Next Article in Journal
Ulva compressa from Copper-Polluted Sites Exhibits Intracellular Copper Accumulation, Increased Expression of Metallothioneins and Copper-Containing Nanoparticles in Chloroplasts
Previous Article in Journal
Characterisation of Gel-Forming Mucins Produced In Vivo and In Ex Vivo Conjunctival Explant Cultures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Rhizosphere Bacteria in Plant Growth Promotion, Biocontrol, and Bioremediation of Contaminated Sites: A Comprehensive Review of Effects and Mechanisms

1
College of Natural Resources and Environment, Northwest Agriculture and Forestry University, Yangling 712100, China
2
College of Life Sciences, Yan’an University, Yan’an 716000, China
3
College of Resources and Environmental Sciences, Gansu Agricultural University, Lanzhou 730070, China
4
Institute of Chemistry and Technology of Environmental Protection, Faculty of Chemistry, Brno University of Technology, Purkynova 118, 612 00 Brno, Czech Republic
5
Institute of Molecular Biology and Biotechnology, The University of Lahore, Defense Road, Lahore 54000, Pakistan
6
Department of Agrochemistry, Soil Science, Microbiology and Plant Nutrition, Faculty of AgriSciences, Mendel University in Brno, Zemedelska 1, 613 00 Brno, Czech Republic
7
Institute of Soil and Environmental Science, University of Agriculture Faisalabad, Faisalabad 38040, Pakistan
8
Agricultural Research, Ltd., Zahradni 400/1, 664 41 Troubsko, Czech Republic
9
School of Environmental and Municipal Engineering, Lanzhou Jiaotong University, Lanzhou 730070, China
10
Biology Center CAS, SoWa RI, Na Sadkach 7, 370 05 České Budějovice, Czech Republic
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(19), 10529; https://doi.org/10.3390/ijms221910529
Submission received: 22 August 2021 / Revised: 23 September 2021 / Accepted: 27 September 2021 / Published: 29 September 2021
(This article belongs to the Section Molecular Plant Sciences)

Abstract

:
Agriculture in the 21st century is facing multiple challenges, such as those related to soil fertility, climatic fluctuations, environmental degradation, urbanization, and the increase in food demand for the increasing world population. In the meanwhile, the scientific community is facing key challenges in increasing crop production from the existing land base. In this regard, traditional farming has witnessed enhanced per acre crop yields due to irregular and injudicious use of agrochemicals, including pesticides and synthetic fertilizers, but at a substantial environmental cost. Another major concern in modern agriculture is that crop pests are developing pesticide resistance. Therefore, the future of sustainable crop production requires the use of alternative strategies that can enhance crop yields in an environmentally sound manner. The application of rhizobacteria, specifically, plant growth-promoting rhizobacteria (PGPR), as an alternative to chemical pesticides has gained much attention from the scientific community. These rhizobacteria harbor a number of mechanisms through which they promote plant growth, control plant pests, and induce resistance to various abiotic stresses. This review presents a comprehensive overview of the mechanisms of rhizobacteria involved in plant growth promotion, biocontrol of pests, and bioremediation of contaminated soils. It also focuses on the effects of PGPR inoculation on plant growth survival under environmental stress. Furthermore, the pros and cons of rhizobacterial application along with future directions for the sustainable use of rhizobacteria in agriculture are discussed in depth.

1. Introduction

Recently, crop production has been facing serious threats due to various biotic and abiotic stresses. Feeding the growing population and enhancing agricultural production on limited land are significant challenges for researchers and farmers in the current era [1]. In addition, present agricultural practices, such as the use of fertilizers, pesticides, herbicides, and irrigation with untreated wastewater, pose serious threats to the environment and cause soil degradation [2]. Moreover, urbanization and industrialization have caused a significant reduction in the agricultural area, which has further motivated scientists to develop sustainable strategies to increase crop yields from the already shrinking cropped area [3]. Increasing the area available for crop production has proven difficult; therefore, strategies should be developed to increase the crop yield per unit area in a way that prevents the further degradation of natural resources [4]. Therefore, adopting alternative approaches in today’s agriculture is necessary for ensuring environmental sustainability and future food security.
Several technical strategies suggested in the past involve enhancing agricultural production by reducing agricultural inputs such as fertilizers and pesticides. In this context, the use of rhizobacteria has been increasingly gaining momentum. Rhizobacteria reside in the rhizosphere, and those having beneficial effects on plants are termed plant growth-promoting rhizobacteria [5,6]. These rhizobacteria are equipped with a number of mechanisms (both direct and indirect) through which they improve plant growth in diverse agricultural settings. Several previous studies have reported the natural enhancement of plant growth of field crops by applying plant growth-promoting rhizobacteria (PGPR). The mechanisms for plant growth promotion used by rhizobacteria, which inhabit the rhizosphere, include metabolic adjustments, adjustments in phytohormone levels, production of exopolysaccharides, root colonization, and enhancement of nutrient availability [7,8,9]. These rhizobacteria also indirectly improve plant growth by inducing plant resistance to various biotic and abiotic stresses, such as pathogen attack and heavy metal contamination, using such mechanisms as the production of antibiotics, induction of induced systemic resistance, rhizosphere competence, and production of antagonistic substances for biocontrol [7,10,11,12,13]. Moreover, the mechanisms used by rhizobacteria for the bioremediation of contaminated soils include the production of biosurfactants and siderophores, biosorption, ACC deaminase activity, and production of polymeric substances [14,15,16].
In this review, we summarize the key knowledge of plant growth promotion resulting from rhizosphere bacterial application in diverse agricultural settings. Here, we synthesize the main findings and highlight the in-depth analysis of mechanisms used by rhizobacteria for plant growth promotion, biocontrol, and bioremediation of contaminated sites (Figure 1) in a comprehensive manner. We then discuss the pros and cons of rhizobacterial application in modern agriculture for improving plant growth, and we then evaluate technical suggestions for the future use of rhizobacteria in agriculture.

2. Plant Growth Promotion by Plant Growth-Promoting Rhizobacteria (PGPR): An Overview

During their development, plants are in intimate and continuous contact with microorganisms present in the root vicinity, known as the rhizosphere. Microbes living in the rhizosphere of several plants and having several positive effects on the host plant through various mechanisms are usually termed plant growth-promoting rhizobacteria (PGPR) [17,18]. In the rhizosphere, plant roots secrete a number of exudates that act as attractants for microbes, which eventually improve the physicochemical properties of the surrounding soil. On the other hand, these exudates maintain the function and structure of microbial communities near plant roots [9,19]. Plants and bacteria form symbiotic associations to alleviate abiotic stresses [20,21,22,23]. PGPR can assist plants in their growth by fixation of atmospheric nitrogen, producing siderophores, generating phytohormones (auxins, gibberellins, cytokinins), solubilizing phosphorus (P), or synthesizing stress-relieving enzymes [24]. Moreover, certain bacteria improve the accessibility of essential nutrients, improve root progression, and lessen stress-induced damage by modifying plant defense systems [25,26]. Furthermore, PGPR indirectly help plant symbionts by initiating induced systemic resistance, exerting an antibiosis effect, and potentially improving the content of plant cell metabolites [25,27]. PGPR can withstand hostile natural conditions such as shortage of water, salt stress, weed invasion, lack of nutrients, and heavy metal pollution [28]. The use of PGPR could help to enhance and improve sustainable agriculture and natural stability. These PGPR can be found in association with roots (in the rhizosphere), which enhance plant growth in the absence of pathogens or lessen the harmful effects of pathogens on crop yield by antibiosis, competition, induced systemic resistance, and siderophore production [29,30,31]. Several mechanisms used by PGPR in plant growth promotion are described in detail in the following section.

3. Mechanisms of Action of PGPR in Plant Growth Promotion

3.1. Biological Nitrogen Fixation

Nitrogen, one of the most vital elements, plays an important role in plant growth and various metabolic activities. Nitrogen (N2) accounts for 78% of total atmospheric gasses, but this form is not available to plants. Conversion of nitrogen to ammonia (NH3) by certain bacteria and archaea using a nitrogenase protein complex [32] is termed biological nitrogen fixation (BNF) [7,33]. Various nitrogen fixers are distributed in nature, and the N2 fixing ability of microbes can be a substitute for commercial fertilizers [34,35]. Nitrogen-fixing microbes can be classified as (i) symbiotic (mutualistic relationship between bacteria and leguminous plants and non-leguminous trees) and (ii) non-symbiotic (free-living and endophytic organisms) [7,36,37,38]. Widely reported symbiotic N2 fixers in soil are Frankia and Rhizobium species, whereas diazotrophic PGPR, including Cyanobacteria, Azospirillum, Pseudomonas, Azotobacter, Acetobacter, and Nostoc, are non-symbiotic nitrogen-fixing microbes [18,36]. Non-symbiotic microbes are important in the natural environment; they fix less nitrogen but provide enough to meet host plant demand. A complex process of infection and establishment between the roots of leguminous plants and symbionts results in the formation of root nodules [39]. Energy in the form of ATP (adenosine triphosphate) is required in the process of nitrogen fixation: for this, bacterial carbon resources undergo oxidative phosphorylation, and the energy is stored in the form of glycogen [40]. The genes for nitrogen fixation in both symbionts and free-living non-symbionts are nif (nitrogenase complex) genes [41]. The nif genes encode enzymes involved in the fixation of atmospheric N2 into a form of nitrogen available to plants. The primary enzyme encoded by nif genes is the nitrogenase complex. In addition to enzymes, nif genes also encode regulatory proteins for N2 fixation. The oxygen (O2) concentration and low level of nitrogen in the root environment of the host plant are responsible for the induction of nif gene expression [40]. Oxygen is a negative regulator of nif genes and has an inhibitory effect on nitrogenase enzyme activity, though it is required for the respiration of Rhizobium and bacteroid species. Bacterial hemoglobin, which binds free oxygen radicals, provides sufficient O2 for bacteroid respiration within root nodules while simultaneously preventing O2 from inhibiting N2 fixation [36]. At a low level of dissolved oxygen, after the transformation of Rhizobium etli with a hemoglobin gene from Vitreoscilla sp. (a Gram –ve bacterium), the respiratory rate of rhizobial cells was increased 2–3-fold compared with non-transformed strains [38].
The nitrogenase complex (nif) consists of regulatory genes, structural genes, and genes involved in the biosynthesis of Fe protein and Fe–molybdenum cofactor activation. In non-symbiotic microbes, nif genes are found in a cluster of around 20–24 kb containing seven operons encoding 20 proteins. [42]. A small rise in the ethylene level of a plant may cause infection and nodulation by Rhizobium species in the roots of leguminous crops [7]. However, the increase in ethylene levels reduces nodulation in legumes [43]. Some rhizobacterium strains have the ability to mitigate the rise in ethylene levels by synthesizing rhizobiotoxine (a small molecule) in the host legume roots [7,14], which inhibits the function of the ACC synthase enzyme (ethylene biosynthetic enzyme). Some ACC deaminase-synthesizing rhizobacteria remove ACC, which is the precursor of ethylene in green plants [43]. The nodulation efficacy of Rhizobium spp. in which ACC deaminase is absent can be increased by inserting isolated genes of ACC deaminase. For instance, a nif gene from R. leguminosarum bv. viciae was isolated and inserted into the DNA of Sinorhizobium meliloti, and it dramatically increased the biomass and nodulation of alfalfa plants [44]. A summary of a range of studies regarding the PGPR-mediated growth enhancement of various crops through BNF is reported in Table 1.

3.2. Phosphorus Solubilization

Phosphorus (P) is the second most vital macroelement after nitrogen and is required for adequate plant nutrition [18]. It has key roles in the processes of photosynthesis, energy transfer, and various plant metabolic processes [67]. In most soils, P is present in excess, but not in a form available to plants [68]: only 0.1% of total soil phosphorus is available for plant use [69,70,71]. Phosphorus becomes immobilized in soil through complex interactions with various cations such as Ca2+ and Mg2+ in alkali soils and Fe3+ and Al+ in acidic soils [7,72]. In most soils, available P is not present in a sufficient quantity for plants, and commercial fertilizers are used to meet plant requirements [73]. According to estimates, about 52.3 billion tons of phosphorus fertilizers are applied to agricultural land every year [74]. Excessive use of synthetic fertilizers not only results in environmental hazards, i.e., global warming due to the production of nitrogen oxides, but is also a source of eutrophication of water bodies [75,76]. Plant growth-promoting microbes, on the other hand, offer a suitable option to enhance crop productivity with no adverse effects on the environment.
These microbes increase phosphorus availability in soil by mediating the mineralization of organic P in soil [77]. Mineralization of organic P by microorganisms requires P solubilization, which greatly depends on soil pH. Such microbes modify the soil pH by producing various organic and inorganic acids and other metabolites through a mechanism known as rhizosphere acidification. Certain Rhizobium, Pseudomonas, and Bacillus species are considered phosphate-solubilizing bacteria (PSB) [18]. However, the exact extent of P solubilization is highly variable among soils and the various bacteria involved and further depends upon the soil conditions. Recently, Muleta et al. [78] reported that various strains of Mesorhizobium sp. caused the pH to drop from 6.9 to 5.2 in Pikovskaya’s medium amended with tri-calcium phosphate, and, as a result, the P solubilization increased as desired, from 1.53 to 138 µg mL−1. In the current situation, it is necessary to explore an economically affordable and ecologically healthy option for P availability in roots and plants. Keeping this scenario in mind, there should be an emphasis on the use of PGPR instead of abundant reliance on the use of commercial phosphorus sources.
Various microbial species, including bacteria and fungi, are involved in P solubilization in soil [69,79]. The integration of organic amendments with a low dose of chemical fertilizer is a strategy to reduce the soil C:P ratio and increase P availability for crop plants [80]. About 50% of all bacteria are able to solubilize P [68]. Bacillus, Rhizobium, Pseudomonas, Azotobacter, Burkholderia, Enterobacter, Microbacterium, Serratia, Burkholderia, and Beijerinckia are the most significant PSB [81]. Some soil bacteria, e.g., actinobacteria, release low-molecular-weight organic solutes and solubilize inorganic phosphate [36]. The principal P solubilization mechanism involves the production of organic acids and OH ions, production of protons or bicarbonate release (cation/anion balance), gaseous exchange of O2/CO2, and production of siderophores by soil microbes [82]. The release of organic acid by microorganisms to solubilize inorganic P bound to soil colloids is an important mechanism in which COO (carboxyl group) and OH (hydroxyl ion) act as chelators of cations such as Fe, Al3+, and Ca2+ and compete for P adsorption sites in soil [67,75,83,84]. Chelators are low-molecular-weight organic acids that compete with Al and Fe oxides for P fixation sites [85]. Gluconic and ketogluconic acids are known to act as Ca chelators in alkaline soils [86]. Organic acids are released by microbial biomass during fermentation or oxidative respiration of organic sources [82]. Phosphate-solubilizing microbes primarily secrete oxalic, malic, fumaric, acetic, tartaric, malonic, glutanic, propoinic, butyric, lactic, gluconic, 2-keto gluconic, glyconic, and oxalic acids [82,87]. These low-molecular-weight organic acids solubilize the fixed inorganic P by lowering the soil pH, chelating cations, and competing with phosphate (PO4) for adsorption sites in the soil [88]. Excretion of root exudates, e.g., ligands, also plays a significant role in altering P concentration in soil solution [89]. Inorganic acids (e.g., hydrochloric acid) can also solubilize PO4 but are less effective than organic acids at the same pH [90]. Hence, phosphorus solubilization is an effective and sustainable approach to enhance the P availability for plants and reduce dependence on unsustainable, costly synthetic fertilizer sources. The integrated use of organic sources together with reduced doses of inorganic fertilizer increases the removal of P fixed with soil colloids and sustains soil fertility [91]. For sustainable soil health and fertility, research is required to enhance the efficiency of P solubilization approaches and identify new bacteria species with more efficiency. A summary of various studies involving P solubilization through microbial candidates and improvement in plant growth is provided in Table 1.

3.3. Potassium Solubilization

Another vital macronutrient after nitrogen and phosphorus is potassium (K), which is the seventh most abundant element in the Earth’s crust. It plays a key role in various metabolic, growth, and development-related processes in plants. Moreover, potassium is responsible for activating more than 80 enzymes involved in photosynthesis activity, nitrate reduction, starch synthesis, and various energy metabolic processes [92,93,94]. An increase in plant vigor against various biotic stresses, diseases, and pests is also regulated by an adequate supply of K [95]. The potassium concentration in soil ranges between 0.04% and 3%, but only 1–2% of the total concentration is available to plants, and 98–99% is fixed in its mineral form [96]. The remaining 90–98% of soil K is fixed with minerals and termed mineral K, which is unavailable to plants. In soil, K is found in various forms, including solution K, mineral K, exchangeable K, and non-exchangeable K [96]. The solution form of K is promptly consumed by plants and microbes and might leach down in some soils. Generally, in agricultural soils, solution K ranges from 2 to 5 mg L−1. Minerals containing K are feldspar and mica. Unexchangeable K reacts with oxygen atoms in the interlayers of certain clay minerals and becomes fixed. This form of K comprises 1–10% of total soil potassium [97].
Nutrient acquisition in soil is enhanced by microbial activity via the processes of decomposition, mineralization, and storage [52]. Integrated use of synthetic and organic sources in various moisture regimes is important to sustain soil physical health, nutrient availability, microbial activity, and crop production [91]. Potassium-solubilizing microbes are considered major contributors to K dissolution in soil, and researchers have also used strains of potassium-solubilizing bacteria to improve nutrient availability and crop yield [98]. Saprophytic bacteria, actinomycetes, and fungi are known as K-solubilizing microbes [99]. Potassium-solubilizing bacteria are found in the root rhizosphere [100]. Potassium-solubilizing bacterial isolates with more efficiency still need to be screened with the use of advanced Aleksandrov medium, which largely depends on the halo zone structure adjoining the bacterial colonies [101,102]. The solubilization of rock silicate by various bacterial species is well known. In soil bacterial groups, Bacillus mucilaginosus, B. edaphicus, and B. circulans are the most efficient K solubilizers. Potassium solubilization by microbes is affected by pH, aeration, the bacterial strains utilized, and the type of K-bearing natural resources. Hence, a slight increase in soil pH dramatically changes the solubilization of silicate [103].
Low-molecular-weight organic acids, including tartaric acid, lactic acid, fumaric acid, glycolic acid, oxalic acid, malic acid, citric acid, gluconic acid, and 2-ketogluconic acid, are produced by KSB and are able to release fixed K from K-bearing minerals [58,99,104]. Gasses such as ammonia (NH3) and H2S (hydrogen sulfide) are released as a result of organic matter decomposition which, after oxidation, form strong acids such as H2SO4 (sulfuric acid) and HNO3 nitric acid. Cations such as K+, Ca2+, Mn2+, and Mg2+ are displaced from the cation exchange site by H+ ions [105]. KSB produce organic acids that help to release K from mineral complexes by forming chelates with Fe2+, Al3+, Si4+, and Ca2+ ions bound to K minerals [106]. According to one study [107], weathering of phlogopite is facilitated by KSB through acidic dissolution of the crystal lattice and Al3+ (aluminum) chelation. Moreover, isolated strains of B. altitudinis can hasten the weathering of K-containing feldspar mineral, change surface morphology, and induce the development of a novel mineral complex. One strain dissolved K in the mineral (feldspar) by producing organic acids and released considerably more Al, Fe, and Si [105]. It has been observed that the assembly of extracellular polymers, such as proteins and polysaccharides, affects mineral dissolution and leads to production of the available form of K from K-bearing soil minerals that can be taken up by plants [108,109]. KSB create a microenvironment by synthesizing biofilms around the microbial cells for the purpose of weathering [106]. Due to biofilm formation, the residence time of water on aluminosilicate will be high compared with the residence time on bare rock, ultimately leading to high mineral weathering. Microbial biofilms hasten the process of weathering in addition to decreasing nutrient losses from minerals by acting as a defensive layer on the mineral. Furthermore, the formation of a biofilm on the mineral surface promotes the decay of K-rich shale and the release of K, Al, and Si in the bacteria–mineral contact model [110]. The use of compost not only increases the activity of potassium-solubilizing microbes but also acts as a source of mineral potassium that is released into soil [111]. According to Wang et al. [112], plant roots release organic acid, which is effective in mobilizing bound K. Other PGPR (e.g., IAA-producing bacteria) may also play a role by providing K to plants through an increase in root exudate formation [113]. A summary of K-solubilizing microbes involved in plant growth promotion is given in Table 1.

3.4. Siderophore Production

Siderophores are small organic molecules formed by microorganisms in iron-deficient conditions that improve the ability to uptake iron [114]. As a micronutrient, iron is important for the survival of all organisms. The predominant form of iron present on Earth is Fe3+, but this form is poorly soluble; therefore, the amount of Fe available to living organisms is very small. Microorganisms and plants need a sufficient amount of Fe, which is of greater concern in the root region, where plants and microbes compete for Fe [115]. PGPR generate siderophores of low molecular weight in an iron-limiting environment in order to make iron available to plants [116]. Thus, they increase plant growth by enhancing iron availability in the rhizospheric zone [117]. The siderophores create Fe competition in the rhizospheric zone, which decreases pathogenic microbe abundance [114,118] and increases plant growth. Siderophores produced by PGPR improve plant Fe uptake and reduce the growth of pathogens by showing a high affinity for rhizospheric Fe3+ and retaining almost all free iron [119].
Siderophores can be divided into four main groups based on their functional groups: hydroxamates, catecholates, carboxylates, and salicylates [120]. The beneficial effects of siderophores produced by PGPR on plant growth have been demonstrated in many studies (Table 1). For example, in some studies, radiolabeled ferric siderophores were used as a source of iron, and plants were able to uptake the labeled iron [116]. When mung bean plants were inoculated with siderophore-producing Pseudomonas strain GRP3 and grown in iron-deficient soils, chlorotic symptoms were reduced, and the chlorophyll level increased as compared to the control plants [121]. In another study, P. fluorescens C7 Fe-pyoverdine was taken up by Arabidopsis thaliana plants, resulting in increased iron content in plant tissues [78]. In ecologically stressed conditions, such as pollution with heavy metals, the iron supply is very important. In such situations, siderophores alleviate the harmful effects of heavy metals on plants [122].

3.5. Production of Phytohormones

Phytohormones or plant growth regulators are organic constituents that promote plant growth [123]. PGPR stimulate the production of phytohormones in plants and hence promote their growth and development. Various bacteria performing this role have been isolated and characterized (Table 1), and promising results in crop growth promotion have been reported. Phytohormones such as gibberellins, cytokinin, abscisic acid, ethylene, and auxin can promote the blooming of root cells via the overproduction of adjacent roots, with a subsequent increase in nutrient and water uptake [124].
The system that plants use to endure stress is complex and intricate. Microorganisms use different biochemical and molecular mechanisms to improve plant growth. PGPR improve plant growth by adjusting hormones and the accessibility of nutrients in plants and stimulating resistance to disease-causing organisms [15,125]. These also generate specific metabolites that control plant pathogens in the root zone. For example, rhizobitoxine increases plant growth and expansion in stressed conditions by limiting ethylene production [126]. In addition, different bacteria have sigma factors that modify the expression of plant genes in a way that increases their survival in stress situations [26].
The most commonly studied phytohormone produced by PGPR is indole-3-acetic acid (IAA), which facilitates plant–microbe association [127,128]. The role of IAA depends on the endogenous IAA levels in plants [111]. Auxins produced by PGPR are involved in plant growth promotion by stimulating transcriptional alterations in hormones [129], enhancing root biomass, shrinking stomata size and density [130], and triggering the expression of auxin response genes [131].
Cytokinins and gibberellins are also produced by various PGPR (Table 1) [26,132]. Several strains of PGPR can stimulate a comparatively huge volume of gibberellins, leading to increased plant shoot growth [133]. The association of these hormones with auxins can modify the root structure [134]. PGPR also produce cytokinins, which increase the production of root exudates by the plant [131], perhaps increasing the presence of PGPR that are associated with the plant.
The role of PGPR in stressed and unstressed environments has been studied and they frequently appear to provide better growth stimulation in stressful situations, such as drought stress [135]. For various PGPR, ethylene plays a key role in increasing plant tolerance under stress conditions [136]. Several studies have shown improved stress tolerance in plants after inoculation with PGPR that produce 1-aminoacyclopropane-1-carboxylase (ACC) deaminase [135,137]. This seems to arise because PGPR are able to prevent ethylene from reaching levels that decrease plant growth [131,138], which has been tested with Camelina sativa [139].
Furthermore, working as biocontrol agents, PGPR protect plants against pathogens by producing biochemical and molecular defense responses in the plant [140]. PGPR can activate ISR within plants, which stimulates the expression pathogenesis-related genes, facilitated via phytohormone signaling pathways and defense regulatory proteins, to protect plants against future pathogen attacks [141].

3.6. Root Colonization and Increased Uptake of Plant Nutrients

Plant roots interact with a wide range of soil microorganisms inhabiting the rhizosphere [9,142]. An important aspect of such associations between the plant roots, rhizosphere, and rhizobacteria is the improvement in root growth and proliferation, which play significant roles in the transfer of nutrients and water to the upper plant parts [9,143,144]. For instance, such an association results in enhanced root exudation which, in turn, attracts microbial candidates toward the vicinity of the root, which eventually results in aggressive root colonization [145,146,147].
The concept of root colonization denotes the multiplication of bacterial populations ectophytically in the rhizosphere and endophytically inside the roots [148]. Root colonization by PGPR is considered essential for plant development. In the process of root colonization, rhizobacteria propagate from the source of the inoculum, such as in seed treatments, to the actively growing root region and proliferate in the rhizosphere [149]. Roots play an important role in the uptake of essential nutrients needed for plant growth and survival. In the rhizosphere, aggressive root colonization enhances plant growth, which is indicative of positive root–rhizosphere–rhizobacterial interaction [119]. Moreover, several rhizobacteria produce one or more types of phytohormones in the rhizosphere (Table 1) that also activate phytohormone-producing genes through organic compounds abundant in the root cap and the elongation region [145,148]. For example, Lakshmanan et al. [150] observed that when B. subtilis strain FB17 was inoculated in the roots of Arabidopsis thaliana, it expressed multiple genes involved in metabolism, stress response, and plant defense during the root colonization process.

