biotechnol. Adv. 1990, 8, 335–346. [Google Scholar] [CrossRef]
  123. Sørensen, J. The rhizosphere as a habitat for soil microorganisms. In Modern Soil Microbiology; van Elsas, J.D., Trevors, J.T., Wellington, E.M.H., Eds.; Marcel Dekker, Inc.: New York, NY, USA, 1997; pp. 21–45. [Google Scholar]
  124. Kennedy, A.C.; De Luna, L.Z. Rhizosphere. In Encyclopedia of Soil as in the Environment; Hillel, D., Ed.; Elsevier: Oxford, UK, 2004; pp. 399–406. [Google Scholar]
  125. Ngullie, E.; Singh, A.K.; Sema, A.; Srivastava, A.K. Citrus growth and rhizosphere properties. Commun. Soil Sci. Plant Anal. 2015, 46, 1540–1550. [Google Scholar] [CrossRef]
  126. Curl, E.A.; Truelove, B. The Rhizosphere; Springer: New York, NY, USA, 1986; p. 289. [Google Scholar]
  127. DeAngelis, K.M.; Silver, W.L.; Thompson, A.W.; Firestone, M.K. Microbial communities acclimate to recurring changes in soil redox potential status. Environ. Microbiol. 2010, 12, 3137–3149. [Google Scholar] [CrossRef]
  128. Bais, H.P.; Weir, T.L.; Perry, L.G.; Gilroy, S.; Vivanco, J.M. The role of root exudates in rhizosphere interactions with plants and other organisms. Annu. Rev. Plant Biol. 2006, 57, 233–266. [Google Scholar] [CrossRef]
  129. Haichar, Z.F.; Marol, C.; Berge, O.; Rangel-Castro, J.I.; Prosser, J.I.; Balesdent, J.; Heulin, T.; Achouak, W. Plant host habitat and root exudates shape soil bacterial community structure. ISME J. 2008, 2, 1221. [Google Scholar] [CrossRef]
  130. Dennis, P.G.; Miller, A.J.; Hirsch, P.R. Are root exudates more important than other sources of rhizodeposits in structuring rhizosphere bacterial communities? FEMS Microbiol. Ecol. 2010, 72, 313–327. [Google Scholar] [CrossRef]
  131. Yadav, A. Exploring the potential of endophytes in agriculture: A minireview. Adv. Plants Agric. Res. 2017, 6, 102–106. [Google Scholar] [CrossRef]
  132. Bhardwaj, D.; Ansari, M.W.; Sahoo, R.K.; Tuteja, N. Biofertilizers function as key player in sustainable agriculture by improving soil fertility, plant tolerance and crop productivity. Microb. Cell Fact. 2014, 13, 66. [Google Scholar] [CrossRef] [PubMed]
  133. Seneviratne, G.; Kulasooriya, S.A. Reinstating soil microbial diversity in agroecosystems: The need of the hour for sustainability and health. Agric. Ecosyst. Environ. 2013, 164, 181–182. [Google Scholar] [CrossRef]
  134. Rout, G.R.; Sahoo, S. Role of iron in plant growth and metabolism. Rev. Agric. Sci. 2015, 3, 1–24. [Google Scholar] [CrossRef]
  135. Hebbern, C.A.; Laursen, K.H.; Ladegaard, A.H.; Schmidt, S.B.; Pedas, P.; Bruhn, D.; Schjoerring, J.K.; Wulfsohn, D.; Husted, S. Latent manganese deficiency increases transpiration in barley (Hordeum vulgare). Physiol. Plant. 2009, 135, 307–316. [Google Scholar] [CrossRef]
  136. Schwarz, G.; Mendel, R.R. Molybdenum cofactor biosynthesis and molybdenum enzymes. Annu. Rev. Plant Biol. 2006, 57, 623–647. [Google Scholar] [CrossRef]
  137. Follmer, C. Insights into the role and structure of plant ureases. Phytochemistry 2008, 69, 18–28. [Google Scholar] [CrossRef]
  138. Kawagashira, N.; Ohtomo, Y.; Murakami, K.; Matsubara, K.; Kawai, J.; Carninci, R.; Hayashizaki, P.; Kikuchi, S.; Higo, K. Multiple zinc finger motifs with comparison of plant and insects. Genom. Inform. 2001, 12, 368–369. [Google Scholar]
