Next Article in Journal
Impacts of Arbuscular Mycorrhizal Fungi on Metabolites of an Invasive Weed Wedelia trilobata
Previous Article in Journal
Pathogenomes of Shiga Toxin Positive and Negative Escherichia coli O157:H7 Strains TT12A and TT12B: Comprehensive Phylogenomic Analysis Using Closed Genomes
Previous Article in Special Issue
Sphingomonas sediminicola Dae20 Is a Highly Promising Beneficial Bacteria for Crop Biostimulation Due to Its Positive Effects on Plant Growth and Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Plant-Growth-Promoting Rhizobacteria in Plant Machinery for Soil Heavy Metal Detoxification

by
Haichen Qin
,
Zixiao Wang
,
Wenya Sha
,
Shuhong Song
,
Fenju Qin
* and
Wenchao Zhang
*
School of Chemistry and Life Sciences, Suzhou University of Science and Technology, Suzhou 215009, China
*
Authors to whom correspondence should be addressed.
Microorganisms 2024, 12(4), 700; https://doi.org/10.3390/microorganisms12040700
Submission received: 23 February 2024 / Revised: 23 March 2024 / Accepted: 26 March 2024 / Published: 29 March 2024

Abstract

:
Heavy metals migrate easily and are difficult to degrade in the soil environment, which causes serious harm to the ecological environment and human health. Thus, soil heavy metal pollution has become one of the main environmental issues of global concern. Plant-growth-promoting rhizobacteria (PGPR) is a kind of microorganism that grows around the rhizosphere and can promote plant growth and increase crop yield. PGPR can change the bioavailability of heavy metals in the rhizosphere microenvironment, increase heavy metal uptake by phytoremediation plants, and enhance the phytoremediation efficiency of heavy-metal-contaminated soils. In recent years, the number of studies on the phytoremediation efficiency of heavy-metal-contaminated soil enhanced by PGPR has increased rapidly. This paper systematically reviews the mechanisms of PGPR that promote plant growth (including nitrogen fixation, phosphorus solubilization, potassium solubilization, iron solubilization, and plant hormone secretion) and the mechanisms of PGPR that enhance plant–heavy metal interactions (including chelation, the induction of systemic resistance, and the improvement of bioavailability). Future research on PGPR should address the challenges in heavy metal removal by PGPR-assisted phytoremediation.

1. Introduction

Since the 20th century, amidst the rapid development of the social economy, humans have wantonly burned fossil fuels, mined gold deposits, and used pesticides, fertilizers, and other chemicals, causing a significant rise in heavy metal pollution in the soil environment, both in terms of intensity and geographical extent [1]. According to the China National Soil Pollution Survey Bulletin, the heavy metals of Cd, Hg, As, Cu, Pb, Cr, Zn, and Ni in soil exceeded the acceptable standard by 7.0%, 1.6%, 2.7%, 2.1%, 1.5%, 1.1%, 0.9%, and 4.8%, respectively. They enter the soil environment through fertilization, irrigation, atmospheric settlement, and other means, causing soil and groundwater pollution [2]. Soil contaminated by heavy metals shows ecological imbalance, biodiversity loss, soil erosion, desertification, acidification, salinization, declining soil fertility, soil consolidation and soil subsidence [3,4]. Moreover, heavy metals can only be transferred from one chemical state to another due to non-biodegradability [5]. Because of their remarkable environmental persistence, heavy metals in soil will be absorbed and accumulate in plants, then pose hazards to humans and animals via the food chain [6]. Hence, it is crucial to remediate these toxic heavy metals in soil.
Although traditional physicochemical treatments such as soil replacement, soil washing, and chemical solidification have partly reduced the migration of soil heavy metals to groundwater, their applications are limited due to high energy consumption, secondary pollution, and the breakdown of soil aggregate structures [7]. In contrast, bioremediation (such as biosorption or bioaccumulation) is a simple, environmentally friendly and sustainable method that has raised significant interest [8]. Bioremediation relies on natural biological processes to eliminate toxic pollutants by means of microorganisms, green plants, or their enzymes. Among the various bioremediation methods, phytoremediation is widely accepted for its economic feasibility and eco-friendliness. However, low biomass yield, sensitivity to polymetallic substances, and shallow root systems limit the efficiency of phytoremediation. A promising method to improve the repair efficiency of plants with high metal concentrations is the employment of plant-growth-promoting rhizobacteria (PGPR) as a tool to alleviate the stress of heavy metals. It serves as a biotechnology and a sustainable and environmentally convenient alternative to enhance agricultural productivity [9].
In recent years, PGPR has garnered significant attention for its role in phytoremediation. According to statistics, the annual number of publications from 1995 to 2023 exhibited an exponential increase (Figure 1a). Notably, the years with high growth rates were 2019 (with a growth rate of 36.33% compared to the previous year) and 2020 (with a growth rate of 31.54% compared to the previous year). Such sharp growth is attributed to the global environmental concern for heavy metal pollution along with the rapid development of industry.
As shown in Figure 1b, the study of interaction of PGPR with various harmful heavy metals (especially Zn and Cd) has also soared in the last 5 years, indicating that considerable efforts have been devoted to the field of heavy metal treatment involving PGPR. PGPR can enhance the absorption and transport of heavy metals in soil plants by supplying nutrients, increasing plant yield, and enhancing stress resistance to heavy metals [10,11]. Utilizing PGPR as a bioinoculant enables nutrient recycling, soil structure stabilization, and the regulation of heavy metal bioavailability and toxicity, thereby promoting plant biomass and root growth [12].
PGPR creates a beneficial relationship with plants and has gained widespread interest in the fields of heavy metal stress management in agricultural and environmental research. In this paper, the detoxification mechanism of PGPR to heavy metals in soil is divided into two aspects: promoting plant growth and interacting with heavy metals, which can also be understood as endogenous accumulation and exogenous decreases in heavy metals via plants, respectively. Its recent applications in agricultural and environmental remediation, as well as the potential prospects and limitations of PGPR-assisted phytoremediation, are also discussed.

2. Plant-Growth-Promoting Rhizobacteria (PGPR)

The term “Rhizosphere” was first proposed by German microbiologist Lorenz Hiltner in 1904 to explain the impact of plant root exudates on soil microorganisms [13]. Rhizosphere microbes are the community of microbes that inhabit this area, mainly including bacteria, actinomycetes, fungi, algae, and viruses [14,15]. PGPR in the plant rhizosphere typically refers to beneficial, free-living bacteria in the soil that are connected to plant roots [16]. Plants raise the pressure on microorganisms to survive through allelopathy, enabling them to selectively favor bacteria that contribute most to their growth [17]. The concentration of rhizospheric bacteria is particularly high in plants because the roots of most plants can exude significant levels of nutrients (especially tiny molecules like amino acids, sugars, and organic acids), which provide energy for the growth and metabolism of bacteria [18].
In 1978, Kloepper and Schroth first introduced the concept of PGPR, defining it as a group of microorganisms that colonize the plant rhizosphere and promote plant growth through the plant rhizosphere [18]. Table 1 shows the typical species and primary mechanisms of PGPR in facilitating heavy metal extraction in plants.
PGPR presented in heavy-metal-contaminated soil can not only promote plant growth and enhance plant biomass, but also interact with heavy metals, thus improving the efficiency of the phytoremediation of heavy-metal-contaminated soil. From these two aspects, we will give a thorough overview of the detoxification mechanisms of PGPR.

3. PGPR Can Promote Plant Growth

The primary mechanism underlying its growth-promoting effects is depicted in Figure 2.

3.1. Biological Nitrogen Fixation

Nitrogen is not only one of the most important nutrition sources for plant growth, flowering and fruiting, but also an essential element of amino acids, protein, chlorophyll, nucleic acids, membrane lipids, ATP, NADH, co-enzymes, etc. [42,43]. Yet, only active types of oxidized (such as NOX) or reduced (such as NH3 and amines) nitrogen can be assimilated by plants [44]. The nitrogen needed for crop growth is usually supplemented by the application of nitrogen fertilizer. However, only 30–50% of the nitrogen fertilizer is absorbed by the plants, and the rest enters the aqueous environment, causing water eutrophication [45]. Therefore, finding an environmentally responsible strategy to ameliorate the situation is essential. PGPR acts as a microorganism that can coexist with plants to fix nitrogen; it has great value in agricultural application and environmental restoration.
Biological nitrogen fixation (BNF) refers to the process of converting nitrogen in the atmosphere into ammonia available to plants through the action of nitrogen-fixing microorganisms [46]. Common nitrogen-fixing microorganisms include Rhizobium spp., Acetobacter spp., Arthrobacter spp., Citrobacter spp., Clostridium spp., Streptomyces spp., and so on [47]. Among them, nitrogen-fixing rhizobia, containing nitrogenase, are a primary type of nitrogen-fixing microorganism, mainly inhabiting the roots of leguminous plants and forming nodules with plant roots. These nodules provide a suitable microenvironment for nitrogenase to convert nitrogen into ammonia to prevent it from being subjected to hypoxia, risking inhibitory activity. Plants provide a good living environment, carbon energy, and essential nutrients for rhizobia, and rhizobia provide nitrogen nutrients for plants [48,49,50]. The symbiosis of rhizobia and legumes is one of the most famous symbiotic relationships in nature. According to research, almost 70% of bio-fixed nitrogen comes from rhizobia and legumes, and rhizobia provide 90% of the nitrogen required by these plants [47]. Inoculation with Bradyrhizobium sp. increased both biomass and nitrogen content, confirming this point [10]. PGPR contains the key enzyme nitrogenase of BNF, with its activity typically regulated by the transcription of the nitrogen fixation gene (nif) [51]. Nitrogen fixation is instrumental in promoting plant growth, highlighting one of the vital roles of PGPR in enhancing plant vitality. Six common PGPR strains exhibited nitrogenase activity and significantly promoted the growth of wheat and spinach in the study by Cakmakcı et al. [52]. It was also verified that the inoculation of PGPR significantly increased the nitrogenase activity of Dalbergia sissoo seedlings and promoted its growth [53].
Typically, heavy metals can adversely affect the growth, nodules, nitrogenase activity, and nitrogen-fixing effects of legumes [54]. This has provided researchers with ideas for the selection of heavy-metal-resistant rhizobia. It was found that excessive amounts of heavy metals such as Cu2+ and Zn2+ reduced nitrogenase activity and nodule formation in alfalfa, while co-inoculation with alfalfa and rhizobium (Agrobacterium tumefaciens) could alleviate heavy metal stress and significantly increase nitrogenase activity and plant biomass [55]. Chen et al. also showed similar results, where inoculation with rhizobia alleviated copper-induced growth inhibition and significantly increased the nitrogen content and copper uptake of the plant [56]. Following their screening of Burkholderia spp., a Cd-resistant nitrogen-fixing bacterium, Shen et al. discovered that inoculating plants with Burkholderia spp. decreased Cd concentrations in their roots and leaves by 58.11% and 64.54%, respectively, and by 72.89% and 70.03%, respectively, when compared to uninoculated plants [57]. Furthermore, PGPR possesses plant-growth-promoting properties, thereby indirectly enhancing nitrogen bioavailability by promoting root surface area and morphology, resulting in higher nitrogen uptake.