4. Plant Growth-Promoting Rhizobacteria (PGPR) as Biocontrol Agents: An Overview

In order of decreasing abundance, bacteria, actinomycetes, fungi, protozoa, nematodes, and microarthropods are among the microbes found in the rhizosphere. The plant rhizosphere is a thin layer of soil that adheres to the root surface [132,151]. Plant-associated microbes impact the growth, yields, and physiological and environmental benefits of the host plant in a variety of ways.
Plant disease outbreaks are a major cause of decreased crop yield, deteriorating production quality, and causing the contamination of food grains. Pesticides have been developed in response to the ever-increasing range and complexity of plant diseases [152,153]. Unfortunately, continued use of these pesticides has resulted in phytopathogen resistance, which raises a number of environmental concerns. Biological control is being considered as an alternative to pesticides for phytopathogen control. [154]. Plant growth and health are assisted by the use of PGPB as biological agents. PGPB have a number of benefits over traditional pest control methods. The use of PGPB in agriculture is both environmentally friendly and non-toxic. PGPB work through a variety of mechanisms to reduce or avoid harm caused by phytopathogens [24].
Plant growth is influenced by PGPR in two ways: indirectly and directly. Direct plant growth promotion by PGPR involves either providing the plant with bacterium-produced compounds, such as phytohormones, or promoting the absorption of certain nutrients from the environment [41]. As PGPR reduce or prevent the negative effects of one or more phytopathogenic species, they indirectly promote plant growth. This can be accomplished by generating antagonistic compounds or inducing pathogen resistance [41]. One or more of these mechanisms can be used by a specific PGPR to influence plant growth and development.
PGPR may function as biocontrol agents via a variety of mechanisms (Table 2) irrespective of their position in plant growth enhancement, such as the establishment of auxin phytohormone development [51], reduction in plant ethylene levels [155], or nitrogen fixation [156]. Plant–PGPR interactions are commercially exploited [157], and they hold great promise for long-term agriculture. A number of commercial and food crops have been studied in relation to these associations [158].
Furthermore, siderophore synthesis is one of the most important PGPR mechanisms for preventing phytopathogen propagation. These siderophores serve as iron chelators, binding the majority of the iron in the rhizosphere. As a result, the lack of iron in the rhizosphere inhibits bacterial and fungal pathogen proliferation [24]. Systemic resistance to phytopathogens is another potential mechanism used by PGPR in pathogen biocontrol.
The use of bacteriophage as a biocontrol agent has been a promising yet uncommon technique in recent years. Phages have the inherent ability to address phage resistance or new bacterial strains and are compatible with a variety of other biocontrol agents. Because of their sensitivity to UV light, they must be sprayed on the plant in the evening [166]. They can be used in phage-based diagnostics of phytopathogenic bacteria in addition to being used as biocontrol agents.
Because of the negative effects of chemical pesticides on both ecosystems and humans, sustainable agriculture has become a global necessity in recent years [10]. Various studies have investigated the use of PGPR in increasing crop production, revealing their potential to improve crop nutrition, yield, and disease management [12,132,167,168]. The use of PGPR has allowed decreasing chemical inputs into the soil and helped to mitigate environmental hazards caused by the overuse of chemicals [169,170]. Singh et al. [167,168,171] have identified PGPR as one of the suitable alternatives for use as growth promoters and biocontrol agents.

5. Mechanisms of Action of PGPR as Biocontrol Agents

5.1. Production of Antibiotics

The key method of plant growth-promoting bacteria for combating phytopathogen-caused damage is the production of antibiotics (Figure 1, Table 2). The biocontrol abilities of Pseudomonas strains are largely dependent on aggressive root colonization, induction of plant systemic resistance, and the development of antifungal antibiotics [172]. The potential of rhizobacteria as biocontrol candidates against phytopathogens is generally associated with the development of one or more antibiotics [155]. Over the last two decades, the concept of antibiosis, or biocontrol based on the synthesis of molecules that destroy or slow the growth of the target pathogen, has become well defined [140,173,174]. Antibiotics are a diverse group of organic, low-molecular-weight compounds that inhibit the growth or metabolic activities of bacteria and other microorganisms [175]. According to [176], six groups of antibiotic compounds are best linked to the biocontrol of root diseases: phenazines, phloroglucinols, pyoluteorin, pyrrolnitrin, cyclic lipopeptides, and hydrogen cyanide (HCN), all of which are diffusible except for HCN, which is volatile. Lipopeptide biosurfactants developed by Pseudomonas and Bacillus species have recently been applied in biocontrol because of their possible beneficial impact on competitive interactions with bacteria, fungi, oomycetes, protozoa, nematodes, and plants [177,178]. Antibiotics produced by Pseudomonads include 2,4-diacetylphloroglucinol (DAPG), amphisin, hydrogen cyanide, phenazine, oomycin A, tropolone, pyoluteorin, tensin, pyrrolnitrin, and cyclic lipopeptides, while Streptomyces, Bacillus, and Stenotrophomonas spp. produce kanosamine, oligomycin A, xanthobaccin, and zwittermicin, which have been identified as antibiotics that have antibacterial, antifungal, antiviral, anthelminthic, antimicrobial, cytotoxic, phytotoxic, antioxidant, and antitumor properties [179].
Antibiotics have been isolated from a wide range of fungal and bacterial species, with mechanisms of action that include inhibiting pathogen cell wall synthesis, influencing cell membrane structures, and inhibiting the development of initiation complexes on the small subunit of the ribosome (Table 2) [180]. Pyrrolnitrin, an antibiotic developed by the P. fluorescens BL915 strain, has been shown to protect cotton plants from Rhizoctonia solani during damping-off [181].
Antibiotics produced by the majority of Bacillus species, such as polymyxin, circulin, and colistin, are effective against Gram-positive and Gram-negative bacteria as well as many pathogenic fungi [180]. The B. cereus UW85 strain contributes to the biocontrol of damping off in alfalfa by suppressing oomycete pathogens and producing the antibiotics zwittermicin A (aminopolyol) and kanosamine (aminoglycoside) [182,183]. Some researchers have highlighted the use of sporulating Gram-positive species, such as Bacillus and Paenibacillus spp., as biocontrol agents to serve as a biological solution, which can confer higher population stability during the formulation and storage of inoculant products [184,185].

5.2. Induction of Plant Systemic Resistance

Non-pathogenic rhizobacteria are said to suppress disease in plants by inducing a resistance mechanism known as induced systemic resistance (ISR) [186]. When plants are properly stimulated, they gain an enhanced defensive capacity, which is known as induced resistance [186]. Van Peer et al. [187] identified ISR in carnation plants that were systemically protected against F. oxysporum by the P. fluorescens strain WCS417r, and in [188], the leaves of cucumber plants with rhizobacterial strains were protected against anthracnose caused by Colletotrichum orbiculare. The pathogens and the studied rhizobacteria were inoculated and remained confined and spatially isolated on the same plant, preventing microbial antagonism and allowing the protective effect to be mediated by the plant.
The ISR is a central mechanism for Pseudomonas, Trichoderma, Bacillus, and mycorrhiza to improve plant protection against various pathogens. ISR is the product of pathogen-specific recognition by plant receptors [141,179]. Salicylic acid is generated by several PGPB, which activate systemic acquired resistance (SAR), a mechanism similar to ISR. Another important biocontrol mechanism is phytopathogen virulence factor detoxification. Toxins formed by Xanthomonas albilineans and Fusarium species can be detoxified by certain PGPR [154]. Quorum sensing is used by many PGPR to control virulence factor development, which inhibit the pathogen’s quorum sensing ability by interfering with autoinducer signals, halting the expression of virulence factors [24].
Rhizobacteria-mediated ISR is similar to pathogen-induced systemic acquired resistance (SAR) in that both types of induced resistance result in uninfected plant parts becoming more resistant to plant pathogens, such as fungal, bacterial, and viral pathogens as well as nematodes and insects [186,189,190,191]. In the same plant, the same strain causes resistance to multiple pathogens [192]. Pseudomonas and Bacillus spp. are the most studied rhizobacteria that cause ISR [193,194,195]. Vleesschauwer and Höfte [192] coined the term ISR to describe induced systemic resistance caused by non-pathogenic rhizobacteria or PGPR, regardless of the signaling mechanism involved, while SAR refers to salicylic acid-dependent induced resistance caused by a localized infection.
In several plant species, the ability to develop ISR in response to specific rhizobacteria has been demonstrated [186] and appears to be dependent on the specificity of the rhizobacteria–plant interaction [196]. Failure to induce ISR in certain hosts may be due to a lack of inducing components being generated in the rhizosphere or a plant species’ inability to perceive certain compounds [196]. The evidence suggests that induction of resistance requires specific identification between the plant and the rhizobacteria. Pseudomonas putida WCS358r and Pseudomonas fluorescens WCS374r, for example, behave differently depending on the plant species: WCS358r elicits ISR in Arabidopsis, but not in radish or carnation plants [187,194,197,198]. Radish plants, on the other hand, react to WCS374r, while Arabidopsis does not [194,198].

5.3. Rhizosphere Competence and Root Colonization

For bacteria to be considered valid PGPR, root colonization is needed, and it is widely assumed that a biocontrol agent must colonize the rhizosphere and the surface of the plant that it protects [199,200,201]. As a result, if a PGPR species does not effectively colonize the roots, it is also ineffective as a biocontrol agent against root disease [202].
The most common root-colonizing PGPR in various crops are Pseudomonas and Bacillus spp. (Table 2). Several members of this group have widespread soil distribution, are effective rhizosphere colonizers, and produce a variety of metabolites that inhibit a wide range of pathogens in plants [203]. Many other root-colonizing PGPR strains have also been discovered to have antifungal properties against a variety of soil pathogens.
Biological regulation of soilborne diseases is also inconsistent among several possible biocontrol strains, such as Pseudomonas and Bacillus spp. Inadequate root colonization by introduced bacteria is one of the major causes of this inconsistency [204]. Mutants of Pseudomonas strains that had lost their biocontrol activity serve in verifying the connection between the poor biocontrol output of a biocontrol agent and inefficient root colonization. Understanding the bacterial traits that lead to root colonization is critical in this regard.
Molecular techniques can now be used to identify and count microorganisms on plant surfaces in real time. One of the methods used in the study of in situ bacterial root colonization was the use of marker genes, such as the gfp gene, which encodes the green fluorescent protein (GFP). Confocal laser scanning microscopy (CLSM) can be used to monitor and visualize GFP-transformed bacteria [204]. This GFP-based technique can be used to research the colonization patterns of various biocontrol agents in addition to visualizing root colonization.

5.4. Outcompetition and Direct Antagonism Against Pathogens

Biocontrol agents are bacteria that minimize the occurrence or severity of plant diseases, while antagonists are bacteria that have antagonistic behavior against a pathogen [205]. The following bacterial antagonistic behaviors and rhizospheric climates can be highlighted: (1) synthesis of hydrolytic enzymes that can lyse pathogenic fungal cells, such as chitinases, glucanases, proteases, and lipases [180,206], (2) competition for nutrients and appropriate colonization of niches at the root surface [207,208,209], (3) control of plant ethylene levels through the ACC deaminase enzyme, which can act to modulate ethylene levels in a plant in response to infection-induced stress [186,210], and (4) development of siderophores and antibiotics. The development of secondary metabolites, such as auxins, IAA, cytokinins, riboflavin, and vitamins, may be a direct result of the PGPR stimulation of plant growth [211]. Cell division and expansion [212] or improved nutrient availability [41,213,214,215] stimulate plant organ development.

5.5. Synthesis of Hydrogen Cyanide (HCN)

Biocontrol mechanisms of bacteria, such as those present in some Pseudomonas strains, are typically dependent on secreted bioactive factors that target the pathogen, such as antibiotics, exoenzymes, or HCN, according to Thomashow and Weller [216]. Dekkers et al. [217] discovered that phenazine-1-carboxamide (oxychlororaphin, or OCP), a phenazine formed by P. chlororaphis PCL1391, prevented F. oxysporum f.sp. radicis-lycopersici from causing tomato root rot. Because of their rapid and violent colonization of plant roots, fluorescent pseudomonads have been considered effective biological control agents against soilborne plant pathogens, according to Lugtenberg et al. [218]. They observed that one process was competition for nutrients in the rhizosphere at favored colonization sites, while another was the development of metabolites, including antibiotics, siderophores, and hydrogen cyanide. According to Kremer and Souissi [219], rhizobacteria strains can synthesize hydrogen cyanide and have an impact on seedling root growth in a variety of plants. They discovered that about 32% of bacteria in a sample of over 2000 isolates were cyanogenic, with HCN levels ranging from trace to >30 nmol/mg cellular protein. Pseudomonads were found to be the most susceptible to cyanogenesis, which was aided by the addition of glycine to the culture medium.

5.6. Synthesis of Cell Wall Degrading Enzymes

The development of cell wall degrading enzymes is one of the main mechanisms used by biocontrol agents to regulate soilborne pathogens [220,221]. Cell wall degrading enzymes secreted by biocontrol strains of PGPR, such as β-1,3-glucanase, chitinase, cellulase, and protease, have a strong inhibitory impact on the hyphal growth of fungal pathogens. Chitinase and β-1,3-glucanase are enzymes that break down chitin, a soluble linear polymer of β-,4-N-acetylglucosamine, which is a major component of fungal cell walls.
Paenibacillus and Streptomyces spp. produce β-1,3-glucanase, which lyses the fungal cell walls of pathogenic F. oxysporum. Bacillus cepacia produces b-1,3-glucanase, which breaks down the cell walls of soilborne pathogens such as R. solani, P. ultimum, and S. rolfsii [48]. B. licheniformis, B. cereus, B. circulans, and B. thuringiensis are all potential biocontrol agents with chitinolytic activity [222]. Serratia marcescens, Enterobacter agglomerans, Pseudomonas aeruginosa, and Pseudomonas fluorescens have been found to have chitinolytic activities among Gram-negative bacteria [223].
The structural integrity of the target pathogen’s cell walls is affected by rhizobacteria cell wall degrading enzymes [224]. The chitinolytic and antifungal activities of S. marcescens B2, a potent biocontrol strain, against the soilborne pathogens R. solani and F. oxysporum were investigated by Someya et al. [225]. The mycelia of fungal pathogens co-inoculated with this strain displayed a variety of abnormalities, including partial swelling in the hyphae and at the tip, hyphal curling, and hyphal tip bursting.
Several bacteria produce enzymes that hydrolyze cellulose, hemicelluloses, chitin, and proteins, which inhibit phytopathogen activity. Serratia plymuthica C48, Serratia marcescens, Paenibacillus sp., Streptomyces sp., and Pseudomonas stutzeri produce chitinase, which degrades the mycelia of different fungal phytopathogens. Streptomyces, Paenibacillus, and Bacillus sp. produce β-1,3-glucanase, which lyses fungal cell walls. PGPR protease and lipase can also degrade proteins and lipids associated with cell walls. Hydrogen cyanide is produced by Pseudomonas, Rhizobium, Bacillus, Alcaligenes, and Aeromonas and improves the antifungal activity of these bacteria [12,24,48,152,179].

6. Plant Growth-Promoting Rhizobacteria (PGPR) in Bioremediation: An Overview

High concentrations of metalloids and trace metals can contaminate soil as a result of pollution from the rapidly developing industrial sector, dumping of petrochemical spillage, trace metal waste, leaded fuel, mine tailings, pesticides, sewage sludge, atmospheric deposition, and coal combustion residues [226,227]. Arsenic, chromium, cadmium, lead, nickel, copper, mercury, and zinc are the most common inorganic pollutants found in industrial wastes [16]. Organic contaminants (i.e., antibiotics such as tetracyclines, sulfonamides, macrolides, and quinolones), pesticides (i.e., bentazon and atrazine), polycyclic aromatic hydrocarbons (PAHs), total petroleum hydrocarbons (TPHs) from the consumption and exploration of fossil fuels, polychlorinated biphenyls (PCBs), halogenated compounds (i.e., trichloroethylene and perchloroethylene), and polychlorinated terphenyls (PCTs) are extensively used in industrial sectors and are considered most resistant [228,229,230,231]. Pesticides also introduce a number of molecular, morphological, physiological, and biochemical changes in plants that adversely affect productivity and growth in addition to the development of pest resistance [232]. Thus, in order to clean up trace metals and polluted environments, bioremediation and remediation technologies are constantly being improved, using naturally occurring or genetically engineered microorganisms [233,234]. According to recent research, among microorganisms used for bioremediation, the utilization of plant growth-promoting rhizobacteria (PGPR) is becoming more common due to their various abilities to detoxify and degrade toxins, as well as their significant effects on plant growth promotion [229,231]. In contrast to physical and chemical remediation approaches, bioremediation with PGPR is gaining more attention for the removal of industrial waste contaminants due to its environmentally friendly nature, lower cost, and demonstrated performance [230]. The ability of PGPR to improve plant growth and overcome trace metal toxicity can be aided by their interactions [227]. Microbes such as PGPR colonize the rhizosphere or live near the surface of roots, and they tend to produce and secrete a variety of regulatory compounds, i.e., metal-binding proteins, phytohormones, and siderophores, to protect the plant from toxicity [233]. Similarly, dark brown sugarcane (Saccharum officinarum L.) molasses, because it is rich in complex organic compounds and trace metals, inhibits seed germination and depletes vegetation on agricultural soils by lowering the alkalinity of soil and the availability of manganese [235,236]. Recently, PGPR were used in the bioremediation of sugarcane molasses-based anaerobically digested distillery effluent [237]. Distillery spent wash disposed on land is considered very hazardous waste, causing a decrease in the alkalinity of soil, seed germination inhibition, and vegetative destruction [231]. Moreover, it also lowers the penetration of sunlight in marine environments, thereby lowering the photosynthetic activity and dissolved oxygen content and harming both aquatic flora and fauna [233,235]. Correspondingly, the ability of grasses and native weeds to remove trace metals by using in situ phytoremediation of distillery waste was also observed [238]. Previously, a number of rhizosphere bacteria were screened and tested for their bioremediation potential against inorganic and organic pollutants from the soil, as shown in (Table 3 and Table 4). Bacterial pretreatment of agricultural effluents mediates the modification and oxidation of inorganic and organic contaminants, making them readily bioavailable to wetland plant roots and rhizosphere microorganisms, the latter of which use the biotransformed materials as biomass, nitrogen, and energy supply [239,240,241]. Rhizosphere bacteria are equipped with many mechanisms, including bioaccumulation (net accumulation of contaminants in microorganism cells), biomineralization (transfer of aqueous contaminants into crystalline or amorphous precipitates), biotransformation (transformation of contaminants from toxic to less toxic forms), and biosorption (binding of contaminants with cation binding proteins present on the cell wall of microorganisms) [3,7,16,231]. To overcome the toxicity of trace metals in the rhizosphere, microorganisms have developed a range of pathways, including (1) transporting them to the exterior of the cell by metal ion pumping [242], (2) sequestration and accumulation of metal ions inside the cell [243], (3) transformation/conversion of toxic metals to a less toxic form [244], and (4) desorption/adsorption of metals [245]. Root exudates play a significant role in bacterial quorum sensing and biofilm formation because they can chemotactically attract rhizobia to plants, resulting in colonization and adherence to the legume roots along with the control of genes involved in rhizobial nodulations (rhizosphere and nod-expressed (rhi)), which enhances the development of plants and bacterial remediation potential [241,246]. Overall, trace metals are mobilized during phytodegradation via acidification, chelation, and protonation, while they are immobilized via alkalinization, complexation, and precipitation [234,237,247]. Through a variety of plant growth pathways, PGPR may be used to maximize crop yields [245]. Environmentally sustainable methods have encouraged the utilization of a wide variety of beneficial, agriculturally significant bacteria, which has resulted in increased nutrient absorption and plant health [235]. PGPR are also important in improving soil fertility, plant health, and pollutant remediation [236]. However, engineering PGPR to produce outcomes in novel agricultural endeavors would be extremely beneficial to the microorganism inoculant industry [229]. One of the extremely important sectors in undeveloped and developed countries is agriculture. The pervasive use of chemicals in modern agriculture has been a source of public concern for the past three decades due to potential negative impacts on the environment, as well as on animal and human health [231]. Agricultural activity is shifting to more sustainable and environmentally responsible strategies around the world, fueled by increasing awareness of the environmental harm and human health risks caused by toxic waste and the overuse of fertilizers and pesticides, prompting increasing demands to address these issues. Therefore, using PGPR to improve plant growth under normal conditions, abiotic stress, and plant pathogen attack is crucial, as these outcomes may not be achieved without PGPR (Figure 2).

7. Mechanisms of Action of PGPR in Bioremediation

7.1. Siderophores and Heavy Metal Removal

Microorganisms produce siderophores, which demonstrate good affinity in chelating iron [304,305]. Siderophores are produced by various plant growth-promoting rhizobacteria (PGPR). The solubility of iron phosphate in soil is influenced by siderophores [306]. Due to the formation of siderophores, which are low-molecular-weight proteins, bacterial activities that enhance mineral nutrient absorption by plants can help plants grow in trace metal-contaminated soils [301,307]. Fungi, bacteria, and plants contain iron-chelating secondary metabolites in iron-limiting environments [305,308]. Plant growth can be stimulated by siderophore-producing PGPR either directly by enhancing plant iron nutrition or indirectly by reducing the activity of plant pathogens in the soil root zone by minimizing their iron supply [309]. Iron is more often found as Fe3+, which is bound to insoluble oxyhydroxides and hydroxides, making it inaccessible to microorganisms and plants. Fe3+ has a greater affinity for siderophores as compared with Fe2+ or other trace metals, i.e., Co, Cu, Cd, Pb, Ni, Cr, and Zn [235,301]. As rhizobacterial siderophores tend to reduce the stress exerted by trace metal pollutants, iron supply to developing plants under trace metal stress becomes more critical [310]. Interestingly, phytosiderophores have a lower affinity for iron than bacterial siderophores but require lower iron content for optimal growth as compared to bacteria [306,311]. Plants belonging to the Poaceae family have significant potential to enhance the iron supply and its uptake in roots [268,311,312]. Growing siderophore-producing grass species in conjunction with accumulator plants will enhance phytoextraction processes [313]. While siderophore production for the remediation of trace metals has significant potential, phytosiderophores gain their specificity by absorption of iron phytosiderophores via a membrane carrier, rather than by directly chelating iron in soils [235]. Phytoextraction is enhanced by improving the nutrition of plants and mobilizing metals [295]. By selectively promoting iron uptake from the reservoir of trace metal cations vying for transport, siderophore-producing bacteria have been shown to enhance the growth, chlorophyll contents, and biomass of various crops cultivated in trace metal-contaminated soils [306]. Furthermore, the complexation of trace metals by rhizobacterial siderophores in the rhizosphere possibly inhibits the formation of oxidative stress and free radicals [261]. Membrane receptor proteins recognize and scavenge the Fe3+ siderophore complexes from the rhizosphere, which are then secreted. They are too large to pass through membrane porins [268]. Instead, ATP-binding cassette transporters mediate transport through the cytoplasmatic membrane, especially in Gram-negative bacteria, and the TonB-dependent transporter proteins transport metabolic energy from the cytoplasm to the outer membrane [305]. However, this type of mechanism is scarcely heard of in Gram-positive bacteria, as they are TonB-dependent transporter proteins [163]. Often, bacteria develop numerous types of siderophores, but the formation of siderophores and various molecules in a contaminated rhizosphere depends on the circumstances, i.e., bacterial strain, soil type, climatic conditions, and pollutant concentration [235,301].