  139. Kramer, U.; Clemens, S. Function and homeostasis of zinc, copper, and nickel in plants. In Molecular Biology of Metal Homeostasis and Detoxification: From Microbes to Man; Tamas, M.J., Martinoia, E., Eds.; Springer-Verlag: Berlin, Germany, 2006; pp. 215–271. [Google Scholar]
  140. Constantin, M.; Chioncel, M.F.; Petrescu, L.; Vrancianu, C.O.; Paun, M.; Cristian, R.E.; Sidoroff, M.; Dionisie, M.V.; Chifiriuc, M.C. From rock to living systems: Lanthanides toxicity and biological interactions. Ecotoxicol. Environ. Saf. 2025, 289, 117494. [Google Scholar] [CrossRef]
  141. Sharma, A.; Johri, B. Growth promoting influence of siderophore-producing Pseudomonas strains GRP3A and PRS9 in maize (Zea mays L.) under iron limiting conditions. Microbiol. Res. 2003, 158, 243–248. [Google Scholar] [CrossRef] [PubMed]
  142. Sah, S.; Singh, N.; Singh, R. Iron acquisition in maize (Zea mays L.) using Pseudomonas siderophore. 3 Biotech 2017, 7, 121. [Google Scholar] [CrossRef] [PubMed]
  143. Zhao, L.; Wang, Y.; Kong, S. Effects of Trichoderma asperellum and its siderophores on endogenous auxin in Arabidopsis thaliana under iron-deficiency stress. Int. Microbiol. 2020, 23, 501–509. [Google Scholar] [CrossRef]
  144. Cui, K.; Xu, T.; Chen, J.; Yang, H.; Liu, X.; Zhuo, R.; Peng, Y.; Tang, W.; Wang, R.; Chen, L.; et al. Siderophores, a potential phosphate solubilizer from the endophyte Streptomyces sp. CoT10, improved phosphorus mobilization for host plant growth and rhizosphere modulation. J. Clean. Prod. 2022, 367, 133110. [Google Scholar] [CrossRef]
  145. Singh, P.; Kumar, R.; Khan, A.; Singh, A.; Srivastava, A. Bacillibactin siderophore induces iron mobilisation responses inside aerobic rice variety through YSL15 transporter. Rhizosphere 2023, 27, 100724. [Google Scholar] [CrossRef]
  146. Wensing, A.; Braun, S.D.; Büttner, P.; Expert, D.; Völksch, B.; Ullrich, M.S.; Weingart, H. Impact of siderophore production by Pseudomonas syringae pv. syringae 22d/93 on epiphytic fitness and biocontrol activity against Pseudomonas syringae pv. glycinea 1a/96. Appl. Environ. Microbiol. 2010, 76, 2704–2711. [Google Scholar] [CrossRef]
  147. Sha, Z.; Watanabe, T.; Chu, Q.; Oka, N.; Osaki, M.; Shinano, T. A reduced phosphorus application rate using a mycorrhizal plant as the preceding crop maintains soybean seeds’ nutritional quality. J. Agric. Food Chem. 2019, 67, 32–42. [Google Scholar] [CrossRef]
  148. Bellenger, J.P.; Wichard, T.; Kustka, A.B.; Kraepiel, A.M.L. Nitrogen fixing soil bacterium uses catechol siderophores for molybdenum and vanadium acquisition. Nat. Geosci. 2008, 1, 243–246. [Google Scholar] [CrossRef]
  149. Bellenger, J.P.; Wichard, T.; Xu, Y.; Kraepiel, A.M. Essential metals for nitrogen fixation in a free-living N2-fixing bacterium: Chelation, homeostasis and high use efficiency. Environ. Microbiol. 2011, 13, 1395–1411. [Google Scholar] [CrossRef]
  150. Holler, T.; Wegener, G.; Knittel, K.; Boetius, A.; Brunner, B.; Kuypersm, M.M.M.; Widdel, F. Substantial 13C/12C and D/H fractionation during anaerobic oxidation of methane by marine consortia enriched in vitro. Environ. Microbiol. Rep. 2009, 1, 370–376. [Google Scholar] [CrossRef]