3.2. Phosphate Solubilization

Phosphorus stands as a crucial nutrient vital for plant metabolism and nutrient cycling, ranking second only to nitrogen in significance [58,59,60]. Although soil harbors abundant phosphorus in both organic and inorganic forms, it often binds with calcium ions, iron ions, aluminum ions, and others, rendering it challenging for plants to uptake and utilize effectively [59]. To bolster phosphorus nutrition for crops and enhance yields, agricultural chemicals such as phosphorus-based fertilizers are commonly employed. However, excessive phosphorus has caused serious environmental pollution issues, including a decline in soil fertility, an imbalance in the soil microbial system, and water eutrophication [61]. Hence, there is an urgent demand for a cost-effective and environmentally responsible alternative to the use of phosphorus fertilizers.
One of the most promising avenues for developing durable and secure technology is the application of phosphorus-solubilizing bacteria (PSB) [62], which plays an important role in phosphorus cycling and promoting plant growth. It has been demonstrated that PGPR functions as a PSB to transform insoluble phosphorus into a form that plants can use [63], mainly by secreting organic acids (like citric acid, oxalic acid, and gluconic acid), chelating metal ions to form soluble complexes (such as phosphates of calcium, iron, and aluminum), and by producing enzymes (pyridoxal phosphatase (PDXP), phytase, C-P lyase) to hydrolyze organic phosphorus in soil into inorganic forms. The phosphorus-dissolving pathways assisted by PGPR are summarized in Figure 3.
Pseudomonas spp., Enterobacter spp., Bacillus spp., Serratia spp., Arthrobacter spp., Burkholderia spp., Aspergillus spp., Gongronella spp., Penicillium spp., Talaromyces spp., and so on, have been identified as PSB [63,64,65]. In addition to providing plants with essential P nutrients, PSB also produce other metabolites, like indole-3-acetic acid (IAA) and siderophore, that are highly effective in promoting plant growth. According to Prakash et al., the PSB Bacillus sp. STJP can promote crop growth by generating IAA, siderophores, and phosphorus solubilization [66].
Furthermore, PSB have demonstrated the ability to facilitate the mineralization of heavy metals and generate stable mineralization products [67]. Teng et al. showed that Pseudomonas L1–5 could transform Pb2+ into lead hydroxyapatite and lead pyroalumite by dissolving phosphorus, showcasing the promising application potential in the bioremediation of lead-contaminated soil [68]. The effectiveness of phosphorus solubilization is influenced by the bacterial species. For instance, when 800 mg/kg of rock phosphate was introduced, Pantoea sp. and Enterobacter sp. significantly enhanced P solubilization by 49.9% and 88.6%, respectively, which led to a 13.7–26.4% increase in the immobilization of Pb [69].
In conclusion, PSB can relieve heavy metal stress in plants in the following ways:
(i) Secreting organic acids and phosphatases to dissolve insoluble phosphorus and provide necessary phosphorus nutrition for plants to promote plant growth; (ii) secreting IAA and siderophores, which provides the potential to promote plant growth; (iii) inducing heavy metal mineralization to fix heavy metals and reduce their bioavailability.

3.3. Siderophore Production

Iron is one of the essential nutrients for plants and is crucial for many cell functions. Despite being widely distributed throughout the Earth’s crust, iron is typically found in an insoluble oxidation state that makes it difficult for plant roots to absorb and use [70]. Under low iron stress, some microorganisms and crops can produce a low-molecular-weight organic molecule called siderophore that has a strong affinity for Fe3+ [71]. Therefore, an iron absorption system is formed, with siderophore serving as the primary material: The siderophore chelates with insoluble iron to form a complex, and the membrane receptor transports the complex across the microbial membrane. Following the enzyme system reaction in the cell, the Fe3+ bound to the iron carrier is re-released; then, Fe3+ is reduced to Fe2+, which can be absorbed and utilized by most plants, promoting plant growth [63]. Furthermore, siderophores can also chelate with other toxic metal ions to reduce their concentration. In this regard, siderophores also protect cells from oxidative damage and promote plant growth by preventing oxidative stress. Studies have shown that intracellular reactive species of superoxide (O2·) or hydrogen peroxide (H2O2) can destroy Fe–S clusters, leading to excess free Fe2+ released from the damage cluster. The released Fe2+ subsequently can be catalyzed by the Fenton reaction to produce harmful hydroxyl radicals (OH·) which can attack DNA, and the siderophores accumulated in cells can chelate the release of Fe2+, thus inhibiting the formation of OH· [72,73]. Pyoverdine (PVDI), a fluorescent iron carrier, can protect Pseudomonas aeruginosa from oxidative damage by chelating with released ferrous, inhibiting the formation of OH· [73]. A ginseng rhizosphere strain, DCY119T, when ingested by iron-stressed seedlings, can produce siderophores and get rid of surplus ROS, thereby lowering oxidative stress [74].
In iron-poor environments, PGPR can outcompete rhizosphere pathogenic microorganisms for scarce iron resources. This competitive advantage inhibits the proliferation and propagation of pathogenic microorganisms, thereby making a substantial contribution to the biological control of fungal infections [70]. Di et al. demonstrated that Aureobasidium pullulans L1 and L8 effectively compete with Monilinia laxa for iron resources through the secretion of siderophores, diminishing the mycelial growth and conidial germination of M. laxa, thus providing protection to peach fruit from post-harvest decay [75]. In addition, compared with plant-derived siderophores, PGPR that can secrete siderophores exhibit greater efficiency in binding Fe3+, attributed to their superior solubility and stronger affinity for metals [76].

3.4. Phytohormone Production

As a natural source of phytohormones, PGPR may produce a variety of phytohormones concurrently, which is crucial for promoting root development, nutrient uptake, biomass synthesis, and other characteristics of the host plants [77]. Phytohormones secreted by PGPR, like auxin (IAA), cytokinin (CTK), gibberellin (GA), and ethylene (ETH) [78], make a significant contribution to plant growth under heavy metal stress conditions [79].

3.4.1. Auxin

Auxin is a kind of low-molecular-weight molecule that naturally exists in plants and is involved in processes like organ development, root hair production, and bud elongation [80,81], such as indole-3-acetic acid (IAA) [82,83]. Many PGPR, including Azospirillum spp., Alcaligenes spp., Klebsiella spp., Enterobacter spp., Acetobacter spp., Bacillus spp., Pseudomonas spp., Xanthomonas spp., Rhizobium spp., Arthrobacter spp., and Bradyrhizobium spp., have been reported to synthesize IAA [84,85]. In heavy metal pollution areas, PGPR can stimulate plant root growth by secreting IAA, which increases the plant root surface area and enhances access to soil nutrients and heavy metal accumulation, so as to improve the efficiency of the plant root absorption of heavy metals [70,86]. Ma et al. found that PGPR increased Cu and Zn content in Brassica oxyrrhina by 146% and 61% [86], and the inoculation of the PGPR strain significantly increased the shoot copper content of maize and sunflower [87]. The higher ability of heavy metal accumulation in plants inoculated with PGPR was related to the high level of IAA produced by bacterial strains, which increased the surface area and length of roots, providing plants with more soil nutrients and heavy metals. Thus, bacteria that produce IAA in rhizosphere soil are considered to play an important role in heavy metal uptake by plants. Carlos et al. reported that stress from Pb, Cu, and As significantly increased the amount of IAA synthesized by Serratia K120, Enterobacter K131, Enterobacter N9, and Escherichia coli N16, promoting the growth of sunflower plants, which has potential application in phytoremediation systems [88]. An experiment in which researchers injected IAA-deficient mutants into wheat (Triticum aestivum) and tracked the growth response showed that wheat infected with IAA-deficient mutants of P. moraviensis grew more slowly and presented smaller roots than wild-type strains, verifying the significance of PGPR in encouraging plant development with the production of IAA [89].

3.4.2. Cytokinin

Cytokinin (CTK) is a hormone widely found in higher plants, algae, and bacteria [70], and is the second most important phytohormone after IAA [90]. Cytokinin stimulates cell division, cell expansion, and tissue expansion, and plays a vital function in promoting chloroplast development, vascular development, and bud differentiation in plants [70,91,92]. Currently, numerous studies have shown that PGPR can produce cytokinins and promote plant growth under heavy metal stress [93]. For example, Wang et al. inoculated Arabidopsis thaliana with a cytokinin-producing PGPR, Bacillus megalosporum, and found that elevated transcript levels of the cytokinin receptors in plant root shoots and roots significantly promoted plant growth [94]. According to Piotrowska, cytokinin significantly alleviated the growth inhibition of the green alga Acutodesmus obliquus and enhanced its ability to combat lead toxicity under lead stress [95]. Similarly, Yu et al. reported that cytokinin plays a crucial role in increasing the biomass and cadmium uptake efficiency of the super-accumulating plant Sedum alfredii [96].
Nieto and Frankenberger demonstrated a significant improvement in the growth state of plants inoculated with Azotobacter chococcum. The dry weight of root and shoot tissue was reported to be 5.57 times and 5.01 times higher, respectively, compared to the blank group. They concluded that the increased plant yield was primarily attributed to the generation of cytokinin by rhizosphere bacteria [97]. The application of thiadiuride and kinetin, cytokinin-like agents, on maize (Zea mays L.) seedlings can alleviate the deterioration effects of heavy metals on seed germination ability, seedling growth, and membrane permeability [98]. Similarly, the application of kinetin increased the photosynthetic rate of pea plants under Cd stress [99]. Therefore, cytokinins also have an important role in alleviating plant toxicity from heavy metals.

3.4.3. Aminocyclopropane-1-Carboxylate (ACC) Deaminase

A key regulator of plant growth, ethylene, is a gaseous plant hormone. Low levels of ethylene regulate bud and root growth, encourage flowering, fruit ripening and shedding, and leaf aging and shedding, and significantly influence how plants react to abiotic and biological stimuli [100,101]. Conversely, high levels of ethylene inhibit normal plant growth and promote plant senescence and even premature death [70]. However, plants can experience large amounts of ethylene production under heavy metal stress. Therefore, the proper regulation of ethylene level is crucial for the growth of plants stressed by heavy metal stress [47].
Ethylene is synthesized by two exclusive enzymatic reactions. In the first step, the substrate S-adenoyl-L-methionine (SAM) is converted by ACC synthetase (ACS) to ACC and 5′-methylthioadenosine (MTA). In the second step, ACC is converted to ethylene, CO2, and cyanide by ACC oxidase (ACO). Among them, ACC is a direct precursor of ethylene, so ethylene levels can be controlled by regulating ACC levels [102]. If reducing the activity of ACS and ACO, ethylene synthesis could be inhibited to some extent. In the Misra study, inoculation with PGPR significantly reduced the ACS and ACO activity of Zea mays [103]. The inoculation of wheat plants with the PGPR strain Anospira brasiliensis FP2 resulted in a drop in ACO expression in vivo, resulting in the reduced ethylene content of wheat roots [104]. Similarly, volatile organic compounds emitted by Bacillus subtilis SYST2 repressed the transcription of the ACO1 gene and caused a decrease in endogenous ethylene content in tomato seedlings [105].
Moreover, there is an enzyme that can decompose ethylene, making a great contribution to reducing ethylene content, namely ACC deaminase (ACCD). This enyzme can decompose the ethylene precursor ACC into α-ketobutyrate and ammonia [47,106,107]. Glick et al. concluded from model prediction analysis that almost all PGPR have ACC deaminase activity [108]. Therefore, PGPR inoculation has a positive impact on the accumulation of heavy metals in plants. Sunflowers grew better and accumulated more Zn and Pb in the presence of PGPR (Bacillus safensis FO-036b(T) and Pseudomonas fluorescens p.f.169), which contains ACC deaminase, according to research by Mousavi et al. [109]. Carlos et al. discovered that the rhizobium strains of Enterobacter N9, Serratia K120, Klebsiella Mc173, and Escherichia coli N16 that produce ACCD dramatically lengthened cauliflower’s shoots and roots, thus accelerating the phytoremediation of metals such as copper, nickel, zinc, lead, and arsenic [88]. In summary, it is evident that PGPR, particularly those capable of producing ACC deaminase, play a significant role in aiding plant growth and improve the accumulation of heavy metals.

3.4.4. Gibberellins

Bioactive gibberellin (GA) is a diterpenoid plant hormone that is involved in several plant developmental processes, including seed dormancy, germination, flowering, fruit ripening, root growth, and root hair enrichment, through complex biosynthesis [70,110,111]. At present, more than 130 kinds of GAs have been found. PGPR mainly produce GA1, GA3, GA4, and GA20, with GA3 being the most prevalent form [70,112]. Typical PGPR species that can produce GA include Acetobacter spp., Bacillus spp., and Azotobacter spp. [70,113].
GA can improve plant adaptation to heavy metal toxicity. One study showed that the application of GA3 alleviated the toxicity of Cu stress to pea (Pisum sativum L.) seedlings [114]. Under Cd stress, GA application on Cyphomandra betacea promoted beetroot seedling growth and increased the biomass, leaf net photosynthetic rate, and carotenoid and soluble sugar content. The Cd content of C. betacea seedlings reduced gradually with increasing concentrations of GA [115]. For hyperaccumulator Sedum alfredii, the application of GA significantly increased the dry biomass of the root, stem, leaf, and shoot. The enhanced accumulation of Cd and Pb in the shoot of S. alfredii demonstrates significant potential for heavy metal phytoremediation [116].