7.2. Biosurfactants in Heavy Metal Removal

Biosurfactants classified as amphiphilic molecules, which have large hydrophobic and hydrophilic groups, are one of the most important agents in the remediation of contaminants in the rhizosphere [164,235]. Surfactants have a hydrophilic component (functional groups) that makes them soluble in water and a hydrophobic part that enables them to form hydrophobic interiors and interact with hydrophobic compounds and adsorb at interfaces (e.g., air/water), thereby lowering the surface tension [314,315]. Biosurfactants, the biobased substitutes of surfactants, have risen to prominence as a result of strong toxic manifestations [316]. Biosurfactants, which are microbial surface-active metabolites, are metal-complexing agents that have been shown to be useful in the remediation of heavy metal-contaminated areas [164,304,317]. Biosurfactants have unique properties that make them a possible replacement for conventional remediation methods [314]. These are less toxic in nature and have higher biodegradability and environmental compatibility [318]. Other benefits include their ability to be made from low-cost agro-based raw materials and organic wastes, as well as their ability to maintain functionality even at extreme salt concentration, temperature, and pH in the rhizosphere [319]. The majority of biosurfactants discovered so far come from terrestrial microorganisms [164]. Biosurfactants are produced by a variety of microbial genera, including yeasts, Enterobacter, Rhodococcus, Bacillus, Halomonas, Pseudomonas, Arthrobacter, and Acinetobacter. Brevibacterium sp. and Ochrobactrum sp. are bacteria that produce biosurfactants in crude oil-contaminated soils [237,320]. Emulsifying agents have been identified as Achromobacter sp., Brevibacillus sp., Pusillimonas sp., Dietzia sp., and Sphingopyxis sp. [321]. In recent studies, it was observed that trace metals, i.e., Cd, Zn, and Pb, had higher affinities for biosurfactants, such as rhamnolipid formed by various Pseudomonas aeruginosa, than for many of the soil components to which they are bound in metal-contaminated soils [322,323]. Heavy metal desorption from solid phases is aided by biosurfactants in two ways [164]. The first is the complexation of free metals in solution. According to Le-Chatelier’s principle, this reduces the metal’s solution phase activity and thus promotes desorption [323]. Secondly, biosurfactants aggregate at the solid–liquid interface and decrease the interfacial stress. The authors of [324] also stated that the size and charge of biosurfactants structures influence the flow of biosurfactant–metal complexes through the soil [235]. The ability of biosurfactants to form complexes with metals is central to their bioremediation of trace metal-contaminated soil [325]. Biosurfactant micelles may also strip metal ions from the rhizosphere [236]. The main effectiveness of metal-activated biosurfactant action is influenced by a series of factors, including soil pH, soil particle size, soil structure, cation-exchange capacity (CEC), trace metal concentration, and climatic conditions [164,321]. Because of the long period of contamination, the trace metal has much time to stabilize, making removal more difficult [322]. Biosurfactants’ wide range of applications in organic and trace metal-contaminated soils can be attributed to their low toxicity, small size, biodegradability, cost-effectiveness, and high specificity [326]. The significance of biosurfactants in facilitating systems as biocontrol agents has yet to be fully investigated and warrants further research. Such research will aid in the replacement of harsh chemical surfactants with green alternatives. There is also more work to be carried out regarding the production costs of green surfactants in order to achieve net economic gain from biosurfactant use in the agricultural sector for the remediation of trace metal-contaminated soils.

7.3. Biosorption

Biosorption is a simple, metabolically passive physicochemical process that involves the binding of metal ions/biosorbate to the surface of a biologically derived biosorbent [327,328]. Biopolymers, industrial or agriculture waste, plant-derived materials, microorganisms, and other biological removal materials are used for biosorption remediation of trace metals [329]. In contrast to oxidation through anaerobic and aerobic metabolism, it is a reversible, rapid process that involves the binding of ions to the functional groups present on the surface of the biosorbent in aqueous solutions through various interactions [330,331]. Operational simplicity, low quantity of sludge generation, high efficacy, the lack of requirements for additional nutrient or increased chemical oxygen in water, regeneration ability of the biosorbent, and low operational cost are recognized as some of the benefits of this process [332]. Biosorption can eliminate contaminants even at low concentration, which is especially important for the removal of trace metals because they are toxic at ppb levels [333]. For the biosorption process, either living or dead microorganisms or other agricultural and industrial byproducts may be used as biosorbents [293]. In the first step of trace metal ion biosorption, the biosorbent should be suspended in a solution containing the biosorbate [334]. Equilibrium is achieved after a defined period of incubation. The metal-enriched biosorbent will be isolated at this point [332,335]. Biosorption is beneficial since it is reversible, does not require nutrients, has a single-stage mechanism of short duration, poses no risk of harmful effects or cellular expansion, allows for intermediate equilibrium concentrations of trace metal ions, and is not regulated by metabolism [336]. The attachment of the sorbate to the biosorbent is a dynamic process that occurs during biosorption [337,338]. Examples of biosorbent mechanisms to remove trace metal ions include physical (van der Waals forces or electrostatic reaction) or chemical (displacement of bound trace metal cations, ion exchange, or protons) binding, complexation, precipitation, reduction, and chelation [293,328,339]. Biosorbents contain functional/chemical groups that can attract and sequester trace metal ions, such as phosphate, imine, phenolic, phosphodiester, carboxyl, sulfhydryl, carbonyl, sulfonate, thioether, imidazole, amide, and amine [329,335]. The biosorption mechanism is influenced by a number of factors, including soil pH, soil temperature, initial metal ion concentration, biomass concentration, agitation speed of the biosorbent, and biosorbent concentration [340,341]. In terms of trace metal detection, major improvements have been made in the last decade with the use of optical chemical sensors, graphene-modified nanomaterials, and other biomonitoring instruments [342]. Given these technological advances and the ability to reuse metal resources, biosorption is a competitive, affordable, and promising mechanism, and it can foster sustainable economic development in countries.

7.4. ACC Deaminase Activity and Reduction in Ethylene Levels

The trace metal stress-induced acceleration of ethylene development in plants is modulated by rhizobacteria with ACC deaminase activity, and this alteration in root architecture and the plant’s root uptake system can result in increased uptake of inorganic contaminants [15,343]. ACC deaminase rhizobacteria have great potential for the development of bacterial inoculation in a trace metal-contaminated rhizosphere to boost plant growth in unfavorable environments, particularly for hyperaccumulators [155,344]. The ACC deaminase derived from a rhizobacterium (Kluyvera ascorbata L. SUD165) has a higher potential to protect tomato (Lycopersicon esculentum L.) and canola (Brassica napus L.) seeds from nickel chloride toxicity when cultivated under gnotobiotic conditions by reducing the levels of ethylene under Ni stress [300]. Ethylene is considered a stress hormone that has been shown to encourage plant growth at low concentrations, but at moderate to high concentrations, it can prevent root elongation and senescence of the plant [345]. In another study, Belimov et al. [316] observed that rhizobacteria ACC deaminase promoted the growth of roots and shoots in canola seedlings grown in soil contaminated with 300 mM CdCl2. A goal of the study was to find a connection between rhizobacteria with in vitro ACC deaminase activity and increased cadmium accumulation in plant tissues due to increased root growth caused by the rhizobacteria [269]. Furthermore, ACC deaminase producing rhizobacteria are used to create plant–inoculant systems for phytoremediation of contaminated soil environments. Similarly, in another experiment, inoculation with a rhizobacterium Pseudomonas putida L. 06909 caused a significant decrease in Cd phytotoxicity and increased metal accumulation in sunflower (Helianthus annus L.) plant root by 40% [279]. It is highly probable that, in addition to other characteristics, rhizobacterium Pseudomonas putida L. 06909 has ACC deaminase activity, which aids in the reduction of trace metal phytotoxicity and increased metal aggregation in plant roots. Li et al. [288] observed that tobacco (Nicotiana tabacum L.) plants inoculated with a rhizobacterium (Pseudomonas putida L. UW4) showed good development and absorbed a significant amount of nickel from nickel-contaminated soil. The rapid growth of hyperaccumulators induced by ACC deaminase rhizobacteria can aid not only in remediation but also in the uptake of trace metals [15,345]. Using transgenic hyperaccumulator plants that express ACC deaminase is a technique for the phytoremediation of trace metal-contaminated agricultural soils [343]. However, further research is required to investigate different facets of this technique in order to improve its efficacy in the remediation of trace metals.

7.5. Production of Exopolysaccharides and Polymeric Substances

Since the discovery of the bacterial exopolysaccharide (EPS) adsorption potential, numerous studies have been published on a diverse range of bacterial species and EPSs with the required capability, i.e., remediation of trace metal-contaminated soils [346,347]. EPS, which has a simple polysaccharide backbone, can be structurally modified by changing the polymeric length or by adding an array of different side chains, functional units, non-carbohydrate substituents, and various bonds and linkages in a combinatorial manner [348,349]. The nature and percentage of the available carbon source, abiotic stress factors, including pH, temperature, and trace metals, and the growth process of rhizobacterium during which synthesis occurs all play roles in determining the polysaccharide composition [350,351]. The utilization of negatively charged EPS (EPS with abundant anionic functional groups) as a viable biosorbent must be emphasized in strategies for trace metal remediation using rhizobacterial EPSs [352,353]. Abundant ionizable and active non-carbohydrate side chains and functional groups, such as acetamido (chitin group), structural polysaccharides (fungi), sulfhydryl, amine, and carboxyl groups present in proteins and phosphodiester, hydroxyl, and phosphate groups in polysaccharides, impart an overall negative charge to the polymer [354]. In addition, unlike homopolysaccharides, extracellular heteropolysaccharides are frequently polyanionic due to the interaction of some of these functional groups with the polysaccharide backbone [355]. Immobilization and sorption occur by various processes, i.e., complexation, ion exchange, and precipitation. Certain documented commercial rhizobacterial EPS strains having the potential of anionicity are Enterobacter sp. A47, Pseudomonas oleovorans L., Xanthomonas campestris L., Streptococci sp., Pasteurella multocida L., Pseudomonas aeruginosa L., Sphingomonas paucimobilis L., and Azotobacter vinelandii L. [347,350,351]. The association between positively charged trace metal ions and negatively charged EPS and cell surfaces causes EPS- or intact microbial cell (dead or alive)-mediated biosorption [356]. Herminiimonasarsenicoxydans L., a rhizobacterium with a Gram-negative phenotype, has been reported not only to cause or activate the formation of biofilm in response to arsenic (As) contamination but also to use EPS to scavenge arsenic ions when exposed to concentrations up to 5 mM [351,357]. This study shows that although the synthesis of rhizobacterial EPS may not be induced in response to trace metal stress, the formed EPS can also adsorb the metal. Similarly, EPSs produced by Marinobacter sp. have been evaluated for the remediation of trace copper (Cu) and lead (Pb) [358]. However, a variety of safer rhizobacteria, which are scattered in the environment, are waiting to be discovered for the remediation of environments contaminated with trace metals. These EPS producers should be investigated for trace metal ion chelation capabilities, as they can develop a potent polysaccharide with anionic moieties.

8. Pros and Cons of Rhizosphere Bacteria for Agricultural Sustainability

Rhizosphere bacterial inoculants are indisputably necessary for the augmentation of plant growth and maintenance of soil output. As reported in the above sections, rhizosphere bacterial inoculants benefit plants through various mechanisms, although some studies indicate adverse effects [5]. In this section, we compare the pros and cons of rhizosphere bacterial biofertilizers, and a comparison of such biofertilizers is presented in Table 5 and demonstrated in Figure 3.

8.1. Pros of Rhizosphere Bacterial Application

Numerous commercially available microbial biofertilizers are sold as dried or liquid cultures under a variety of trade names, as listed in Table 5. The application of such rhizosphere bacterial biofertilizers could have an impact on agricultural sustainability and phytopathogen biocontrol and sustain soil and plant production by improving nutrient availability and reducing the application of chemical fertilizers and pesticides. The bioremediation and biodegradation of hazardous substances of biological or anthropogenic origin could also be improved [375]. Various beneficial rhizosphere bacterial inoculants could be formulated in the form of products, including biofertilizers. Such products are viable sources of nutrients that could act as alternatives to chemical fertilizers, stimulate plant growth, remediate heavy metal-contaminated environments, and mitigate environmental stresses [376,377].
The application of rhizosphere bacterial biofertilizers has many advantages over the conventional fertilization system. Initially, a small amount of rhizosphere bacterial inoculum is sufficient for the preparation of biofertilizer. Their application does not require energy sources for their survival under field conditions, as they possess saprophytic nutritional requirements that make their large-scale application feasible [378]. Their application is simple in terms of seed, soil, and/or root treatments. Crop seeds are treated with solid or liquid formulations of rhizosphere bacterial inoculants with or without a carrier material. Such formulations may also be directly applied in the field along with compost fertilizers. The inoculated bacteria can survive in the vicinity of the rhizosphere and colonize plant roots, where they have a beneficial influence on plants [379]. These bacterial inoculants spread along with the root system and improve plant growth through their versatile metabolism, as briefly discussed in the above sections.
The application of such bacterial biofertilizers sustains soil health and productivity by improving the soil bacterial community structure and composition and affecting afforestation [380]. Such biofertilizers benefit plants by facilitating renewable nutrient availability via atmospheric N fixation and solubilization of nutrients. The three types of N-fixing biofertilizers, namely, free-living, symbiotic/endophytic, and rhizosphere bacteria, were reported to enhance N availability to crops [381]. For example, rhizobia biofertilizers can fix 50–300 kg N ha−1, increase yield by 10–35%, protect soil fertility, and improve residual N availability for subsequent crops [382]. Various nutrient-solubilizing rhizosphere bacterial biofertilizers were also reported to increase the availability, uptake, translocation, and accumulation of P, K, Zn, Fe, Se, Mn, and Si by solubilizing them from minerals and producing various organic and inorganic acids [383]. Such biofertilizers are nonspecific and can be used on any crop. They produce enzymes to convert insoluble organic P to a soluble form and have increased crop yield by 10–30% [68]. These nutrient-solubilizing rhizosphere bacteria improve nutrient uptake, accumulation, and translocation into cereal grains and could be a promising option for the biofortification of cereals [80]. Rhizosphere bacterial inoculants mediate photostimulation by communicating with plants via quorum sensing, which enables bacteria to synchronize the expression of genes and behavior. Bacterial inoculants produce N-acyl-L-homoserine lactones, cyclodipeptides, and various phytohormones that are involved in the stimulation of plant growth and defense [384]. The application of rhizosphere bacterial inoculants could modulate systemic mechanisms in a plant to enhance its defense against adverse abiotic and biotic stress conditions. Resistance to biotic stress is elevated through the production of secondary compounds called allelochemicals that induce immunity against pathogen attack. Additionally, such bioinoculants play a multifaceted role in the alleviation of abiotic stress due to climate alterations and in the restoration of natural soil against a variety of toxic metals [87].

8.2. Cons of Rhizosphere Bacterial Application

Recently, there has been growing interest in the use of biofertilizer products. However, their application faces serious constraints at various levels, from laboratory screening to field applications. We classify all biofertilizer constraints broadly into research and development, regulatory and marketing, and field-level constraints. Such constraints are described in detail in the following subsections.

8.2.1. Research and Development Constraints

The biofertilizer constraints encountered in research and development can be classified into biological, technical, quality control, carrier, and biosafety constraints. Among the biological constraints, we may often fail to select a potential PGP rhizosphere bacterial strain, as this is a difficult task. For biofertilizer development, the initial evaluation of rhizosphere bacterial isolates for PGP characteristics under laboratory conditions does not ensure their efficacy in field conditions. For example, isolates from pure cultures demonstrating less in vitro growth-promoting activity showed greater plant growth promotion under field conditions. The potential isolated strains occasionally pose a challenge to screening due to the incomplete understanding of potential PGP mechanisms. As a result, such potentially useful strains with PGP mechanisms are occasionally rejected owing to their unsatisfactory results during in vitro experiments [371]. The efficiency of applied inoculants deteriorated in the presence of many other microorganisms under field conditions. Thus, target strains should be chosen based on their performance under field conditions, and they should be applied to various crops and grown in a variety of soil types [385]. The selected bioinoculant in a biofertilizer could be less effective at replacing native ineffective strains, less compatible with a crop, and less capable of colonizing host plant roots and thriving in soil [367].
The shelf life of biofertilizers and the commercialization of a successful rhizosphere bacterial inoculant remain as major challenges [386]. Biofertilizers with a short shelf life need to be recycled before expiration, which results in financial losses for the associated firm. Their storage and transportation require additional caution because biofertilizers are composed of live bacterial cells and can deteriorate under harsh environmental conditions [367]. A mutation in a bioinoculant may create a serious problem if it decreases its efficiency and thus raises the cost of production. To increase the shelf life of biofertilizers, a suitable carrier is needed for field application. The lack of a suitable carrier is a significant restriction on its widespread usage in fields. Peat, charcoal, and lignite are considered excellent carriers for biofertilizer processing; however, the majority are in short supply in developing countries, and mining of these carriers has been downscaled in developed countries. A potential carrier should be inexpensive, nearly sterile, and free from moisture and toxic substances in addition to having both high organic matter content and water-holding capacity [372]. At the moment, there are no quality control procedures for biofertilizers. It is necessary to develop quality control standards for biofertilizers to demonstrate their efficacy in promoting plant growth on a field scale [367].

8.2.2. Regulatory and Marketing Cnstraints

Biofertilizer production and quality control require advanced technology and highly qualified and trained human resources. The major infrastructure limitations are a lack of advanced technology, necessary technical support and equipment, trained workforce, and skilled technical personnel [367]. Regulatory constraints include difficulties in registering biofertilizer products and filing patents due to frequently changing and inconsistent regulations among regions and countries. The entire regulatory process of developing a potential biofertilizer, from registration to commercialization, is lengthy, potentially taking several years, and quite complex [385]. Financial constraints are a serious hindrance to large-scale biofertilizer production. After manufacturing the biofertilizer, small producers lack the financial resources necessary to distribute it independently. As a result of a delay in its distribution, the product quality deteriorates, which reduces its potential [367]. The lack of adequate transportation and storage facilities is a significant impediment to commercialization. Farmers either lack or have insufficient knowledge about the sustainable agriculture benefits of biofertilizers over hazardous agrochemicals. As a result, demand for such green products declines. Due to a shortage of qualified technical personnel, the establishment of extension centers has little effect on increasing farmer awareness [367]. During packing and marketing, the rhizosphere bacterial bioinoculants are exposed to high temperatures (≥40 °C), which may result in their inactivation or death, thereby reducing their value as biofertilizers. As a result, these low-quality packets will be detrimental to farmers and efforts to increase the overall crop yield.

8.2.3. Field–Level Constraints

Farmers are skeptical of biofertilizers due to the extremely slow and frequently unsuccessful crop responses to applied biofertilizers, as the inoculum requires a longer time to colonize the roots and for the effective concentration to be established. Biofertilizer efficacy is reduced in the field due to the residual properties of harmful chemicals [367]. Environmental stresses contribute significantly to the reduction in biological activity in some areas. Several other factors contributing to the poor performance of biofertilizers include nutrient availability, soil acidity and alkalinity, high and low temperatures, pesticide application, radiation, and high nitrate concentrations in the soil, which limit the bioinoculants’ ability to fix atmospheric N, solubilize nutrients, and interact with indigenous soil microbiota, which influences the presence and survivability of rhizosphere bacteria and the host plant [387]. Numerous soils are contaminated with heavy metals as well as deficient in other critical nutrients, which reduce the biological potential of rhizosphere bacterial inoculants in biofertilizers [388]. Region-specific rhizosphere bacterial biofertilizers should be identified to optimize the effectiveness of the used strains. Soil fumigation with broad-spectrum biocidal fumigants has a deleterious effect on the soil microbial community [389]. The inconsistent application of biofertilizer limits the presence of viable bacterial populations, which results in their inefficiency in promoting the growth of agronomic crops. Typically, farmers expect rapid, visible outcomes from a single application of biofertilizers, which represents another serious limitation to their wide-scale application. The limited application of biofertilizer could be due to the lack of awareness in farmers about the concentration, time, and method of biofertilizer application. Repeated applications of biofertilizer are needed to maintain the bacterial numbers and ensure a viable PGP rhizobacterial population in soil for alleviating various environmental stresses. However, the literature is limited in specifying the required biofertilizer dosage. Therefore, extensive research is required to evaluate the optimal dose of biofertilizers and their effects on crop productivity and stress alleviation.

9. Future Directions

The application of rhizosphere bacterial strains represents a sustainable technology in agriculture being embraced in many developed countries. Native plant microbiomes play an important role in increasing plant survival under abiotic and biotic stresses [390]. To promote crop productivity, the desired plant growth-promoting traits could be inserted into rhizosphere bacterial strains through recent innovative genetic engineering technologies. It is difficult to screen the bulk of rhizosphere bacterial strains to identify advantageous traits under stress conditions; however, a limited number of beneficial bacterial strains can be isolated, engineered, and applied to crops [373]. The importance of identifying and evaluating novel beneficial rhizosphere bacterial strains is emphasized. Core plant microbiomes can be attained through screening and signal transmission during various growth stages, including seed transfer and plant germination [391]. The intrinsic plant ability to stimulate the rhizosphere bacterial community can be useful in selecting the desirable strains. Rhizosphere bacterial strains can be subjected to whole-genome sequencing to differentiate their functional capabilities. Native plant rhizosphere bacterial diversity can be in situ modified through prevailing biochemical and molecular-based approaches [374]. Collective efforts are needed to realize the potential impact of rhizosphere bacteria on environmental restoration by broadening their applications. Public misunderstanding and misinformation must be overcome by educating the public about the advantages of rhizosphere bacteria. Novel strategies must be explored for the formulation and transfer of rhizosphere bacterial biofertilizer at the laboratory, greenhouse, field, and market scales to sustain agricultural productivity and the environment. The laboratory support of detailed mechanistic studies on the ability to stimulate growth and bioremediate environments contaminated with pollutants is essential. More mechanistically effective strains might be genetically engineered and regulated according to approved policies to prevent future hazards.