  151. Wichard, T.; Mishra, B.; Myneni, S.C.B.; Bellenger, J.P.; Kraepiel, A.M.L. Storage and bioavailability of molybdenum in soils increased by organic matter complexation. Nat. Geosci. 2009, 2, 625. [Google Scholar] [CrossRef]
  152. Srivastava, S.; Dong, H.; Baars, O.; Sheng, Y. Bioavailability of mineral-associated trace metals as cofactors for nitrogen fixation by Azotobacter vinelandii. Geobiology 2023, 21, 507–519. [Google Scholar] [CrossRef] [PubMed]
  153. Yadav, S.; Kaushik, R.; Saxena, A.K.; Arora, D.K. Diversity and phylogeny of plant growth-promoting bacilli from moderately acidic soil. J. Basic Microbiol. 2011, 51, 98–106. [Google Scholar] [CrossRef]
  154. Van Scholl, L.; Kuyper, T.W.; Smits, M.M.; Landeweert, R.; Hoffland, E.; van Breemen, N. Rock-eating mycorrhizas: Their role in plant nutrition and biogeochemical cycles. Plant Soil 2008, 303, 35–47. [Google Scholar] [CrossRef]
  155. Loper, J.E.; Buyer, J.S. Siderophores in microbial interactions on plant surfaces. Mol. Plant Microbe Interact. 1991, 4, 5–13. [Google Scholar] [CrossRef]
  156. Raaijmakers, J.M.; Sluis, L.V.D.; Bakker, P.A.; Schippers, B.; Koster, M.; Weisbeek, P.J. Utilisation of heterologous siderophores and rhizosphere competence of fluorescent Pseudomonas spp. Can. J. Microbiol. 1995, 41, 126–135. [Google Scholar] [CrossRef]
  157. Glick, B.R. Bacteria with ACC deaminase can promote plant growth and help to feed the world. Microbiol. Res. 2014, 169, 30–39. [Google Scholar] [CrossRef]
  158. Ma, Y.; Oliveira, R.S.; Freitas, H.; Zhang, C. Biochemical and molecular mechanisms of plant-microbe-metal interactions: Relevance for phytoremediation. Front. Plant Sci. 2016, 7, 918. [Google Scholar] [CrossRef]
  159. Gu, S.; Yang, T.; Shao, Z.; Wang, T.; Cao, K.; Jousset, A.; Friman, V.P.; Mallon, C.; Mei, X.; Wei, Z.; et al. Siderophore-mediated interactions determine the disease suppressiveness of microbial consortia. mSystems 2020, 5, e00811-19. [Google Scholar] [CrossRef]
  160. Gu, S.; Wei, Z.; Shao, Z.; Friman, V.P.; Cao, K.; Yang, T.; Kramer, J.; Wang, X.; Li, M.; Mei, X.; et al. Competition for iron drives phytopathogen control by natural rhizosphere microbiomes. Nat. Microbiol. 2020, 5, 1002–1010. [Google Scholar] [CrossRef]
  161. Maurhofer, M.; Hase, C.; Meuwly, P.; Metraux, J.P.; Defago, G. Induction of systemic resistance of tobacco to tobacco necrosis virus by the root-colonizing Pseudomonas fluorescens strain CHA0: Influence of the gacA gene and of pyoverdine production. Phytopathology 1994, 84, 139–146. [Google Scholar] [CrossRef]
  162. Leeman, M.; Den Ouden, F.M.; Van Pelt, J.A.; Dirkx, F.P.M.; Steijl, H.; Bakker, P.A.H.M.; Schippers, B. Iron availability affects induction of systemic resistance to Fusarium wilt of radish by Pseudomonas fluorescens. Phytopathology 1996, 86, 149–155. [Google Scholar] [CrossRef]
  163. Michavila, G.; Adler, C.; De Gregorio, P.R.; Lami, M.J.; Caram-Di Santo, M.C.; Zenoff, A.M.; de Cristobal, R.E.; Vincent, P.A. Pseudomonas protegens CS 1 from the lemon phyllosphere as a candidate for citrus canker biocontrol agent. Plant Biol. 2017, 19, 608–617. [Google Scholar] [CrossRef] [PubMed]