4. Interactions between PGPR and Heavy Metals

4.1. Chelation

4.1.1. Exopolysaccharide Production

Exopolysaccharides (EPSs) are a kind of high-molecular-weight natural polymer secreted by microorganisms, and include sugars, proteins, amino acids, lipids, and other substances. They typically build up on the surface of bacterial cells and serve as a barrier against harmful external factors such as phagocytosis, pathogen attachment, drying, pH, and heavy metals [78,117,118]. According to numerous studies, EPSs can reduce toxicity by adsorbing heavy metals through electrostatic contact, ion exchange, complexation, surface precipitation, redox, and other interactions [119,120]. The mechanism illustrating how EPSs protect cells from heavy metal ions is depicted in Figure 4. Without EPS protection, metal ions will directly contact the cell and react with specific proteins or enzymes within the cell, causing a loss of their activity and eventual cell damage.
Electrostatic interactions between functional groups like hydroxyl, acetylamino, and amino groups in EPSs and positively charged metal cations (Cd2+, Pb2+, Ni2+, Co2+, and Cr5+) result in the formation of organometallic complexes on the surface of cells [120]. Ion exchange is a popular mechanism to explain the metal bio-sorption by EPSs [121,122]. The heavy metals chromium and cadmium form a combination with the EPSs generated by PGPR (Azotobacter spp.), which decreases their ability to migrate and lessens the pressure on wheat growth. Higher pH values typically result in fewer protons in solution being used to compete with metals for binding locations in EPSs, leading to increased metal adsorbing in the ion exchange equilibrium [123]. Additionally, EPSs can play a role in mitigating the toxicity of heavy metals through surface precipitation or redox reactions. The chemical properties of the solution, especially the pH level, have a profound impact on the shape and solubility of metals in aqueous solutions. As the pH rises, most metal ions undergo transformation from hydrated metal cations to hydroxylated monomers and polymers. Over time, this process leads to the formation of crystalline oxide precipitation [124,125].
It has been reported that EPS components facilitate the enzymatic reduction in Cr (VI) by interacting with the toxic group of carboxyl and hydroxyl groups, thereby binding Cr (VI). Additionally, protein components within EPSs may also contribute to reductions in Cr (VI). Furthermore, primary functional groups, such as COO- and -OH groups in EPSs, can bind Cr (VI) to shield cells from the toxic effects of Cr (VI) [126,127]. EPSs may also resist toxic metals by trapping them outside cells or forming biofilms. These two pathways can effectively mitigate the environmental toxicity of metals and enhance the symbiotic development and growth of legumes in metal-contaminated soils [128].

4.1.2. Metallothionein Production

Metallothioneins (MTs) are a family of universal low-molecular-weight, cysteine-rich (about a third of their amino acid content) proteins [129], widely distributed in eukaryotes such as fungi, plants, animals, and some prokaryotes [130,131]. They can form sulfur-based metal clusters by binding to a large number of metal ions, such as Zn2+, Cd2+, and Cu2+, through the thiol group of their cysteine residues, which play an important role in regulating the dynamic balance and detoxification of plant metal ions [132,133].
Recombinant strains with MTs were found to contribute to plants binding heavy metals in soil and acted as free-radical scavengers [134]. Murthy et al. treated plants with PGPR strains with bacterial MTs and found that plants inoculated with PGPR enhanced the bioremediation process in metal-infected soils and had positive results for the removal of heavy metals such as Pb2+ [135]. In tobacco leaves, MTs can regulate the responses of Silene to Cu stress [136].

4.1.3. Soil Organic Acid Production

Some PGPR can secrete organic acids, including formic acid, acetic acid, tartaric acid (TA), succinic acid, oxalic acid (OA), citric acid (CA), and gluconic acid (GA) [137,138]. For some super-enriched plants, the release of organic acids is one of the mechanisms for the migration of heavy metals (by altering their mobility) [139]. For instance, soil PSB strains can secrete large amounts of gluconic acid, thereby improving cadmium mobility and increasing the bioavailability [140]. Organic acids can also enhance the metal mobility in soil by reducing soil pH and forming complexes with heavy metals [141], thus increasing the absorption of heavy metals in the rhizosphere soil by some plants. Previous studies have shown that low-molecular-weight organic acids (LMWOAs) secreted by plant roots can form soluble Cd-LMWOA complexes with Cd, making it more easily absorbed by plants [142]. Similarly, treatment with tartaric acid, malic acid, oxalic acid, and citrate acid considerably raised root and stem Cd concentrations compared to the control, having a favorable impact on the uptake of total Cd in the soil [143]. Some studies used Brassica juncea L. to explore the effects of different doses of CA on heavy metal accumulation and stress tolerance and showed that the exogenous application of CA in growth medium containing Cd significantly reduced the harmful effects of Cd on plants [144].
As early as 1998, Hassen et al. reported that the presence of citric acid increased the biosorption of Cr by Pseudomonas aeruginosa and Cu by Bacillus thuringiensis [145]. Farid et al. found that the combined application of citric acid and 5-aminolevulinic acid could improve the biomass, photosynthesis, and gas exchange characteristics of sunflowers in Cr-contaminated soil [146]. Chai et al. showed that the content of oxalic acid in the inflorescence, stem, and fine root of Saussurea involucrata was positively correlated with the bioaccumulation of Cd [147]. Similarly, Chen et al. found that tartaric acid or malic acid can effectively improve the growth potential of hybrid flowers under Cd stress [148]. Tomato seedlings inoculated with Pseudomonas aeruginosa and Burnetidia eriagladioli reduced Cd-induced toxicity by upregulating the levels of LMWOAs (fumaric acid, malic acid, succinic acid, and citric acid), which further proves the important role of PGPR in the secretion of organic acids in plants to alleviate heavy metal stress [149].

4.2. Induced Systematic Resistance

Induced systematic resistance (ISR) is a state of enhanced defense formed in plants by activating potential resistance, and is induced by various factors, such as rhizosphere bacteria [150]. ISR initiates multiple potential defense mechanisms, including increasing the activity of chitinase, β-1,3 glucanase, and peroxidase; accumulating antimicrobial low-molecular substances such as phytolexins; and forming protective biopolymers, viz. lignin, callose, and hydroxyproline-rich glycoprotein [151]. Given that the efficacy of ISR hinges on defense mechanisms triggered by inducers, employing natural PGPR strains as inducers of plant defense responses could enhance their practicality and effectiveness.
It was found that PGPR can complete the ISR process by increasing the activity of antioxidant enzymes [152,153]. As shown in Figure 5, under heavy metal stress, plant cells will release excessive ROS, resulting in oxidative damage. However, PGPR promote the synthesis of a variety of antioxidant enzymes by plant cells to resist this oxidative damage, such as superoxide dismutase (SOD), peroxidase (POD), catalase (CAT), ascorbic peroxidase (APX), etc. Thich can effectively remove ROS from plant cells and is conducive to plant growth [154]. Gururani et al. isolated two PGPR strains from the potato rhizosphere, Bacillus pumilus strain DH-11 and Bacillus firmus strain 40, which improved the Zn tolerance of potato by up-regulating the gene transcription levels of ROS-scavenging enzymes (APX, SOD, CAT, DHAR, and GR) in potato plants, contributing to increased plant tolerance to Zn [155]. According to Ju et al., PGPR and rhizobium-inoculated alfalfa could decrease the accumulation of ROS, increase the activity of antioxidant enzymes, and enhance the copper tolerance of alfalfa [156]. Sunflower and tomato can increase the synthesis of SOD and CAT, reduce the production of ROS, and enhance Cr tolerance after being inoculated with Pseudomonas sp. [157]. PGPR significantly improved the antioxidant system produced by Cd induction by upregulating the mRNA expression of SOD, POD, and PPO genes in tomato [158].

4.3. Transform Toxic Heavy Metals

Bacteria in the soil can convert toxic heavy metals into forms that are easily absorbed by plant roots [159], thereby improving the efficiency of phytoremediation. PGPR can enhance the bioavailability of heavy metals, transforming them from insoluble forms to soluble forms [160]. This process is also known biodegradation [161]. For example, the conversion of selenate to organic Se in the presence of bacteria can increase selenium accumulation in plants [162]. Jeong et al. showed that Bacillus megaliium significantly increased the amount of exchangeable Cd in rhizosphere soil [163]. The inoculation of sunflower with Micrococcus sp. MU1 and Klebsiella sp. BAM1 (Cd-tolerant PGPR strain) could increase the concentration of water-soluble Cd in rhizosphere soil, thus enhancing the absorption of Cd by sunflower roots [164]. For instance, biosurfactants produced by Pseudomonas aeruginosa BS2 lead to the solubilization of Pb and Cd [165]. Bacillus subtilis, B. cereus, Flavobacterium sp., and Pseudomonas aeruginosa increased the availability of water-soluble Zn in soil and Zn accumulation by plants [166].

5. Conclusions and Future Perspectives

Heavy metals have led to severe environmental pollution problems, prompting extensive research into heavy metal detoxification methods. Considering the economic and environmental benefits, the PGPR-assisted phytoremediation of heavy metal pollution emerges as a promising environmentally friendly strategy. By fixing nitrogen, dissolving phosphorus, secreting plant growth hormones, and increasing the bioavailability of heavy metals, PGPR can stimulate plant growth through their metabolic processes, enhancing the effectiveness of soil heavy metal remediation. At the same time, PGPR affect plant cell physiological activities, induce ISR, activate plant antioxidant enzymes, increase iron supply by secreting high-affinity siderophores, competitively inhibit the uptake of heavy metals by plant roots, and influence the absorption, transportation, and intracellular distribution of heavy metals. These actions help alleviate heavy metal stress, enhance heavy metal tolerance, and ultimately improve the efficiency of phytoremediation. Under various heavy metal stress environments, PGPR play an important role in the migration and transformation of heavy metals in rhizosphere soil and in helping plants to absorb and accumulate heavy metals or immobilize the soil.
Future research directions for PGPR include the following: further expanding the PGPR strain bank; devising strategies for the synergistic use of multiple PGPR strains; sourcing mixtures of multiple PGPR strains directly from natural environments; and developing macromolecular materials such as biochar combined with PGPR for phytoremediation. Moreover, there is potential to broaden the application scope of PGPR to areas such as sludge treatment, sewage purification, mine reclamation, sediment restoration, and more. It is imperative for researchers to continue advancing their understanding of PGPR and promptly translate this knowledge into practical applications.