Author Contributions

Conceptualization, Q.S. and A.M.; methodology, W.X.; software, F.U.H.; validation, J.K., M.B., and M.Z.M.; formal analysis, A.M.; resources, J.H.; writing—original draft preparation, Q.S., M.E., and M.N. (Munaza Naseem); writing—review and editing, A.M., A.K., M.N. (Muhammad Naveed), and W.X.; funding acquisition, W.X. and M.B. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by Ministry of Education, Youth and Sports of the Czech Republic, grant number FCH-S-20-6446.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors gratefully acknowledge the Environmental Science Laboratory, Institute of Soil and Environmental Sciences, University of Agriculture, Faisalabad, Pakistan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bharti, N.; Barnawal, D. Amelioration of salinity stress by PGPR: ACC deaminase and ROS scavenging enzymes activity. In PGPR Amelioration in Sustainable Agriculture; Singh, A.K., Kumar, A., Singh, P.K., Eds.; Woodhead Publishing: Cambridge, UK, 2019; pp. 85–106. [Google Scholar] [CrossRef]
  2. Barrow, C. Biochar: Potential for countering land degradation and for improving agriculture. Appl. Geogr. 2012, 34, 21–28. [Google Scholar] [CrossRef]
  3. Niamat, B.; Naveed, M.; Ahmad, Z.; Yaseen, M.; Ditta, A.; Mustafa, A.; Rafique, M.; Bibi, R.; Sun, N.; Xu, M. Calcium-Enriched Animal Manure Alleviates the Adverse Effects of Salt Stress on Growth, Physiology and Nutrients Homeostasis of Zea mays L. Plants 2019, 8, 480. [Google Scholar] [CrossRef] [Green Version]
  4. Godfray, H.C.J.; Beddington, J.R.; Crute, I.R.; Haddad, L.; Lawrence, D.; Muir, J.F.; Pretty, J.; Robinson, S.; Thomas, S.M.; Toulmin, C. Food security: The challenge of feeding 9 billion people. Science 2010, 327, 812–818. [Google Scholar] [CrossRef] [Green Version]
  5. Saharan, B.S.; Nehra, V. Plant growth promoting rhizobacteria: A critical review. Life Sci. Med. Res. 2011, 21, 1–30. [Google Scholar]
  6. Danish, S.; Zafar-Ul-Hye, M.; Mohsin, F.; Hussain, M. ACC-deaminase producing plant growth promoting rhizobacteria and biochar mitigate adverse effects of drought stress on maize growth. PLoS ONE 2020, 15, e0230615. [Google Scholar] [CrossRef] [Green Version]
  7. Ali, M.A.; Naveed, M.; Mustafa, A.; Abbas, A. The good, the bad, and the ugly of rhizosphere microbiome. In Probiotics and Plant Health; Springer: Singapore, 2017; pp. 253–290. [Google Scholar]
  8. Ismail, M.A.; Amin, M.A.; Eid, A.M.; Hassan, S.E.D.; Mahgoub, H.A.; Lashin, I.; Abdelwahab, A.T.; Azab, E.; Gobouri, A.A.; Elkelish, A. Comparative Study between Exogenously Applied Plant Growth Hormones versus Metabolites of Microbial Endophytes as Plant Growth-Promoting for Phaseolus vulgaris L. Cells 2021, 10, 1059. [Google Scholar] [CrossRef]
  9. Khan, N.; Ali, S.; Shahid, M.A.; Mustafa, A.; Sayyed, R.Z.; Curá, J.A. Insights into the Interactions among Roots, Rhizosphere, and Rhizobacteria for Improving Plant Growth and Tolerance to Abiotic Stresses: A Review. Cells 2021, 10, 1551. [Google Scholar] [CrossRef]
  10. Mustafa, A.; Naveed, M.; Saeed, Q.; Ashraf, M.N.; Hussain, A.; Abbas, T.; Kamran, M.; Minggang, X. Application potentials of plant growth promoting rhizobacteria and fungi as an alternative to conventional weed control methods. In Sustainable Crop Production; IntechOpen: London, UK, 2019. [Google Scholar]
  11. Mustafa, A.; Naveed, M.; Abbas, T.; Saeed, Q.; Hussain, A.; Ashraf, M.N.; Minggang, X. Growth response of wheat and associated weeds to plant antagonistic rhizobacteria and fungi. Ital. J. Agron. 2019, 14, 191–198. [Google Scholar] [CrossRef]
  12. Abbas, T.; Zahir, Z.A.; Naveed, M.; Abbas, S.; Alwahibi, M.S.; Elshikh, M.S.; Mustafa, A. Large scale screening of rhizospheric allelopathic bacteria and their potential for the biocontrol of wheat-associated weeds. Agronomy 2020, 10, 1469. [Google Scholar] [CrossRef]
  13. Ray, P.; Lakshmanan, V.; Labbé, J.L.; Craven, K.D. Microbe to microbiome: A paradigm shift in the application of microorganisms for sustainable agriculture. Front. Microbiol. 2020, 11, 3323. [Google Scholar] [CrossRef]
  14. Ahmad, M.; Naseer, I.; Hussain, A.; Zahid Mumtaz, M.; Mustafa, A.; Hilger, T.H.; Ahmad Zahir, Z.; Minggang, X. Appraising endophyte–plant symbiosis for improved growth, nodulation, nitrogen fixation and abiotic stress tolerance: An experimental investigation with chickpea (Cicer arietinum L.). Agronomy 2019, 9, 621. [Google Scholar] [CrossRef] [Green Version]
  15. Nazli, F.; Mustafa, A.; Ahmad, M.; Hussain, A.; Jamil, M.; Wang, X.; Shakeel, Q.; Imtiaz, M.; El-Esawi, M.A. A Review on Practical Application and Potentials of Phytohormone-Producing Plant Growth-Promoting Rhizobacteria for Inducing Heavy Metal Tolerance in Crops. Sustainability 2020, 12, 9056. [Google Scholar] [CrossRef]
  16. Haider, F.U.; Ejaz, M.; Cheema, S.A.; Khan, M.I.; Zhao, B.; Cai, L.; Salim, M.A.; Naveed, M.; Khan, N.; Núñez-Delgado, A.; et al. Phytotoxicity of petroleum hydrocarbons: Sources, impacts and remediation strategies. Environ. Res. 2021, 197, 111031. [Google Scholar] [CrossRef]
  17. Goudaa, S.; Kerryb, R.G.; Dasc, G.; Paramithiotisd, S.; Shine, 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]
  18. Umar, W.; Ayub, M.A.; ur Rehman, M.Z.; Ahmad, H.R.; Farooqi, Z.U.R.; Shahzad, A.; Rehman, U.; Mustafa, A.; Nadeem, M. Nitrogen and phosphorus use efficiency in agroecosystems. In Resources Use Efficiency in Agriculture; Springer: Singapore, 2020; pp. 213–257. [Google Scholar]
  19. Pandey, P.; Irulappan, V.; Bagavathiannan, M.V.; Senthil-Kumar, M. Impact of combined abiotic and biotic stresses on plant growth and avenues for crop improvement by exploiting physio-morphological traits. Front. Plant Sci. 2017, 8, 537. [Google Scholar] [CrossRef] [Green Version]
  20. Raza, W.; Ling, N.; Yang, L.; Huang, Q.; Shen, Q. Response of tomato wilt pathogen Ralstonia solanacearum to the volatile organic compounds produced by a biocontrol strain Bacillus amyloliquefaciens SQR-9. Sci. Rep. 2016, 6, 24856. [Google Scholar] [CrossRef] [Green Version]
  21. Raza, W.; Yousaf, S.; Rajer, F.U. Plant growth promoting activity of volatile organic compounds produced by Bio-control strains. Sci. Lett. 2016, 4, 40–43. [Google Scholar]
  22. Bharti, N.; Pandey, S.S.; Barnawal, D.; Patel, V.K.; Kalra, A. Plant growth promoting rhizobacteria Dietzianatronolimnaea modulates the expression of stress responsive genes providing protection of wheat from salinity stress. Sci. Rep. 2016, 6, 34768. [Google Scholar] [CrossRef] [Green Version]
  23. Ramakrishna, W.; Rathore, P.; Kumari, R.; Yadav, R. Brown gold of marginal soil: Plant growth promoting bacteria to overcome plant abiotic stress for agriculture, biofuels and carbon sequestration. Sci. Total Environ. 2020, 711, 135062. [Google Scholar] [CrossRef]
  24. Olanrewaju, O.S.; Glick, B.R.; Babalola, O.O. Mechanisms of action of plant growth promoting bacteria. World J. Microbiol. Biotechnol. 2017, 33, 197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Khan, N.; Bano, A.; Rahman, M.A.; Guo, J.; Kang, Z.; Babar, M.A. Comparative physiological and metabolic analysis reveals a complex mechanism involved in drought tolerance in Chickpea (Cicer arietinum L.) induced by PGPR and PGRs. Sci. Rep. 2019, 9, 2097. [Google Scholar] [CrossRef]
  26. Gupta, A.; Gopal, M.; Thomas, G.V.; Manikandan, V.; Gajewski, J.; Thomas, G.; Seshagiri, S.; Schuster, S.C.; Rajesh, P.; Gupta, R. Whole genome sequencing and analysis of plant growth promoting bacteria isolated from the rhizosphere of plantation crops coconut, cocoa and arecanut. PLoS ONE 2014, 9, e104259. [Google Scholar] [CrossRef] [Green Version]
  27. 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; Springer: Singapore, 2019; pp. 281–311. [Google Scholar]
  28. Kumari, A.; Kumar, R. Exploring phyllosphere bacteria for growth promotion and yield of potato (Solanum tuberosum L.). Int. J. Curr. Microbiol. Appl. Sci. 2018, 7, 1065–1071. [Google Scholar] [CrossRef]
  29. Shivakumar, B. Biocontrol Potential and Plant Growth Promotional Activity of Fluorescent Pseudomonads of Western Ghats. Ph.D. Dissertation, UAS, Dharwad, India, 2007. [Google Scholar]
  30. Mohammed, A.F.; Oloyede, A.R.; Odeseye, A.O. Biological control of bacterial wilt of tomato caused by Ralstonia solanacearum using Pseudomonas species isolated from the rhizosphere of tomato plants. Arch. Phytopathol. Plant Prot. 2020, 53, 1–16. [Google Scholar] [CrossRef]
  31. Larkin, R.P. Biological control of soil borne diseases in organic potato production using hypovirulent strains of Rhizoctonia solani. Biol. Agric. Hortic. 2020, 36, 1–11. [Google Scholar] [CrossRef]
  32. Dixon, R.; Kahn, D. Genetic regulation of biological nitrogen fixation. Nat. Rev. Microbiol. 2004, 2, 621–631. [Google Scholar] [CrossRef]
  33. Kim, K.Y.D.; Jordan, D.; McDonald, G.A. Solubilization of hydroxyapatite. Enterobacter agglomerans and cloned Escherichia coli in culture medium. Biol. Fertil. Soils 1997, 24, 347–352. [Google Scholar] [CrossRef]
  34. Ladha, J.K.; de Bruijn, F.J.; Malik, K.A. Introduction: Assessing opportunities for nitrogen fixation in rice-a frontier project. Plant Soil 1997, 124, 1–10. [Google Scholar] [CrossRef]
  35. Raymond, J.; Siefert, J.L.; Staples, C.R.; Blankenship, R.E. The natural history of nitrogen fixation. Mol. Biol. Evol. 2004, 21, 541–554. [Google Scholar] [CrossRef] [Green Version]
  36. Bhattacharyya, P.N.; Jha, D.K. Plant growth-promoting rhizobacteria (PGPR): Emergence in agriculture. World J. Microbiol. Biotechnol. 2012, 28, 1327–1350. [Google Scholar] [CrossRef]
  37. Ahmad, M.; Khan, M.S. Effects of pesticides on plant growth promoting traits of Mesorhizobium strain MRC4. J. Saudi Soc. Agric. Sci. 2012, 11, 63–71. [Google Scholar]
  38. Zahran, H.H. Rhizobia from wild legumes: Diversity, taxonomy, ecology, nitrogen fixation and biotechnology. J. Biotechnol. 2001, 91, 143–153. [Google Scholar] [CrossRef]
  39. Giordano, W.; Hirsch, A.M. The expression of MaEXP1, a Melilotus alba expansin gene, is upregulated during the sweetclover-Sinorhizobium meliloti interaction. Mol. Plant Microbe Interact. 2004, 17, 613–622. [Google Scholar] [CrossRef] [Green Version]
  40. Marroquí, S.; Zorreguieta, A.; Santamaría, C. Enhanced symbiotic performance by Rhizobium tropici glycogen synthase mutants. J. Bacteriol. 2001, 183, 854–864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Glick, B.R. The enhancement of plant growth by free living bacteria. Can. J. Microbiol. 1995, 41, 109114. [Google Scholar] [CrossRef]
  42. Parray, J.A.; Jan, S.; Kamili, A.N.; Qadri, R.A.; Egamberdieva, D.; Ahmad, P. Current perspectives on plant growth-promoting rhizobacteria. J. Plant Growth Regul. 2016, 35, 877–902. [Google Scholar] [CrossRef]
  43. Ma, W.; Penrose, D.M.; Glick, B.R. Strategies used by rhizobia to lower plant ethylene levels and increase nodulation. Can. J. Microbiol. 2002, 48, 947–954. [Google Scholar] [CrossRef]
  44. Ma, W.; Charles, T.C.; Glick, B.R. Expression of an exogenous 1 aminocyclopropane-1-carboxylate deaminase gene in Sinorhizobium meliloti increases its ability to nodulate alfalfa. Appl. Environ. Microbiol. 2004, 100, 5891–5897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Dubeikovsky, A.N.; Mordukhova, E.A.; Kochetkov, V.V.; Polikarpova, F.Y.; Boronin, A.M. Growth promotion of blackcurrant softwood cuttings by recombinant strain Pseudomonas uorescens BSP53a synthesizing an increased amount of indole3-acetic acid. Soil Biol. Biochem. 1993, 25, 1277–1281. [Google Scholar] [CrossRef]
  46. Ke, X.; Feng, S.; Wang, J.; Lu, W.; Zhang, W.; Chen, M.; Lin, M. Effect of inoculation with nitrogen-fixing bacterium Pseudomonas stutzeri A1501 on maize plant growth and the microbiome indigenous to the rhizosphere. Syst. Appl. Microbiol. 2019, 42, 248–260. [Google Scholar] [CrossRef] [PubMed]
  47. Ramírez, M.; Valderrama, B.; Arredondo-Peter, R.; Soberón, M.; Mora, J.; Hernández, G. Rhizobium etli genetically engineered for the heterologous expression of Vitreoscilla sp. hemoglobin: Effects on free-living and symbiosis. Mol. Plant-Microbe Interact. 1999, 12, 1008–1015. [Google Scholar] [CrossRef] [Green Version]
  48. Liu, D.; Yan, R.; Fu, Y.; Wang, X.; Zhang, J.; Xiang, W. Antifungal, Plant Growth-Promoting, and Genomic Properties of an Endophytic Actinobacterium Streptomyces sp. NEAU-S7GS2. Front. Microbiol. 2019, 10, 2077. [Google Scholar] [CrossRef] [PubMed]
  49. Rima, F.S.; Biswas, S.; Sarker, P.K. Bacteria endemic to saline coastal belt and their ability to mitigate the effects of salt stress on rice growth and yields. Ann. Microbiol. 2018, 68, 525–535. [Google Scholar] [CrossRef]
  50. Tewari, S.; Arora, N.K. Fluorescent Pseudomonas sp. PF17 as an efficient plant growth regulator and biocontrol agent for sunflower crop under saline conditions. Symbiosis 2016, 68, 99–108. [Google Scholar] [CrossRef]
  51. Patten, C.L.; Glick, B.R. Role of Pseudomonas putida indoleacetic acid in development of the host plant root system. Appl. Environ. Microbiol. 2002, 68, 3795–3801. [Google Scholar] [CrossRef] [Green Version]
  52. Parmar, P.; Sindhu, S.S. Potassium solubilization by rhizosphere bacteria: Influence of nutritional and environmental conditions. J. Microbiol. Res. 2013, 3, 25–31. [Google Scholar]
  53. 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]
  54. Robin, A.; Mougel, C.; Siblot, S.; Vansuyt, G.; Mazurier, S.; Lemanceau, P. Effect of ferritin overexpression in tobacco on the structure of bacterial and pseudomonad communities associated with the roots. FEMS Microbiol. Ecol. 2006, 58, 492–502. [Google Scholar] [CrossRef]
  55. Vansuyt, G.; Robin, A.; Briat, J.F.; Curie, C.; Lemanceau, P. Iron acquisition from Fe-pyoverdine by Arabidopsis thaliana. Mol. Plant-Microbe Interact. 2007, 20, 441–447. [Google Scholar] [CrossRef] [Green Version]
  56. Mounir, A.M.; Osman, Y.M.M.; Khalil, O.A. Impact of potassium solubilizing bacteria on growth and yield of garlic. Plant Arch. 2020, 20, 8374–8388. [Google Scholar]
  57. Elkhatib, H.A.; Saed, M.G.; Ramadan, A.M.; Ahmed, F.M. Impact of potassium fertilizer rates and potassium solubilizing bacteria inoculation on the growth, yield and quality of potato (Solanum tuberosum L.). Alex. Sci. Exch. J. 2019, 40, 127–141. [Google Scholar]
  58. Liu, D.; Lian, B.; Dong, H. Isolation of Paenibacillus sp. and assessment of its potential for enhancing mineral weathering. Geomicrobiol. J. 2012, 29, 413–421. [Google Scholar] [CrossRef]
  59. Tewari, S.; Arora, N.K. Role of salicylic acid from Pseudomonas aeruginosa PF23EPS+ in growth promotion of sunflower in saline soils infested with phytopathogen Macrophomina phaseolina. Environ. Sustain. 2018, 1, 49–59. [Google Scholar] [CrossRef]
  60. Egamberdieva, D.; Wirth, S.; Jabborova, D.; Räsänen, L.A.; Liao, H. Coordination between Bradyrhizobium and Pseudomonas alleviates salt stress in soybean through altering root system architecture. J. Plant Interact. 2017, 12, 100–107. [Google Scholar] [CrossRef] [Green Version]
  61. Asari, S.; Danuse, T.; Jakub, R.; Ondrej, N.; David, V.P.; Sarosh, B.; Johan, M. Analysis of plant growth-promoting properties of Bacillus amyloliquefaciens UCMB5113 using Arabidopsis thaliana as host plant. Planta 2017, 245, 15–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. El-Esawi, M.A.; Alaraidh, I.A.; Alsahli, A.A.; Alzahrani, S.A.; Ali, H.M.; Alayafi, A.A. Serratia liquefaciens KM4 improves salt stress tolerance in maize by regulating redox potential, ion homeostasis, leaf gas exchange and stress-related gene expression. Int. J. Mol. Sci. 2018, 19, 3310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. López-Reyes, L.; Carcaño-Montiel, M.G.; Lilia, T.L.; Medina-de la Rosa, G.; Armando, T.H.R. Antifungal and growth-promoting activity of Azospirillumbrasilense in Zea mays L. ssp. mexicana. Arch. Phytopathol. Plant Prot. 2017, 50, 727–743. [Google Scholar] [CrossRef]
  64. Cavite, H.J.M.; Mactal, A.G.; Evangelista, E.V. Growth and Yield Response of Upland Rice to Application of Plant Growth-Promoting Rhizobacteria. J. Plant Growth Regul. 2021, 40, 494–508. [Google Scholar] [CrossRef]
  65. Sheng, M.; Jia, H.; Zhang, G.; Zeng, L.; Zhang, L.; Long, Y.; Lan, J.; Hu, Z.; Zeng, Z.; Wang, B.; et al. Siderophore Production by Rhizosphere Biological Control Bacteria Brevibacillus brevis GZDF3 of Pinellia ternata and Its Antifungal Effects on Candida albicans. J. Microbiol. Biotechnol. 2020, 30, 689–699. [Google Scholar] [CrossRef]
  66. 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. 2021, 203, 1195–1209. [Google Scholar] [CrossRef]
  67. Billah, M.; Khan, M.; Bano, A.; Hassan, T.U.; Munir, A.; Gurmani, A.R. Phosphorus and phosphate solubilizing bacteria: Keys for sustainable agriculture. Geomicrobiol. J. 2019, 36, 904–916. [Google Scholar] [CrossRef]
  68. Sharma, S.B.; Sayyed, R.Z.; Trivedi, M.H.; Gobi, T.A. Phosphate solubilizing microbes: Sustainable approach for managing phosphorus deficiency in agricultural soils. SpringerPlus 2013, 2, 587. [Google Scholar] [CrossRef] [Green Version]
  69. Alori, E.T.; Glick, B.R.; Babalola, O.O. Microbial Phosphorus Solubilization and Its Potential for Use in Sustainable Agriculture. Front. Microbiol. 2017, 8, 971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Zhu, J.; Li, M.; Whelan, M. Phosphorus activators contribute to legacy phosphorus availability in agricultural soils: A review. Sci. Total Environ. 2018, 612, 522–537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Adnan, M.; Fahad, S.; Khan, I.A.; Saeed, M.; Ihsan, M.Z.; Saud, S.; Riaz, M.; Wang, D.; Wu, C. Integration of poultry manure and phosphate solubilizing bacteria improved availability of Ca bound P in calcareous soils. 3 Biotech 2019, 9, 1–10. [Google Scholar] [CrossRef]
  72. Alaylar, B.; Egamberdieva, D.; Gulluce, M.; Karadayi, M.; Arora, N.K. Integration of molecular tools in microbial phosphate solubilization research in agriculture perspective. World J. Microbiol. Biotechnol. 2020, 36, 1–12. [Google Scholar] [CrossRef]
  73. Pereira, S.I.; Castro, P.M. Phosphate-solubilizing rhizobacteria enhance Zea mays growth in agricultural P-deficient soils. Ecol. Eng. 2014, 73, 526–535. [Google Scholar] [CrossRef]
  74. FAO. World Fertilizer Trends and Outlook to 2020; Food and Agriculture Organization of the United Nations: Rome, Italy, 2017. [Google Scholar]
  75. Dash, S.; Borah, S.S.; Kalamdhad, A.S. Study of the limnology of wetlands through a one-dimensional model for assessing the eutrophication levels induced by various pollution sources. Ecol. Model. 2020, 416, 108907. [Google Scholar] [CrossRef]
  76. Younas, F.; Mustafa, A.; Farooqi, Z.U.R.; Wang, X.; Younas, S.; Mohy-Ud-Din, W.; Ashir Hameed, M.; Mohsin Abrar, M.; Maitlo, A.A.; Noreen, S.; et al. Current and Emerging Adsorbent Technologies for Wastewater Treatment: Trends, Limitations, and Environmental Implications. Water 2021, 13, 215. [Google Scholar] [CrossRef]
  77. Loper, J.E.; Schroth, M.N. Influence of bacterial sources of indole-3-acetic acid on root elongation of sugar beet. Phytopathology 1986, 76, 386–389. [Google Scholar] [CrossRef]
  78. Muleta, A.; Tesfaye, K.; Selassie, T.H.; Cook, D.R.; Assefa, F. Phosphate solubilization and multiple plant growth promoting properties of Mesorhizobium species nodulating chickpea from acidic soils of Ethiopia. Arch. Microbiol. 2021, 203, 2129–2137. [Google Scholar] [CrossRef]
  79. Kafle, A.; Cope, K.; Raths, R.; Krishna, Y.J.; Subramanian, S.; Bucking, H.; Garcia, K. Harnessing Soil Microbes to Improve Plant Phosphate Efficiency in Cropping Systems. Agronomy 2019, 9, 127. [Google Scholar] [CrossRef] [Green Version]
  80. Singh, D.; Prasanna, R. Potential of microbes in the biofortification of Zn and Fe in dietary food grains. A review. Agron. Sustain. Dev. 2020, 40, 1–21. [Google Scholar] [CrossRef]
  81. Kalayu, G. Phosphate Solubilizing Microorganisms: Promising Approach as Biofertilizers. Int. J. Agron. 2019, 2019, 4917256. [Google Scholar] [CrossRef]
  82. Satyaprakash, M.; Nikitha, T.; Reddi, E.U.B.; Sadhana, B.; Vani, S.S. Phosphorous and phosphate solubilising bacteria and their role in plant nutrition. Int. J. Curr. Microbiol. Appl. Sci. 2017, 6, 2133–2144. [Google Scholar]
  83. Patel, S.T.; Minocheherhomji, F.P. Review: Plant Growth Promoting Rhizobacteria: Blessing to Agriculture. Int. J. Pure Appl. Biosci. 2018, 6, 481–492. [Google Scholar] [CrossRef]
  84. Pradhan, A.; Pahari, A.; Mohapatra, S.; Mishra, B.B. Phosphate-Solubilizing Microorganisms in Sustainable Agriculture: Genetic Mechanism and Application. In Advances in Soil Microbiology: Recent Trends and Future Prospects; Springer Nature Singapore Pte Ltd.: Singapore, 2017; pp. 81–97. [Google Scholar] [CrossRef]
  85. Castagno, L.N.; Sannazzaro, A.I.; Gonzalez, M.E.; Pieckenstain, F.L.; Estrella, M.J. Phosphobacteria as key actors to overcome phosphorus deficiency in plants. Ann. Appl. Biol. 2021, 178, 256–267. [Google Scholar] [CrossRef]
  86. Walpola, B.C.; Yoon, M.H. Isolation and characterization of phosphate solubilizing bacteria and their co-inoculation efficiency on tomato plant growth and phosphorous uptake. Afr. J. Microbiol. Res. 2013, 7, 266–275. [Google Scholar]
  87. Kumar, A.; Singh, V.K.; Tripathi, V.; Singh, P.P.; Singh, A.K. Plant growth-promoting rhizobacteria (PGPR): Perspective in agriculture under biotic and abiotic stress. In Crop Improvement through Microbial Biotechnology; Elsevier: Amsterdam, The Netherlands, 2018; pp. 333–342. [Google Scholar]
  88. Nahas, E. Factors determining rock phosphate solubilization by microorganism isolated from soil. World J. Microbiol. Biotechnol. 1996, 12, 18–23. [Google Scholar] [CrossRef] [PubMed]
  89. Hinsinger, P. Bioavailability of soil inorganic P in the rhizosphere as affected by rootinduced chemical changes: A review. Plant Soil 2001, 237, 173–195. [Google Scholar] [CrossRef]
  90. Kim, J.; Rees, D.C. Nitrogenase and biological nitrogen fixation. Biochemistry 1994, 33, 389–397. [Google Scholar] [CrossRef]