  164. Dhanya, S.; Sherin, V.; Divya, K.; Sreekumar, J.; Jisha, M.S. Pseudomonas taiwanensis (MTCC11631) mediated induction of systemic resistance in Anthurium andreanum L. against blight disease and visualization of defense related secondary metabolites using confocal laser scanning microscopy. Biocatal. Agric. Biotechnol. 2020, 24, 101561. [Google Scholar] [CrossRef]
  165. Ghazy, N.; El-Nahrawy, S. Siderophore production by Bacillus subtilis MF497446 and Pseudomonas koreensis MG209738 and their efficacy in controlling Cephalosporium maydis in maize plant. Arch. Microbiol. 2020, 203, 1195–1209. [Google Scholar] [CrossRef]
  166. Gupta, V.; Kumar, G.N.; Buch, A. Colonization by multi-potential Pseudomonas aeruginosa P4 stimulates peanut (Arachis hypogea L.) growth, defense physiology and root system functioning to benefit the root-rhizobacterial interface. J. Plant Physiol. 2020, 248, 153144. [Google Scholar] [CrossRef]
  167. Arora, N.; Kang, S.; Maheshwari, D. Isolation of siderophore-producing strains of Rhizobium meliloti and their biocontrol potential against Macrophomina phaseolina that causes charcoal rot of groundnut. Curr. Sci. 2001, 81, 673–677. [Google Scholar]
  168. Verma, V.C.; Singh, S.K.; Prakash, S. Bio-control and plant growth promotion potential of siderophore producing endophytic Streptomyces from Azadirachta indica A. Juss. J. Basic Microbiol. 2011, 51, 550–556. [Google Scholar] [CrossRef]
  169. Myo, E.M.; Liu, B.; Ma, J.; Shi, L.; Jiang, M.; Zhang, K.; Ge, B. Evaluation of Bacillus velezensis NKG-2 for bio-control activities against fungal diseases and potential plant growth promotion. Biol. Control 2019, 134, 23–31. [Google Scholar] [CrossRef]
  170. Di Francesco, A.; Sciubba, L.; Bencivenni, M.; Marzadori, C.; Baraldi, E. Application of Aureobasidium pullulans in iron-poor soil. Can the production of siderophores improve iron bioavailability and yeast antagonistic activity? Ann. Appl. Biol. 2022, 180, 398–406. [Google Scholar] [CrossRef]
  171. Seregin, I.V.; Kozhevnikova, A.D. Low-molecular-weight ligands in plants: Role in metal homeostasis and hyperaccumulation. Photosynth. Res. 2021, 150, 51–96. [Google Scholar] [CrossRef] [PubMed]
  172. Murata, Y.; Murata, J.; Namba, K. Unraveling the new biological roles and possible applications of phytosiderophores in plants and mammals. Met. Res. 2022, 2, MR202202. [Google Scholar]
  173. Carrillo, J.T.; Borthakur, D. Methods for metal chelation in plant homeostasis: Review. Plant Physiol. Biochem. 2021, 163, 95–107. [Google Scholar] [CrossRef]
  174. Delaporte-Quintana, P.; Lovalsa, N.C.; Raspisarda, V.A.; Pedraza, R.O. The plant growth promoting bacteria Glauconobacter diazotrophicus and Azospirillum brasiliense contribute to the iron nutrition of strawberry plants through siderophores production. Plant Growth Regul. 2020, 91, 185–199. [Google Scholar] [CrossRef]
  175. Engelbrecht, A.; Saad, H.; Gross, H.; Kaysser, L. Natural products from Nocardia and their role in pathogenicity. Microb. Physiol. 2021, 31, 217–232. [Google Scholar] [CrossRef]
  176. Barra-Caracciolo, A.; Terenzi, V. Rhizosphere microbial communities and heavy metals. Microorganisms 2021, 9, 1462. [Google Scholar] [CrossRef]
  177. Ahmed, E.; Holmström, S.J. Siderophores in environmental research: Roles and applications. Microb. Biotechnol. 2014, 7, 196–208. [Google Scholar] [CrossRef]