6. Problems in the Practical Application of PGPR in Environmental Remediation

In view of the previous research progress and looking forward to research work in the future, more attention should be paid to the following aspects.
First, the use of PGPR in agricultural management is constrained by factors like short survival times, low survival rates, and the uniqueness of effective strains in actual environments. Mathematical modeling and computer-based simulation results indicate that the competition for limited resources between PGPR populations and resident microorganisms is the most important factor for determining the survival of PGPR. The most effective PGPR application is in organic and mineral-poor soil or stressed soil, because the development of resident microflora is inhibited [167]. Currently, the majority of PGPR utilized in research are isolated from contaminated areas, which is time-consuming and laborious if they need to be separated and screened from the target area before each governance. This greatly limits the application of PGPR. It is worth noting whether high-performance strains can be developed in the future, or introducing other organisms or vectors to enhance the survival rate and efficiency of PGPR.
Second, given that various strains of plants have varying capacities for tolerating certain heavy metals, care should be taken while applying medications in real agricultural applications to prevent unneeded losses. To maximize their tolerance to heavy metals through domestication, multiple PGPR repair combinations should be developed to receive multiple PGPR mixtures directly from the environment.
Third, current studies primarily focus on the individual management of heavy metal stress within laboratory settings. However, actual heavy-metal-polluted land is often accompanied by chemical pollution, drought, salt stress, and other harsh conditions. Moreover, plant hormone regulation and interactions vary across different stress environments. Therefore, future research on the mechanism of PGPR in improving plant resistance to heavy metal toxicity should be further explored in field studies, and the different responses of plants under common stresses of two or more kinds of pollution should be focused on. And on the basis of screening excellent functional PGPR strains, the development of special PGPR strains suitable for different crops, different geologies, and different temperatures could play an important role in dealing with different soil types and climatic conditions, which is the direction and goal of PGPR strain research and development.
Fourth, there is currently limited research on the long-term outcomes for plants after the phytoremediation of heavy metal pollution. Most studies tend to focus on the restored soil, which represents a drawback of phytoremediation. Certain heavy metals like As, Se, and Hg can be absorbed by accumulator plants and subsequently volatilized into the atmosphere as gases, a process known as phytovolatilization [168]. In this process, the metal content in plants remains relatively low, suggesting a long-term remediation strategy. However, for plants unable to undergo phytovolatilization to accumulate excessive heavy metals, incineration might be the optimal solution, by which heavy metal oxides can be recovered and plant biochar—an environmentally friendly material—can be produced, thus achieving a green cycle.
Other suitable application of PGPR should also be paid attention to. PGPR are employed to safeguard plants from pathogens and mitigate environmental stress through a range of biological control mechanisms, including the secretion of antibiotics, the induction of systemic resistance, and the induction of systemic tolerance, among others. In agricultural production, plant diseases exert a serious impact on the growth and development of plants. However, the adoption of this technology is currently limited, primarily due to a lack of awareness among most farmers and variations in farmland practices across different locations [169]. In light of the aforementioned discrepancies, researchers need to devise more user-friendly strategies for the application of PGPR. Farmers are more prone to employ the process if it is straightforward and demonstrably effective.
In conclusion, PGPR are anticipated to replace chemical fertilizers, insecticides and synthetic growth regulators in the future, thereby fostering sustainable agricultural growth and facilitating the bioremediation of heavy metal contamination in the environment.