  91. Zulfiqar, U.; Farooq, M.; Hussain, S.; Maqsood, M.; Hussain, M.; Ishfaq, M.; Ahmad, M.; Anjum, M.Z. Lead toxicity in plants: Impacts and remediation. J. Environ. Manag. 2019, 250, 109557. [Google Scholar] [CrossRef]
  92. Wang, Y.; Wu, W.H. Regulation of potassium transport and signaling in plants. Curr. Opin. Plant Biol. 2017, 39, 123–128. [Google Scholar] [CrossRef]
  93. Adams, E.; Shin, R. Transport, signaling, and homeostasis of potassium and sodium in plants. J. Integr. Plant Biol. 2014, 56, 231–249. [Google Scholar] [CrossRef]
  94. Zahoor, R.; Zhao, W.; Dong, H.; Snider, J.L.; Abid, M.; Iqbal, B.; Zhou, Z. Potassium improves photosynthetic tolerance to and recovery from episodic drought stress in functional leaves of cotton (Gossypium hirsutum L.). Plant Physiol. Biochem. 2017, 119, 21–32. [Google Scholar] [CrossRef]
  95. Filho, A.B.C.; Dutra, A.F.; Da Silva, G.S. Phosphate and potassium fertilization for radish grown in a latosol with a high content of these nutrients. Rev. Caatinga 2017, 30, 412–419. [Google Scholar] [CrossRef]
  96. Sparks, D.L.; Huang, P.M. Physical chemistry of soil potassium. In Potassium in Agriculture; Soil Science Society of America: Madison, WI, USA, 1985; pp. 201–276. [Google Scholar]
  97. Radulov, I.; Berbecea, A.; Imbrea, F.; Lato, A.; Crista, F.; Mergheş, P. Potassium in soil-plant-human system. Res. J. Agric. Sci. 2014, 46, 47–52. [Google Scholar]
  98. Sindhu, S.S.; Parmar, P.; Phour, M. Nutrient cycling: Potassium solubilization by microorganisms and improvement of crop growth. In Geomicrobiology and Biogeochemistry; Springer: Berlin/Heidelberg, Germany, 2014; pp. 175–198. [Google Scholar]
  99. Saiyad, S.A.; Jhala, Y.K.; Vyas, R.V. Comparative efficiency of five potash and phosphate solubilizing bacteria and their key enzymes useful for enhancing and improvement of soil fertility. Int. J. Sci. Res. 2015, 5, 1–6. [Google Scholar]
  100. Archana, D.S.; Nandish, M.S.; Savalagi, V.P.; Alagawadi, A.R. Characterization of potassium solubilizing bacteria (KSB) from rhizosphere soil. Bioinfolet-A Q. J. Life Sci. 2013, 10, 248–257. [Google Scholar]
  101. Rajawat, M.V.S.; Singh, S.; Tyagi, S.P.; Saxena, A.K. A modified plate assay for rapid screening of potassium-solubilizing bacteria. Pedosphere 2016, 26, 768–773. [Google Scholar] [CrossRef]
  102. Hu, X.; Chen, J.; Guo, J. Two phosphate-and potassium-solubilizing bacteria isolated from Tianmu Mountain, Zhejiang, China. World J. Microbiol. Biotechnol. 2006, 22, 983–990. [Google Scholar] [CrossRef]
  103. Sheng, X.F. Growth promotion and increased potassium uptake of cotton and rape by a potassium releasing strain of Bacillus edaphicus. Soil Biol. Biochem. 2005, 37, 1918–1922. [Google Scholar] [CrossRef]
  104. Prajapati, K.; Sharma, M.C.; Modi, H.A. Growth promoting effect of potassium solubilizing microorganisms on okra (Abelmoscus Esculantus). Int. J. Agri. Sci. Res. 2013, 1, 181–188. [Google Scholar]
  105. Huang, Z.; He, L.; Sheng, X.; He, Z. Weathering of potash feldspar by Bacillus sp. L11. Wei sheng wuxue bao. Acta Microbiol. Sin. 2013, 53, 1172–1178. [Google Scholar]
  106. Meena, V.S.; Maurya, B.R.; Verma, J.P. Does a rhizospheric microorganism enhance K+ availability in agricultural soils? Microbiol. Res. 2014, 169, 337–347. [Google Scholar] [CrossRef]
  107. Abou-el-Seoud, I.; Abdel-Megeed, A. Impact of rock materials and biofertilizations on P and K availability for maize (Zea Maize) under calcareous soil conditions. Saudi J. Biol. Sci. 2012, 19, 55–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Shelobolina, E.; Xu, H.; Konishi, H.; Kukkadapu, R.; Wu, T.; Blöthe, M.E. Microbial lithotrophic oxidation of structural Fe (II) in biotite. Appl. Environ. Microbiol. 2012, 78, 5746–5752. [Google Scholar] [CrossRef] [Green Version]
  109. Sheng, X.F.; Gong, J.X. Increased degradation of phenanthrene in soil by Pseudomonas sp. GF3 in the presence of wheat. Soil Biol. Biochem. 2006, 38, 2587–2592. [Google Scholar] [CrossRef]
  110. Padhan, D.; Sen, A.; Adhikary, S.; Kundu, R.; Yadav, V.K. Chapter-3 Potassium Solubilisation in Soils: Mechanisms, Effect on Plant Growth and Future Prospects. Curr. Res. Soil Fertil. 2019, 19, 37. [Google Scholar]
  111. Spaepen, S.; Vanderleyden, J. Auxin and plant-microbe interactions. Cold Spring Harb. Perspect. Biol. 2011, 3, a001438. [Google Scholar] [CrossRef] [Green Version]
  112. Wang, J.G.; Zhang, F.S.; Zhang, X.L.; Cao, Y.P. Release of potassium from K-bearing minerals: Effect of plant roots under P deficiency. Nutr. Cycl. Agroecosyst. 2000, 56, 45–52. [Google Scholar] [CrossRef]
  113. Etesami, H.; Alikhani, H.A.; Hosseini, H.M. Indole-3-acetic acid and 1-aminocyclopropane-1-carboxylate deaminase: Bacterial traits required in rhizosphere, rhizoplane and/or endophytic competence by beneficial bacteria. In Bacterial Metabolites in Sustainable Agroecosystem; Springer: Berlin/Heidelberg, Germany, 2015; pp. 183–258. [Google Scholar]
  114. 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]
  115. Hider, R.C.; Kong, X. Chemistry and biology of siderophores. Nat. Prod. Rep. 2010, 27, 637–657. [Google Scholar] [CrossRef] [PubMed]
  116. Rasouli-Sadaghiani, M.; Malakouti, M.J.; Khavazi, K.; Miransari, M. Siderophore Efficacy of Fluorescent Pseudomonades Affecting Labeled Iron (59Fe) Uptake by Wheat (Triticumaestivum L.) Genotypes Differing in Fe Efficiency. In Use of Microbes for the Alleviation of Soil Stresses; Springer: New York, NY, USA, 2014; pp. 121–132. [Google Scholar]
  117. Subramanium, N.; Sundaram, L. Siderophore producing Pseudomonas spp. isolated from rhizospheric soil and enhancing iron content in Arachis hypogaea L. plant. J. Agric. Technol. 2020, 16, 429–442. [Google Scholar]
  118. Reed, S.C.; Yang, X.; Thornton, P.E. Incorporating phosphorus cycling into global modeling efforts: A worthwhile, tractable endeavor. New Phytol. 2015, 208, 324–329. [Google Scholar] [CrossRef] [PubMed]
  119. Hakim, S.; Naqqash, T.; Nawaz, M.S.; Laraib, I.; Siddique, M.J.; Zia, R.; Mirza, M.S.; Imran, A. Rhizosphere Engineering with Plant Growth-Promoting Microorganisms for Agriculture and Ecological Sustainability. Front. Sustain. Food Syst. 2021, 5, 617157. [Google Scholar] [CrossRef]
  120. Grobelak, A.; Hiller, J. Bacterial siderophores promote plant growth: Screening of catechol and hydroxamate siderophores. Int. J. Phytorem. 2017, 19, 825–833. [Google Scholar] [CrossRef]
  121. Sharma, A.; Johri, B.; Sharma, A.; Glick, B. Plant growth-promoting bacterium Pseudomonas spp. strain GRP3 influences iron acquisition in mung bean (Vigna radiata L. Wilzeck). Soil Biol. Biochem. 2003, 35, 887–894. [Google Scholar] [CrossRef]
  122. Braud, A.; Jézéquel, K.; Léger, M.A. Lebeau T, Siderophore production by using free and immobilized cells of two pseudomonads cultivated in a medium enriched with Fe and/or toxic metals (Cr, Hg, Pb). Biotechnol. Bioeng. 2006, 94, 1080–1088. [Google Scholar] [CrossRef] [PubMed]
  123. Damam, M.; Kaloori, K.; Gaddam, B.; Kausar, R. Plant growth promoting substances (phytohormones) produced by rhizobacterial strains isolated from the rhizosphere of medicinal plants. Int. J. Pharm. Sci. Rev. 2016, 37, 130–136. [Google Scholar]
  124. Sureshbabu, K.; Amaresan, N.; Kumar, K. Amazing multiple function properties of plant growth promoting rhizobacteria in the rhizosphere soil. Int. J. Curr. Microbiol. Appl. Sci. 2016, 5, 661–683. [Google Scholar] [CrossRef]
  125. Spence, C.; Bais, H. Role of plant growth regulators as chemical signals in plant–microbe interactions: A double edged sword, Curr. Opin. Plant Biol. 2015, 27, 52–58. [Google Scholar] [CrossRef]
  126. Kumar, K.V.; Srivastava, S.; Singh, N.; Behl, H. Role of metal resistant plant growth promoting bacteria in ameliorating fly ash to the growth of Brassica juncea. J. Hazard. Mater. 2009, 170, 51–57. [Google Scholar] [CrossRef]
  127. Ahemad, M.; Kibret, M. Mechanisms and applications of plant growth promoting rhizobacteria: Current perspective. J. King Saud Univ. Sci. 2014, 26, 1–20. [Google Scholar] [CrossRef] [Green Version]
  128. Afzal, I.; Shinwari, Z.K.; Iqrar, I. Selective isolation and characterization of agriculturally beneficial endophytic bacteria from wild hemp using canola. Pak. J. Bot. 2015, 47, 1999–2008. [Google Scholar]
  129. Spaepen, S.; Bossuyt, S.; Engelen, K.; Marchal, K.; Vanderleyden, J. Phenotypical and molecular responses of A rabidopsis thaliana roots as a result of inoculation with the auxin-producing bacterium A zospirillumbrasilense. New Phytol. 2014, 201, 850–861. [Google Scholar] [CrossRef]
  130. Llorente, B.E.; Alasia, M.A.; Larraburu, E.E. Biofertilization with Azospirillumbrasilense improves in vitro culture of Handroanthusochraceus, a forestry, ornamental and medicinal plant. New Biotechnol. 2016, 33, 32–40. [Google Scholar] [CrossRef]
  131. Ruzzi, M.; Aroca, R. Plant growth-promoting rhizobacteria act as biostimulants in horticulture. Sci. Hortic. 2015, 196, 124–134. [Google Scholar] [CrossRef]
  132. Kumar, A.; Vandana, R.S.; Singh, M.; Pandey, K.D. Plant growth promoting rhizobacteria (PGPR). A promising approach for disease management. In Microbes and Environmental Management; Singh, J.S., Singh, D.P., Eds.; Studium Press: New Delhi, India, 2015; pp. 195–209. [Google Scholar]
  133. Jha, C.K.; Saraf, M. Plant growth promoting rhizobacteria (PGPR): A review. E3 J. Agric. Res. Dev. 2015, 5, 108–119. [Google Scholar]
  134. Vacheron, J.; Desbrosses, G.; Bouffaud, M.L.; Touraine, B.; Moenne-Loccoz, Y.; Muller, D. Plant growth-promoting rhizobacteria and root system functioning. Front. Plant Sci. 2013, 4, 356. [Google Scholar] [CrossRef] [Green Version]
  135. Vejan, P.; Abdullah, R.; Khadiran, T.; Ismail, S.; Nasrulhaq Boyce, A. Role of plant growth promoting rhizobacteria in agricultural sustainability–A review. Molecules 2016, 21, 573. [Google Scholar] [CrossRef]
  136. Nadeem, S.M.; Ahmad, M.; Zahir, Z.A.; Javaid, A.; Ashraf, M. The role of mycorrhizae and plant growth promoting rhizobacteria (PGPR) in improving crop productivity under stressful environments. Biotechnol. Adv. 2014, 32, 429–448. [Google Scholar] [CrossRef]
  137. 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]
  138. Pérez-Montaño, F.; Alías-Villegas, C.; Bellogín, R.; Del Cerro, P.; Espuny, M.; Jiménez-Guerrero, I. Plant growth promotion in cereal and leguminous agricultural important plants: From microorganism capacities to crop production. Microbiol. Res. 2014, 169, 325–336. [Google Scholar] [CrossRef] [Green Version]
  139. Heydarian, Z.; Yu, M.; Gruber, M.; Glick, B.R.; Zhou, R.; Hegedus, D.D. Inoculation of soil with plant growth promoting bacteria producing 1-aminocyclopropane-1-carboxylate deaminase or expression of the corresponding acds gene in transgenic plants increases salinity tolerance in camelina sativa. Front. Microbiol. 2016, 7, 1966. [Google Scholar] [CrossRef]
  140. Lugtenberg, B.; Kamilova, F. Plant-growth-promoting rhizobacteria. Annu. Rev. Microbiol. 2009, 63, 541–556. [Google Scholar] [CrossRef] [Green Version]
  141. Pieterse, C.M.; Zamioudis, C.; Berendsen, R.L.; Weller, D.M.; Van Wees, S.C.; Bakker, P.A. Induced systemic resistance by beneficial microbes. Annu. Rev. Phytopathol. 2014, 52, 347–375. [Google Scholar] [CrossRef] [Green Version]
  142. Zhang, R.; Vivanco, J.M.; Shen, Q. The unseen rhizosphere root–soil–microbe interactions for crop production. Curr. Opin. Microbiol. 2017, 37, 8–14. [Google Scholar] [CrossRef]
  143. Von Wiren, N.; Lauter, F.R.; Ninnemann, O.; Gillissen, B.; Walch-Liu, P.; Engels, C. Differential regulation of three functional ammonium transporter genes by nitrogen in root hairs and by light in leaves of tomato. Plant J. 2000, 21, 167–175. [Google Scholar] [CrossRef]
  144. Ahn, S.J.; Shin, R.; Schachtman, D.P. Expression of KT/KUP genes in Arabidopsis and the role of root hairs in K+ uptake. Plant Physiol. 2004, 134, 1135–1145. [Google Scholar] [CrossRef] [Green Version]
  145. Desbrosses, G.J.; Stougaard, J. Root nodulation: A paradigm for how plant-microbe symbiosis influences host developmental pathways. Cell Host Microbe 2011, 10, 348–358. [Google Scholar] [CrossRef] [Green Version]
  146. Traxler, M.F.; Kolter, R. Natural products in soil microbe interactions and evolution. Nat. Prod. Rep. 2015, 32, 956–970. [Google Scholar] [CrossRef]
  147. Combes-Meynet, E.; Pothier, J.F.; Moënne-Loccoz, Y.; Prigent-Combaret, C. The Pseudomonas secondary metabolite 2,4-diacetylphloroglucinol is a signal inducing rhizoplane expression of Azospirillum genes involved in plant-growth promotion. Mol. Plant-Microbe Interact. 2011, 24, 271–284. [Google Scholar] [CrossRef] [Green Version]
  148. Hassan, M. The Role of Pectin Utilization in Root Colonization and Plant Growth-Promotion by Bacillus amyloliquefaciens subsp. plantarum (Bap). Master’s Thesis, Auburn University, Auburn, ME, USA, 2016. [Google Scholar]
  149. Ramaekers, L.; Remans, R.; Rao, I.M.; Blair, M.W.; Vanderleyden, J. Strategies for improving phosphorus acquisition efficiency of crop plants. Field Crops Res. 2010, 117, 169–176. [Google Scholar] [CrossRef]
  150. Lakshmanan, V.; Castaneda, R.; Rudrappa, T.; Bais, H.P. Root transcriptome analysis of Arabidopsis thaliana exposed to beneficial Bacillus subtilis FB17 rhizobacteria revealed genes for bacterial recruitment and plant defense independent of malate. Planta 2013, 238, 657–668. [Google Scholar] [CrossRef]
  151. Hiltner, L. About recent experiences and problems the field of soil bacteriology with special consideration of green manure and fallow. Arbeiten der Deutschen Landwirtschaftlichen Gesellschaft 1904, 98, 59–78. [Google Scholar]
  152. Guo, R.F.; Yuan, G.F.; Wang, Q.M. Effect of NaCl treatments on glucosinolate metabolism in broccoli sprouts. J. Zhejiang Univ. Sci. B 2013, 14, 124. [Google Scholar] [CrossRef]
  153. He, D.C.; Zhan, J.S.; Xie, L.H. Problems, challenges and future of plant disease management: From an ecological point of view. J. Integr. Agric. 2016, 15, 705–715. [Google Scholar] [CrossRef]
  154. Compant, S.; Reiter, B.; Sessitsch, A.; Nowak, J.; Clément, C.; Barka, E.A. Endophytic colonization of Vitis vinifera L. by plant growth-promoting bacterium Burkholderia sp. strain PsJN. Appl. Environ. Microbiol. 2005, 71, 1685–1693. [Google Scholar] [CrossRef] [Green Version]
  155. Glick, B.R.; Cheng, Z.; Czarny, J.; Duan, J. Promotion of plant growth by ACC deaminase-containing soil bacteria. Eur. J. Plant Pathol. 2007, 119, 329–339. [Google Scholar] [CrossRef]
  156. Dobereiner, J. Review article History and New Perspectives of Diazotrophs in Association with Non-Leguminous Plants. Symbiosis 1992, 13, 1–13. [Google Scholar]
  157. Podile, A.R.; Kishore, G.K. Plant growth-promoting rhizobacteria. In Plant-Associated Bacteria; Springer: Dordrecht, The Netherlands, 2007; pp. 195–230. [Google Scholar]
  158. Gray, E.J.; Smith, D.L. Intracellular and extracellular PGPR: Commonalities and distinctions in the plant–bacterium signaling processes. Soil Biol. Biochem. 2005, 37, 395–412. [Google Scholar] [CrossRef]
  159. Cook, R.J.; Weller, D.M.; El-Banna, A.Y.; Vakoch, D.; Zhang, H. Yield responses of direct-seeded wheat to rhizobacteria and fungicide seed treatments. Plant Dis. 2002, 86, 780–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Amein, T.; Omer, Z.; Welch, C. Application and evaluation of Pseudomonas strains for biocontrol of wheat seedling blight. Crop Prot. 2008, 27, 532–536. [Google Scholar] [CrossRef]
  161. Ryder, M.H.; Yan, Z.; Terrace, T.E.; Rovira, A.D.; Tang, W.; Correll, R.L. Use of strains of Bacillus isolated in China to suppress take-all and rhizoctonia root rot, and promote seedling growth of glasshouse-grown wheat in Australian soils. Soil Biol. Biochem. 1999, 31, 19–29. [Google Scholar] [CrossRef]
  162. Cavaglieri, L.; Orlando, J.R.M.I.; Rodríguez, M.I.; Chulze, S.; Etcheverry, M. Biocontrol of Bacillus subtilis against Fusarium verticillioides in vitro and at the maize root level. Res. Microbiol. 2005, 156, 748–754. [Google Scholar] [CrossRef]
  163. Moustafa, H.E.; Abo-Zaid, G.A.; Abd-Elsalam, H.E.; Hafez, E.E. Antagonistic and inhibitory effect of Bacillus subtilis against certain plant pathogenic fungi, I. Biotechnology 2009, 8, 53–61. [Google Scholar]
  164. Das, A.J.; Lal, S.; Kumar, R.; Verma, C. Bacterial biosurfactants can be an ecofriendly and advanced technology for remediation of heavy metals and co-contaminated soil. Int. J. Environ. Sci. Technol. 2016, 14, 1343–1354. [Google Scholar] [CrossRef]
  165. Khan, M.R.; Fischer, S.; Egan, D.; Doohan, F.M. Biological control of Fusarium seedling blight disease of wheat and barley. Phytopathology 2006, 96, 386–394. [Google Scholar] [CrossRef] [Green Version]
  166. Buttimer, C.; McAuliffe, O.; Ross, R.P.; Hill, C.; O’Mahony, J.; Coffey, A. Bacteriophages and bacterial plant diseases. Front. Microbiol. 2017, 8, 34. [Google Scholar] [CrossRef] [Green Version]
  167. Singh, V.K.; Singh, A.K.; Kumar, A. Disease management of tomato through PGPB: Current trends and future perspective. 3 Biotech 2017, 7, 255. [Google Scholar] [CrossRef] [PubMed]
  168. Singh, M.; Kumar, A.; Singh, R.; Pandey, K.D. Endophytic bacteria: A new source of bioactive compounds. 3 Biotech 2017, 7, 315. [Google Scholar] [CrossRef] [PubMed]
  169. Kumar, A.; Verma, H.; Singh, V.K.; Singh, P.P.; Singh, S.K.; Ansari, W.A.; Yadav, A.; Singh, P.K.; Pandey, K.D. Role of Pseudomonas sp. in sustainable agriculture and disease management. In Agriculturally Important Microbes for Sustainable Agriculture; Meena, V., Mishra, P., Bisht, J., Pattanayak, A., Eds.; Springer: Singapore, 2017; pp. 195–215. [Google Scholar]
  170. Kumar, A.; Singh, A.K.; Kaushik, M.S.; Mishra, S.K.; Raj, P.; Singh, P.K.; Pandey, K.D. Interaction of turmeric (Curcuma longa L.) with beneficial microbes: A review. 3 Biotech 2017, 7, 1–8. [Google Scholar] [CrossRef] [PubMed]
  171. Singh, R.; Pandey, D.K.; Kumar, A.; Singh, M. PGPR isolates from the rhizosphere of vegetable crop Momordica charantia: Characterization and application as biofertilizer. Int. J. Curr. Microbiol. Appl. Sci. 2017, 6, 1789–1802. [Google Scholar]
  172. 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] [PubMed]
  173. Dowling, D.N.; O’Gara, F. Metabolites of Pseudomonas involved in the biocontrol of plant disease. Trends Biotechnol. 1994, 12, 133–141. [Google Scholar] [CrossRef]
  174. Whipps, J.M. Microbial interactions and biocontrol in the rhizosphere. J. Exp. Bot. 2001, 52, 487–511. [Google Scholar] [CrossRef]
  175. Duffy, B. Pathogen self-defense: Mechanisms to counteract microbial antagonism. Annu. Rev. Phytopathol. 2003, 41, 501–538. [Google Scholar] [CrossRef] [Green Version]
  176. Haas, D.; Défago, G. Biological control of soil-borne pathogens by fluorescent pseudomonads. Nat. Rev. Microbiol. 2005, 3, 307–319. [Google Scholar] [CrossRef]
  177. de Bruijn, I.; de Kock, M.J.D.; Yang, M.; de Waard, P.; van Beek, T.A.; Raaijmakers, J.M. Genome-based discovery, structure prediction and functional analysis of cyclic lipopeptide antibiotics in Pseudomonas species. Mol. Microbiol. 2007, 63, 417–428. [Google Scholar] [CrossRef]
  178. Raaijmakers, J.M.; de Bruijn, I.; Nybroe, O.; Ongena, M. Natural functions of lipopeptides from Bacillus and Pseudomonas: More than surfactants and antibiotics. FEMS Microbiol. Rev. 2010, 34, 1037–1062. [Google Scholar] [CrossRef] [Green Version]
  179. Abbas, T.; Zahir, Z.A.; Naveed, M.; Kremer, R.J. Limitations of existing weed control practices necessitate development of alternative techniques based on biological approaches. Adv. Agron. 2018, 147, 239–280. [Google Scholar]
  180. Maksimov, I.V.; Abizgil’dina, R.R.; Pusenkova, L.I. Plant growth promoting rhizobacteria as alternative to chemical crop protectors from pathogens (Review). Appl. Biochem. Microbiol. 2011, 47, 333–345. [Google Scholar] [CrossRef]
  181. Hill, D.S.; Stein, J.I.; Torkewitz, N.R.; Morse, A.M.; Howell, C.R.; Pachlatko, J.P.; Becker, J.O.; Ligon, J.M. Cloning of genes involved in the synthesis of pyrrolnitrin from Pseudomonas fluorescens and role of pyrrolnitrin synthesis in biological control of plant disease. Appl. Environ. Microbiol. 1994, 60, 78–85. [Google Scholar] [CrossRef] [Green Version]
  182. Silo-Suh, L.A.; Lethbridge, B.J.; Raffel, S.J.; He, H.; Clardy, J.; Handelsman, J. Biological activities of two fungistatic antibiotics produced by Bacillus cereus UW85. Appl. Environ. Microbiol. 1994, 60, 2023–2030. [Google Scholar] [CrossRef] [Green Version]
  183. He, H.; Silo-Suh, L.A.; Handelsman, J.; Clardy, J. Zwittermicin A, an antifungal and plant protection agent from Bacillus cereus. Tetrahedron Lett. 1994, 35, 2499–2502. [Google Scholar] [CrossRef]
  184. Emmert, E.A.; Handelsman, J. Biocontrol of plant disease: A (Gram-) positive perspective. FEMS Microbiol. Lett. 1999, 171, 1–9. [Google Scholar] [CrossRef] [PubMed]
  185. Kokalis-Burelle, N.; Kloepper, J.W.; Reddy, M.S. Plant growth-promoting rhizobacteria as transplant amendments and their effects on indigenous rhizosphere microorganisms. Appl. Soil Ecol. 2005, 31, 91–100. [Google Scholar] [CrossRef]
  186. Van Loon, L.C.; Bakker, P.A.H.M.; Pieterse, C.M.J. Systemic resistance induced by rhizosphere bacteria. Annu. Rev. Phytopathol. 1998, 36, 453–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Van Peer, R.; Niemann, G.J.; Schippers, B. Induced resistance and phytoalexin accumulation in biological control of Fusarium wilt of carnation by Pseudomonas sp. strain WCS 417 r. Phytopathology 1991, 81, 728–734. [Google Scholar] [CrossRef]
  188. Wei, G.; Kloepper, J.W.; Tuzun, S. Induction of systemic resistance of cucumber to Colletotrichum orbiculare by select strains of plant growth-promoting rhizobacteria. Phytopathology 1991, 81, 1508–1512. [Google Scholar] [CrossRef]
  189. Durrant, W.E.; Dong, X. Systemic acquired resistance. Annu. Rev. Phytopathol. 2004, 42, 185–209. [Google Scholar] [CrossRef]