  178. Mohanty, S.K.; Gonneau, C.; Salamatipour, A.; Pietrofesa, R.A.; Casper, B.; Christofidou-Solomidou, M.; Willenbring, J.K. Siderophore-mediated iron removal from chrysotile: Implications for asbestos toxicity reduction and bioremediation. J. Hazard. Mater. 2018, 341, 290–296. [Google Scholar] [CrossRef]
  179. Roskova, Z.; Skarohlid, R.; McGachy, L. Siderophores: An alternative bioremediation strategy? Sci. Total Environ. 2022, 819, 153144. [Google Scholar] [CrossRef]
  180. Tripathi, M.; Munor, H.P.; Shouch, Y.; Meyer, J.M. Isolation and functional characterization of siderophore producing lead and cadmium resistant Pseudomonas putida KNP9. Curr. Microbiol. 2005, 50, 233–237. [Google Scholar] [CrossRef]
  181. Dimkpa, C.O.; Merten, D.; Svatoš, A.; Büchel, G.; Kothe, E. Metal-induced oxidative stress impacting plant growth in contaminated soil is alleviated by microbial siderophores. Soil Biol. Biochem. 2009, 41, 154–162. [Google Scholar] [CrossRef]
  182. Singh, A.; Kaushik, M.S.; Srivastava, M.; Tiwari, D.N.; Mishra, A.K. Siderophore mediated attenuation of cadmium toxicity by paddy field cyanobacterium Anabaena oryzae. Algal Res. 2016, 16, 63–68. [Google Scholar] [CrossRef]
  183. Whiting, S.N.; de Souza, M.P.; Terry, N. Rhizosphere bacteria mobilize Zn for hyperaccumulation by Thlaspi caerulescens. Environ. Sci. Technol. 2001, 35, 3144–3150. [Google Scholar] [CrossRef] [PubMed]
  184. Chang, J.; Kim, D.D.; Semrau, J.D.; Lee, J.Y.; Heo, H.; Gu, W.; Yoon, S. Enhancement of nitrous oxide emissions in soil microbial consortia via copper competition between proteobacterial methanotrophs and denitrifiers. Appl. Environ. Microbiol. 2021, 87, e02301. [Google Scholar] [CrossRef]
  185. Chang, J.; Peng, P.; Ul-Haque, M.F.; Hira, A.; Dispirito, A.A.; Semrau, J.D. Inhibition of nitrous oxide reduction in forest soil microcosms by different forms of methanobactin. Environ. Microbiol. 2023, 25, 2338–2350. [Google Scholar] [CrossRef]
  186. Ferousi, C.; Majer, S.H.; DiMucci, I.M.; Lancaster, K.M. Biological and bioinspired inorganic N–N bond-forming reactions. Chem. Rev. 2020, 120, 5252–5307. [Google Scholar] [CrossRef]
  187. Chang, J.; Gu, W.; Park, D.; Semrau, J.D.; Dispirito, A.A.; Yoon, S. Methanobactin from Methylosinus trichosporium OB3b inhibits N2O reduction in denitrifiers. ISME J. 2018, 12, 2086–2089. [Google Scholar] [CrossRef]
  188. Chang, J.; Peng, P.; DiSpirito, A.A.; Semrau, J.D. Variable inhibition of nitrous oxide reduction in denitrifying bacteria by different forms of methanobactin. Appl. Environ. Microbiol. 2022, 88, e02346. [Google Scholar] [CrossRef]
  189. Lee, A.; Winther, M.; Priemé, A.; Blunier, T.; Christensen, S. Hot spots of N2O emission move with the seasonally mobile oxic-anoxic interface in drained organic soils. Soil Biol. Biochem. 2017, 115, 178–186. [Google Scholar] [CrossRef]
  190. Sultana, S.; Alamb, S.; Karimc, M.M. Screening of siderophore-producing salt-tolerant rhizobacteria suitable for supporting plant growth in saline soils with iron limitation. J. Agri. Food Res. 2021, 4, 100150. [Google Scholar] [CrossRef]
  191. Sadeghi, A.; Karimi, E.; Dahaji, P.A.; Javid, M.G.; Dalvand, Y.; Askar, H. Plant growth promoting activity of an auxin and siderophore producing isolate of Streptomyces under saline soil conditions. World J. Microbiol. Biotechnol. 2012, 28, 1503–1509. [Google Scholar] [CrossRef] [PubMed]