Author Contributions

H.Q.: conceptualization, methodology, writing—original draft preparation, writing—reviewing and editing. Z.W.: investigation, literature search. W.S.: investigation. S.S.: investigation. F.Q.: writing—reviewing and editing. W.Z.: conceptualization, writing—reviewing and editing, supervision, coordination, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (grant number 22006110), the Natural Science Foundation of Jiangsu Province of China (grant number BK20200987), and the Science and Technology Planning Project of the Ministry of Housing, Urban-Rural Development of Jiangsu province (grant number 2020ZD12).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, Q.; Wang, C. Natural and human factors affect the distribution of soil heavy metal pollution: A review. Water Air Soil Pollut. 2020, 231, 350. [Google Scholar] [CrossRef]
  2. Hou, D. Sustainable remediation in China: Elimination, immobilization, or dilution. Environ. Sci. Technol. 2021, 55, 15572–15574. [Google Scholar] [CrossRef] [PubMed]
  3. Keesstra, S.; Mol, G.; De Leeuw, J.; Okx, J.; De Cleen, M.; Visser, S. Soil-related sustainable development goals: Four concepts to make land degradation neutrality and restoration work. Land 2018, 7, 133. [Google Scholar] [CrossRef]
  4. Vaverková, M.D.; Maxianová, A.; Winkler, J.; Adamcová, D.; Podlasek, A. Environmental consequences and the role of illegal waste dumps and their impact on land degradation. Land Use Policy 2019, 89, 104234. [Google Scholar] [CrossRef]
  5. Naila, A.; Meerdink, G.; Jayasena, V.; Sulaiman, A.Z.; Ajit, A.B.; Berta, G. A review on global metal accumulators—Mechanism, enhancement, commercial application, and research trend. Environ. Sci. Pollut. Res. 2019, 26, 26449–26471. [Google Scholar] [CrossRef] [PubMed]
  6. Leong, Y.K.; Chang, J.-S. Bioremediation of heavy metals using microalgae: Recent advances and mechanisms. Bioresour. Technol. 2020, 303, 122886. [Google Scholar] [CrossRef]
  7. Padhan, D.; Rout, P.P.; Kundu, R.; Adhikary, S.; Padhi, P.P. Bioremediation of heavy metals and other toxic substances by microorganisms. In Soil Bioremediation: An Approach towards Sustainable Technology; John Wiley & Sons: Hoboken, NJ, USA, 2021; pp. 285–329. [Google Scholar]
  8. Dixit, R.; Malaviya, D.; Pandiyan, K.; Singh, U.B.; Sahu, A.; Shukla, R.; Singh, B.P.; Rai, J.P.; Sharma, P.K.; Lade, H. Bioremediation of heavy metals from soil and aquatic environment: An overview of principles and criteria of fundamental processes. Sustainability 2015, 7, 2189–2212. [Google Scholar] [CrossRef]
  9. Pantoja-Guerra, M.; Valero-Valero, N.; Ramírez, C.A. Total auxin level in the soil–plant system as a modulating factor for the effectiveness of PGPR inocula: A review. Chem. Biol. Technol. Agric. 2023, 10, 6. [Google Scholar] [CrossRef]
  10. Dary, M.; Chamber-Pérez, M.; Palomares, A.; Pajuelo, E. “In situ” phytostabilisation of heavy metal polluted soils using Lupinus luteus inoculated with metal resistant plant-growth promoting rhizobacteria. J. Hazard. Mater. 2010, 177, 323–330. [Google Scholar] [CrossRef]
  11. De Souza, L.A.; de Andrade, S.A.L.; de Souza, S.C.R.; Schiavinato, M.A. Arbuscular mycorrhiza confers Pb tolerance in Calopogonium mucunoides. Acta Physiol. Plant. 2012, 34, 523–531. [Google Scholar] [CrossRef]
  12. Ahemad, M. Remediation of metalliferous soils through the heavy metal resistant plant growth promoting bacteria: Paradigms and prospects. Arab. J. Chem. 2019, 12, 1365–1377. [Google Scholar] [CrossRef]
  13. Hartmann, A.; Rothballer, M.; Schmid, M. Lorenz Hiltner, a pioneer in rhizosphere microbial ecology and soil bacteriology research. Plant Soil 2008, 312, 7–14. [Google Scholar] [CrossRef]
  14. Mccully, M.; Harper, J.; An, M.; Wu, H.; Kent, J.H. The rhizosphere: The key functional unit in plant/soil/microbial interactions in the field. implications for the understanding of allelopathic effects. Pol. J. Vet. Sci. 2005, 15, 493–498. [Google Scholar]
  15. Jha, C.K.; Saraf, M. Plant growth promoting rhizobacteria (PGPR): A review. J. Agric. Res. Dev. 2015, 5, 108–119. [Google Scholar]
  16. Barriuso, J.; Ramos Solano, B.; Lucas, J.A.; Lobo, A.P.; García-Villaraco, A.; Gutiérrez Mañero, F.J. Ecology, genetic diversity and screening strategies of plant growth promoting rhizobacteria (PGPR). In Plant-Bacteria Interactions: Strategies and Techniques to Promote Plant Growth; John Wiley & Sons: Hoboken, NJ, USA, 2008; pp. 1–17. [Google Scholar]
  17. Penrose, D.M.; Glick, B.R. Levels of ACC and related compounds in exudate and extracts of canola seeds treated with ACC deaminase-containing plant growth-promoting bacteria. Can. J. Microbiol. 2001, 47, 368–372. [Google Scholar] [CrossRef] [PubMed]
  18. Kloepper, J.W. Plant growth-promoting rhizobacteria on radishes. In Proceedings of the 4th International Conference on Plant Pathogenic Bacter, Station de Pathologie Vegetale et Phytobacteriologie, INRA, Angers, France, 2 September–27 August 1978; pp. 879–882. [Google Scholar]
  19. Liu, A.; Wang, W.; Zheng, X.; Chen, X.; Fu, W.; Wang, G.; Ji, J.; Jin, C.; Guan, C. Improvement of the Cd and Zn phytoremediation efficiency of rice (Oryza sativa) through the inoculation of a metal-resistant PGPR strain. Chemosphere 2022, 302, 134900. [Google Scholar] [CrossRef]
  20. Asadullah, A.B.; Javed, H. PGPR assisted bioremediation of heavy metals and nutrient accumulation in Zea mays under saline sodic soil. Pak. J. Bot. 2021, 53, 31–38. [Google Scholar]
  21. Wu, J.; Kamal, N.; Hao, H.; Qian, C.; Liu, Z.; Shao, Y.; Zhong, X.; Xu, B. Endophytic Bacillus megaterium BM18-2 mutated for cadmium accumulation and improving plant growth in Hybrid Pennisetum. Biotechnol. Rep. 2019, 24, e00374. [Google Scholar] [CrossRef] [PubMed]
  22. Kamal, N.; Liu, Z.; Qian, C.; Wu, J.; Zhong, X. Improving hybrid Pennisetum growth and cadmium phytoremediation potential by using Bacillus megaterium BM18-2 spores as biofertilizer. Microbiol. Res. 2021, 242, 126594. [Google Scholar] [CrossRef]
  23. Tirry, N.; Kouchou, A.; El Omari, B.; Ferioun, M.; El Ghachtouli, N. Improved chromium tolerance of Medicago sativa by plant growth-promoting rhizobacteria (PGPR). J. Genet. Eng. Biotechnol. 2021, 19, 149. [Google Scholar] [CrossRef]
  24. Wang, T.; Wang, S.; Tang, X.; Fan, X.; Yang, S.; Yao, L.; Li, Y.; Han, H. Isolation of urease-producing bacteria and their effects on reducing Cd and Pb accumulation in lettuce (Lactuca sativa L.). Environ. Sci. Pollut. Res. 2020, 27, 8707–8718. [Google Scholar] [CrossRef] [PubMed]
  25. Danish, S.; Kiran, S.; Fahad, S.; Ahmad, N.; Ali, M.A.; Tahir, F.A.; Rasheed, M.K.; Shahzad, K.; Li, X.; Wang, D. Alleviation of chromium toxicity in maize by Fe fortification and chromium tolerant ACC deaminase producing plant growth promoting rhizobacteria. Ecotoxicol. Environ. Saf. 2019, 185, 109706. [Google Scholar] [CrossRef]
  26. Samreen, T.; Zahir, Z.A.; Naveed, M.; Asghar, M. Boron tolerant phosphorus solubilizing Bacillus spp. MN-54 improved canola growth in alkaline calcareous soils. Int. J. Agric. Biol. 2019, 21, 538–546. [Google Scholar]
  27. Abdelkrim, S.; Jebara, S.H.; Saadani, O.; Chiboub, M.; Abid, G.; Jebara, M. Effect of Pb-resistant plant growth-promoting rhizobacteria inoculation on growth and lead uptake by Lathyrus sativus. J. Basic Microbiol. 2018, 58, 579–589. [Google Scholar] [CrossRef]
  28. Gupta, P.; Kumar, V.; Usmani, Z.; Rani, R.; Chandra, A. Phosphate solubilization and chromium (VI) remediation potential of Klebsiella sp. strain CPSB4 isolated from the chromium contaminated agricultural soil. Chemosphere 2018, 192, 318–327. [Google Scholar] [CrossRef] [PubMed]
  29. Mitra, S.; Pramanik, K.; Ghosh, P.K.; Soren, T.; Sarkar, A.; Dey, R.S.; Pandey, S.; Maiti, T.K. Characterization of Cd-resistant Klebsiella michiganensis MCC3089 and its potential for rice seedling growth promotion under Cd stress. Microbiol. Res. 2018, 210, 12–25. [Google Scholar] [CrossRef]
  30. Biswas, J.K.; Mondal, M.; Rinklebe, J.; Sarkar, S.K.; Chaudhuri, P.; Rai, M.; Shaheen, S.M.; Song, H.; Rizwan, M. Multi-metal resistance and plant growth promotion potential of a wastewater bacterium Pseudomonas aeruginosa and its synergistic benefits. Environ. Geochem. Health 2017, 39, 1583–1593. [Google Scholar] [CrossRef]
  31. Hansda, A.; Kumar, V. Cu-resistant Kocuria sp. CRB15: A potential PGPR isolated from the dry tailing of Rakha copper mine. 3 Biotech 2017, 7, 132. [Google Scholar] [CrossRef]
  32. Khan, W.U.; Yasin, N.A.; Ahmad, S.R.; Ali, A.; Ahmed, S.; Ahmad, A. Role of Ni-tolerant Bacillus spp. and Althea rosea L. in the phytoremediation of Ni-contaminated soils. Int. J. Phytoremediat. 2017, 19, 470–477. [Google Scholar] [CrossRef]
  33. Sobariu, D.L.; Fertu, D.I.T.; Diaconu, M.; Pavel, L.V.; Hlihor, R.-M.; Drăgoi, E.N.; Curteanu, S.; Lenz, M.; Corvini, P.F.-X.; Gavrilescu, M. Rhizobacteria and plant symbiosis in heavy metal uptake and its implications for soil bioremediation. New Biotechnol. 2017, 39, 125–134. [Google Scholar] [CrossRef]
  34. Hassan, W.; Bashir, S.; Ali, F.; Ijaz, M.; Hussain, M.; David, J. Role of ACC-deaminase and/or nitrogen fixing rhizobacteria in growth promotion of wheat (Triticum aestivum L.) under cadmium pollution. Environ. Earth Sci. 2016, 75, 267. [Google Scholar] [CrossRef]
  35. Pandey, N.; Bhatt, R. Role of soil associated Exiguobacterium in reducing arsenic toxicity and promoting plant growth in Vigna radiata. Eur. J. Soil Biol. 2016, 75, 142–150. [Google Scholar] [CrossRef]
  36. Ma, Y.; Rajkumar, M.; Rocha, I.; Oliveira, R.S.; Freitas, H. Serpentine bacteria influence metal translocation and bioconcentration of Brassica juncea and Ricinus communis grown in multi-metal polluted soils. Front. Plant Sci. 2015, 5, 757. [Google Scholar] [CrossRef] [PubMed]
  37. Guo, J.; Chi, J. Effect of Cd-tolerant plant growth-promoting rhizobium on plant growth and Cd uptake by Lolium multiflorum Lam. and Glycine max (L.) Merr. in Cd-contaminated soil. Plant Soil 2014, 375, 205–214. [Google Scholar] [CrossRef]
  38. Islam, F.; Yasmeen, T.; Ali, Q.; Ali, S.; Arif, M.S.; Hussain, S.; Rizvi, H. Influence of Pseudomonas aeruginosa as PGPR on oxidative stress tolerance in wheat under Zn stress. Ecotoxicol. Environ. Saf. 2014, 104, 285–293. [Google Scholar] [CrossRef] [PubMed]
  39. Islam, F.; Yasmeen, T.; Riaz, M.; Arif, M.S.; Ali, S.; Raza, S.H. Proteus mirabilis alleviates zinc toxicity by preventing oxidative stress in maize (Zea mays) plants. Ecotoxicol. Environ. Saf. 2014, 110, 143–152. [Google Scholar] [CrossRef] [PubMed]
  40. Pandey, S.; Ghosh, P.K.; Ghosh, S.; De, T.K.; Maiti, T.K. Role of heavy metal resistant Ochrobactrum sp. and Bacillus spp. strains in bioremediation of a rice cultivar and their PGPR like activities. J. Microbiol. 2013, 51, 11–17. [Google Scholar] [CrossRef] [PubMed]
  41. Zaidi, S.; Usmani, S.; Singh, B.R.; Musarrat, J. Significance of Bacillus subtilis strain SJ-101 as a bioinoculant for concurrent plant growth promotion and nickel accumulation in Brassica juncea. Chemosphere 2006, 64, 991–997. [Google Scholar] [CrossRef]
  42. Wagner, S.C. Biological Nitrogen Fixation. Nat. Educ. Knowl. 2011, 3, 15. [Google Scholar]
  43. Ohyama, T. Nitrogen as a major essential element of plants. Nitrogen Assim. Plants 2010, 37, 1–17. [Google Scholar]
  44. Stefan, B.; Rubio, L.M. State of the Art in Eukaryotic Nitrogenase Engineering. FEMS Microbiol. Lett. 2018, 365, fnx274. [Google Scholar]
  45. Hodge, A.; Robinson, D.; Fitter, A. Are microorganisms more effective than plants at competing for nitrogen? Trends Plant Sci. 2000, 5, 304–308. [Google Scholar] [CrossRef] [PubMed]