  190. Beneduzi, A.; Ambrosini, A.; Passaglia, L.M. Plant growth-promoting rhizobacteria (PGPR): Their potential as antagonists and biocontrol agents. Genet. Mol. Biol. 2012, 35, 1044–1051. [Google Scholar] [CrossRef] [Green Version]
  191. Van der Ent, S.; Verhagen, B.W.; Van Doorn, R.; Bakker, D.; Verlaan, M.G.; Pel, M.J.; Joosten, R.G.; Proveniers, M.C.; Van Loon, L.C.; Ton, J.; et al. MYB72 is required in early signaling steps of rhizobacteria-induced systemic resistance in Arabidopsis. Plant Physiol. 2008, 146, 1293–1304. [Google Scholar] [CrossRef] [Green Version]
  192. Somers, E.; Vanderleyden, J.; Srinivasan, M. Rhizosphere bacterial signalling: A love parade beneath our feet. Crit. Rev. Microbiol. 2004, 30, 205–240. [Google Scholar] [CrossRef]
  193. Kloepper, J.W.; Ryu, C.M.; Zhang, S. Induced systemic resistance and promotion of plant growth by Bacillus spp. Phytopathology 2004, 94, 1259–1266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Van Wees, S.C.; Van der Ent, S.; Pieterse, C.M. Plant immune responses triggered by beneficial microbes. Curr. Opin. Plant Biol. 2008, 11, 443–448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Vleesschauwer, D.; Höfte, M. Rhizobacteria-induced systemic resistance. Adv. Bot. Res. 2009, 51, 223–281. [Google Scholar]
  196. Van Loon, L.C. Plant responses to plant growth-promoting rhizobacteria. In New Perspectives and Approaches in Plant Growth-Promoting Rhizobacteria Research; Springer: Dordrecht, The Netherlands, 2007; pp. 243–254. [Google Scholar]
  197. Van Peer, R.; Schippers, B. Lipopolysaccharides of plant-growth promoting Pseudomonas sp. strain WCS417r induce resistance in carnation to Fusarium wilt. Neth. J. Plant Pathol. 1992, 98, 129–139. [Google Scholar] [CrossRef]
  198. Leeman, M.; Van Pelt, J.A.; Den Ouden, F.M.; Heinsbroek, M.; Bakker, P.A.H.M.; Schippers, B. Induction of systemic resistance by Pseudomonas fluorescens in radish cultivars differing in susceptibility to fusarium wilt, using a novel bioassay. Eur. J. Plant Pathol. 1995, 101, 655–664. [Google Scholar] [CrossRef]
  199. Silva, H.S.A.; Romeiro, R.D.S.; Mounteer, A. Development of a root colonization bioassay for rapid screening of rhizobacteria for potential biocontrol agents. J. Phytopathol. 2003, 151, 42–46. [Google Scholar] [CrossRef]
  200. Handelsman, J.; Stabb, E.V. Biocontrol of soilborne plant pathogens. Plant Cell 1996, 8, 1855. [Google Scholar] [CrossRef] [PubMed]
  201. Benizri, E.; Baudoin, E.; Guckert, A. Root colonization by inoculated plant growth-promoting rhizobacteria. Biocontrol Sci. Technol. 2001, 11, 557–574. [Google Scholar] [CrossRef]
  202. Montealegre, J.R.; Reyes, R.; Pérez, L.M.; Herrera, R.; Silva, P.; Besoain, X. Selection of bioantagonistic bacteria to be used in biological control of Rhizoctonia solani in tomato. Electron. J. Biotechnol. 2003, 6, 115–127. [Google Scholar] [CrossRef]
  203. Rangarajan, S.; Saleena, L.M.; Vasudevan, P.; Nair, S. Biological suppression of rice diseases by Pseudomonas spp. under saline soil conditions. Plant Soil 2003, 251, 73–82. [Google Scholar] [CrossRef]
  204. Bloemberg, G.V.; Lugtenberg, B.J. Molecular basis of plant growth promotion and biocontrol by rhizobacteria. Curr. Opin. Plant Biol. 2001, 4, 343–350. [Google Scholar] [CrossRef]
  205. Beattie, G.A. Plant-associated bacteria: Survey, molecular phylogeny, genomics and recent advances. In Plant-Associated Bacteria; Springer: Dordrecht, The Netherlands, 2007; pp. 1–56. [Google Scholar]
  206. Neeraja, C.; Anil, K.; Purushotham, P.; Suma, K.; Sarma, P.V.S.R.N.; Moerschbacher, B.M.; Podile, A.R. Biotechnological approaches to develop bacterial chitinases as a bioshield against fungal diseases of plants. Crit. Rev. Biotechnol. 2010, 30, 231–241. [Google Scholar] [CrossRef]
  207. Stephens, P.M.; Crowley, J.J.; O’Connell, C. Selection of pseudomonad strains inhibiting Pythium ultimum on sugarbeet seeds in soil. Soil Biol. Biochem. 1993, 25, 1283–1288. [Google Scholar] [CrossRef]
  208. Kamilova, F.; Validov, S.; Azarova, T.; Mulders, I.; Lugtenberg, B. Enrichment for enhanced competitive plant root tip colonizers selects for a new class of biocontrol bacteria. Environ. Microbiol. 2005, 7, 1809–1817. [Google Scholar] [CrossRef]
  209. Glick, B.R.; Bashan, Y. Genetic manipulation of plant growth-promoting bacteria to enhance biocontrol of phytopathogens. Biotechnol. Adv. 1997, 15, 353–378. [Google Scholar] [CrossRef] [Green Version]
  210. Dakora, F.D. Defining new roles for plant and rhizobial molecules in sole and mixed plant cultures involving symbiotic legumes. New Phytol. 2003, 158, 39–49. [Google Scholar] [CrossRef]
  211. Campanoni, P.; Blasius, B.; Nick, P. Auxin transport synchronizes the pattern of cell division in a tobacco cell line. Plant Physiol. 2003, 133, 1251–1260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Chabot, R.; Antoun, H.; Cescas, M.P. Growth promotion of maize and lettuce by phosphate-solubilizing Rhizobium leguminosarum biovar. phaseoli. Plant Soil 1996, 184, 311–321. [Google Scholar] [CrossRef]
  213. Yanni, Y.G.; Rizk, R.Y.; Corich, V.; Squartini, A.; Ninke, K.; Philip-Hollingsworth, S.; Orgambide, G.; De Bruijn, F.; Stoltzfus, J.; Buckley, D.; et al. Natural endophytic association between Rhizobium leguminosarum bv. trifolii and rice roots and assessment of its potential to promote rice growth. In Opportunities for Biological Nitrogen Fixation in Rice and Other Non-Legumes; Springer: Dordrecht, The Netherlands, 1997; pp. 99–114. [Google Scholar]
  214. Validov, S. Biocontrol of Tomato Foot and Root Rot by Pseudomonas bacteria in Stonewool. Ph.D. Thesis, Leiden University, Leiden, The Netherlands, 2007. [Google Scholar]
  215. Thomashow, L.S.; Weller, D.M. Current concepts in the use of introduced bacteria for biological disease control: Mechanisms and antifungal metabolites. In Plant-Microbe Interactions; Springer: Boston, MA, USA, 1996; pp. 187–235. [Google Scholar]
  216. Dekkers, L.C.; Mulders, I.H.; Phoelich, C.C.; Chin-A-Woeng, T.F.; Wijfjes, A.H.; Lugtenberg, B.J. The sss colonization gene of the tomato-Fusarium oxysporum f. sp. radicis-lycopersici biocontrol strain Pseudomonas fluorescens WCS365 can improve root colonization of other wild-type Pseudomonas spp. bacteria. Mol. Plant Microbe Interact. 2000, 13, 1177–1183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Lugtenberg, B.J.; Dekkers, L.; Bloemberg, G.V. Molecular determinants of rhizosphere colonization by Pseudomonas. Annu. Rev. Phytopathol. 2001, 39, 461–490. [Google Scholar] [CrossRef] [PubMed]
  218. Kremer, R.J.; Souissi, T. Cyanide production by rhizobacteria and potential for suppression of weed seedling growth. Curr. Microbiol. 2001, 43, 182–186. [Google Scholar] [CrossRef] [PubMed]
  219. Chet, I.; Ordentlich, A.; Shapira, R.; Oppenheim, A. Mechanisms of biocontrol of soil-borne plant pathogens by rhizobacteria. Plant Soil 1990, 129, 85–92. [Google Scholar] [CrossRef]
  220. Kobayashi, D.Y.; Reedy, R.M.; Bick, J.; Oudemans, P.V. Characterization of a chitinase gene from Stenotrophomonas maltophilia strain 34S1 and its involvement in biological control. Appl. Environ. Microbiol. 2002, 68, 1047–1054. [Google Scholar] [CrossRef] [Green Version]
  221. Sadfi, N.; Cherif, M.; Fliss, I.; Boudabbous, A.; Antoun, H. Evaluation of bacterial isolates from salty soils and Bacillus thuringiensis strains for the biocontrol of Fusarium dry rot of potato tubers. J. Plant Pathol. 2001, 83, 101–117. [Google Scholar]
  222. Neiendam Nielsen, M.; Sørensen, J. Chitinolytic activity of Pseudomonas fluorescens isolates from barley and sugar beet rhizosphere. FEMS Microbiol. Ecol. 1999, 30, 217–227. [Google Scholar] [CrossRef]
  223. Budi, S.W.; van Tuinen, D.; Arnould, C.; Dumas-Gaudot, E.; Gianinazzi-Pearson, V.; Gianinazzi, S. Hydrolytic enzyme activity of Paenibacillus sp. strain B2 and effects of the antagonistic bacterium on cell integrity of two soil-borne pathogenic fungi. Appl. Soil Ecol. 2000, 15, 191–199. [Google Scholar] [CrossRef]
  224. Someya, N.; Kataoka, N.; Komagata, T.; Hirayae, K.; Hibi, T.; Akutsu, K. Biological control of cyclamen soilborne diseases by Serratia marcescens strain B2. Plant Dis. 2000, 84, 334–340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Kurek, E.; Jaroszuk-Ściseł, J. Rye (Secale cereale) growth promotion by Pseudomonas fluorescens strains and their interactions with Fusarium culmorum under various soil conditions. Biol. Control 2003, 26, 48–56. [Google Scholar] [CrossRef]
  226. Mulla, S.I.; Bharagava, R.N.; Belhaj, D.; Saratale, G.D.; Kumar, A.; Rajasekar, A.; Tallur, P.N.; Edalli, V.A.; Hu, A.; Yu, C.P. Microbial degradation of phenolic compounds. In Microbes and Enzymes in Soil Health and Bioremediation; Microorganisms for sustainability; Kumar, A., Sharma, S., Eds.; Springer: Singapore, 2019; Volume 16. [Google Scholar] [CrossRef]
  227. Phulpoto, A.H.; Maitlo, M.A.; Kanhar, N.A. Culture dependent to culture independent approaches for the bioremediation of paints: A review. Int. J. Environ. Sci. Technol. 2021, 18, 241–262. [Google Scholar] [CrossRef]
  228. Muratova, A.Y.; Turkovskaya, O.V.; Antonyuk, L.P.; Makarov, O.E.; Pozdnyakova, L.I.; Ignatov, V.V. Oil-oxidizing potential of associative rhizobacteria of the genus Azospirillum. Microbiology 2005, 74, 210–215. [Google Scholar] [CrossRef]
  229. Ali, M.H.; Sattar, M.T.; Khan, M.I.; Naveed, M.; Rafique, M.; Alamri, S.; Siddiqui, M.H. Enhanced Growth of Mungbean and Remediation of Petroleum Hydrocarbons by Enterobacter sp. MN17 and Biochar Addition in Diesel Contaminated Soil. Appl. Sci. 2020, 10, 8548. [Google Scholar]
  230. Ejaz, M.; Zhao, B.; Wang, X.; Bashir, S.; Haider, F.U.; Aslam, Z.; Khan, M.I.; Shabaan, M.; Naveed, M.; Mustafa, A. Isolation and characterization of oil degrading Enterobacter sp. from naturally hydrocarbon contaminated soils and their potential against bioremediation of crude oil. Appl. Sci. 2021, 11, 3504. [Google Scholar] [CrossRef]
  231. Haider, F.U.; Liqun, C.; Coulter, J.A.; Alam Cheema, S.; Wu, J.; Zhang, R.; Wenjun, M.; Farooq, M. Cadmium toxicity in plants: Impacts and remediation strategies. Ecotoxicol. Environ. Saf. 2021, 211, 111887. [Google Scholar] [CrossRef]
  232. Varjani, S.; Kumar, G.; Rene, E.R. Developments in biochar application for pesticide remediation: Current knowledge and future research directions. J. Environ. Manag. 2019, 232, 505–513. [Google Scholar] [CrossRef]
  233. Mahjoubi, M.; Cappello, S.; Souissi, Y.; Jaouani, A.; Cherif, A. Microbial bioremediation of petroleum hydrocarbon–contaminated marine environments. In Recent Insights in Petroleum Science and Engineering; Zoveidavianpoor, M., Ed.; IntechOpen: London, UK, 2017. [Google Scholar] [CrossRef] [Green Version]
  234. Lu, L.; Zhang, J.; Peng, C. Shift of Soil Polycyclic Aromatic Hydrocarbons (PAHs) dissipation pattern and microbial community composition due to rhamnolipid supplementation. Water Air Soil Pollut. 2019, 230, 107. [Google Scholar] [CrossRef]
  235. Kumar, V.; Chandra, R. Bacterial-assisted phytoextraction mechanism of heavy metals by native hyperaccumulator plants from distillery waste–contaminated site for eco-restoration. In Microbes for Sustainable Development and Bioremediation; CRC Press: Boca Raton, FL, USA, 2020. [Google Scholar] [CrossRef]
  236. Macías-Almazán, A.; Lois-Correa, J.A.; Domínguez-Crespo, M.A.; López-Oyama, A.B.; Torres-Huerta, A.M.; Brachetti-Sibaja, S.B.; Rodríguez-Salazar, A.E. Influence of operating conditions on proton conductivity of nanocellulose films using two agroindustrial wastes: Sugarcane bagasse and pinewood sawdust. Carbohydr. Polym. 2020, 238, 116171. [Google Scholar] [CrossRef]
  237. Russo, F.; Ceci, A.; Pinzari, F.; Siciliano, A.; Guida, M.; Malusà, E.; Tartanus, M.; Miszczak, A.; Maggi, O.; Persiani, A.M. Bioremediation of dichlorodiphenyltrichloroethane (DDT)-contaminated agricultural soils: Potential of two autochthonous saprotrophic fungal strains. Appl. Environ. Microbiol. 2019, 85, 01720-19. [Google Scholar] [CrossRef]
  238. Chandra, R.; Kumar, V. Phytoextraction of heavy metals by potential native plants and their microscopic observation of root growing on stabilised distillery sludge as a prospective tool for in situ phytoremediation of industrial waste. Environ. Sci. Pollut. Res. 2017, 24, 2605–2619. [Google Scholar] [CrossRef]
  239. Chouychai, W.; Thongkukiatkul, A.; Upatham, S.; Pokethitiyook, P.; Kruatrachue, M.; Lee, H. Effect of corn plant on survival and phenanthrene degradation capacity of Pseudomonas sp. UG14Lr in two soils. Int. J. Phytoremediat. 2012, 14, 585–595. [Google Scholar] [CrossRef] [PubMed]
  240. Malik, Z.A.; Ahmed, S. Degradation of petroleum hydrocarbons by oil field isolated bacterial consortium. Afr. J. Biotechnol. 2012, 11, 650–658. [Google Scholar]
  241. Thomas, F.; Cébron, A. Short-term rhizosphere effect on available carbon sources, phenanthrene degradation, and active microbiome in an aged contaminated industrial soil. Front. Microbiol. 2016, 7, 92. [Google Scholar] [CrossRef] [PubMed]
  242. Ahmad, I.; Akhtar, M.J.; Zahir, Z.A.; Naveed, M.; Mitter, B.; Sessitsch, A. Cadmium-tolerant bacteria induce metal stress tolerance in cereals. Environ. Sci. Pollut. Res. 2014, 21, 11054–11065. [Google Scholar] [CrossRef] [PubMed]
  243. Lata, S.; Kaur, H.P.; Mishra, T. Cadmium bioremediation: A review. Int. J. Pharm. Sci. Res. 2019, 10, 4120–4128. [Google Scholar] [CrossRef]
  244. Yaghoubian, Y.; Siadat, S.A.; Telavat, M.R.M.; Pirdashti, H.; Yaghoubia, I. Bio-removal of cadmium from aqueous solutions by filamentous fungi: Trichoderma spp. and Piriformospora indica. Environ. Sci. Pollut. Res. 2019, 26, 7863–7872. [Google Scholar] [CrossRef] [PubMed]
  245. Zhang, F.; Liu, M.; Li, Y.; Che, Y.; Xiao, Y. Effects of arbuscularmycorrhizal fungi, biochar and cadmium on the yield and element uptake of Medicago sativa. Sci. Total Environ. 2019, 655, 1150–1158. [Google Scholar] [CrossRef] [PubMed]
  246. Panchenko, L.; Muratova, A.; Dubrovskaya, E.; Golubev, S.; Turkovskaya, O. Dynamics of natural revegetation of hydrocarbon-contaminated soil and remediation potential of indigenous plant species in the steppe zone of the southern Volga Uplands. Environ. Sci. Pollut. Res. 2018, 25, 3260–3274. [Google Scholar] [CrossRef]
  247. Kumar, A.; Sharma, G.; Naushad, M.; Ala’a, H.; Kumar, A.; Hira, I.; Ahamad, T.; Ghfar, A.A.; Stadler, F.J. Visible photodegradation of ibuprofen and 2, 4-D in simulated waste water using sustainable metal free-hybrids based on carbon nitride and biochar. J. Environ. Manag. 2019, 231, 1164–1175. [Google Scholar] [CrossRef]
  248. Valderrama, J.A.; Durante-Rodríguez, G.; Blázquez, B.; García, J.L.; Carmona, M.; Díaz, E. Bacterial Degradation of Benzoate cross-regulation between aerobic and anaerobic pathways. J. Biol. Chem. 2012, 287, 10494–10508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Johnson, D.L.; Maguire, K.L.; Anderson, D.R.; McGrath, S.P. Enhanced dissipation of chrysene in planted soil: The impact of a rhizobial inoculum. Soil Biol. Biochem. 2004, 36, 33–38. [Google Scholar] [CrossRef]
  250. Al-Wasify, R.S.; Hamed, S.R. Bacterial biodegradation of crude oil using local isolates. Int. J. Bacteriol. 2014, 2014, 863272. [Google Scholar] [CrossRef] [PubMed]
  251. Ichor, T.; Okerentugba, P.O.; Okpokwasili, G.C. Biodegradation of total petroleum hydrocarbon by aerobic heterotrophic bacteria isolated from crude oil contaminated brackish waters of Bodo Creek. J. Bioremed. Biodeg. 2014, 5, 1000236. [Google Scholar] [CrossRef]
  252. Jilani, S. Bioremediation application for textile effluent treatment. Middle East J. Sci. Res. 2015, 23, 26–34. [Google Scholar] [CrossRef]
  253. Verma, J.P.; Jaiswal, D.K.; Sagar, R. Pesticide relevance and their microbial degradation: A-state-of-art. Rev. Environ. Sci. Biotechnol. 2014, 13, 429–466. [Google Scholar] [CrossRef]
  254. Koksoy, H.; Aslim, B. Determination of herbicide resistance in aquatic cyanobacteria by probit analysis. J. Appl. Biol. Sci. 2013, 7, 37–41. [Google Scholar]
  255. Begum, S.S.; Arundhati, A. A study of bioremediation of methyl parathion in vitro India. Int. J. Curr. Microbiol. App. Sci. 2016, 5, 464–474. [Google Scholar] [CrossRef]
  256. Baraldi, E.A.; Damianovic, M.; Manfio, G.R.; Foresti, E.; Vazoller, R.F. Performance of a horizontal-flow anaerobic immobilized biomass (HAIB) reactor and dynamics of the microbial community during degradation of pentachlorophenol (PCP). Anaerobe 2008, 14, 268–274. [Google Scholar] [CrossRef]
  257. Tripathi, A.; Dixit. S. Bioremediation of phenolic compounds by higher fungi—A review. Int. J. Curr. Adv. Res. 2016, 4, 14–35. [Google Scholar] [CrossRef] [Green Version]
  258. Thakurta, G.S.; Aakula, M.; Chakrabarty, J.; Dutta, S. Bioremediation of phenol from synthetic and real wastewater using Leptolyngbya sp.: A comparison and assessment of lipid production. 3 Biotech 2018, 8, 206. [Google Scholar] [CrossRef]
  259. Sheng, X.F.; He, L.Y. Solubilization of potassium-bearing minerals by a wild-type strain of Bacillus edaphicus and its mutants and increased potassium uptake by wheat. Can. J. Microbiol. 2006, 52, 66–72. [Google Scholar] [CrossRef] [PubMed]
  260. Li, J.H.; Gao, Y.; Wu, S.C.; Cheung, K.C.; Wang, X.R.; Wong, M.H. Physiological and biochemical responses of rice (Oryza sativa L.) to phenanthrene and pyrene. Int. J. Phytorem. 2008, 10, 106–118. [Google Scholar] [CrossRef] [PubMed]
  261. Chouychai, W.; Thongkukiatkul, A.; Upatham, S.; Lee, H.; Pokethitiyook, P.; Kruatrachue, M. Plant-enhanced phenanthrene and pyrene biodegradation in acidic soil. J. Environ. Biol. 2009, 30, 139–144. [Google Scholar] [PubMed]
  262. Lechner, U.; Turkowsky, D.; Dinh, T.T.H.; Al-Fathi, H.; Schwoch, S.; Franke, S.; Gerlach, M.S.; Koch, M.; Bergen, M.V.; Jehmlich, N.; et al. Desulfitobacterium contributes to the microbial transformation of 2,4,5-T by methanogenic enrichment cultures from a Vietnamese active landfill. Microb. Biotechnol. 2018, 11, 1137–1156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Mulligan, C.N.; Yong, R.N.; Gibbs, B.F. Surfactant enhanced remediation of contaminated soil: A review. Eng. Geol. 2001, 60, 371–380. [Google Scholar] [CrossRef]
  264. Kaczorek, E.; Chrzanowski, Ł.; Pijanowska, A.; Olszanowski, A. Yeast and bacteria cell hydrophobicity and hydrocarbon biodegradation in the presence of natural surfactants: Rhamnolipides and saponins. Bioresour. Technol. 2008, 99, 4285–4291. [Google Scholar] [CrossRef]
  265. Mohanty, S.; Mukherji, S. Surfactant-aided biodegradation of NAPLs by Burkholderia multivorans: Comparison between Triton X-100 and rhamnolipid JBR 515. Colloids Surf. B 2012, 102, 644–652. [Google Scholar] [CrossRef]
  266. Selberg, A.; Juuram, K.; Budashova, J.; Tenno, T. Biodegradation and leaching of surfactants during surfactant-amended bioremediation of oil-polluted soil. In Applied Bioremediation—Active and Passive Approaches; Patil, Y.B., Rao, P., Eds.; IntechOpen: London, UK, 2013. [Google Scholar] [CrossRef] [Green Version]
  267. Babu, A.G.; Shim, J.; Shea, P.J.; Oh, B.T. Penicillium aculeatum PDR-4 and Trichoderma sp. PDR-16 promote phytoremediation of mine tailing soil and bioenergy production with sorghum-sudan grass. Ecol. Eng. 2014, 69, 186–191. [Google Scholar] [CrossRef]
  268. Mahajan, P.; Kaushal, J. Role of phytoremediation in reducing cadmium toxicity in soil and water. Hindawi J. Toxicol. 2018, 2018, 4864365. [Google Scholar] [CrossRef] [Green Version]
  269. Belimov, A.A.; Hontzeas, N.; Safronova, V.I.; Demchinskaya, S.V.; Piluzza, G.; Bullitta, S.; Glick, B.R. Cadmium-tolerant plant growth promoting rhizobacteria associated with the roots of Indian mustard (Brassica juncea L. Czern.). Soil Biol. Biochem. 2005, 37, 241–250. [Google Scholar] [CrossRef]
  270. Robinson, B.; Russell, C.; Hedley, M.J.; Clothier, B. Cadmium adsorption by rhizobacteria: Implications for New Zealand Pastureland. Agric. Ecosyst. Environ. 2001, 87, 315–321. [Google Scholar] [CrossRef]
  271. Gupta, A.; Meyer, J.M.; Goel, R. Development of heavy metal resistant mutants of phosphate solubilizing Pseudomonas sp. NBRI4014 and their characterization. Curr. Microbiol. 2002, 45, 323–332. [Google Scholar] [CrossRef] [PubMed]
  272. Marques, G.C.; Moreira, H.; Franco, A.R.; Rangel, A.O.S.S.; Castro, P.M.L. Inoculating Helianthus annuus (sunflower) grown in zinc and cadmium contaminated soils with plant growth promoting bacteria—Effects on phytoremediation strategies. Chemosphere 2013, 92, 74–83. [Google Scholar] [CrossRef] [PubMed]
  273. Prapagdee, B.; Khonsue, N. Bacterial-assisted cadmium phytoremediation by Ocimum gratissimum L. in polluted agricultural soil: A field trial experiment. Int. J. Environ. Sci. Technol. 2015, 12, 3843–3852. [Google Scholar] [CrossRef] [Green Version]
  274. Yuan, Z.; Yi, H.; Wang, T.; Zhang, Y.; Zhu, X.; Yao, J. Application of phosphate solubilizing bacteria in immobilization of Pb and Cd in soil. Environ. Sci. Pollut. Res. 2017, 27, 21877–21884. [Google Scholar] [CrossRef]
  275. Martín-González, A.; Díaz, S.; Borniquel, S.; Gallego, A.; Gutiérrez, J.C. Cytotoxicity and bioaccumulation of heavy metals by ciliated protozoa isolated from urban wastewater treatment plants. Res. Microbiol. 2006, 157, 108–118. [Google Scholar] [CrossRef]
  276. Saranya, K.; Sundaramanickam, A.; Shekhar, S.; Meena, M.; Sathishkumar, R.S.; Balasubramanian, T. Biosorption of multiheavy metals by coral associated phosphate solubilising bacteria Cronobacter muytjensii KSCAS2. J. Environ. Manag. 2018, 222, 396–401. [Google Scholar] [CrossRef]
  277. Choińska-Pulit, A.; Sobolczyk-Bednarek, J.; Łaba, W. Optimization of copper, lead and cadmium biosorption onto newly isolated bacterium using a Box-Behnken design. Ecotoxicol. Environ Saf. 2018, 149, 275–283. [Google Scholar] [CrossRef]
  278. Vishan, I.; Saha, B.; Sivaprakasam, S.; Kalamdhad, A. Evaluation of Cd(II) biosorption in aqueous solution by using lyophilized biomass of novel bacterial strain Bacillus badius AK: Biosorption kinetics, thermodynamics and mechanism. Environ. Technol. Innov. 2019, 14, 100323. [Google Scholar] [CrossRef]