  192. Córdova-Albores, L.C.; Zelaya-Molina, L.X.; Ávila-Alistac, N.; Valenzuela-Ruíz, V.; Cortés-Martínez, N.E.; Parra-Cota, F.I.; Burgos-Canul, Y.Y.; Chávez-Díaz, I.F.; Fajardo-Franco, M.L.; de los Santos-Villalobos, S. Omics sciences potential on bioprospecting of biological control microbial agents: The case of Mexican agro-biotechnology. Mex. J. Phytopathol. 2021, 39, 147–184. [Google Scholar] [CrossRef]
  193. Kong, H.; Cheng, W.; Wei, H.; Yuan, Y.; Yang, Z.; Zhang, X. An overview of recent progress in siderophore-antibiotic conjugates. Eur. J. Med. Chem. 2019, 182, 111615. [Google Scholar] [CrossRef] [PubMed]
  194. Negash, K.H.; Norris, J.K.S.; Hodgkinson, J.T. Siderophore-antibiotic conjugate design: New drugs for bad bugs? Molecules 2019, 24, 3314. [Google Scholar] [CrossRef]
  195. Ezzeddine, Z.; Ghssein, G. Towards new antibiotics classes targeting bacterial metallophores. Microb. Pathog. 2023, 182, 106221. [Google Scholar] [CrossRef]
  196. Ezzeddine, Z.; Ghssein, G. Metallophores as promising chelates for heavy metals removal from polluted water. Bioremediat. J. 2024, 1–3. [Google Scholar] [CrossRef]
  197. Wang, L.; Yan, W.; Chen, J.; Huang, F.; Gao, P. Function of the ironbinding chelator produced by Coriolus versicolor in lignin biodegradation. Sci. China Ser. C Life Sci. 2008, 51, 214–221. [Google Scholar] [CrossRef]
  198. Winkelmann, G. The history of siderophores. Curr. Trends Microbiol. 2022, 16, 87–92. [Google Scholar]
  199. Atugoda, D.R.A.M.T.R.; Mandakini, L.L.U.; Bandara, N.J.G.J.; Gunaguardana, D. How to taxonomically-ambiguous cyanobiont and vanadete assist in the phytoremediation of cadmium by Azolla pinnata: Implications for CKDu. Environ. Pollut. 2018, 7, 53–65. [Google Scholar] [CrossRef]
  200. Moscatello, N.; Swayambhu, G.; Jones, C.H.; Xu, J.; Dai, N.; Pfeifer, B.A. Continuous removal of copper, magnesium and nickel from industrial wastewater utilizing the natural product yersiniabactin immobilized withing a packed-bed column. Chem. Eng. 2018, 343, 173–179. [Google Scholar] [CrossRef]
  201. Wong, M.Y.; Neill, C.; Marino, R.; Silvério, D.; Howarth, R.W. Molybdenum phosphorus, and pH constratin nitrogen fixation in a tropical forest in the southeastern Amazon. Ecology. 2021, 102, e03211. [Google Scholar] [CrossRef] [PubMed]
  202. Sidiqi, M.H.; Erber, W.N. Red blood cell disorders. In Clinical Pharmacology, 12th ed.; Brown, M.J., Sharma, P., Mir, F.A., Bennett, P.N., Eds.; Elsevier: Edinburgh, UK, 2018; pp. 528–542. [Google Scholar]
  203. El-Lababidi, R.M.; Rizk, J.G. Cefiderocol: A siderophore cephalosporin. Ann. Pharmacother. 2020, 54, 1215–1231. [Google Scholar] [CrossRef] [PubMed]
  204. Soares, E.V. Perspective on the biotechnological production of bacterial siderophores and their use. Appl. Microbiol. Biotechnol. 2022, 106, 3985–4004. [Google Scholar] [CrossRef]
Figure 1. Metallophore excretion and uptake transport systems. ABC: ATP-Binding Cassette Transporters; TnSS: Type n Secretion System; MSF: Major Facilitator Superfamily Transporters; RND: Resistance–Nodulation–Division Transporters; TRAP: Tripartite ATP-Independent Periplasmic Transporters; TBDT: TonB- or TonR-Dependent Transporters.