  46. Sukul, P.; Kumar, J.; Rani, A.; Abdillahi, A.M.; Rakesh, R.B.; Kumar, M.H. Functioning of plant growth promoting rhizobacteria (PGPR) and their mode of actions: An overview from chemistry point of view. Plant Arch. 2021, 21, 628–634. [Google Scholar] [CrossRef]
  47. Oleńska, E.; Małek, W.; Wójcik, M.; Swiecicka, I.; Thijs, S.; Vangronsveld, J. Beneficial features of plant growth-promoting rhizobacteria for improving plant growth and health in challenging conditions: A methodical review. Sci. Total Environ. 2020, 743, 140682. [Google Scholar] [CrossRef] [PubMed]
  48. Rovira, A. Microbial inoculation of plants: I. Establishment of free-living nitrogen-fixing bacteria in the rhizosphere and their effects on maize, tomato, and wheat. Plant Soil 1963, 19, 304–314. [Google Scholar] [CrossRef]
  49. Baldani, J.I.; Baldani, V.L. History on the biological nitrogen fixation research in graminaceous plants: Special emphasis on the Brazilian experience. An. Acad. Bras. Ciênc. 2005, 77, 549–579. [Google Scholar] [CrossRef]
  50. Mus, F.; Crook, M.B.; Garcia, K.; Costas, A.G.; Peters, J.W. Symbiotic Nitrogen Fixation and Challenges to Extending it to Non-Legumes. Appl. Environ. Microbiol. 2016, 82, 3698–3710. [Google Scholar] [CrossRef]
  51. Santos, K.F.; Moure, V.; Hauer, V.; Santos, A.; Donatti, L.; Galvão, C.; Pedrosa, F.; Souza, E.; Wassem, R.; Steffens, M. Wheat colonization by an Azospirillum brasilense ammonium-excreting strain reveals upregulation of nitrogenase and superior plant growth promotion. Plant Soil 2017, 415, 245–255. [Google Scholar] [CrossRef]
  52. Çakmakçı, R.; Erat, M.; Erdoğan, Ü.; Dönmez, M.F. The influence of plant growth–promoting rhizobacteria on growth and enzyme activities in wheat and spinach plants. J. Plant Nutr. Soil Sci. 2007, 170, 288–295. [Google Scholar] [CrossRef]
  53. Dhiman, V.K.; Rana, N.; Dhiman, V.K.; Pandey, H.; Verma, P.; Singh, D. Effect of rhizobial isolates and nitrogen fertilizers on nursery performance, nodulation behavior and nitrogenase activity of Dalbergia sissoo Roxb. seedlings. Plant Stress 2022, 4, 100080. [Google Scholar] [CrossRef]
  54. Hao, X.; Taghavi, S.; Xie, P.; Orbach, M.; Alwathnani, H.; Rensing, C.; Wei, G. Phytoremediation of heavy and transition metals aided by legume-rhizobia symbiosis. Int. J. Phytoremediat. 2014, 16, 179–202. [Google Scholar] [CrossRef] [PubMed]
  55. Jian, L.; Bai, X.; Zhang, H.; Song, X.; Li, Z. Promotion of growth and metal accumulation of alfalfa by coinoculation with Sinorhizobium and Agrobacterium under copper and zinc stress. PeerJ 2019, 7, e6875. [Google Scholar] [CrossRef]
  56. Chen, J.; Liu, Y.-Q.; Yan, X.-W.; Wei, G.-H.; Zhang, J.-H.; Fang, L.-C. Rhizobium inoculation enhances copper tolerance by affecting copper uptake and regulating the ascorbate-glutathione cycle and phytochelatin biosynthesis-related gene expression in Medicago sativa seedlings. Ecotoxicol. Environ. Saf. 2018, 162, 312–323. [Google Scholar] [CrossRef] [PubMed]
  57. Shen, S.; Li, Y.; Chen, M.; Huang, J.; Liu, F.; Xie, S.; Kong, L.; Pan, Y. Reduced cadmium toxicity in rapeseed via alteration of root properties and accelerated plant growth by a nitrogen-fixing bacterium. J. Hazard. Mater. 2023, 449, 131040. [Google Scholar] [CrossRef] [PubMed]
  58. Patel, K.; Goswami, D.; Dhandhukia, P.; Thakker, J. Techniques to study microbial phytohormones. In Bacterial Metabolites in Sustainable Agroecosystem; Springer International Publishing: Berlin/Heidelberg, Germany, 2015; pp. 1–27. [Google Scholar]
  59. Mehta, P.; Sharma, R.; Putatunda, C.; Walia, A. Endophytic fungi: Role in phosphate solubilization. In Advances in Endophytic Fungal Research: Present Status and Future Challenges; Springer International Publishing: Berlin/Heidelberg, Germany, 2019; pp. 183–209. [Google Scholar]
  60. De Zutter, N.; Ameye, M.; Debode, J.; De Tender, C.; Ommeslag, S.; Verwaeren, J.; Vermeir, P.; Audenaert, K.; De Gelder, L. Shifts in the rhizobiome during consecutive in planta enrichment for phosphate-solubilizing bacteria differentially affect maize P status. Microb. Biotechnol. 2021, 14, 1594–1612. [Google Scholar] [CrossRef] [PubMed]
  61. Schmidt, W. Iron solutions: Acquisition strategies and signaling pathways in plants. Trends Plant Sci. 2003, 8, 188–193. [Google Scholar] [CrossRef]
  62. 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] [PubMed]
  63. Saha, M.; Sarkar, S.; Sarkar, B.; Sharma, B.K.; Bhattacharjee, S.; Tribedi, P. Microbial siderophores and their potential applications: A review. Environ. Sci. Pollut. Res. 2016, 23, 3984–3999. [Google Scholar] [CrossRef]
  64. Sabaté, D.C.; Brandan, C.P.; Petroselli, G.; Erra-Balsells, R.; Audisio, M.C. Biocontrol of Sclerotinia sclerotiorum (Lib.) de Bary on common bean by native lipopeptide-producer Bacillus strains. Microbiol. Res. 2018, 211, 21–30. [Google Scholar] [CrossRef]
  65. Tank, N.; Saraf, M. Enhancement of plant growth and decontamination of nickel-spiked soil using PGPR. J. Basic Microbiol. 2009, 49, 195–204. [Google Scholar] [CrossRef]
  66. Tanimoto, E. Regulation of root growth by plant hormones—Roles for auxin and gibberellin. Crit. Rev. Plant Sci. 2005, 24, 249–265. [Google Scholar] [CrossRef]
  67. Ashrafuzzaman, M.; Hossen, F.A.; Ismail, M.R.; Hoque, A.; Islam, M.Z.; Shahidullah, S.; Meon, S. Efficiency of plant growth-promoting rhizobacteria (PGPR) for the enhancement of rice growth. Afr. J. Biotechnol. 2009, 8, 1247–1252. [Google Scholar]
  68. Teng, Z.; Shao, W.; Zhang, K.; Huo, Y.; Li, M. Characterization of phosphate solubilizing bacteria isolated from heavy metal contaminated soils and their potential for lead immobilization. J. Environ. Manag. 2019, 231, 189–197. [Google Scholar] [CrossRef] [PubMed]
  69. Goswami, D.; Thakker, J.N.; Dhandhukia, P.C. Portraying mechanics of plant growth promoting rhizobacteria (PGPR): A review. Cogent Food Agric. 2016, 2, 1127500. [Google Scholar] [CrossRef]
  70. Amara, U.; Khalid, R.; Hayat, R. Soil bacteria and phytohormones for sustainable crop production. In Bacterial Metabolites in Sustainable Agroecosystem; Springer International Publishing: Berlin/Heidelberg, Germany, 2015; pp. 87–103. [Google Scholar]
  71. Kudoyarova, G.; Arkhipova, T.; Korshunova, T.; Bakaeva, M.; Dodd, I.C. Phytohormone Mediation of Interactions between Plants and Non-Symbiotic Growth Promoting Bacteria under Edaphic Stresses. Front. Plant Sci. 2019, 10, 483140. [Google Scholar] [CrossRef] [PubMed]
  72. Halliwell, B.; Gutteridge, J.M. Biologically relevant metal ion-dependent hydroxyl radical generation An update. FEBS Lett. 1992, 307, 108–112. [Google Scholar] [CrossRef] [PubMed]
  73. Jin, Z.; Li, J.; Ni, L.; Zhang, R.; Xia, A.; Jin, F. Conditional privatization of a public siderophore enables Pseudomonas aeruginosa to resist cheater invasion. Nat. Commun. 2018, 9, 1383. [Google Scholar] [CrossRef] [PubMed]
  74. Huo, Y.; Kang, J.P.; Ahn, J.C.; Kim, Y.J.; Piao, C.H.; Yang, D.U.; Yang, D.C. Siderophore-producing rhizobacteria reduce heavy metal-induced oxidative stress in Panax ginseng Meyer. J. Ginseng Res. 2021, 45, 218–227. [Google Scholar] [CrossRef]
  75. Di Francesco, A.; Baraldi, E. How siderophore production can influence the biocontrol activity of Aureobasidium pullulans against Monilinia laxa on peaches. Biol. Control 2021, 152, 104456. [Google Scholar] [CrossRef]
  76. Khanna, K.; Kohli, S.K.; Kaur, R.; Handa, N.; Bakshi, P.; Sharma, P.; Ohri, P.; Bhardwaj, R. Reconnoitering the Efficacy of Plant Growth Promoting Rhizobacteria in Expediting Phytoremediation Potential of Heavy Metals. J. Plant Growth Regul. 2022, 42, 6474–6502. [Google Scholar] [CrossRef]
  77. Taghavi, S.; Garafola, C.; Monchy, S.; Newman, L.; Hoffman, A.; Weyens, N.; Barac, T.; Vangronsveld, J.; van der Lelie, D. Genome survey and characterization of endophytic bacteria exhibiting a beneficial effect on growth and development of poplar trees. Appl. Environ. Microbiol. 2009, 75, 748–757. [Google Scholar] [CrossRef] [PubMed]
  78. Manoj, S.R.; Karthik, C.; Kadirvelu, K.; Arulselvi, I.; Rajkumar, M. Understanding the molecular mechanisms for the enhanced phytoremediation of heavy metals through plant growth promoting rhizobacteria: A review. J. Environ. Manag. 2020, 254, 109779. [Google Scholar] [CrossRef]
  79. Gavrilescu, M. Enhancing phytoremediation of soils polluted with heavy metals. Curr. Opin. Biotechnol. 2022, 74, 21–31. [Google Scholar] [CrossRef] [PubMed]
  80. Du, M.; Spalding, E.P.; Gray, W.M. Rapid Auxin-Mediated Cell Expansion. Annu. Rev. Plant Biol. 2020, 71, 379–402. [Google Scholar] [CrossRef] [PubMed]
  81. Chen, B.; Luo, S.; Wu, Y.; Ye, J.; Wang, Q.; Xu, X.; Pan, F.; Khan, K.Y.; Feng, Y.; Yang, X. The effects of the endophytic bacterium Pseudomonas fluorescens Sasm05 and IAA on the plant growth and cadmium uptake of Sedum alfredii Hance. Front. Microbiol. 2017, 8, 2538. [Google Scholar] [CrossRef] [PubMed]
  82. Gomes, G.L.B.; Scortecci, K.C. Auxin and its role in plant development: Structure, signalling, regulation and response mechanisms. Plant Biol. 2021, 23, 894–904. [Google Scholar] [CrossRef] [PubMed]
  83. Ali, S.; Charles, T.C.; Glick, B.R. Endophytic phytohormones and their role in plant growth promotion. In Functional Importance of the Plant Microbiome: Implications for Agriculture, Forestry and Bioenergy; Springer International Publishing: Berlin/Heidelberg, Germany, 2017; pp. 89–105. [Google Scholar]
  84. Geries, L.; Elsadany, A.Y. Maximizing growth and productivity of onion (Allium cepa L.) by Spirulina platensis extract and nitrogen-fixing endophyte Pseudomonas stutzeri. Arch. Microbiol. 2021, 203, 169–181. [Google Scholar] [CrossRef] [PubMed]
  85. Lebrazi, S.; Fadil, M.; Chraibi, M.; Fikri-Benbrahim, K. Screening and optimization of indole-3-acetic acid production by Rhizobium sp. strain using response surface methodology. J. Genet. Eng. Biotechnol. 2020, 18, 21. [Google Scholar] [CrossRef]
  86. Ma, Y.; Rajkumar, M.; Zhang, C.; Freitas, H. Inoculation of Brassica oxyrrhina with plant growth promoting bacteria for the improvement of heavy metal phytoremediation under drought conditions. J. Hazard. Mater. 2016, 320, 36–44. [Google Scholar] [CrossRef]
  87. Abbaszadeh-Dahaji, P.; Atajan, F.A.; Omidvari, M.; Tahan, V.; Kariman, K. Mitigation of copper stress in maize (Zea mays) and sunflower (Helianthus annuus) plants by copper-resistant Pseudomonas strains. Curr. Microbiol. 2021, 78, 1335–1343. [Google Scholar] [CrossRef]
  88. Carlos, M.-H.J.; Stefani, P.-V.Y.; Janette, A.-M.; Melani, M.-S.S.; Gabriela, P.-O. Assessing the effects of heavy metals in ACC deaminase and IAA production on plant growth-promoting bacteria. Microbiol. Res. 2016, 188, 53–61. [Google Scholar] [CrossRef] [PubMed]
  89. Hassan, T.U. Construction of IAA-Deficient Mutants of Pseudomonas moraviensis and Their Comparative Effects with Wild Type Strains as Bio-inoculant on Wheat in Saline Sodic Soil. Geomicrobiol. J. 2019, 36, 376–384. [Google Scholar] [CrossRef]
  90. Li, S.M.; Zheng, H.X.; Zhang, X.S.; Sui, N. Cytokinins as central regulators during plant growth and stress response. Plant Cell Rep. 2020, 40, 271–282. [Google Scholar] [CrossRef] [PubMed]
  91. Hayat, R.; Ali, S.; Amara, U.; Khalid, R.; Ahmed, I. Soil beneficial bacteria and their role in plant growth promotion: A review. Ann. Microbiol. 2010, 60, 579–598. [Google Scholar] [CrossRef]
  92. Kieber, J.J.; Schaller, G.E. Cytokinin signaling in plant development. Development 2018, 145, dev149344. [Google Scholar] [CrossRef] [PubMed]