  279. Wu, S.C.; Cheung, K.C.; Luo, Y.M.; Wong, M.H. Effects of inoculation of plant growth-promoting rhizobacteria on metal uptake by Brassica juncea. Environ. Pollut. 2006, 140, 124–135. [Google Scholar] [CrossRef] [PubMed]
  280. Arunakumara, K.K.; Walpola, B.C.; Song, J.S.; Shin, M.; Lee, C.; Yoon, M. Phytoextraction of heavy metals induced by bioaugmentation of a phosphate solubilising bacterium. Korean J. Environ. Agric. 2014, 33, 220–230. [Google Scholar] [CrossRef] [Green Version]
  281. Faisal, M.; Hasnain, S. Bacterial Cr (VI) reduction concurrently improves sunflower (Helianthus annuus L.) growth. Biotechnol. Lett. 2005, 27, 943–947. [Google Scholar] [CrossRef] [PubMed]
  282. Abou-Shanab, R.A.; Ghanem, K.; Ghanem, N.; Al-Kolaibe, A. The role of bacteria on heavy metal extraction and uptake by plants growing on multi-metal-contaminated soils. World J. Microbiol. Biotechnol. 2008, 24, 253–262. [Google Scholar] [CrossRef]
  283. Oves, M.; Khan, M.S.; Zaidi, A. Chromium reducing and plant growth promoting novel strain Pseudomonas aeruginosa OSG41 enhance chickpea growth in chromium amended soils. Eur. J. Soil Biol. 2013, 56, 72–83. [Google Scholar] [CrossRef]
  284. 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] [Green Version]
  285. Khan, M.U.; Sessitsch, A.; Harris, M.; Fatima, K.; Imran, A.; Arslan, M.; Shabir, G.; Khan, Q.M.; Afzal, M. Cr-resistant rhizo-and endophytic bacteria associated with Prosopis juliflora and their potential as phytoremediation enhancing agents in metal-degraded soils. Front. Plant Sci. 2015, 5, 755. [Google Scholar] [CrossRef] [Green Version]
  286. Srinath, T.; Verma, T.; Ramteke, P.; Garg, S. Chromium (VI) biosorption and bioaccumulation by chromate resistant bacteria. Chemosphere 2002, 48, 427–435. [Google Scholar] [CrossRef]
  287. Ksheminska, H.; Fedorovych, D.; Babyak, L.; Yanovych, D.; Kaszycki, P.; Koloczek, H. Chromium (III) and (VI) tolerance and bioaccumulation in yeast: A survey of cellular chromium content in selected strains of representative genera. Process Biochem. 2005, 40, 1565–1572. [Google Scholar] [CrossRef]
  288. Li, K.; Ramakrishna, W. Effect of multiple metal resistant bacteria from contaminated lake sediments on metal accumulation and plant growth. J. Hazard. Mater. 2011, 189, 531–539. [Google Scholar] [CrossRef] [PubMed]
  289. Jing, Y.D.; He, Z.L.; Yang, X.E. Role of soil rhizobacteria in phytoremediation of heavy metal contaminated soils. J. Zhejiang Univ. Sci. B 2007, 8, 192–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  290. Wenzel, W.W. Rhizosphere processes and management in plant-assisted bioremediation (phytoremediation) of soils. Plant Soil 2009, 321, 385–408. [Google Scholar] [CrossRef]
  291. Płociniczak, T.; Sinkkonen, A.; Romantschuk, M.; Sułowicz, S.; Piotrowska-Seget, Z. Rhizospheric bacterial strain Brevibacterium casei MH8a colonizes plant tissues and enhances Cd, Zn, Cu phytoextraction by white mustard. Front. Plant Sci. 2016, 16, 101. [Google Scholar] [CrossRef]
  292. Lin, M.; Jin, M.; Xu, K.; He, L.; Cheng, D. Phosphate solubilising bacteria improve the phytoremediation efficiency of Wedeliatrilobata for Cu-contaminated soil. Int. J. Phytoremediat. 2018, 20, 813–822. [Google Scholar] [CrossRef]
  293. Iram, S.; Shabbir, R.; Zafar, H.; Javaid, M. Biosorption and bioaccumulation of copper and lead by heavy metal-resistant fungal isolates. Arab. J. Sci. Eng. 2015, 40, 1867–1873. [Google Scholar] [CrossRef]
  294. De-Silóniz, M.-I.; Balsalobre, L.; Alba, C.; Valderrama, M.-J.; Peinado, J.M. Feasibility of copper uptake by the yeast Pichia guilliermondii isolated from sewage sludge. Res. Microbiol. 2002, 153, 173–180. [Google Scholar] [CrossRef]
  295. Ma, Y.; Rajkumar, M.; Rocha, I.; Oliveira, R.; Freitas, H. Serpentine bacteria influence metal translocation and bioconcentration of Brassica juncea and Ricinus communis growing in multi-pollute soils. Front. Plant Sci. 2015, 5, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. 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]
  297. Becerra-Castro, C.; Prieto-Fernandez, A.; Alvarez-Lopez, V.; Monterroso, C.; Cabello-Conejo, M.I.; Acea, M.J. Nickel solubilising capacity of rhizobacteria isolated from hyperaccumulating and non-hyperaccumulating subspecies of Alyssum serpyllifolium. Int. J. Phytoremediat. 2011, 13, 229–244. [Google Scholar] [CrossRef] [PubMed]
  298. Aboudrar, W.; Schwartz, C.; Morel, J.L.; Boularbah, A. Effect of nickel-resistant rhizosphere bacteria on the uptake of nickel by the hyperaccumulator Noccaea caerulescens under controlled conditions. J. Soils Sediments 2013, 13, 501–507. [Google Scholar] [CrossRef]
  299. Hassan, M.U.; Chattha, M.U.; Khan, I.; Chattha, M.B.; Aamer, M.; Nawaz, M.; Ali, A.; Khan, M.A.U.; Khan, T.A. Nickel toxicity in plants: Reasons, toxic effects, tolerance mechanisms, and remediation possibilities—A review. Environ. Sci. Pollut. Res. 2019, 26, 12673–12688. [Google Scholar] [CrossRef]
  300. Burd, G.I. Plant growth-promoting bacteria that decrease heavy metal toxicity in plants. Can. J. Microbiol. 2000, 46, 237–245. [Google Scholar] [CrossRef]
  301. Govarthanan, M.; Mythili, R.; Selvakumar, T.; Kamala-Kannan, S.; Kim, H. Myco-phytoremediation of arsenic- and lead-contaminated soils by Helianthus annuus and wood rot fungi, Trichoderma sp. isolated from decayed wood. Ecotoxicol. Environ. Saf. 2018, 151, 279–284. [Google Scholar] [CrossRef]
  302. Mohapatra, R.K.; Parhi, P.K.; Pandey, S.; Bindhani, B.K.; Thatoi, H.; Panda, C.R. Active and passive biosorption of Pb(II)using live and dead biomass of marine bacterium Bacillus xiamenensis PbRPSD202: Kinetics and isotherm studies. J. Environ. Manag. 2019, 247, 121–134. [Google Scholar] [CrossRef]
  303. Zhang, X.; Chen, Z.; Ma, Y.; Zhou, Y.; Zhao, S.; Wang, L.; Zhai, H. Influence of elevated Zn (II) on Anammox system: Microbial variation and zinc tolerance. Bioresour. Technol. 2018, 251, 108–113. [Google Scholar] [CrossRef]
  304. Kramer, J.; Özkaya, Ö.; Kümmerli, R. Bacterial siderophores in community and host interactions. Nat. Rev. Microbiol. 2020, 18, 152–163. [Google Scholar] [CrossRef]
  305. Prabhu, N.; Borkar, S.; Garg, S. Phosphate solubilization by microorganisms. In Advances in Biological Science Research; Academic Press: Cambridge, MA, USA, 2019; pp. 161–176. [Google Scholar] [CrossRef]
  306. Rajkumar, M.; Ae, N.; Prasad, M.N.; Freitas, H. Potential of siderophore-producing bacteria for improving heavy metal phytoextraction. Trends Biotechnol. 2010, 28, 142–149. [Google Scholar] [CrossRef]
  307. Gaonkar, T.; Bhosle, S. Effect of metals on a siderophore producing isolate and its implications on microbial assisted bioremediation of metal contaminated soils. Chemosphere 2013, 93, 9. [Google Scholar] [CrossRef] [PubMed]
  308. Chandra, R.; Saxena, G.; Kumar, V. Phytoremediation of environmental pollutants: An eco-sustainable green technology to environmental management. In Advances in Biodegradation and Bioremediation of Industrial Waste; Chandra, R., Ed.; CRC Press: Boca Raton, FL, USA, 2015; pp. 1–29. [Google Scholar]
  309. Ma, Y.; Prasad, M.N.V.; Rajkumar, M.; Freitas, H. Plant growth promoting rhizobacteria and endophytes accelerate phytoremediation of metalliferous soils. Biotechnol. Adv. 2011, 29, 248–258. [Google Scholar] [CrossRef]
  310. Gamalero, E.; Glick, B.R. Mechanisms used by plant growth-promoting bacteria. In Bacteria in Agrobiology: Plant Nutrient Management; Maheshwari, D.K., Ed.; Springer: Berlin/Heidelberg, Germany, 2011; pp. 17–46. [Google Scholar]
  311. Violante, A.; Zhu, J.; Pigna, M.; Jara, A.; Cozzolino, V.; Mora, M.L. Role of Biomolecules in Influencing Transformation Mechanisms of Metals and Metalloids in Soil Environments. In Molecular Environmental Soil Science; Springer: Dordrecht, The Netherlands, 2013; pp. 167–191. [Google Scholar]
  312. Meyer, J.M. Pyoverdines: Pigments siderophores and potential taxonomic markers of fluorescent Pseudomonas species. Arch. Microbiol. 2000, 174, 135–142. [Google Scholar] [CrossRef]
  313. Jing, Y.X.; Yan, J.L.; He, H.D.; Yang, D.J.; Xiao, L. Characterization of bacteria in the rhizosphere soils of Polygonum pubescens and their potential in promoting growth and Cd, Pb, Zn uptake by Brassica napus. Int. J. Phytoremediat. 2014, 16, 321–333. [Google Scholar] [CrossRef] [PubMed]
  314. Sifour, M.; Al-Jilawi, M.H.; Aziz, G.M. Emulsification properties of biosurfactant produced from Pseudomonas aeruginosa RB 28. Pak. J. Biol. Sci. 2007, 10, 1331–1335. [Google Scholar] [CrossRef] [Green Version]
  315. Reddy, M.S.; Naresh, B.; Leela, T.; Prashanthi, M.; Madhusudhan, N.C.; Dhanasri, G. Biodegradation of phenanthrene with biosurfactant production by a new strain of Brevibacillus sp. Bioresour. Technol. 2010, 101, 79807983. [Google Scholar] [CrossRef] [PubMed]
  316. Geys, R.; Soetaert, W.; Bogaer, I.V. Biotechnological opportunities in biosurfactant production. Curr. Opin. Biotechnol. 2014, 30, 66–72. [Google Scholar] [CrossRef]
  317. Ławniczak, Ł.; Marecik, R.; Chrzanowski, Ł. Contributions of biosurfactants to natural or induced bioremediation. Appl. Microbiol. Biotechnol. 2013, 97, 2327–2339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Patil, S.N.; Aglave, B.A.; Pethkar, A.V.; Gaikwad, V.B. Stenotrophomonas koreensis a novel biosurfactant producer for abatement of heavy metals from the environment. Afr. J. Microbiol. Res. 2012, 6, 51735178. [Google Scholar]
  319. Lima, T.M.S.; Procopio, L.C.; Brandao, F.D.; Carvalho, A.M.X.; Totola, M.R.; Borges, A.C. Simultaneous phenanthrene and cadmium removal from contaminated soil by a ligand/biosurfactant solution. Biodegradation 2011, 22, 10071015. [Google Scholar] [CrossRef]
  320. Qazi, M.A.; Malik, Z.A.; Qureshi, G.D.; Hameed, A.; Ahmed, S. Yeast extract as the most preferable substrate for optimized biosurfactant production by rhlB gene positive Pseudomonas putida SOL-10 isolate. J. Bioremed. Biodeg. 2013, 4, 2. [Google Scholar]
  321. Tambekar, D.H.; Gadakh, P.V. Biochemical and molecular detection of biosurfactant producing bacteria from soil. Int. J. Life Sci. Biotech. Pharma Res 2013, 2, 204–211. [Google Scholar]
  322. Juwarkar, A.A.; Nair, A.; Dubey, K.V.; Singh, S.K.; Devotta, S. Biosurfactant technology for remediation of cadmium and lead contaminated soils. Chemosphere 2007, 68, 1996–2002. [Google Scholar] [CrossRef]
  323. Chakraborty, J.; Das, S. Biosurfactant-based bioremediation of toxic metals. In Microbial Biodegradation and Bioremediation; Elsevier: Amsterdam, The Netherlands, 2014; pp. 167–201. [Google Scholar] [CrossRef]
  324. Miller, R.M. Biosurfactant facilitated remediation of contaminated soil. Environ. Health Perspect. 1995, 103, 59–62. [Google Scholar] [PubMed]
  325. Kang, S.W.; Kim, Y.B.; Shin, J.D.; Kim, E.K. Enhanced biodegradation of hydrocarbons in soil by microbial biosurfactant, sophorolipid. Appl. Biochem. Biotechnol. 2010, 160, 780790. [Google Scholar] [CrossRef]
  326. Kapadia, S.G.; Yagnik, B.N. Current trend and potential for microbial biosurfactants. Asian J. Exp. Biol. Sci. 2013, 4, 1–8. [Google Scholar]
  327. Mrvčić, J.; Stanzer, D.; Šolić, E.; Stehlik-Tomas, V. Interaction of lactic acid bacteria with metal ions: Opportunities for improving food safety and quality. World J. Microbiol. Biotechnol. 2012, 28, 2771–2782. [Google Scholar] [CrossRef] [PubMed]
  328. Ahalya, N.; Ramachandra, T.V.; Kanamadi, R.D. Biosorption of heavy metals. Res. J. Chem. Environ. 2003, 7, 71–79. [Google Scholar]
  329. Escudero, L.B.; Quintas, P.Y.; Wuilloud, R.G.; Dotto, G.L. Recent advances on elemental biosorption. Environ. Chem. Lett. 2019, 17, 409–427. [Google Scholar] [CrossRef]
  330. Davis, T.A.; Volesky, B.; Mucci, A. A review of the biochemistry of heavy metal biosorption by brown algae. Water Res. 2003, 37, 4311–4330. [Google Scholar] [CrossRef]
  331. Volesky, B. Biosorption for the next century. In Process Metallurgy; Elsevier: Amsterdam, The Netherlands, 1999; Volume 9, pp. 161–170. [Google Scholar] [CrossRef]
  332. Chojnacka, K. Biosorption and bioaccumulation–the prospects for practical applications. Environ. Int. 2010, 36, 299–307. [Google Scholar] [CrossRef]
  333. Yilmazer, P.; Saracoglu, N. Bioaccumulation and biosorption of copper (II) and chromium (III) from aqueous solutions by Pichia stipitis yeast. J. Chem. Technol. Biotechnol. 2009, 84, 604–610. [Google Scholar] [CrossRef]
  334. Ramrakhiani, L.; Majumder, R.; Khowala, S. Removal of hexavalent chromium by heat inactivated fungal biomass of Termitomycesclypeatus: Surface characterization and mechanism of biosorption. Chem. Eng. J. 2011, 171, 1060–1068. [Google Scholar] [CrossRef]
  335. Ding, Y.; Jing, D.; Gong, H.; Zhou, L.; Yang, X. Biosorption of aquatic cadmium (II) by unmodified rice straw. Bioresour. Technol. 2012, 114, 20–25. [Google Scholar] [CrossRef] [PubMed]
  336. Chojnacka, K. Biosorption and Bioaccumulation in Practice; Nova Science Publishers: Hauppauge, New York, USA, 2009; p. 137. [Google Scholar]
  337. Chojnacka, K.; Chojnacki, A.; Gorecka, H. Biosorption of Cr3+, Cd2+ and Cu2+ ions by blue–green algae Spirulina sp.: Kinetics, equilibrium and the mechanism of the process. Chemosphere 2005, 59, 75–84. [Google Scholar] [CrossRef] [PubMed]
  338. Kanamarlapudi, S.L.R.K.; Chintalpudi, V.K.; Muddada, S. Application of biosorption for removal of heavy metals from wastewater. Biosorption 2018, 18, 69. [Google Scholar] [CrossRef] [Green Version]
  339. Calderón, O.A.R.; Abdeldayem, O.M.; Pugazhendhi, A.; Rene, E.R. Current updates and perspectives of biosorption technology: An alternative for the removal of heavy metals from wastewater. Curr. Pollut. Rep. 2020, 6, 8–27. [Google Scholar] [CrossRef]
  340. Ahemad, M.; Kibret, M. Recent trends in microbial biosorption of heavy metals: A review. Biochem. Mol. Biol. 2013, 1, 19–26. [Google Scholar] [CrossRef]
  341. Villen-Guzman, M.; Gutierrez-Pinilla, D.; Gomez-Lahoz, C.; Vereda-Alonso, C.; Rodriguez-Maroto, J.M.; Arhoun, B. Optimization of Ni (II) biosorption from aqueous solution on modified lemon peel. Environ. Res. 2019, 179, 108849. [Google Scholar] [CrossRef]
  342. El-Naggar, N.E.A.; Hamouda, R.A.; Mousa, I.E.; Abdel-Hamid, M.S.; Rabei, N.H. Biosorption optimization, characterization, immobilization and application of Gelidiumamansii biomass for complete Pb2+ removal from aqueous solutions. Sci. Rep. 2018, 8, 13456. [Google Scholar] [CrossRef]
  343. 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]
  344. Belimov, A.A.; Safronova, V.I.; Sergeyeva, T.A.; Egorova, T.N.; Matveyeva, V.A.; Tsyganov, V.E.; Borisov, A.Y.; Tikhonovich, I.A.; Kluge, C.; Preisfeld, A.; et al. Characterization of plant growth promoting rhizobacteria isolated from polluted soils and containing 1-aminocyclopropane-1-carboxylate deaminase. Can. J. Microbiol. 2001, 47, 642–652. [Google Scholar] [CrossRef] [PubMed]
  345. Rodriguez, H.; Vessely, S.; Shah, S.; Glick, B.R. Effect of a nickel-tolerant ACC deaminase-producing Pseudomonas strain on growth of non transformed and transgenic canola plants. Curr. Microbiol. 2008, 57, 170–174. [Google Scholar] [CrossRef]
  346. Vu, B.; Chen, M.; Crawford, R.J.; Ivanova, E.P. Bacterial extracellular polysaccharides involved in biofilm formation. Molecules 2009, 14, 2535–2554. [Google Scholar] [CrossRef]
  347. Strong, P.J.; Xie, S.; Clarke, W.P. Methane as a resource: Can the methanotrophs add value? Environ. Sci. Technol. 2015, 49, 4001–4018. [Google Scholar] [CrossRef]
  348. Wilshusen, J.H.; Hettiaratchi, J.P.; De Visscher, A.; Saint-Fort, R. Methane oxidation and formation of EPS in compost: Effect of oxygen concentration. Environ. Pollut. 2004, 129, 305–314. [Google Scholar] [CrossRef]
  349. Glick, B. Plant growth-promoting bacteria: Mechanisms and applications. Scientifica 2012, 2012, 1–15. Available online: https://www.scielo.br/scielo.php?script=sci_nlinks&pid=S1415-4757201500040040100093&lng=en (accessed on 1 October 2012). [CrossRef] [Green Version]
  350. Etesami, H.; Emami, S.; Alikhani, H.A. Potassium solubilizing bacteria (KSB): Mechanisms, promotion of plant growth, and future prospects: A review. J. Soil Sci. Plant Nutr. 2017, 17, 897–911. [Google Scholar] [CrossRef]
  351. Etesami, H. Bacterial mediated alleviation of heavy metal stress and decreased accumulation of metals in plant tissues: Mechanisms and future prospects. Ecotoxicol. Environ. Saf. 2018, 147, 175–191. [Google Scholar] [CrossRef]
  352. 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]
  353. Etesami, H.; Maheshwari, D.K. Use of plant growth promoting rhizobacteria (PGPRs) with multiple plant growth promoting traits in stress agriculture: Action mechanisms and future prospects. Ecotoxicol. Environ. Saf. 2018, 156, 225–246. [Google Scholar] [CrossRef] [PubMed]
  354. Gupta, P.; Diwan, B. Bacterial exopolysaccharide mediated heavy metal removal: A review on biosynthesis, mechanism and remediation strategies. Biotechnol. Rep. 2017, 13, 58–71. [Google Scholar] [CrossRef]
  355. Vurukonda, S.S.K.P.; Vardharajula, S.; Shrivastava, M.; SkZ, A. Enhancement of drought stress tolerance in crops by plant growth promoting rhizobacteria. Microbiol. Res. 2016, 184, 13–24. [Google Scholar] [CrossRef]
  356. Rolli, E.; Marasco, R.; Vigani, G.; Ettoumi, B.; Mapelli, F.; Deangelis, M.L.; Gandolfi, C.; Casati, E.; Previtali, F.; Gerbino, R. Improved plant resistance to drought is promoted by the root-associated microbiome as a water stress-dependent trait. Environ. Microbiol. 2015, 17, 316–331. [Google Scholar] [CrossRef] [PubMed]
  357. Sandhya, V.; Grover, M.; Reddy, G.; Venkateswarlu, B. Alleviation of drought stress effects in sunflower seedlings by the exopolysaccharides producing Pseudomonas putida strain GAP-P45. Biol. Fertil. Soils 2009, 46, 17–26. [Google Scholar] [CrossRef]
  358. Bhaskar, P.V.; Bhosle, N.B. Bacterial extracellular polymeric substance (EPS): A carrier of heavy metals in the marine food-chain. Environ. Int. 2006, 32, 191–198. [Google Scholar] [CrossRef] [PubMed]
  359. Artyszak, A.; Gozdowski, D. The effect of growth activators and plant growth-promoting rhizobacteria (PGPR) on the soil properties, root yield, and technological quality of sugar beet. Agronomy 2020, 10, 1262. [Google Scholar] [CrossRef]
  360. Mącik, M.; Gryta, A.; Frąc, M. Biofertilizers in agriculture: An overview on concepts, strategies and effects on soil microorganisms. In Advances in Agronomy; Sparks, D.L., Ed.; Academic Press Inc.: Cambridge, MA, USA, 2020; Volume 162, pp. 31–87. [Google Scholar]
  361. Uribe, D.; Sánchez-Nieves, J.; Vanegas, J. Role of microbial biofertilizers in the development of a sustainable agriculture in the Tropics. In Soil Biology and Agriculture in the Tropics; Dion, P., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; pp. 235–250. [Google Scholar]
  362. Mehnaz, S. An overview of globally available bioformulations. In Bioformulations: For Sustainable Agriculture; Arora, N., Mehnaz, S., Balestrini, R., Eds.; Springer: New Delhi, India, 2016; pp. 268–281. [Google Scholar]
  363. García-Fraile, P.; Menéndez, E.; Celador-Lera, L.; Díez-Méndez, A.; Jiménez-Gómez, A.; Marcos-García, M.; Cruz-González, X.A.; Martínez-Hidalgo, P.; Mateos, P.F.; Rivas, R. Bacterial probiotics: A truly green revolution. In Probiotics and Plant Health; Kumar, V., Kumar, M., Sharma, S., Prasad, R., Eds.; Springer: Singapore, 2017; pp. 131–162. [Google Scholar]
  364. Adeleke, R.A.; Raimi, A.R.; Roopnarain, A.; Mokubedi, S.M. Status and prospects of bacterial inoculants for sustainable management of agroecosystems. In Biofertilizers for Sustainable Agriculture and Environment; Giri, B., Prasad, R., Wu, Q.-S., Varma, A., Eds.; Springer: Cham, Switzerland, 2019; pp. 137–172. [Google Scholar]
  365. Aloo, B.N.; Makumba, B.A.; Mbega, E.R. Plant growth promoting rhizobacterial biofertilizers for sustainable crop production: The past, present, and future. Preprints 2020, 2020090650. [Google Scholar] [CrossRef]
  366. Dal Cortivo, C.; Ferrari, M.; Visioli, G.; Lauro, M.; Fornasier, F.; Barion, G.; Panozzo, A.; Vamerali, T. Effects of seed-applied biofertilizers on rhizosphere biodiversity and growth of common wheat (Triticum aestivum L.) in the field. Front. Plant Sci. 2020, 11, 72. [Google Scholar] [CrossRef] [Green Version]
  367. Mahajan, A.; Gupta, R.D. Bio-fertilizers: Their kinds and requirement in India. In Integrated Nutrient Management (INM) in a Sustainable Rice—Wheat Cropping System; Mahajan, A., Gupta, R.D., Eds.; Springer: Dordrecht, The Netherlands, 2009; pp. 75–100. [Google Scholar]
  368. Malathi, S.; Sriramulu, M. Laboratory efficacy of biotic insecticides against lepidopterous pests fed on treated cabbage leaves. Shashpa 2000, 7, 63–66. [Google Scholar]
  369. Apurva, K.; Singh, C.; Kumar, V.; Jha, V.B. Bacterial biopesticides and their use in agricultural production. In Biofertilizers and Biopesticides in Sustainable Agriculture; Apple Academic Press: London, UK, 2019; pp. 23–42. [Google Scholar]
  370. Schwenk, V.; Riegg, J.; Lacroix, M.; Märtlbauer, E.; Jessberger, N. Enteropathogenic potential of Bacillus thuringiensis isolates from soil, animals, food and biopesticides. Foods 2020, 9, 1484. [Google Scholar] [CrossRef]
  371. Cardinale, M.; Ratering, S.; Suarez, C.; Zapata Montoya, A.M.; Geissler-Plaum, R.; Schnell, S. Paradox of plant growth promotion potential of rhizobacteria and their actual promotion effect on growth of barley (Hordeum vulgare L.) under salt stress. Microbiol. Res. 2015, 181, 22–32. [Google Scholar] [CrossRef] [PubMed]
  372. Bashan, Y.; de-Bashan, L.E.; Prabhu, S.R.; Hernandez, J.P. Advances in plant growth-promoting bacterial inoculant technology: Formulations and practical perspectives (1998–2013). Plant Soil 2014, 378, 1–33. [Google Scholar] [CrossRef] [Green Version]