Figure 1. Metallophore excretion and uptake transport systems. ABC: ATP-Binding Cassette Transporters; TnSS: Type n Secretion System; MSF: Major Facilitator Superfamily Transporters; RND: Resistance–Nodulation–Division Transporters; TRAP: Tripartite ATP-Independent Periplasmic Transporters; TBDT: TonB- or TonR-Dependent Transporters.
Microbiolres 16 00067 g001
Figure 2. Role of metallophores in microbial and plant interactions. Metallophores are molecules that enable microorganisms to interact with their environment, other microorganisms, and plants. In the soil, they contribute to mineralization and solubilization, facilitating the uptake of essential ions. In the rhizosphere, they mobilize nutrients for plant bioavailability and participate in competition and biocontrol, as some microorganisms sequester metallophore-bound nutrients produced by others, depriving the original producers of access to these essential elements. They also play a role in bioremediation through interactions with specific plants. In the phyllosphere, metallophores produced by beneficial microorganisms are generally inaccessible to phytopathogens, contributing to competition and biological control, with some exhibiting antimicrobial activity. In the endosphere, they act as elicitors of systemic resistance and function as plant growth promoters.
Figure 2. Role of metallophores in microbial and plant interactions. Metallophores are molecules that enable microorganisms to interact with their environment, other microorganisms, and plants. In the soil, they contribute to mineralization and solubilization, facilitating the uptake of essential ions. In the rhizosphere, they mobilize nutrients for plant bioavailability and participate in competition and biocontrol, as some microorganisms sequester metallophore-bound nutrients produced by others, depriving the original producers of access to these essential elements. They also play a role in bioremediation through interactions with specific plants. In the phyllosphere, metallophores produced by beneficial microorganisms are generally inaccessible to phytopathogens, contributing to competition and biological control, with some exhibiting antimicrobial activity. In the endosphere, they act as elicitors of systemic resistance and function as plant growth promoters.
Microbiolres 16 00067 g002
Figure 3. Current and potential applications of metallophores in the industrial, medical, and agricultural sectors.
Figure 3. Current and potential applications of metallophores in the industrial, medical, and agricultural sectors.
Microbiolres 16 00067 g003
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

Zelaya-Molina, L.X.; Chávez-Díaz, I.F.; Urrieta-Velázquez, J.A.; Aragón-Magadan, M.A.; Puente-Valenzuela, C.O.; Blanco-Camarillo, M.; de los Santos-Villalobos, S.; Ramos-Garza, J. Microbial Metallophores in the Productivity of Agroecosystems. Microbiol. Res. 2025, 16, 67. https://doi.org/10.3390/microbiolres16030067

AMA Style

Zelaya-Molina LX, Chávez-Díaz IF, Urrieta-Velázquez JA, Aragón-Magadan MA, Puente-Valenzuela CO, Blanco-Camarillo M, de los Santos-Villalobos S, Ramos-Garza J. Microbial Metallophores in the Productivity of Agroecosystems. Microbiology Research. 2025; 16(3):67. https://doi.org/10.3390/microbiolres16030067

Chicago/Turabian Style

Zelaya-Molina, Lily X., Ismael F. Chávez-Díaz, José A. Urrieta-Velázquez, Marco A. Aragón-Magadan, Cristo O. Puente-Valenzuela, Mario Blanco-Camarillo, Sergio de los Santos-Villalobos, and Juan Ramos-Garza. 2025. "Microbial Metallophores in the Productivity of Agroecosystems" Microbiology Research 16, no. 3: 67. https://doi.org/10.3390/microbiolres16030067

APA Style

Zelaya-Molina, L. X., Chávez-Díaz, I. F., Urrieta-Velázquez, J. A., Aragón-Magadan, M. A., Puente-Valenzuela, C. O., Blanco-Camarillo, M., de los Santos-Villalobos, S., & Ramos-Garza, J. (2025). Microbial Metallophores in the Productivity of Agroecosystems. Microbiology Research, 16(3), 67. https://doi.org/10.3390/microbiolres16030067

Article Metrics

Back to TopTop