  93. Tassi, E.; Pouget, J.; Petruzzelli, G.; Barbafieri, M. The effects of exogenous plant growth regulators in the phytoextraction of heavy metals. Chemosphere 2008, 71, 66–73. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, J.; Zhang, Y.; Jin, J.; Li, Q.; Zhao, C.; Nan, W.; Wang, X.; Ma, R.; Bi, Y. An intact cytokinin-signaling pathway is required for Bacillus sp. LZR216-promoted plant growth and root system architecture altereation in Arabidopsis thaliana seedlings. Plant Growth Regul. 2018, 84, 507–518. [Google Scholar] [CrossRef]
  95. Piotrowska-Niczyporuk, A.; Bajguz, A.; Zambrzycka-Szelewa, E.; Bralska, M. Exogenously applied auxins and cytokinins ameliorate lead toxicity by inducing antioxidant defence system in green alga Acutodesmus obliquus. Plant Physiol. Biochem. 2018, 132, 535–546. [Google Scholar] [CrossRef] [PubMed]
  96. Yu, S.; Zehra, A.; Sahito, Z.A.; Wang, W.; Chen, S.; Feng, Y.; He, Z.; Yang, X. Cytokinin-mediated shoot proliferation and its correlation with phytoremediation effects in Cd-hyperaccumulator ecotype of Sedum alfredii. Sci. Total Environ. 2024, 912, 168993. [Google Scholar] [CrossRef]
  97. Nieto, K.F.; Frankenberger, W.T. Influence of adenine, isopentyl alcohol and Azotobacter chroococcum on the vegetative growth of Zea mays. Plant Soil 1991, 135, 213–221. [Google Scholar] [CrossRef]
  98. Lukatkin, A.S.; Gracheva, N.V.; Grishenkova, N.N.; Dukhovskis, P.V.; Brazaitite, A.A. Cytokinin-like growth regulators mitigate toxic action of zinc and nickel ions on maize seedlings. Russ. J. Plant Physiol. 2007, 54, 381–387. [Google Scholar] [CrossRef]
  99. Al-Hakimi, A. Modification of cadmium toxicity in pea seedlings by kinetin. Plant Soil Environ. 2007, 53, 129–135. [Google Scholar] [CrossRef]
  100. Lin, Z.; Zhong, S.; Grierson, D. Recent advances in ethylene research. J. Exp. Bot. 2009, 60, 3311–3336. [Google Scholar] [CrossRef] [PubMed]
  101. Van de Poel, B.; Smet, D.; Van Der Straeten, D. Ethylene and Hormonal Cross Talk in Vegetative Growth and Development. Plant Physiol. 2015, 169, 61–72. [Google Scholar] [CrossRef] [PubMed]
  102. Pattyn, J.; Vaughan-Hirsch, J.; Poel, B.V.D. The regulation of ethylene biosynthesis: A complex multilevel control circuitry. New Phytol. 2021, 229, 770–782. [Google Scholar] [CrossRef]
  103. Misra, S.; Chauhan, P.S. ACC deaminase-producing rhizosphere competent Bacillus spp. mitigate salt stress and promote Zea mays growth by modulating ethylene metabolism. 3 Biotech 2020, 10, 119. [Google Scholar] [CrossRef] [PubMed]
  104. Camilios-Neto, D.; Bonato, P.; Wassem, R.; Tadra-Sfeir, M.Z.; Brusamarello-Santos, L.C.; Valdameri, G.; Donatti, L.; Faoro, H.; Weiss, V.A.; Chubatsu, L.S. Dual RNA-seq transcriptional analysis of wheat roots colonized by Azospirillum brasilense reveals up-regulation of nutrient acquisition and cell cycle genes. BMC Genom. 2014, 15, 378. [Google Scholar] [CrossRef] [PubMed]
  105. Tahir, H.A.; Wu, H.; Wu, L.; Gao, X. Plant growth promotion by volatile organic compounds produced by Bacillus subtilis SYST2. Front. Microbiol. 2017, 8, 236091. [Google Scholar] [CrossRef]
  106. Singh, R.P.; Shelke, G.M.; Anil, K.; Jha, P.N. Biochemistry and genetics of ACC deaminase: A weapon to “stress ethylene” produced in plants. Front. Microbiol. 2015, 6, 937. [Google Scholar]
  107. Glick, B.R.; Cheng, Z.; Czarny, J.; Duan, J. Promotion of plant growth by ACC deaminase-producing soil bacteria. In New Perspectives and Approaches in Plant Growth-Promoting Rhizobacteria Research; Springer: Dordrecht, The Netherlands, 2007; pp. 329–339. [Google Scholar]
  108. Glick, B.R.; Penrose, D.M.; Li, J. A model for the lowering of plant ethylene concentrations by plant growth-promoting bacteria. J. Theor. Biol. 1998, 190, 63–68. [Google Scholar] [CrossRef]
  109. Mousavi, S.M.; Motesharezadeh, B.; Hosseini, H.M.; Alikhani, H.; Zolfaghari, A.A. Root-induced changes of Zn and Pb dynamics in the rhizosphere of sunflower with different plant growth promoting treatments in a heavily contaminated soil. Ecotoxicol. Environ. Saf. 2017, 147, 206. [Google Scholar] [CrossRef] [PubMed]
  110. Yamaguchi, S. Gibberellin metabolism and its regulation. Annu. Rev. Plant Biol. 2008, 59, 225–251. [Google Scholar] [CrossRef] [PubMed]
  111. Nagel, R.; Bieber, J.E.; Schmidt-Dannert, M.G.; Nett, R.S.; Peters, R.J. A Third Class: Functional Gibberellin Biosynthetic Operon in Beta-Proteobacteria. Front. Microbiol. 2018, 9, 423779. [Google Scholar] [CrossRef] [PubMed]
  112. Joo, G.J.; Kang, S.M.; Hamayun, M.; Kim, S.K.; Na, C.I.; Shin, D.H.; Lee, I.J. Burkholderia sp. KCTC 11096BP as a newly isolated gibberellin producing bacterium. J. Microbiol. 2009, 47, 167–171. [Google Scholar] [CrossRef] [PubMed]
  113. Bastián, F.; Cohen, A.; Piccoli, P.; Luna, V.; Bottini, R.; Baraldi, R.; Bottini, R. Production of indole-3-acetic acid and gibberellins A1 and A3 by Acetobacter diazotrophicus and Herbaspirillum seropedicae in chemically-defined culture media. Plant Growth Regul. 1998, 24, 7–11. [Google Scholar] [CrossRef]
  114. Massoud, M.B.; Sakouhi, L.; Karmous, I.; Zhu, Y.; El Ferjani, E.; Sheehan, D.; Chaoui, A. Protective role of exogenous phytohormones on redox status in pea seedlings under copper stress. J. Plant Physiol. 2018, 221, 51–61. [Google Scholar] [CrossRef] [PubMed]
  115. Yang, L.; Huan, Y.; Sun, J.; Lin, L.; Liao, M.a.; Wang, Z.; Liang, D.; Xia, H.; Lv, X.; Wang, J. Effects of exogenous gibberellic acid on growth and cadmium accumulation in Cyphomandra betacea seedlings. Environ. Prog. Sustain. Energy 2021, 40, e13655. [Google Scholar] [CrossRef]
  116. Liang, Y.; Xiao, X.; Guo, Z.; Peng, C.; Zeng, P.; Wang, X. Co-application of indole-3-acetic acid/gibberellin and oxalic acid for phytoextraction of cadmium and lead with Sedum alfredii Hance from contaminated soil. Chemosphere 2021, 285, 131420. [Google Scholar] [CrossRef] [PubMed]
  117. Liu, H.; Fang, H.H.P. Characterization of electrostatic binding sites of extracellular polymers by linear programming analysis of titration data. Biotechnol. Bioeng. 2010, 80, 806–811. [Google Scholar] [CrossRef]
  118. Naseem, H.; Bano, A. Role of plant growth-promoting rhizobacteria and their exopolysaccharide in drought tolerance of maize. J. Plant Interact. 2014, 9, 689–701. [Google Scholar] [CrossRef]
  119. Vimalnath, S.; Subramanian, S. Studies on the biosorption of Pb(II) ions from aqueous solution using extracellular polymeric substances (EPS) of Pseudomonas aeruginosa. Colloids Surf. B Biointerfaces 2018, 172, 60–67. [Google Scholar] [CrossRef] [PubMed]
  120. Dobrowolski, R.; Szcze, A.; Czemierska, M.; Jarosz-Wiko Azka, A. Studies of cadmium(II), lead(II), nickel(II), cobalt(II) and chromium(VI) sorption on extracellular polymeric substances produced by Rhodococcus opacus and Rhodococcus rhodochrous. Bioresour. Technol. 2017, 225, 113–120. [Google Scholar] [CrossRef]
  121. Yin, Y.; Hu, Y.; Xiong, F. Sorption of Cu(II) and Cd(II) by extracellular polymeric substances (EPS) from Aspergillus fumigatus. Int. Biodeterior. Biodegrad. 2011, 65, 1012–1018. [Google Scholar] [CrossRef]
  122. Li, W.W.; Yu, H.Q. Insight into the roles of microbial extracellular polymer substances in metal biosorption. Bioresour. Technol. 2014, 160, 15–23. [Google Scholar] [CrossRef] [PubMed]
  123. Comte, S.; Guibaud, G.; Baudu, M. Biosorption properties of extracellular polymeric substances (EPS) resulting from activated sludge according to their type: Soluble or bound. Process Biochem. 2006, 41, 815–823. [Google Scholar] [CrossRef]
  124. Dong, B.; Liu, X.; Dai, L.; Dai, X. Changes of heavy metal speciation during high-solid anaerobic digestion of sewage sludge. Bioresour. Technol. 2013, 131, 152–158. [Google Scholar] [CrossRef]
  125. Kushwaha, S.; Sreedhar, B.; Sudhakar, P.P. A spectroscopic study for understanding the speciation of Cr on palm shell based adsorbents and their application for the remediation of chrome plating effluents. Bioresour. Technol. 2012, 116, 15–23. [Google Scholar] [CrossRef] [PubMed]
  126. Zhang, B.; Wang, Z.; Shi, J.; Dong, H. Sulfur-based mixotrophic bio-reduction for efficient removal of chromium(VI) in groundwater. Geochim. Cosmochim. Acta 2019, 268, 296–309. [Google Scholar] [CrossRef]
  127. Lu, Y.-Z.; Chen, G.-J.; Bai, Y.-N.; Fu, L.; Qin, L.-P.; Zeng, R.J. Chromium isotope fractionation during Cr(VI) reduction in a methane-based hollow-fiber membrane biofilm reactor. Water Res. 2018, 130, 263–270. [Google Scholar] [CrossRef]
  128. Nocelli, N.; Bogino, P.C.; Banchio, E.; Giordano, W. Roles of Extracellular Polysaccharides and Biofilm Formation in Heavy Metal Resistance of Rhizobia. Materials 2016, 9, 418. [Google Scholar] [CrossRef]
  129. Bataineh, Z.M.; Heidger, P.M.; Thompson, S.A.; Timms, B.G. Immunocytochemical localization of metallothionein in the rat prostate gland. Prostate 1986, 9, 397–410. [Google Scholar] [CrossRef] [PubMed]
  130. Du, J.; Yang, J.-L.; Li, C.-H. Advances in metallotionein studies in forest trees. Plant Omics 2012, 5, 46–51. [Google Scholar]
  131. Cai, L.; Xu, Z.; Ren, M.; Guo, Q.; Hu, X.; Hu, G.; Wan, H.; Peng, P. Source identification of eight hazardous heavy metals in agricultural soils of Huizhou, Guangdong Province, China. Ecotoxicol. Environ. Saf. 2012, 78, 2–8. [Google Scholar] [CrossRef] [PubMed]
  132. Capdevila, M.; Atrian, S. Metallothionein protein evolution: A miniassay. JBIC J. Biol. Inorg. Chem. 2011, 16, 977–989. [Google Scholar] [CrossRef]
  133. Si, M.; Lang, J. The roles of metallothioneins in carcinogenesis. J. Hematol. Oncol. 2018, 11, 1–20. [Google Scholar] [CrossRef]
  134. Ehsanpour, A.A.; Zarei, S.; Abbaspour, J. The role of over expression of P5CS gene on proline, catalase, ascorbate peroxidase activity and lipid peroxidation of transgenic tobacco (Nicotiana tabacum L.) plant under in vitro drought stress. J. Cell Mol. Res. 2012, 4, 43–49. [Google Scholar]
  135. Murthy, S.; Bali, G.; Sarangi, S. Effect of lead on metallothionein concentration in leadresistant bacteria Bacillus cereus isolated from industrial effluent. Afr. J. Biotechnol. 2011, 10, 15966–15972. [Google Scholar] [CrossRef]
  136. Hussain, I.; Afzal, S.; Ashraf, M.A.; Rasheed, R.; Saleem, M.H.; Alatawi, A.; Ameen, F.; Fahad, S. Effect of metals or trace elements on wheat growth and its remediation in contaminated soil. J. Plant Growth Regul. 2023, 42, 2258–2282. [Google Scholar] [CrossRef]
  137. Israr, D.; Mustafa, G.; Khan, K.S.; Shahzad, M.; Ahmad, N.; Masood, S. Interactive effects of phosphorus and Pseudomonas putida on chickpea (Cicer arietinum L.) growth, nutrient uptake, antioxidant enzymes and organic acids exudation. Plant Physiol. Biochem. 2016, 108, 304–312. [Google Scholar] [CrossRef]
  138. Li, W.C.; Ye, Z.H.; Wong, M.H. Metal mobilization and production of short-chain organic acids by rhizosphere bacteria associated with a Cd/Zn hyperaccumulating plant, Sedum alfredii. Plant Soil 2010, 326, 453–467. [Google Scholar] [CrossRef]
  139. Tong, B.; Sun, T.; Sun, L. Low molecular weight organic acids in root exudates and cadmium accumulation in cadmium hyperaccumulator Solanum nigrum L. and nonhyperaccumulator Solanum lycopersicum L. Afr. J. Biotechnol. 2011, 10, 17180–17185. [Google Scholar]
  140. Yang, P.; Zhou, X.-F.; Wang, L.-L.; Li, Q.-S.; Zhou, T.; Chen, Y.-K.; Zhao, Z.-Y.; He, B.-Y. Effect of phosphate-solubilizing bacteria on the mobility of insoluble cadmium and metabolic analysis. Int. J. Environ. Res. Public Health 2018, 15, 1330. [Google Scholar] [CrossRef] [PubMed]
  141. Renella, G.; Landi, L.; Nannipieri, P. Degradation of low molecular weight organic acids complexed with heavy metals in soil. Geoderma 2004, 122, 311–315. [Google Scholar] [CrossRef]
  142. Krishnamurti, G.; Cieslinski, G.; Huang, P.; Van Rees, K. Kinetics of cadmium release from soils as influenced by organic acids: Implication in cadmium availability. J. Environ. Qual. 1997, 26, 271–277. [Google Scholar] [CrossRef]