  373. Peiffer, J.A.; Spor, O.; Koren, A.; Jin, Z.; Tringe, S.G.; Dangl, J.L.; Buckler, E.S.; Ley, R.E. Diversity and heritability of the maize rhizosphere microbiome under field conditions. Proc. Natl. Acad. Sci. USA 2013, 110, 6548–6553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  374. Bender, S.F.; Wagg, C.; van der Heijden, M.G. An underground revolution: Biodiversity and soil ecological engineering for agricultural sustainability. Trends Ecol. Evolut. 2016, 31, 440–452. [Google Scholar] [CrossRef]
  375. Khatoon, Z.; Huang, S.; Rafique, M.; Fakhar, A.; Kamran, M.A.; Santoyo, G. Unlocking the potential of plant growth-promoting rhizobacteria on soil health and the sustainability of agricultural systems. J. Environ. Manag. 2020, 273, 111118. [Google Scholar] [CrossRef]
  376. Kumar, S.S.; Kadier, A.; Malyan, S.K.; Ahmad, A.; Bishnoi, N.R. Phytoremediation and rhizoremediation: Uptake, mobilization and sequestration of heavy metals by plants. In Plant-Microbe Interactions in Agro-Ecological Perspectives; Springer: New Delhi, India, 2017; pp. 367–394. [Google Scholar]
  377. Kour, D.; Rana, K.L.; Yadav, A.N.; Yadav, N.; Kumar, M.; Kumar, V.; Vyas, P.; Dhaliwal, H.S.; Saxena, A.K. Microbial biofertilizers: Bioresources and eco-friendly technologies for agricultural and environmental sustainability. Biocatal. Agric. Biotechnol. 2020, 23, 101487. [Google Scholar] [CrossRef]
  378. Raj, S.N.; Shetty, H.S.; Reddy, M.S. Plant growth promoting rhizobacteria: Potential green alternative for plant productivity. In PGPR: Biocontrol and Biofertilization; Springer: Dordrecht, The Netherlands, 2005; pp. 197–216. [Google Scholar]
  379. Son, J.S.; Sumayo, M.; Hwang, Y.J.; Kim, B.S.; Ghim, S.Y. Screening of plant growth-promoting rhizobacteria as elicitor of systemic resistance against gray leaf spot disease in pepper. Appl. Soil Ecol. 2014, 73, 1–8. [Google Scholar] [CrossRef]
  380. Liu, J.; Dang, P.; Gao, Y.; Zhu, H.; Zhu, H.; Zhao, F.; Zhao, Z. Effects of tree species and soil properties on the composition and diversity of the soil bacterial community following afforestation. For. Ecol. Manag. 2018, 427, 342–349. [Google Scholar] [CrossRef]
  381. Bhattacharjee, R.B.; Singh, A.; Mukhopadhyay, S.N. Use of nitrogen-fixing bacteria as biofertiliser for non-legumes: Prospects and challenges. Appl. Microbiol. Biotechnol. 2008, 80, 199–209. [Google Scholar] [CrossRef]
  382. Chen, J.H. The combined use of chemical and organic fertilizers and/or biofertilizer for crop growth and soil fertility. In Proceedings of the International Workshop on Sustained Management of the Soilrhizosphere System for Efficient Crop Production and Fertilizer Use, Land Development Department, Bangkok, Thailand, 16–20 October 2006; Volume 16, p. 20. [Google Scholar]
  383. Kumawat, N.; Kumar, R.; Kumar, S.; Meena, V.S. Nutrient solubilizing microbes (NSMs): Its role in sustainable crop production. In Agriculturally Important Microbes for Sustainable Agriculture; Springer: Singapore, 2017; pp. 25–61. [Google Scholar]
  384. Ortiz-Castro, R.; López-Bucio, J. Phytostimulation and root architectural responses to quorum-sensing signals and related molecules from rhizobacteria. Plant Sci. 2019, 284, 135–142. [Google Scholar] [CrossRef]
  385. Meena, M.; Swapnil, P.; Divyanshu, K.; Kumar, S.; Tripathi, Y.N.; Zehra, A.; Marwal, A.; Upadhyay, R.S. PGPR-mediated induction of systemic resistance and physiochemical alterations in plants against the pathogens: Current perspectives. J. Basic Microbiol. 2020, 60, 828–861. [Google Scholar] [PubMed]
  386. Zandi, P.; Basu, S.K. Role of plant growth-promoting rhizobacteria (PGPR) as biofertilizers in stabilizing agricultural ecosystems. Org. Farming Sustain. Agric. 2016, 71–87. [Google Scholar] [CrossRef]
  387. Mahanty, T.; Bhattacharjee, S.; Goswami, M.; Bhattacharyya, P.; Das, B.; Ghosh, A.; Tribedi, P. Biofertilizers: A potential approach for sustainable agriculture development. Environ. Sci. Pollut. Res. 2017, 24, 3315–3335. [Google Scholar] [CrossRef] [PubMed]
  388. Aka, R.J.N.; Babalola, O.O. Effect of bacterial inoculation of strains of Pseudomonas aeruginosa, Alcaligenes feacalis and Bacillus subtilis on germination, growth and heavy metal (Cd, Cr, and Ni) uptake of Brassica juncea. Int. J. Phytoremediat. 2016, 18, 200–209. [Google Scholar]
  389. Dangi, S.R.; Tirado-Corbalá, R.; Gerik, J.; Hanson, B.D. Effect of long-term continuous fumigation on soil microbial communities. Agronomy 2017, 7, 37. [Google Scholar] [CrossRef] [Green Version]
  390. Dupuy, L.X.; Mimault, M.; Patko, D.; Ladmiral, V.; Ameduri, B.; MacDonald, M.P.; Ptashnyk, M. Micromechanics of root development in soil. Curr. Opin. Genet. Dev. 2018, 51, 18–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  391. Gundel, P.E.; Omacini, M.; Sadras, V.O.; Ghersa, C.M. The interplay between the effectiveness of the grass-endophyte mutualism and the genetic variability o the host plant. Evol. Appl. 2010, 3, 538–546. [Google Scholar] [CrossRef] [PubMed]
Figure 1. An overview of the mechanisms employed by rhizobacteria for plant growth promotion, biocontrol of plant pests, and bioremediation of contaminated soils. BNF, biological nitrogen fixation; ISR, induced systemic resistance; HC, hydrogen cyanide; ACC, 1-aminocyclopropane, 1-carboxylic acid.
Figure 1. An overview of the mechanisms employed by rhizobacteria for plant growth promotion, biocontrol of plant pests, and bioremediation of contaminated soils. BNF, biological nitrogen fixation; ISR, induced systemic resistance; HC, hydrogen cyanide; ACC, 1-aminocyclopropane, 1-carboxylic acid.
Ijms 22 10529 g001
Figure 2. Consequences of stresses on plants without PGPR (left side) and with PGPR (right side). The bottom part shows the most studied microbial candidates involved in plant growth promotion, biocontrol, and bioremediation processes.
Figure 2. Consequences of stresses on plants without PGPR (left side) and with PGPR (right side). The bottom part shows the most studied microbial candidates involved in plant growth promotion, biocontrol, and bioremediation processes.
Ijms 22 10529 g002
Figure 3. Comparison of pros and cons of rhizobacterial application for sustainable agriculture with possible future directions [359,360,371,372,373,374].
Figure 3. Comparison of pros and cons of rhizobacterial application for sustainable agriculture with possible future directions [359,360,371,372,373,374].
Ijms 22 10529 g003
Table 1. Summary of studies on rhizobacterial mechanisms involved in plant growth promotion.
Table 1. Summary of studies on rhizobacterial mechanisms involved in plant growth promotion.
PGPR CandidateHost PlantGrowth ConditionProposed Mechanism(s)Plant ResponseReferences
Rhizobium tropiciBeanPotN2 fixation and IAA productionIncrease in number of nodules and plant dry weight[45]
Pseudomonas stutzeri A1501Maize FieldN2 fixation and ACC deaminase production Better growth and more nitrogen accumulation [46]
Rhizobium etliBeanPotNitrogenase activityA 16% increase in the N2 content of the resultant seeds and a 25–30% increase in leaf nitrogen[47]
Pseudomonas fluorescens BSP53Blackcurrant cuttingsPot and fieldIAA overproduction Root development[45]
Streptomyces sp. NEAU-S7GS2SoybeanPotIAA production and inorganic phosphate solubilizationA 77% increase in plant growth and a 38% decrease in sclerotinia stem rot of soybean [48]
Halobacillus dabanensis strain SB-26 and Halobacillus sp. GSP 34RicePotIAA production and nitrogen fixationEnhanced plant height and yield[49]
Fluorescent Pseudomonas sp. PF17SunflowerPotIAA production and phosphate solubilizationPromoted overall plant growth[50]
wild-type P. putida GR12-2Mung beanPotIAA-overproducing and pH-lowering mechanismRoot and shoot development[51]
R. leguminosarum bv. ViciaeAlfalfaPotACC deaminase Increase in nodule number and biomass [44]
Bacillus sp. HWP47Wheat Pot ACC deaminase activity and chelation mechanismCaused 51.46% increase in root dry weight (RDW) [52]
Rhizobacteria sp. Solanum tuberosumPotACC deaminase activity and phytohormone productionROS antioxidant enzyme expression and improved photosynthesis[53]
PseudomonadsTobaccoFieldFerric siderophore productionAdventitious root development[54]
Pseudomonas fluorescens C7Arabidopsis thalianaPot and Field Trials Fe-pyoverdine complexIncrease in iron inside plant tissues and hence enhanced plant growth [55]
Bacillus mucilaginosus HQ013329GarlicPot Various extracellular polymers are produced (primarily proteins and polysaccharides)Promoted garlic plant growth characteristics and avoided environmental pollution hazards caused by heavy application of potassium fertilizers[56]
Bacillus circulansPepper
Potato
Pot Acidolysis of the surrounding area of microorganismHighest total yield of pepper and potato tuber yield per plant and average tuber weight[57]
B. mucilaginosus strain RCBC13Tomato Pot Production of α-ketogluconic acid125% increase in biomass, whereas K and P uptake were more than 150% in tomato [58]
Pseudomonas aeruginosa PF23EPS+SunflowerPot and field Salicylic acid productionImproved growth of crop in salinized soil[59]
Bradyrhizobium japonicum USDA 110 with salt-tolerant Pseudomonas putida TSAU1SoybeanHydroponicFormation of nodulesIncreased plant growth and nitrogen and phosphorus uptake [60]
Bacillus amyloliquefaciens subsp. plantarum UCMB5113Arabidopsis thalianaLabImproved root growth and configurationIncreased plant growth and reduced disease effect[61]
Bacillus firmus SW5Soybean PotImproved root structureIncreased biomass, nutrient uptake, and gas exchange parameters[62]
Azospirillum brasilenseMaizePot and fieldSiderophore productionIncreased biomass and improved plant health[63]
Acidovorax delafieldiiRicePot Siderophore production and tricalcium phosphate solubilizationEnhanced biomass, grain yield, and NPK uptake[64]
Brevibacillus brevis GZDF3Pinellia ternata (an important herb in traditional Chinese medicine)Lab Siderophore productionHelpful in the development of new biological agent[65]
Bacillus subtilis MF497646 and Pseudomonas koreensis MG209738MaizePot and fieldSiderophore productionIncreased antioxidant activities and yield parameters[66]
Table 2. Summary of studies on rhizobacterial mechanisms involved in biocontrol.
Table 2. Summary of studies on rhizobacterial mechanisms involved in biocontrol.
Biocontrol AgentPlant PathogenHost PlantProposed Mechanism(s)Reference
Pseudomonas fluorescensFusarium culmorumRyeFe(III)-chelating compounds (including siderophores)[159]
Acinetobacter, Pseudomonas, Staphylococcus, Bacillus, Enterobacter, Pantoea, AlcaligenesFusarium oxysporum, Alternaria alternate, F. culmorum, F. solani, Botrytis cinerea, Pythium ultimum, Phytophthora cryptogeaWheatAntagonism and growth promotion[155]
Bacillus sp. L324-92Gaeumannomyces graminis var tritici, Rhizoctonia root rot, R. solani AG8, Pythium root rot, Pythium irregulare P. ultimum.WheatNot specified[90]
Bacillus sp., Pseudomonas fluorescensR. oryzae, P. ultimum, G. graminis, R. solaniWheatNot specified[160]
Pseudomonas fluorescensMicroconidium nivale/ Fusarium nivaleWheatGrowth promotion, siderophore production, in vitro antibiosis[161]
Bacillus subtilis and B. cereusTake all (G. graminis var tritici) Rhizoctonia root rot (R. solaniAG8)WheatGrowth promotion[162]
Bacillus subtilis CE1Fusarium verticillioidesMaizeNot specified[163]
Pseudomonas chlororaphisMacrophomina phaseolina (charcoal rot of sorghum)SorghumExtracellular antibiotics, production of volatiles, siderophores, effective root colonization[164]
Pseudomonas fluorescens MKB 100 and MKB 249, P. frederiksbergensis 202, Pseudomonas spp. MKB 158Fusarium culmorumWheat and barleyInduced resistance, antibiotic production, pathogenesis-related proteins (induced resistance) in wheat[165]
Table 3. Summary of studies on the effects of rhizobacterial strains involved in the bioremediation of polyaromatic hydrocarbon-contaminated soils.
Table 3. Summary of studies on the effects of rhizobacterial strains involved in the bioremediation of polyaromatic hydrocarbon-contaminated soils.
PGPR Candidate PollutantsPollutant
Concentration
Remediation Efficiency (%)References
Azoarcus sp., Escherichia coli L.Benzoates 1 mM90.0[248]
Pseudomonas fluorescens L. and Pseudomonas putida L. Benzoates and related substances1 mM60.0[227]
Rhizobium leguminosarum L.Chrysene500 mg kg-128.0[249]
Azospirillum lipoferum L. and Azospirillum brasilense L. Crude oil1% v/v57.0[228]
Acinetobacter lwoffii L., Bacillus subtilis L., and Pseudomonas aeruginosa L.Crude oil1% v/v89.0[250]
Alcaligenes faecalis L., Citrobacter murliniae L., Dietzia papillomatosis L., Nocardioides deserti L.Crude oil1174 mg/L90.0[251]
Enterobacter sp. MN17Crude oil5000 mg kg−163.0[229]
Enterobacter cloacaeCrude oil2000 ppm54.0[230,231]
Pseudomonas cepacia L. and Arthrobacter sp.2,4-Dichlorphenoxyessigsäure (kurz 2,4-D) 5% v/v80.0[236,252]
Rhizopus arrhizus L. and Pseudomonas aeruginosa L.Dichlorodiphenyltrichloroethane (DDT)1 mg L−134.0[237,253]
Chroococcus sp. and Synechocystis sp.Linurin1 mg L−1>98[254]
Pseudomonas sp., Pseudomonas aeruginosa L., Pseudomonas stutzeri L.Parathion3800 µg/mL>49[255]
Archaea sp., Acinetobacter sp., Achromobacter sp., Alcaligenes sp., Aspergillus sp., Arthrobacter sp., Azotobacter sp., Bacillus cereus sp., Flavobacterium sp., Neurospora sp.Phenolic compounds13.0 mg L−199.0[256,257]
Leptolyngbya sp., Penicillium sp., Candida tropicalis L., Debaryomyces subglobosus L., Pseudomonas aeruginosa L., Pseudomonas putida L., and Trichosporon cutaneum L.Phenolic compounds150.0 mg L−199.0[258]
Pseudomonas sp.Phenanthrene200 mg kg-1>62[259]
Actinobacteria sp.Phenanthrene100 mg L−150.0[260]
Pseudomonas putida L.Phenanthrene100 mg kg-189.0[261]
Pseudomonas sp.Phenanthrene100 mg kg−173.0[228]
Pseudomonas sp.Phenanthrene100 mg kg-150.0[239]
Actinobacteria sp., Caulobacterales sp., Rhizobiales sp., Rhodococcus sp., Xanthomonadales sp.Phenanthrene1260 mg kg−148.0[241]
Pseudomonas cepacia L.2,4,5-Trichlorophenoxyacetic acid (2,4,5-T)200 µM27.0[262]
Alcaligene sp., Bacillus sp., Citrobacter sp., Corynebacterium sp., Flavobacterium sp., and Pseudomonas sp.Surfactants200 mg L−150.0[263]
Bacillus sp., Candia sp., Pseudomonas sp.Surfactants300 mg L-1 [264]
Burkholderia multivorans L.Surfactants150 mg L-141.0[265]
Pseudomonas sp.Surfactants2% v/v94.0[266]
Table 4. Summary of studies on rhizobacterial mechanisms involved in the bioremediation of heavy metal-contaminated soils.
Table 4. Summary of studies on rhizobacterial mechanisms involved in the bioremediation of heavy metal-contaminated soils.
Heavy Metals Microbes Host Plant(s)Proposed
Mechanism(s)
References
AsPenicillium aculeatum PDR-4SunflowerPhosphatase, siderophore [267]
CdBacillus sp., Endophytic sp., Rahnella sp., Enterobacter sp., Klebsiella spp., Pseudomonas sp., Arthrobacter sp. TISTR 2220, Agromyces AR33,
Brevundimonas Kro13, Ralstonia sp. TISTR 2219, Streptomyces AR17, Variovorax paradoxus L.
Wheat, Mustard, Soybean, Mung bean, Willow, Amaranthus, Sunflower, Basil, Goat willowExopolysaccharide production, siderophores [231,268,269,270,271,272,273,274]
CdBacillus badius L., Cronobacter muytjensii L., Drepanomonas revolute L., Euplotes sp., Pseudomonas azotoformans L., Uronema nigricans L.Biosorption[275,276,277,278]
CdAlcaligenes sp., Bacillus sp., Pseudomonas sp., Rhodococcus sp., and Variovorax sp. Canola, Sunflower ACC deaminase activity and reduction in ethylene levels[68,279]
CoEnterobacter ludwigii L.SunflowerExopolysaccharide production[280]
CrAerococcus sp., Bacillus sp., Endophytic sp., Rhizobacteria sp., Staphylococcus sp., Ochrobactrum intermedium L., Pseudomonas aeruginosa L., Pseudomonas
pseudoalcaligenes L.
Chickpea, Mesquite, Prosopis, Maize, Sorghum, Wheat, Mung bean, Soybean, SunflowerExopolysaccharide production[271,281,282,283,284,285]
CrBacillus circulans L., Bacillus megaterium L., Cronobacter muytjensii L., Saccharomyces cerevisiae L. Biosorption[276,286,287]
Cu Pseudomonas sp., Rhizobacteria sp., Bacillus subtilis L., Brevibacterium halotolerans L., Bacillus
pumilus L., Brevibacterium casei MH8a, Paenibacillus polymyxa L., Pseudomonas pseudoalcaligenes L.
Maize, Sunflower, Sorghum, Mustard, Trailing daisy Exopolysaccharide production, siderophores[282,288,289,290,291,292]
CuAspergillus flavus L., Cronobacter muytjensii L., Pseudomonas azotoformans L., Drepanomonas revolute L., Euplotes sp., Uronema nigricans L., Pichia guilliermondii L. Maize, Wheat, SorghumBiosorption[275,276,277,293,294]
FePseudomonas sp., Psychrobacter sp.
Castor beanExopolysaccharide production, siderophores[295]
NiMicrococcus sp., Pseudomonas sp., Psychrobacter sp., Bacillus subtilis SJ-101, Bacillus pumilus L., Rhodococcus erythropolis X79289, Rhodococcus globerulus X80619 Mustard, Alyssum sp., Alpine pennycress, Castor bean, Mung bean, Wheat, Soybean Exopolysaccharide production, siderophores[271,295,296,297,298,299]
NiKluyvera ascorbate L. (SUD165), Pseudomonas putida L. (UW4)Tomato, Canola, Tobacco ACC deaminase activity and reduction in ethylene levels[288,300]
PbEnterobacter sp., Klebsiella sp., Enterobacter ludwigii L., AgromycesAR33, Bacillus subtilis L., Brevibacterium halotolerans L., Bacillus megaterium HKP-1, Bacillus
pumilus L., Penicillium aculeatum PDR-4, Pseudomonas pseudoalcaligenes L., Streptomyces AR17
Sunflower, Mustard, Maize, Sorghum, Goat willowExopolysaccharide production, siderophores[91,231,272,274,
279,280,282,301]
PbAspergillus niger L., Bacillus xiamenensis L., Pseudomonas azotoformans L. Biosorption[277,293,302]
ZnEnterobacter sp., Klebsiella sp., Pseudomonas sp., Rahnella sp., Rhizobacteria sp., Psychrobacter sp.
Agromyces AR33, Bacillus megaterium HKP-1, Bacillus subtilis L., Brevibacterium halotolerans L., BrevibacteriumcaseiMH8a, Bacilluspumilus L., Enterobacter ludwigii L., Streptomyces AR17
Mustard, Sunflower, Castor bean, Maize, Sorghum, Goat willow,Exopolysaccharide production, siderophores[272,279,280,282,291,295,303]
ZnCronobacter muytjensii L., Drepanomonas revolute L., Euplotes sp., Uronema nigricans L. Biosorption[275,276]
Table 5. Pros of biofertilizers and their composition.
Table 5. Pros of biofertilizers and their composition.
Pro’s TypesProduct (Bacterial Composition)References
Nitrogen fixersAgriLife NitroFix (A. chroococcum, A. vinelandii, A. diazotrophicus, A. lipoferum, R. japonicum), Ajay Azospirillum (Azospirillum sp.), Azofer (A. brasilense), Azo-N (A. brasilense + A. lipoferum), Azo-N Plus (A. brasilense + A. lipoferum + A. chroococcum), Azoter (A. chroococcum, A. brasilense, B. megaterium), Azotobacterin (A. brasilense B-4485), BactoFil A10 (B. Megaterium + A. brasilense, A. vinelandii), BactoFil Soya (B. japonicum), BiAgro 10 (B. japonicum), Bioboots (Bradyrhizobium sp. + D. acidovorans), Biofix (Rhizobia), BioGro (C. freundii, K. pneumoniae, P. fluorescens), Bio-N (Azospirillum spp.), Cell-Tech (Rhizobia), Custom N2 (P. polymyxa), Dimargon (A. chroococcum), Legume Fix (B. japonicum + Rhizobium sp.), Mamezo (Rhizobia), Nitragin Gold (Rhizobia), Nitrasec (Rhizobium sp.), Nitrofix (Azospirillum sp.), Nodulator (B. Japonicum), Nodulator PRO (B. subtilis + B. Japonicum), Nodulest 10 (B. japonicum), Nodumax (Bradyrhizobium spp.), Phylazonit M (A. chroococcum + B. megaterium), Rhizofer (R. etli), Rhizosum Aqua (Azospirillum sp.), Rhizosum N (A. vinelandii + R. irregularis), Rizo-Liq (Bradyrhizobium sp., + M. ciceri, + Rhizobium spp.), Rizo-Liq Top (B. japonicum), Symbion N (Azospirillum sp. + Rhizobium sp. + Acetobacter sp. + Azotobacter sp.), TagTeam (P. bilaii + Rhizobia), TwinN and TripleN (Azorhizobium sp. + Azoarcus sp. + Azospirillum sp.), Zadspirillum (A. brasilense), [11,359,360,361,362,363,364,365,366]
Nutrient solubilizersBio Phos (B. megaterium), Biozink (PGPR consortia), CataPult (Bacillus spp. + G. intraradices), CBF (B. mucilaginosus, + B. subtilis), Fosforina (P. Fluorescens), K Sol B (F. aurantia), P Sol B (P. striata, + B. polymyxa, B. megaterium), Phosphobacterin (B. megaterium), Rhizosum K (F. aurantia), Rhizosum PK (B. megaterium, + F. aurantia, + R. irregularis), Symbion van Plus (B. megaterium), Zn Sol B (T. thiooxidans), [360,362,363,366,367]
BiopesticidesBiobit, Dipel, and Delfin (Bacillus thuringiensis var. kurstaki), Certan (Bacillus thuringiensis var. aizawai), Acrobe, Skectal, Vectobac (Bacillus thuringiensis var. israelensis), Trident, Novodor (Bacillus thuringiensis var. tenebrionis), Ciba-Foil, Agree, Cutlass (Bacillus thuringiensis var. conjugates), MVP, M-Trak (Pseudomonas fluorescens (Bt toxin), Doom (Bacillus papilliae), Invade (Serratia entomophila)[368,369,370]
Other biofertilizersAmase (P. azotoformans), Bioativo (PGPR consortia), EVL Coating (PGPR consortia), Biotilis (B. subtilis), Cedomon (P. chlororaphis), Cedress (P. chlororaphis)[360,362]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Saeed, Q.; Xiukang, W.; Haider, F.U.; Kučerik, J.; Mumtaz, M.Z.; Holatko, J.; Naseem, M.; Kintl, A.; Ejaz, M.; Naveed, M.; et al. Rhizosphere Bacteria in Plant Growth Promotion, Biocontrol, and Bioremediation of Contaminated Sites: A Comprehensive Review of Effects and Mechanisms. Int. J. Mol. Sci. 2021, 22, 10529. https://doi.org/10.3390/ijms221910529

AMA Style

Saeed Q, Xiukang W, Haider FU, Kučerik J, Mumtaz MZ, Holatko J, Naseem M, Kintl A, Ejaz M, Naveed M, et al. Rhizosphere Bacteria in Plant Growth Promotion, Biocontrol, and Bioremediation of Contaminated Sites: A Comprehensive Review of Effects and Mechanisms. International Journal of Molecular Sciences. 2021; 22(19):10529. https://doi.org/10.3390/ijms221910529

Chicago/Turabian Style

Saeed, Qudsia, Wang Xiukang, Fasih Ullah Haider, Jiří Kučerik, Muhammad Zahid Mumtaz, Jiri Holatko, Munaza Naseem, Antonin Kintl, Mukkaram Ejaz, Muhammad Naveed, and et al. 2021. "Rhizosphere Bacteria in Plant Growth Promotion, Biocontrol, and Bioremediation of Contaminated Sites: A Comprehensive Review of Effects and Mechanisms" International Journal of Molecular Sciences 22, no. 19: 10529. https://doi.org/10.3390/ijms221910529

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