  143. Yu, G.; Liu, J.; Long, Y.; Chen, Z.; Sunahara, G.I.; Jiang, P.; You, S.; Lin, H.; Xiao, H. Phytoextraction of cadmium-contaminated soils: Comparison of plant species and low molecular weight organic acids. Int. J. Phytoremediat. 2020, 22, 383–391. [Google Scholar] [CrossRef] [PubMed]
  144. Al Mahmud, J.; Hasanuzzaman, M.; Nahar, K.; Bhuyan, M.B.; Fujita, M. Insights into citric acid-induced cadmium tolerance and phytoremediation in Brassica juncea L.: Coordinated functions of metal chelation, antioxidant defense and glyoxalase systems. Ecotoxicol. Environ. Saf. 2018, 147, 990–1001. [Google Scholar] [CrossRef] [PubMed]
  145. Hassen, A.; Saidi, N.; Cherif, M.; Boudabous, A. Effects of heavy metals on Pseudomonas aeruginosa and Bacillus thuringiensis. Bioresour. Technol. 1998, 65, 73–82. [Google Scholar] [CrossRef]
  146. Farid, M.; Ali, S.; Saeed, R.; Rizwan, M.; Bukhari, S.A.H.; Abbasi, G.H.; Hussain, A.; Ali, B.; Zamir, M.S.I.; Ahmad, I. Combined application of citric acid and 5-aminolevulinic acid improved biomass, photosynthesis and gas exchange attributes of sunflower (Helianthus annuus L.) grown on chromium contaminated soil. Int. J. Phytoremediat. 2019, 21, 760–767. [Google Scholar] [CrossRef]
  147. Chai, M.-W.; Li, R.-L.; Shi, F.-C.; Liu, F.-C.; Pan, X.; Cao, D.; Wen, X. Effects of cadmium stress on growth, metal accumulation and organic acids of Spartina alterniflora Loisel. Afr. J. Biotechnol. 2012, 11, 6091–6099. [Google Scholar]
  148. Chen, H.-C.; Zhang, S.-L.; Wu, K.-J.; Li, R.; He, X.-R.; He, D.-N.; Huang, C.; Wei, H. The effects of exogenous organic acids on the growth, photosynthesis and cellular ultrastructure of Salix variegata Franch. Under Cd stress. Ecotoxicol. Environ. Saf. 2020, 187, 109790. [Google Scholar] [CrossRef]
  149. Khanna, K.; Jamwal, V.L.; Sharma, A.; Gandhi, S.G.; Ohri, P.; Bhardwaj, R.; Al-Huqail, A.A.; Siddiqui, M.H.; Ali, H.M.; Ahmad, P. Supplementation with plant growth promoting rhizobacteria (PGPR) alleviates cadmium toxicity in Solanum lycopersicum by modulating the expression of secondary metabolites. Chemosphere 2019, 230, 628–639. [Google Scholar] [CrossRef] [PubMed]
  150. Loon, L.; Bakker, P.; Pieterse, C. Systemic resistance induced by rhizosphere bacteria. Annu. Rev. Phytopathol. 1998, 36, 453–483. [Google Scholar] [CrossRef]
  151. Annapurna, K.; Kumar, A.; Kumar, L.V.; Govindasamy, V.; Bose, P.; Ramadoss, D. PGPR-induced systemic resistance (ISR) in plant disease management. In Bacteria in Agrobiology: Disease Management; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013; pp. 405–425. [Google Scholar]
  152. Jorquera, M.A.; Maruyama, F.; Ogram, A.V.; Navarrete, O.U.; Lagos, L.M.; Inostroza, N.G.; Acuña, J.J.; Rilling, J.I.; de La Luz Mora, M. Rhizobacterial Community Structures Associated with Native Plants Grown in Chilean Extreme Environments. Microb. Ecol. 2016, 72, 633–646. [Google Scholar] [CrossRef] [PubMed]
  153. Gopalakrishnan, S.; Arumugam, S.; Vijayabharathi, R.; Krishnamurthy, L. Plant growth promoting rhizobia: Challenges and opportunities. 3 Biotech 2014, 5, 355–377. [Google Scholar] [CrossRef] [PubMed]
  154. Miller, G.; Suzuki, N.; Ciftci-Yilmaz, S.; Mittler, R. Reactive oxygen species homeostasis and signalling during drought and salinity stresses. Plant Cell Environ. 2010, 33, 453–467. [Google Scholar] [CrossRef]
  155. Gururani, M.A.; Upadhyaya, C.P.; Baskar, V.; Venkatesh, J.; Nookaraju, A.; Park, S.W. Plant growth-promoting rhizobacteria enhance abiotic stress tolerance in Solanum tuberosum through inducing changes in the expression of ROS-scavenging enzymes and improved photosynthetic performance. J. Plant Growth Regul. 2013, 32, 245–258. [Google Scholar] [CrossRef]
  156. Ju, W.; Liu, L.; Jin, X.; Duan, C.; Fang, L. Co-inoculation effect of plant-growth-promoting rhizobacteria and rhizobium on EDDS assisted phytoremediation of Cu contaminated soils. Chemosphere 2020, 254, 126724. [Google Scholar] [CrossRef] [PubMed]
  157. Gupta, P.; Rani, R.; Chandra, A.; Kumar, V. Potential applications of Pseudomonas sp. (strain CPSB21) to ameliorate Cr6+ stress and phytoremediation of tannery effluent contaminated agricultural soils. Sci. Rep. 2018, 8, 4860. [Google Scholar] [CrossRef] [PubMed]
  158. Khanna, K.; Jamwal, V.L.; Kohli, S.K.; Gandhi, S.G.; Ohri, P.; Bhardwaj, R.; Abdjtllah, E.F.; Hashem, A.; Ahmad, P. Plant growth promoting rhizobacteria induced Cd tolerance in Lycopersicon esculentum through altered antioxidative defense expression. Chemosphere 2019, 217, 463–474. [Google Scholar] [CrossRef]
  159. Jing, Y.D.; He, Z.L.; Yang, X.E. Role of soil rhizobacteria in phytoremediation of heavy metal contaminated soils. J. Zhejiang Univ. B 2007, 8, 192–207. [Google Scholar] [CrossRef]
  160. Hamidpour, M.; Nemati, H.; Abbaszadeh Dahaji, P.; Roosta, H.R. Effects of plant growth-promoting bacteria on EDTA-assisted phytostabilization of heavy metals in a contaminated calcareous soil. Environ. Geochem. Health 2020, 42, 2535–2545. [Google Scholar] [CrossRef] [PubMed]
  161. Gupta, R.; Khan, F.; Alqahtani, F.M.; Hashem, M.; Ahmad, F. Plant growth–promoting Rhizobacteria (PGPR) assisted bioremediation of Heavy Metal Toxicity. Appl. Biochem. Biotechnol. 2023, 1–29. [Google Scholar] [CrossRef] [PubMed]
  162. Zayed, A.; Terry, L.N. Accumulation and volatilization of different chemical species of selenium by plants. Planta 1998, 206, 284–292. [Google Scholar] [CrossRef]
  163. Jeong, S.; Moon, H.S.; Nam, K.; Kim, J.Y.; Kim, T.S. Application of phosphate-solubilizing bacteria for enhancing bioavailability and phytoextraction of cadmium (Cd) from polluted soil. Chemosphere 2012, 88, 204–210. [Google Scholar] [CrossRef] [PubMed]
  164. Prapagdee, B.; Chanprasert, M.; Mongkolsuk, S. Bioaugmentation with cadmium-resistant plant growth-promoting rhizobacteria to assist cadmium phytoextraction by Helianthus annuus. Chemosphere 2013, 92, 659–666. [Google Scholar] [CrossRef] [PubMed]
  165. Juwarkar, A.A.; Nair, A.; Dubey, K.V.; Singh, S.; Devotta, S. Biosurfactant technology for remediation of cadmium and lead contaminated soils. Chemosphere 2007, 68, 1996–2002. [Google Scholar] [CrossRef]
  166. He, C.Q.; Tan, G.; Liang, X.; Du, W.; Chen, Y.; Zhi, G.; Zhu, Y. Effect of Zn-tolerant bacterial strains on growth and Zn accumulation in Orychophragmus violaceus. Appl. Soil Ecol. 2010, 44, 1–5. [Google Scholar] [CrossRef]
  167. Strigul, N.S.; Kravchenko, L.V. Mathematical modeling of PGPR inoculation into the rhizosphere. Environ. Model. Softw. 2006, 21, 1158–1171. [Google Scholar] [CrossRef]
  168. Guarino, F.; Miranda, A.; Castiglione, S.; Cicatelli, A. Arsenic phytovolatilization and epigenetic modifications in Arundo donax L. assisted by a PGPR consortium. Chemosphere 2020, 251, 126310. [Google Scholar] [CrossRef]
  169. Kaushal, M.; Wani, S.P. Plant-growth-promoting rhizobacteria: Drought stress alleviators to ameliorate crop production in drylands. Ann. Microbiol. 2016, 66, 35–42. [Google Scholar] [CrossRef]
Figure 1. Number of papers on PGPR included in Web of Science from 1995 to 2023. (a) Search for the keywords “PGPR” and “PGPR, phytoremediation”. (b) Search for keywords “PGPR, Zn”, “PGPR, Cd”, “PGPR, Cu”, “PGPR, Pb”, and “PGPR, Cr”.
Figure 1. Number of papers on PGPR included in Web of Science from 1995 to 2023. (a) Search for the keywords “PGPR” and “PGPR, phytoremediation”. (b) Search for keywords “PGPR, Zn”, “PGPR, Cd”, “PGPR, Cu”, “PGPR, Pb”, and “PGPR, Cr”.
Microorganisms 12 00700 g001
Figure 2. Growth-promoting mechanisms of PGPR.
Figure 2. Growth-promoting mechanisms of PGPR.
Microorganisms 12 00700 g002
Figure 3. Mechanism of PSB in bio-dissolving insoluble phosphates.
Figure 3. Mechanism of PSB in bio-dissolving insoluble phosphates.
Microorganisms 12 00700 g003
Figure 4. Schematic diagram of mechanism of metal–EPS interactions.
Figure 4. Schematic diagram of mechanism of metal–EPS interactions.
Microorganisms 12 00700 g004
Figure 5. Systemic resistance induced by PGPR.
Figure 5. Systemic resistance induced by PGPR.
Microorganisms 12 00700 g005
Table 1. How PGPR assists in the phytoextraction of heavy metals.
Table 1. How PGPR assists in the phytoextraction of heavy metals.
Target
Heavy
Metals
PGPRTest PlantsPlant-Growth-Promoting Traits/MechanismsReferences
Cd, ZnBacillus sp.Oryza sativaSecretes indole-3-acetic acid (IAA), 1-aminocyclopropane-1-carboxylate (ACC) deaminase, and siderophores; phosphate solubilization.[19]
Ni, Pb, Cd and CrPseudomonas putidaZea maysIncreases the availability of Fe, Zn, K, and Ca.[20]
CdBacillus megateriumHybrid PennisetumSecretes IAA, siderophores, ACC deaminase (ACCD); phosphate solubilization and nitrogen fixation.[21,22]
CrPseudomonas sp.Medicago sativaSecretes IAA and siderophores; produces ammonia, cellulase, pectinase, chitinase, and ACCD; phosphate solubilization; nitrogen fixation.[23]
Cd, PbEnterobacter bugandensis and Bacillus megateriumLactuca sativa L.Secrete IAA and siderophores.[24]
CrAgrobacterium fabrumZea maysSecretes siderophores, IAA, and potassium; phosphate solubilization.[25]
CaBacillus spp. Brassica napus L.Secretes IAA, siderophores and ACCD; phosphate solubilization. [26]
PbLuteibacter sp. and
Variovorax sp.
Lathyrus sativus L.Secrete IAA, siderophores, and HCN; phosphate solubilization.[27]
CrKlebsiella sp.-Secretes IAA, ammonia, siderophores, and HCN.[28]
Cd, Pb and AsKlebsiella michiganensisOryza sativaSecretes IAA and ACCD; nitrogen fixation; phosphate solubilization. [29]
As, Cd and
Cr
Pseudomonas sp. Lens culinarisSecretes IAA.[30]
CuKocuria sp.Saccharum spontaneumSecretes IAA, product ammonia, and hydrogen cyanide (HCN); phosphate solubilization.[31]
NiBacillus spp.Althea rosea L.Secretes IAA; siderophore production; phosphate solubilization. [32]
Cr, CdAzotobacter sp.Lepidium sativumSolubilizing of phosphorus; improves the dissolution and retention of iron in the growth medium; nitrogen fixation; produces plant hormones. [33]
CdAzotobacter sp.Triticum aestivum L.Secretes IAA and ACCD; nitrogen fixation; phosphate solubilization.[34]
AsExiguobacterium sp.Vigna radiataSecretes IAA and EPS.[35]
Ni, Zn and FePsychrobacter sp. and Pseudomonas sp. Brassica juncea and Ricinus communisSecrete siderophores, ACCD, and IAA; phosphate solubilization.[36]
CdBradyrhizobium sp.Lolium multiflorum Lam.Secretes IAA, siderophores, and ACCD; phosphate solubilization.[37]
ZnPseudomonas aeruginosaTriticum aestivum L.Secretes IAA, ACCD, and siderophores; phosphate solubilization.[38]
ZnProteus mirabilisZea maysSecretes IAA, siderophore, and ACCD; phosphate solubilization.[39]
CdOchrobactrum sp.Oryza sativaSecrete siderophores and ACCD.[40]
Pb, AsBacillus sp.
NiBacillus subtilisBrassica junceaSecretes IAA; phosphate solubilization.[41]
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

Qin, H.; Wang, Z.; Sha, W.; Song, S.; Qin, F.; Zhang, W. Role of Plant-Growth-Promoting Rhizobacteria in Plant Machinery for Soil Heavy Metal Detoxification. Microorganisms 2024, 12, 700. https://doi.org/10.3390/microorganisms12040700

AMA Style

Qin H, Wang Z, Sha W, Song S, Qin F, Zhang W. Role of Plant-Growth-Promoting Rhizobacteria in Plant Machinery for Soil Heavy Metal Detoxification. Microorganisms. 2024; 12(4):700. https://doi.org/10.3390/microorganisms12040700

Chicago/Turabian Style

Qin, Haichen, Zixiao Wang, Wenya Sha, Shuhong Song, Fenju Qin, and Wenchao Zhang. 2024. "Role of Plant-Growth-Promoting Rhizobacteria in Plant Machinery for Soil Heavy Metal Detoxification" Microorganisms 12, no. 4: 700. https://doi.org/10.3390/microorganisms12040700

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop