Open access peer-reviewed chapter

The Plant-Rhizobial Symbiotic Interactions Provide Benefits to the Host beyond Nitrogen Fixation That Promote Plant Growth and Productivity

Written By

Jemaneh Z. Habtewold and Ravinder K. Goyal

Submitted: 07 December 2022 Reviewed: 12 December 2022 Published: 01 February 2023

DOI: 10.5772/intechopen.109464

From the Edited Volume

Symbiosis in Nature

Edited by Everlon Cid Rigobelo

Chapter metrics overview

157 Chapter Downloads

View Full Metrics

Abstract

Rhizobial symbiotic interactions are known for nitrogen fixation, providing commercial crops and other plants with self-sufficiency in nitrogen requirements. An enormous contribution from nitrogen fixation is vital to the global nitrogen cycle. The symbiotic nitrogen reduces the carbon footprint of crop cultivation, which underlines its importance in agricultural sustainability. Extensive research efforts have been made to understand the symbiotic relationship at molecular, physiological, and ecological levels. This led to the isolation and modification of symbiotic strains for enhanced nitrogen efficiency. During the evaluation of strains for nitrogen fixation in exchange for supporting the bacterium in terms of space and resources, it has been observed that the accrued benefits to the host plants extend well beyond the nitrogen fixation. The symbiotic interaction has been advantageous to the host for better growth and development, tolerating a stressful environment, and even keeping the pathogenic microbial enemies at bay. Additionally, it enabled the availability of the mineral nutrients, which otherwise were inaccessible to the host. In this chapter, we bring together the information with a focus on the role of rhizobial symbiotic interactions that promote plant growth and productivity through phytohormone synthesis, by facilitating the availability of mineral nutrients, and by improving the plant tolerance to sub-optimal growth conditions.

Keywords

  • rhizobia
  • legumes
  • symbiosis
  • growth
  • plant hormones
  • environmental adaptability
  • stress tolerance
  • heavy metals

1. Introduction

Nitrogen (N) is an essential macro-nutrient that is needed for plant growth. Plants are unable to use N in its gaseous state, which is freely and abundantly available in the atmosphere. The atmospheric nitrogen (N2) must first be converted chemically or naturally into NO3 or NH4+ before plants can assimilate it. Nature has given many microorganisms the ability to convert N2 into a usable form. In microorganisms, the metabolic conversion of N2 takes place either in their free-living state or through symbiosis with the host plant. Legumes have the unique ability to form symbiotic interactions with rhizobia, which helps meet the plant’s N requirement. The symbiotic interaction is very complex, beginning with a chemical conversation between the rhizobia and host plant [1]. The host plant synthesizes specific flavonoids that are perceived through NodD genes expressed by a symbiont. Subsequently, through an array of events, a mutualistic relationship is established in which the bacterium imports carbon metabolites from the host in exchange for a usable form of N, or fixed N. Biological N fixation has been studied extensively to understand the evolutionary logistics, chemistry, regulation, and ecological perspectives [2]. The complexity of symbiotic mechanisms indicates the coevolution of a process that led to the development of N-fixing nodules [3]. The question of whether plants selected bacteria, whether bacteria influenced evolution, or whether legumes and rhizobia coevolved is being debated. Unlike mitochondrial and chloroplast evolution, the loss of genes from the rhizobia to the host nucleus, which would have compromised bacterial survival as a free-living organism, did not occur [4]. Nonetheless, the symbiotic arrangement has evolved over time into a secure partnership.

Symbiotically fixed N plays a key role in agricultural sustainability, especially when many of the agricultural practices, including N supplementation, that were instrumental in enhancing productivity and food security may not have been in tune with safeguarding the environment and preserving our ecosystems. Legumes are high in protein content and contribute roughly one-quarter of total grain production; their N self-sufficiency and increasing the N content of soil for subsequent crops have a significant impact on reducing the need for N-fertilizers [5]. Easing the pressure on chemical-N demand is of environmental significance because the majority of the N in N-fertilizers becomes a source of aquatic system pollution [6] or a substrate for nitrous oxide (N2O) gas in the environment through denitrification. The enormous carbon footprint of fertilizer synthesis and subsequent N-pollution endangers the sustainability of agriculture and the environment. This has generated a significant interest in improving the efficiency of both symbiotic and biological N fixation in legumes and non-legumes, respectively. Many strategies have been employed to improve the symbiotic productivity notwithstanding the challenges of competitiveness of the native strains [7, 8].

Nodulation and N fixation may have been an important factor in the diversification of legumes, enabling them to be one of the largest families of widely distributed plant species [3]. The symbionts developed the capability to synthesize plant hormones that are critical to the growth and development of plants [9]. The biosynthetic pathways of many of the hormones are long and complex, requiring dedicated genetic and cellular machinery. Their involvement in the regulation of nodule development [10] underscores the interdependence of symbiotic partners. The symbiotic interactions are known to enhance plant growth and biomass and increase the availability of micro-elements like iron (Fe) and phosphorus (P) [11]. They enable the host plants to tolerate stressful growth conditions [12], which increases the plant survivability in adverse environmental regimes. It is likely that the enhanced environmental outreach has played a role in the diversification of symbiotic hosts. Although N fixation could be a primary reason for symbiotic evolution, it is not known if the additional beneficial activities were a contributing factor at the early stage. This chapter discusses the benefits of legume-rhizobia symbiosis in plant growth and development, in addition to N fixation, through the production of hormones and increased nutrient availability. The role of symbiosis in stress tolerance and plant defense against pathogens and how this function of symbionts can be improved through molecular tools are also highlighted.

Advertisement

2. Plant growth and development

Plant growth and development, which is greatly influenced by the plant’s capacity to carry out photosynthetic activity in favorable and adverse regimes of nutrient availability and environmental conditions, determines the plant productivity. The symbiotic relationship with rhizobia is known to promote plant growth characteristics in legume crops (ref. in [13]). Although N supply is an important factor in plant growth, numerous studies have shown that rhizobia promote plant growth by producing plant hormones. These phytohormones have a profound effect on plant cellular processes and play a critical role in plant development [14]. The symbiotic interactions may act at molecular, physiological, and cellular levels to provide such benefits [15]. Several hormones, such as jasmonates, brassinosteroids, salicylic acid, nitric oxide, strigolactones, etc., that significantly influence plant signaling or other cellular responses are not discussed in this chapter as their impact on the symbiotic process or nodulation has previously been discussed [10]. The other phytohormones, which have been traditionally recognized for their role in dictating plant growth and development, are discussed. Their representative forms are depicted in Figure 1.

Figure 1.

The representative structure of some plant hormones. Source: PubChem.

Auxins: These are a group of phytohormones that regulate processes like growth, seed germination, cell division, elongation and differentiation, root development, and stress response [16]. Plant-associated microbes, including Rhizobium, Microbacterium, Mycobacterium, Bradyrhizobium and Sphingomonas are among the active producers of indole-3-acetic acid (IAA), which is the major form of auxins in plants [17, 18]. Increased production of IAA by Bradyrhizobium japonicum was associated with enhanced soybean yield and survival [19]. A similar stimulatory effect was observed in a different host-rhizobial symbiosis. The IAA-overproducing Ensifer meliloti (formerly known as Sinorhizobium meliloti) strain was linked with the increased plant biomass and photosynthetic activity in Medicago sativa. There was higher nitrogenase activity, which possibly increased the demand for carbohydrates to balance the C/N ratio for an overall positive effect on plant growth [20]. The overproduction of IAA through genetic modification of a Rhizobium leguminosarum strain also affected nodule development and increased N fixation [21]. The positive effects of IAA-producing rhizobia on plant growth and productivity have been observed in non-legumes. The inoculation of rice (Oryza sativa L.) with Rhizobium bacteria led to about a 20% increase in the dry weight of rice along with enhancements in plant growth characteristics [22, 23]. Many of the symbiotic bacteria have developed genetic infrastructure to synthesize IAA, which clearly promotes the interests of the host plant. What could be the benefit of synthesizing IAA for the symbiotic bacteria other than that the host is healthy? Studies indicate that IAA is involved in nodule development [24, 25]. Nodulated roots are found to be enriched in IAA compared to non-nodulated roots [26]. The exact role of IAA in nodule formation is not known, but the hormone may act as a signal molecule that promotes host-bacterial interaction and attachment of the latter to the host roots [27, 28]. This is supported by an earlier observation that bacteria with a higher level of IAA are found closer to the root rhizosphere than the non-producers that are distantly distributed in the soil [29]. Trehalose (α-D-glucopyranosyl-1,1-α-D-glucopyranoside) synthesis and degradation are tightly regulated during nodule development, suggesting its involvement in nodule formation [30, 31]. The inoculation of soybean with IAA-producing Rhizobium upregulated the expression of the otsA gene along with an increase in nodule biomass [32]. The gene encodes trehalose-6-phosphate synthase, which synthesizes the precursor of trehalose. Similarly, the IAA-overproducing RD64 strain of E. meliloti accumulated more trehalose in the nodules of Medicago truncatula [33]. The mutualistic benefit of IAA production by rhizobia is also observed through the regulation of N fixation genes by the hormone. There was an up-regulation of the N fixation apparatus by the IAA-overproducing endophytes [34]. A higher level of IAA in the bacteroids was associated with more abundance of transcripts of a N fixation gene, nifH [20].

Cytokinins (CKs): They are another group of plant hormones that regulate various aspects of plant growth and differentiation. They are implicated in cell division, seed germination, the formation of roots and shoots, the development of reproductive organs, apical dominance, senescence, stress tolerance, and more, covering almost all aspects of the plant life cycle. Earlier CKs were thought to be synthesized only in roots, but now it has been demonstrated that many plant parts synthesize the hormone. In legume roots, they are involved in nodule organogenesis. The exogenous application of CKs induced nodule-like structures and mutation in CK receptors, histidine kinase in Lotus japonicus hit1-1 (lhk1) and M. truncatula Cytokinin Response 1 (CRE1) inhibited the development of such structures (reviewed in ref. [10, 35]). The M. truncatula genome contains many ATP-binding cassette (ABCG) transporter proteins that enable the movement of hormones to an extracellular environment. One such transporter, MtABCG56, showed root and nodule specific expression, and the transcript level of the gene increased significantly after inoculation with E. meliloti, pointing to its involvement in the initial stages of legume-rhizobia symbiosis [36]. CK production was observed in B. japonicum [37] and secretion of their bioactive forms was demonstrated in E. meliloti [38]. Cytokinins are recognized as signal molecule for nodule development, but they seem to be involved in auto regulation of nodule number (AON). Plants sense the N availability in soil and control nodule formation through AON mechanism. Low N promotes nodulation, however when in abundance, the plant avoids energy intensive symbiotic N and reduce the number of nodules. A reduction in nodule number by a high concentration of CK treatment suggests a role of this hormone in AON [39], but such a possibility needs to be confirmed through further investigation. CKs are synthesized by rhizobia, but the exact role of rhizobial-derived cytokinins is unknown. It has been demonstrated that rhizobial synthesized cytokinins contribute to symbiotic interaction but were not critically required for the interaction as the Bradyrhizobium sp. strain ORS285 mutant deficient in CK production was able to form nodules [40]. The role of CKs and other hormones in nodule development has been comprehensively reviewed [41]. Although hormones enhance plant growth in soybean and non-leguminous crops (ref. in [42]), it is not clearly established that rhizobial-derived cytokinins are transported to the plant tissue where they influence growth and differentiation.

Gibberellins (GAs): There is a large family of GAs that possess a diterpenoid structure. More than 125 GAs have been identified in plants and other organisms such as fungi and bacteria, but only a small portion of them is biologically active. GAs participates in the regulation of a wide range of plant developmental processes. The most notable are breaking the seed dormancy and enabling its germination; stem elongation; changing the vegetative phase of the plant to the reproductive stage; flower and fruit development; seed development; and senescence. The effect of GAs is mediated by their interaction with DELLA proteins. The hormone is involved in DELLA protein degradation. A certain level of these proteins is required to promote infection through an interaction with the Nod factor mediated pathway. Several mutant studies that interfered with GA biosynthesis, DELLA protein functionality, and the application of the hormone suggested that the GAs can have opposite effects on nodule development depending on the dose and mode of application [43]. In pea, a low concentration of GA application had a positive effect on nodule formation, while a higher amount inhibited the infection thread, which is one of the initial structural changes in the root to acquire the bacterium [44]. The identification of cytochromes P450 in B. japonicum indicated the presence of a GA biosynthetic capability in symbiotic organisms, though it was not found to be essential for a symbiotic function [45]. The protein catalyzes the steps of the pathways of many metabolites, including GAs. Subsequently, the genomic analysis of the bacterium identified diterpenoid operon which plays an important role in GA biosynthesis [46]. Later, many rhizobial species were reported to synthesize bioactive GAs (ref. in [47]). A cooperative action of both the symbiont and host led to GA biosynthesis in Bradyrhizobium diazoefficiens, where a precursor of a GA is converted to its bioactive form, GA9, by the host GA-3 oxidase [48]. It is speculated that rhizobial GA biosynthesis in cooperation with the host was an evolutionary event that influences nodulation in a mutually beneficial manner. Besides GAs effect on nodule formation, the exogenous application of GA3 in pea delayed the nodule senescence [49]. It is not known if the role of rhizobial-driven GAs is restricted to nodule development and senescence or if they influence other aspects of plant development.

Abscisic acid (ABA) and ethylene: Both of these hormones are central to plant defense mechanism against a variety of stresses. ABA regulates seed dormancy, senescence, stomatal opening and closing, and many other aspects of plant development. It is synthesized as a stress signal, particularly during abiotic stresses, and plays a critical role in downstream responses. Many transcription factors, including DREB2A and 2B, AREB1, RD22BP1 and MYC/ MYB, are known to regulate ABA-responsive gene expression through interactions with cis-acting elements on the DNA. The hormone alters the root architecture under water deficient conditions and modulates the ion influx to change root osmotic potential, providing drought tolerance to the plant [50]. Indirect evidence suggests that ABA might be involved in the host’s drought tolerance. The inoculation of chickpea with R. leguminosarum led to an enhanced tolerance against drought, which was accompanied by higher synthesis of auxins and GAs with a lower level of ABA [51]. Similarly, in a symbiotic interaction of alfalfa with a IAA-overproducing Rhizobium strain, the host plant showed better adaptation to drought conditions that correlated with a higher level of IAA and downregulation of ABA biosynthesis genes [34]. ABA is known to affect nodule development. Exogenous ABA inhibited nodule formation, which appears to be mediated early in a symbiotic interaction [10]. In a report, GmbZIP1 has been shown to negatively regulate ABA-induced inhibition of nodulation by targeting GmENOD40-1 in soybean [52]. GmbZIP1 is a transcription factor that belongs to the AREB/ABF family, and GmENOD40-1 is a marker gene for nodule formation.

The role of ethylene in nodule formation has been reviewed previously [7, 53]. Ethylene is a negative regulator of nodulation, and interference in ethylene signaling led to stimulation of nodulation in alfalfa [54]. Further, several mutants and transgenic plants with altered ethylene levels and signaling showed variation in nodule size and number [53]. Ethylene is synthesized from its immediate substrate, 1-aminocyclopropane-1-carboxylic acid (ACC), through oxidation. The hormone level could be reduced by the degradation of ACC through a reaction catalyzed by acdS encoded ACC deaminase. A higher level of acdS expression in B. japonicum strains was linked with higher numbers of nodules, higher biomass, and delayed senescence of the nodules. The negative effect of ethylene was also demonstrated through over-expressing the ACC deaminase gene in pea and alfalfa (ref. in [7]).

Advertisement

3. Mineral acquisition

Most nutrients available in natural ecosystems are minimally bioavailable to plants since they are bound to inorganic molecules. Microorganisms like bacteria and fungi have the machinery necessary to depolymerize and mineralize the organic forms of these nutrients. With that, inorganic N, P, and S, including ionic species like ammonium, nitrate, phosphate, and sulfate, become available to the plant [55]. Besides the ability to fix N, soil microorganisms also provide essential nutrients by metabolizing recalcitrant forms of N, P and S. As reviewed in [56], these nutrients are needed in several metabolic processes of microorganisms including protein depolymerization and urea catabolism, phosphate and sulfate ester cleavage and phosphonate and sulfonate breakdown. Here, we discuss the role of rhizobial symbionts in P and Fe acquisition by legumes (Figure 2).

Figure 2.

An illustration depicting the role of root nodulating rhizobia in phosphate and iron acquisition in the rhizosphere. The solubilization process may involve a cooperative action of phosphate-solubilizing rhizosphere microbes (PSRM) and phosphate solubilizing endomycorrhizal fungi (PSEF) with rhizobia, using different organic acids, phytases, and extracellular polymeric substances (EPS). The red and blue arrows show phosphate and iron uptake by plant roots, respectively.

P availability: P is among the macronutrients essential for plant growth and development; legumes may require a greater amount of P to support N fixation in the nodules. This is likely related to the additional requirement for various metabolic processes, membrane phospholipid biosynthesis, and energy transformations in the nodules initiated by the symbiotic interaction [57]. P limitation can alter the distribution of nodule phosphatidylcholine [58], an essential component of membrane phospholipids [59]. It has also been associated with small nodules and reduced activity of Fe–S cluster enzymes [60], which are important in various regulatory systems for environmental signals [61]. The amount of plant-available P in the soil is usually low. Thus, owing to the high P requirements of legumes along with the low availability of soluble P in the soil, additional P solubilization and uptake strategies are vital for effective symbiosis. Like many rhizosphere microbes [62], rhizobial symbionts are often associated with P solubilizations [63, 64, 65, 66, 67, 68]. By determining the expression level of high-affinity phosphate transporter and mutant analysis, the role of P uptake has been implicated in nitrogen fixation in soybean-Sinorhizobium fredii symbiosis and terminal differentiation of nodules in other rhizobial strains [69]. These studies suggest that rhizobial symbionts are important in increasing the availability and uptake of P during legume-rhizobium symbiosis.

Rhizobia-induced phosphate solubilization may involve similar mechanisms that are deployed by other rhizosphere microorganisms, such as the production of acids and phytases, proton extrusion, and extracellular oxidation [70] (Figure 2). Although an external application of succinic acid inhibits P solubilization in legumes (e.g., chickpea) [71], several organic acids of microbial origin have been shown to participate in P solubilization in soil [70, 72]. Gluconic acid is one of the P-solubilizing organic acids produced by root-associated rhizobia [63, 68, 73, 74]. Its production by non-rhizobial bacteria including Azospirillum and Nguyenibacter is involved in the solubilization of calcium phosphate and aluminum phosphate [73, 74]. Besides gluconic acid, the solubilization of Ca and Fe phosphates could also be mediated through other organic acids [68, 75]. Alkaline phosphatases (e.g., Sinorhizobium meliloti 104A14) are the other means to solubilize organic phosphates [64, 76]. It is important to note that the presence of non-rhizobial phosphate-solubilizing microorganisms in the rhizosphere can also have a significant impact on the availability of P to legumes, thereby contributing to their increased productivity [77]. Co-inoculation of Pseudomonas putida strains and B. japonicum TIIIB resulted in greater root and shoot biomass in soybean compared to the rhizobial inoculation alone [78]. Regardless of the role of non-rhizobial microbes, the phosphate solubilization ability of symbiotic rhizobia ensures P availability to support N fixation. During symbiosis, rhizobia might have developed regulation strategies for demand-based P uptake. For example, using a PhoCDET phosphate transport system that involves periplasmic binding proteins, S. meliloti may ensure P availability when environmental phosphate concentration is below 1 μM [79]. Taken together, rhizobial symbionts are effective in phosphate solubilization, which in turn ensures the availability of P to legume plants.

Iron availability: Fe is an essential element in several metabolic processes in plants such as chlorophyl synthesis, DNA synthesis, and respiration. Despite being abundant in soil, an insoluble form of Fe requires plant strategies to make it absorbable [80]. The limitation of Fe in soil has been shown to affect the growth of many plants including legumes [81]. In legume-rhizobium symbiosis, Fe is important not only for each partner’s metabolism but also for maintaining effective symbiotic interaction. Because Fe is a component of several enzymes involved in N-fixation, this mineral is required in a higher amount in a symbiotic interaction than that of the free-living partners. This is evident from the highly sensitive nature of nodulation to Fe limitation [82] and the presence of Fe in higher concentration in nodules than roots, leaves, and seeds [83]. Therefore, legumes have developed strategies for Fe transport into nodules. To meet the high demand of Fe, each of the symbiotic partners is playing a role in Fe uptake. S. meliloti was shown to produce volatile organic compounds that function as signals for Medicago truncatula to increase its Fe uptake rate [84]. Rhizobial symbionts, on the other hand, released Fe reducing enzymes to acidify the rhizosphere, which facilitated Fe-uptake by the plant [85]. Recently, a study on M. truncatula observed an interesting phenomenon: the plant expressed protein signals, which activated the Fe import systems of S. meliloti by overriding the Fe-regulating system, sufT [86].

One of the common Fe import systems in rhizobia and many other microorganisms is the secretion of siderophores, which are low molecular-mass non-protein molecules that can scavenge Fe3+ (Figure 2). Under Fe deficiency, rhizobia can secrete siderophores, which form a soluble complex with Fe3+ and are imported into the cells via receptors present on the outer membrane surfaces of this Gram-negative bacteria. Rhizobia express genes encoding enzymes involved in siderophore production as well as outer membrane receptor proteins of siderophore-Fe complexes when Fe is limited. For example, enzymes involved in the biosynthesis of a catechol siderophore are encoded by sid genes of a rhizobial strain that can nodulate chickpea [87]. Soybean-nodulating B. japonicum was also shown to express the FegA gene, which encodes an outer membrane receptor protein for siderophore-Fe complexes [88]. From the success in root colonization, biomass, and chlorophyll content enhancement with genetically improved rhizobia, it may now be possible to improve strains having high-affinity siderophore-receptor proteins to increase the Fe uptake in legumes [89].

Advertisement

4. Environmental adaptability

Besides nutrient limitation, other unfavorable conditions such as, high concentrations of heavy metals, drought, soil salinity, extreme pH conditions, and diseases, negatively affect the growth and development of plants, resulting in a significant reduction in the yield of agricultural produce. With increasing anthropogenic activities, the environmental conditions are becoming unfavorable for plant growth. The chemical industries, for example, generate a tremendous amount of heavy metal waste (e.g., Ni, As, Hg, Cd, Pb, Cr, Mn, Si, Fe, and Cu). When left untreated, industrial waste becomes a source of heavy metal contamination in the environment. The negative effects of heavy-metal-contaminated soils (HMCS) on plant productivity are evident. In legumes, it resulted in poor nodulation and plant growth [90]. Nevertheless, some legumes can effectively nodulate and grow in HMCS, suggesting a protective effect of rhizobial symbionts against many of these metals. The Rhizobium-induced improvements in nodulation, root and shoot lengths, N-fixation, and seed yield were observed in legumes grown in soil contaminated with Ni (290 mg Ni kg−1 soil) [91], As [92, 93, 94], Hg [95], Cd [96], Cr [97], Cu [98], Al [99], and Pb [100].

The rhizobia display a natural adaptation to stress, as many of the isolates in HMCS showed a significant potential for heavy metal stress tolerance [101]. A number of the rhizobial strains isolated from Hg-contaminated soils showed tolerance to the metal [102]. The variation in tolerance across the strains was linked with the mercuric reductase activity that converts Hg2+ into its less toxic and a volatile form, Hg0. The adaptive capacity is enhanced through cooperative action among different suites of the microflora. This has been demonstrated in faba bean plants, where a combination of Bacillus strains with those of Rhizobium increased the biosorption of Ni and improved plant growth [103]. Similarly, co-inoculation of rhizobia and non-rhizobial bacteria has been shown to improve tolerance against Pb [104], As [94, 105], and Cu [106, 107]. The tolerance mechanism is not well understood, but metal biosorption by rhizobia, accumulation of metals in the roots, volatilization, conversion to less toxic forms, and rhizobia-induced root uptake regulation are some of the ways the symbionts are known to alleviate the metal toxicity. Recently, a molecular investigation into metal tolerance mechanisms revealed changes in m6A RNA methylation in response to Cd stress [101, 108]. The genes that were differentially methylated and expressed were linked to Ca2+ homeostasis, ROS pathways, polyamine metabolism, MAPK signaling, hormones, and biotic stress responses. In another study, the primary targets of Al were genes involved in membrane biogenesis, metal ions binding and transport, carbohydrate, and amino acid metabolism and transport [99]. The Rhizobium strain B3, which was tolerant to Al toxicity, bound less of the ion compared to the wild-type and maintained membrane integrity. Through recombinant technology, the incorporation of mechanisms of heavy metal tolerance from unrelated species can lead to genetic improvements in the symbiotic rhizobia [109].

Drought stress can reduce nodulation, biomass, and chlorophyll contents of legumes while accumulating ROS that disrupt the structure and function of different biomolecules including DNA [110, 111, 112]. Soil salinity negatively impacts rhizobial infections and root nodulations in several legumes: bean [113], soybean [114], pea [115], and chickpea [116]. Some legumes can cope with the adverse conditions, suggesting a positive role for rhizobia in the plant’s adaptation to stress. The observed protection against salinity stress by symbionts was due to the production of osmoprotectant molecules (e.g., glutamine, serine, glutamate, and proline) [117], antioxidants [118] and by changing the xylem osmotic potential and amount of aquaporins [119]. Both salinity and drought stress limit the availability of water to plants. Like salinity stress, there are many examples where the protective effect of rhizobia-host symbiosis is evident in the management of drought stress [120]. Both salinity and drought stress responses are complex and may involve non-rhizobial plant growth promoting rhizobacteria and endomycorrhiza [116, 117, 118, 119]. On the other hand, extreme pH can have a direct impact on the establishment of symbiosis and thus on the N fixation and productivity of legumes [121, 122]. Soil pH can also influence the diversity and structure of the microbial community around the root rhizosphere, which could affect the legume-rhizobia symbiosis [123, 124, 125]. In some legumes, acidic conditions induced the expression of different rhizobial genes (e.g., actA, typA, atvA, lpiA, and ubiF), which may be playing a role in acid stress tolerance mechanism [126, 127, 128, 129, 130].

Symbiotic interactions are known to enhance growth and survival of the host plant against fungal and bacterial pathogens. The anti-phytopathogenic activities of different rhizobial species such as those observed in chickpea against the oomycete pathogen Phytophthora medicaginis [131], pigeon pea against Fusarium wilt diseases [132], and soybean against root rot diseases [133] have been reported.

Advertisement

5. Strain improvement for enhanced productivity

Advances in molecular biology have allowed genetic improvements in rhizobial strains to increase the symbiotic benefits. Rhizobial strains have been modified to increase N fixation in legumes and some cereals [7, 134] as well as to improve the growth, nutrient supply, and stress tolerance [135, 136]. The overproduction of IAA in S. meliloti RD64 by introducing a chimeric operon containing genes from Agrobacterium rhizogenes, Pseudomonas syringae pv. savastanoi, and Agrobacterium tumefaciens [137] was associated with the increased production of acid phosphatase and organic acids, resulting in a higher growth of M. truncatula under P limitation [138]. Another Rhizobium meliloti strain, when modified with extra copies of the nifA and dctABD genes increased the alfalfa biomass by 13% [139]. The nifA gene is involved in nitrogen fixation and dctABD gene product is a dicarboxylate transporter. The latter is a primary carbon used by bacteria for energy and other cellular functions. An improvement in symbiotic efficiency can also lead to enhanced plant growth. This has been shown by transferring Smed_3503, Smed_5985, and Smed_6456, into the S. meliloti Rm1021 strain, which led to a 60% increase in M. truncatula biomass [140].

Besides growth and yield, many studies have successfully demonstrated the potential of genetically improved rhizobial symbionts in alleviating adverse environmental conditions in the host legumes. The tolerance mechanisms of these microbes could be exploited for genetic engineering approaches in symbionts. By using this strategy, a Rhizobium leguminosarum strain transformed with an arsenite S-adenosylmethionine methyltransferase gene of algae, effectively methylated, and volatilized As, to a tolerance against this metal [109]. Similarly, manipulating a superoxide dismutase level reduced Cd stress and increased root and shoot length, biomass, and N content in soybean [135]. The strains have been genetically modified to overproduce cytokinin, IAA, trehalose-6-phosphate synthase, 1-aminocyclopropane-1-carboxylic acid deaminase, high-affinity cytochrome cbb3-type oxidase, etc., to increase salt or drought tolerance or to overcome the waterlogging stress in legumes [32, 34, 54, 137, 138, 141, 142]. Strain improvement followed by extensive evaluation of the strains for survival under field conditions is also critical, as gene modification may reduce the strains’ competitiveness against the indigenous strains and hence their survival rate in the nodules [143].

Advertisement

6. Future perspectives

The rhizobial symbiotic interactions are extremely valuable for the sustainability of crop cultivation. This has gained more importance in view of climate change and its imminent threat to the ecological balance. Leguminous crops make a significant contribution to reducing the use of chemical N-fertilizers. The positive impact of commercial and wild symbiotic hosts is enhanced by their growing abilities in diverse environments. The different rhizobial species indirectly increase the N fixation productivity by promoting the growth and productivity of the host plants. A substantial number of benefits other than N fixation point to close coordination between the host and the symbiont while acquiring this capacity. The horizontal transfer of genes appears to have played an important role in improving the symbiotic functionality of the bacterium [144, 145]. Gaining insight into these evolutionary events could provide a broader base for strain improvement. The genetic engineering tools have demonstrated how specific bacterial functions can be transferred across even unrelated species. Further understanding of a symbiotic interaction can bring us closer to developing tailor-made strains.

Advertisement

Acknowledgments

The funding provided by Alberta Pulse Growers and Alberta Results Driven Agriculture Research is acknowledged.

Advertisement

Conflict of interest

The authors do not have any conflict of interests.

References

  1. 1. Wang Q , Liu J, Zhu H. Genetic and molecular mechanisms underlying symbiotic specificity in legume-rhizobium interactions. Frontiers in Plant Science. 2018;9:313. DOI: 10.3389/fpls.2018.00313
  2. 2. Bruijn F, J: Biological nitrogen fixation. Hoboken. New Jersey: Wiley-Blackwell; 2015
  3. 3. Martínez-Romero E. Coevolution in rhizobium-legume symbiosis? DNA and Cell Biology. 2009;28(8):361-370. DOI: 10.1089/dna.2009.0863
  4. 4. Coba de la Peña T, Fedorova E, Pueyo JJ, Lucas MM. The symbiosome: Legume and rhizobia co-evolution toward a nitrogen-fixing organelle? Front. Plant Science. 2017;8:2229. DOI: 10.3389/fpls.2017.02229
  5. 5. Goyal R, Mattoo A, Schmidt A. Rhizobial–host interactions and symbiotic nitrogen fixation in legume crops toward agriculture sustainability. Frontiers in Microbiology. 2021;12:669404. DOI: 10.3389/fmicb.2021.669404
  6. 6. Raun WR, Johnson GV. Improving nitrogen use efficiency for cereal production. Journal of Agronomy. 1999;91(3):357-363
  7. 7. Goyal R, Schmidt A, Hynes M. Molecular biology in the improvement of biological nitrogen fixation by rhizobia and extending the scope to cereals. Microorganisms. 2021;9:125. DOI: 10.3390/microorganisms9010125
  8. 8. Nag P, Shriti S, Das S. Microbiological strategies for enhancing biological nitrogen fixation in nonlegumes. Journal of Applied Microbiology. 2020;129(2):186-198. DOI: 10.1111/jam.14557
  9. 9. Patten CL, Glick BR. Role of pseudomonas putida indoleacetic acid in development of the host plant root system. Applied and Environmental Microbiology. 2002;68(8):3795-3801. DOI: 10.1128/aem.68.8.3795-3801.2002
  10. 10. Ferguson BJ, Mathesius U. Phytohormone regulation of legume-rhizobia interactions. Journal of Chemical Ecology. 2014;40(7):770-790. DOI: 10.1007/s10886-014-0472-7
  11. 11. Rodríguez H, Fraga R. Phosphate solubilizing bacteria and their role in plant growth promotion. Biotechnology Advances. 1999;17(4-5):319-339. DOI: 10.1016/s0734-9750(99)00014-2
  12. 12. Errickson W, Huang B. Roles and Mechanisms of Rhizobacteria in Regulating Plant Tolerance to Abiotic Stress. Boca Raton: CRC Press; 2020. pp. 25-39
  13. 13. Jaiswal SK, Mohammed M, Ibny FYI, Dakora FD. Rhizobia as a source of plant growth-promoting molecules: Potential applications and possible operational mechanisms. Frontiers in Sustainable Food Systems. 2020;4:619676
  14. 14. Patil A, Kale A, Ajane G, Sheikh R, Patil S. Plant growth-promoting rhizobium: Mechanisms and biotechnological prospective. In: Hansen AP, Choudhary DK, Agrawal PK, Varma A, editors. Rhizobium Biology and Biotechnology. Cham: Springer International Publishing; 2017. pp. 105-134
  15. 15. Kasim W, Osman M, Omar N, Abd El-Daim I, Bejai S, Meijer J. Control of drought stress in wheat using plant-growth-promoting bacteria. Journal of Plant Growth Regulation. 2012;32. DOI: 10.1007/s00344-012-9283-7
  16. 16. Ullah A, Nisar M, Ali H, Hazrat A, Hayat K, Keerio AA, et al. Drought tolerance improvement in plants: An endophytic bacterial approach. Applied Microbiology and Biotechnology. 2019;103(18):7385-7397. DOI: 10.1007/s00253-019-10045-4
  17. 17. Etminani F, Harighi B. Isolation and identification of endophytic bacteria with plant growth promoting activity and biocontrol potential from wild pistachio trees. Plant Pathology Journal. 2018;34(3):208-217. DOI: 10.5423/ppj.Oa.07.2017.0158
  18. 18. Tsavkelova EA, Cherdyntseva TA, Klimova SY, Shestakov AI, Botina SG, Netrusov AI. Orchid-associated bacteria produce indole-3-acetic acid, promote seed germination, and increase their microbial yield in response to exogenous auxin. Archives of Microbiology. 2007;188(6):655-664. DOI: 10.1007/s00203-007-0286-x
  19. 19. Masciarelli O, Llanes A, Luna V. A new PGPR co-inoculated with Bradyrhizobium japonicum enhances soybean nodulation. Microbiological Research. 2014;169(7):609-615. DOI: 10.1016/j.micres.2013.10.001
  20. 20. Defez R, Andreozzi A, Romano S, Pocsfalvi G, Fiume I, Esposito R, et al. Bacterial IAA-delivery into Medicago root nodules triggers a balanced stimulation of c and n metabolism leading to a biomass increase. Microorganisms. 2019;7(10):403. DOI: 10.3390/microorganisms7100403
  21. 21. Camerini S, Senatore B, Lonardo E, Imperlini E, Bianco C, Moschetti G, et al. Introduction of a novel pathway for IAA biosynthesis to rhizobia alters vetch root nodule development. Archives of Microbiology. 2008;190(1):67-77. DOI: 10.1007/s00203-008-0365-7
  22. 22. Chi F, Shen SH, Cheng HP, Jing YX, Yanni YG, Dazzo FB. Ascending migration of endophytic rhizobia, from roots to leaves, inside rice plants and assessment of benefits to rice growth physiology. Applied and Environmental Microbiology. 2005;71(11):7271-7278. DOI: 10.1128/aem.71.11.7271-7278.2005
  23. 23. Singh R, Mishra RPN, Jaiswal HK. Role of rhizobial endophytes as nitrogen fixer in promoting plant growth and productivity of Indian cultivated upland rice (Oryza Sativa L.). Plants. In: Wang YP, Lin M, Tian ZX, Elmerich C, Newton WE, editors. Biological Nitrogen Fixation, Sustainable Agriculture and the Environment. Current Plant Science and Biotechnology in Agriculture. Dordrecht: Springer; 2005;41. ISBN 978-1-4020-3570-8. DOI: 10.1007/1-4020-3570-5_73
  24. 24. Pacios-Bras C, Schlaman HRM, Boot K, Admiraal P, Mateos Langerak J, Stougaard J, et al. Auxin distribution in Lotus japonicus during root nodule development. Plant Molecular Biology. 2003;52(6):1169-1180. DOI: 10.1023/b:plan.0000004308.78057.f5
  25. 25. Subramanian S, Stacey G, Yu O. Distinct, crucial roles of flavonoids during legume nodulation. Trends in Plant Science. 2007;12(7):282-285. DOI: 10.1016/j.tplants.2007.06.006
  26. 26. Ghosh S, Basu PS. Production and metabolism of indole acetic acid in roots and root nodules of Phaseolus mungo. Microbiological Research. 2006;161(4):362-366. DOI: 10.1016/j.micres.2006.01.001
  27. 27. Spaepen S, Das F, Luyten E, Michiels J, Vanderleyden J. Indole-3-acetic acid-regulated genes in rhizobium etli CNPAF512. FEMS Microbiology Letters. 2009;291(2):195-200. DOI: 10.1111/j.1574-6968.2008.01453.x
  28. 28. Torres D, Benavidez I, Donadio F, Mongiardini E, Rosas S, Spaepen S, et al. New insights into auxin metabolism in Bradyrhizobium japonicum. Research in Microbiology. 2018;169(6):313-323. DOI: 10.1016/j.resmic.2018.04.002
  29. 29. Sarwar M, Kremer RJ. Determination of bacterially derived auxins using a microplate method. Letters in Applied Microbiology. 1995;20(5):282-285. DOI: 10.1111/j.1472-765X.1995.tb00446.x
  30. 30. Bargaz A, Lazali M, Amenc L, Abadie J, Ghoulam C, Farissi M, et al. Differential expression of trehalose 6-P phosphatase and ascorbate peroxidase transcripts in nodule cortex of Phaseolus vulgaris and regulation of nodule O2 permeability. Planta. 2013;238(1):107-119. DOI: 10.1007/s00425-013-1877-1
  31. 31. Barraza A, Estrada-Navarrete G, Rodriguez-Alegria ME, Lopez-Munguia A, Merino E, Quinto C, et al. Down-regulation of PvTRE1 enhances nodule biomass and bacteroid number in the common bean. The New Phytologist. 2013;197(1):194-206. DOI: 10.1111/nph.12002
  32. 32. Suárez R, Wong A, Ramírez M, Barraza A, Orozco Mdel C, Cevallos MA, et al. Improvement of drought tolerance and grain yield in common bean by overexpressing trehalose-6-phosphate synthase in rhizobia. Molecular Plant-Microbe Interactions. 2008;21(7):958-966. DOI: 10.1094/mpmi-21-7-0958
  33. 33. Bianco C, Defez R. Medicago truncatula improves salt tolerance when nodulated by an indole-3-acetic acid-overproducing Sinorhizobium meliloti strain. Journal of Experimental Botany. 2009;60(11):3097-3107. DOI: 10.1093/jxb/erp140
  34. 34. Defez R, Andreozzi A, Dickinson M, Charlton A, Tadini L, Pesaresi P, et al. Improved drought stress response in alfalfa plants nodulated by an IAA over-producing rhizobium strain. Frontiers in Microbiology. 2017;8:2466. DOI: 10.3389/fmicb.2017.02466
  35. 35. Gamas P, Brault M, Jardinaud MF, Frugier F. Cytokinins in symbiotic nodulation: When, where, what for? Trends in Plant Science. 2017;22(9):792-802. DOI: 10.1016/j.tplants.2017.06.012
  36. 36. Jarzyniak K, Banasiak J, Jamruszka T, Pawela A, Di Donato M, Novák O, et al. Early stages of legume-rhizobia symbiosis are controlled by ABCG-mediated transport of active cytokinins. Nature Plants. 2021;7(4):428-436. DOI: 10.1038/s41477-021-00873-6
  37. 37. Sturtevant D, Taller B. Cytokinin Production by Bradyrhizobium japonicum. Plant Physiology. 1989;89:1247-1252. DOI: 10.1104/pp.89.4.1247
  38. 38. Kisiala A, Laffont C, Emery RJ, Frugier F. Bioactive cytokinins are selectively secreted by Sinorhizobium meliloti nodulating and nonnodulating strains. Molecular Plant-Microbe Interactions. 2013;26(10):1225-1231. DOI: 10.1094/mpmi-02-13-0054-r
  39. 39. Mens C, Li D, Haaima LE, Gresshoff PM, Ferguson BJ. Local and systemic effect of cytokinins on soybean nodulation and regulation of their isopentenyl transferase (IPT) biosynthesis genes following rhizobia inoculation. Frontiers in Plant Science. 2018;9:1150. DOI: 10.3389/fpls.2018.01150
  40. 40. Podlešáková K, Fardoux J, Patrel D, Bonaldi K, Novák O, Strnad M, et al. Rhizobial synthesized cytokinins contribute to but are not essential for the symbiotic interaction between photosynthetic Bradyrhizobia and Aeschynomene legumes. Molecular Plant-Microbe Interactions. 2013;26(10):1232-1238. DOI: 10.1094/mpmi-03-13-0076-r
  41. 41. Velandia K, Reid JB, Foo E. Right time, right place: The dynamic role of hormones in rhizobial infection and nodulation of legumes. Plant Communications. 2022;3(5):100327. DOI: 10.1016/j.xplc.2022.100327
  42. 42. Swarnalakshmi K, Yadav V, Tyagi D, Dhar DW, Kannepalli A, Kumar S. Significance of plant growth promoting rhizobacteria in grain legumes: Growth promotion and crop production. Plants (Basel). 2020;9(11):1596. DOI: 10.3390/plants9111596
  43. 43. McGuiness PN, Reid JB, Foo E. The role of gibberellins and brassinosteroids in nodulation and arbuscular mycorrhizal associations. Frontiers in Plant Science. 2019;10:269. DOI: 10.3389/fpls.2019.00269
  44. 44. Ferguson BJ, Ross JJ, Reid JB. Nodulation phenotypes of gibberellin and brassinosteroid mutants of pea. Plant Physiology. 2005;138(4):2396-2405. DOI: 10.1104/pp.105.062414
  45. 45. Tully RE, Keister DL. Cloning and mutagenesis of a cytochrome P-450 locus from Bradyrhizobium japonicum that is expressed anaerobically and symbiotically. Applied and Environmental Microbiology. 1993;59(12):4136-4142. DOI: 10.1128/aem.59.12.4136-4142.1993
  46. 46. Kaneko T, Nakamura Y, Sato S, Minamisawa K, Uchiumi T, Sasamoto S, et al. Complete genomic sequence of nitrogen-fixing symbiotic bacterium Bradyrhizobium japonicum USDA110. DNA Research. 2002;9(6):189-197. DOI: 10.1093/dnares/9.6.189
  47. 47. Keswani C, Singh SP, García-Estrada C, Mezaache-Aichour S, Glare TR, Borriss R, et al. Biosynthesis and beneficial effects of microbial gibberellins on crops for sustainable agriculture. Journal of Applied Microbiology. 2022;132(3):1597-1615. DOI: 10.1111/jam.15348
  48. 48. Nett RS, Bender KS, Peters RJ. Production of the plant hormone gibberellin by rhizobia increases host legume nodule size. The ISME Journal. 2022;16(7):1809-1817. DOI: 10.1038/s41396-022-01236-5
  49. 49. Serova TA, Tsyganova AV, Tikhonovich IA, Tsyganov VE. Gibberellins inhibit nodule senescence and stimulate nodule meristem bifurcation in pea (Pisum sativum L.). front. Plant Science. 2019;10:285. DOI: 10.3389/fpls.2019.00285
  50. 50. Sah SK, Reddy KR, Li J. Abscisic acid and abiotic stress tolerance in crop plants. Frontiers in Plant Science. 2016;7:571. DOI: 10.3389/fpls.2016.00571
  51. 51. Bano A, Batool R, Dazzo FB. Adaptation of chickpea to desiccation stress is enhanced by symbiotic rhizobia. Symbiosis. 2010;50:129-133
  52. 52. Xu S, Song S, Dong X, Wang X, Wu J, Ren Z, et al. GmbZIP1 negatively regulates ABA-induced inhibition of nodulation by targeting GmENOD40-1 in soybean. BMC Plant Biology. 2021;21(1):35. DOI: 10.1186/s12870-020-02810-9
  53. 53. Guinel FC. Ethylene, a hormone at the center-stage of nodulation. Frontiers in Plant Science. 2015;6:1121. DOI: 10.3389/fpls.2015.01121
  54. 54. Peters NK, Crist-Estes DK. Nodule formation is stimulated by the ethylene inhibitor aminoethoxyvinylglycine. Plant Physiology. 1989;91(2):690-693. DOI: 10.1104/pp.91.2.690
  55. 55. Van Der Heijden MGA, Bardgett RD, Van Straalen NM. The unseen majority: Soil microbes as drivers of plant diversity and productivity in terrestrial ecosystems. Ecology Letters. 2008;11(3):296-310. DOI: 10.1111/j.1461-0248.2007.01139.x
  56. 56. Jacoby R, Peukert M, Succurro A, Koprivova A, Kopriva S. The role of soil microorganisms in plant mineral nutrition-current knowledge and future directions. Frontiers in Plant Science. 2017;8:1617. DOI: 10.3389/fpls.2017.01617
  57. 57. Mitran T, Meena RS, Lal R, Layek J, Kumar S, Datta R. In: Meena A, Das GSY, Lal R, editors. Role of Soil Phosphorus on Legume Production, in Legumes for Soil Health and Sustainable Management, R.S. Springer Singapore: Singapore; 2018. pp. 487-510
  58. 58. Dokwal D, Romsdahl TB, Kunz DA, Alonso AP, Dickstein R. Phosphorus deprivation affects composition and spatial distribution of membrane lipids in legume nodules. Plant Physiology. 2021;185(4):1847-1859. DOI: 10.1093/plphys/kiaa115
  59. 59. Botella C, Jouhet J, Block MA. Importance of phosphatidylcholine on the chloroplast surface. Progress in Lipid Research. 2017;65:12-23. DOI: 10.1016/j.plipres.2016.11.001
  60. 60. Qin L, Wang M, Chen L, Liang X, Wu Z, Lin Z, et al. Soybean Fe-S cluster biosynthesis regulated by external iron or phosphate fluctuation. Plant Cell Reports. 2015;34(3):411-424. DOI: 10.1007/s00299-014-1718-0
  61. 61. Mettert EL, Kiley PJ. Fe-S proteins that regulate gene expression. Biochimica et Biophysica Acta. 2015;1853(6):1284-1293. DOI: 10.1016/j.bbamcr.2014.11.018
  62. 62. Kucey RMN, Janzen HH, Leggett ME. Microbially mediated increases in plant-available phosphorus. Advances in Agronomy. 1989;42:198-228
  63. 63. Alikhani HA, Saleh-Rastin N, Antoun H. Phosphate solubilization activity of rhizobia native to Iranian soils. Plant and Soil. 2006;287(1):35. DOI: 10.1007/s11104-006-9059-6
  64. 64. Deng S, Elkins JG, Da LH, Botero LM, McDermott TR. Cloning and characterization of a second acid phosphatase from Sinorhizobium meliloti strain 104A14. Archives of Microbiology. 2001;176(4):255-263. DOI: 10.1007/s002030100311
  65. 65. Kang X, Yu X, Zhang Y, Cui Y, Tu W, Wang Q , et al. Inoculation of Sinorhizobium saheli YH1 leads to reduced metal uptake for Leucaena leucocephala grown in mine tailings and metal-polluted soils. Frontiers in Microbiology. 2018;9:1853. DOI: 10.3389/fmicb.2018.01853
  66. 66. Kumar G, Ram R. Phosphate solubilizing rhizobia isolated from Vigna trilobata. American Journal of Microbiological Research. 2014;2:105-109. DOI: 10.12691/ajmr-2-3-4
  67. 67. Sridevi M, Mallaiah KV. Phosphate solubilization by rhizobium strains. Indian Journal of Microbiology. 2009;49(1):98-102. DOI: 10.1007/s12088-009-0005-1
  68. 68. Yadav A, Singh RP, Singh AL, Singh M. Identification of genes involved in phosphate solubilization and drought stress tolerance in chickpea symbiont Mesorhizobium ciceri Ca181. Archives of Microbiology. 2021;203(3):1167-1174. DOI: 10.1007/s00203-020-02109-1
  69. 69. Hu Y, Jiao J, Liu LX, Sun YW, Chen WF, Sui XH, et al. Evidence for phosphate starvation of rhizobia without terminal differentiation in legume nodules. Molecular Plant-Microbe Interactions. 2018;31(10):1060-1068. DOI: 10.1094/MPMI-02-18-0031-R
  70. 70. Rawat P, Das S, Shankhdhar D, Shankhdhar SC. Phosphate-solubilizing microorganisms: Mechanism and their role in phosphate solubilization and uptake. Journal of Soil Science and Plant Nutrition. 2020;21(1):49-68. DOI: 10.1007/s42729-020-00342-7
  71. 71. Iyer B, Rajput MS, Rajkumar S. Effect of succinate on phosphate solubilization in nitrogen fixing bacteria harbouring chickpea and their effect on plant growth. Microbiological Research. 2017;202:43-50. DOI: 10.1016/j.micres.2017.05.005
  72. 72. Alori ET, Glick BR, Babalola OO. Microbial phosphorus solubilization and its potential for use in sustainable agriculture. Frontiers in Microbiology. 2017;8:971. DOI: 10.3389/fmicb.2017.00971
  73. 73. Li XL, Zhao XQ , Dong XY, Ma JF, Shen RF. Secretion of gluconic acid from Nguyenibacter sp. l1 is responsible for solubilization of aluminum phosphate. Frontiers in Microbiology. 2021;12:784025. DOI: 10.3389/fmicb.2021.784025
  74. 74. Rodriguez H, Gonzalez T, Goire I, Bashan Y. Gluconic acid production and phosphate solubilization by the plant growth-promoting bacterium Azospirillum spp. Die Naturwissenschaften. 2004;91(11):552-555. DOI: 10.1007/s00114-004-0566-0
  75. 75. Marra LM. Fonsêca Sousa Soares C R, de Oliveira S M, Avelar Ferreira P a, Lima Soares B, de Fráguas Carvalho R, de Lima J M, and de Souza Moreira F M: Biological nitrogen fixation and phosphate solubilization by bacteria isolated from tropical soils. Plant and Soil. 2012;357(1-2):289-307. DOI: 10.1007/s11104-012-1157-z
  76. 76. Cerecetto V, Beyhaut E, Amenc L, Trives C, Altier N, Drevon J-J. Contrasting expression of rhizobial phytase in nodules of two soybean cultivars grown under low phosphorus availability. Frontiers in Sustainable Food Systems. 2021;4. DOI: 10.3389/fsufs.2020.607678
  77. 77. Prabhu N, Borkar S, Garg S. Phosphate solubilization by microorganisms: Overview, mechanisms, applications and advances (Chapter 11). Advances in Biological Science Research: A Practical Approach. 2019;161-176. DOI: 10.1016/B978-0-12-817497-5.00011-2
  78. 78. Rosas SB, Andrés JA, Rovera M, Correa NS. Phosphate-solubilizing pseudomonas putida can influence the rhizobia–legume symbiosis. Soil Biology and Biochemistry. 2006;38(12):3502-3505. DOI: 10.1016/j.soilbio.2006.05.008
  79. 79. Bardin S, Dan S, Osteras M, Finan TM. A phosphate transport system is required for symbiotic nitrogen fixation by rhizobium meliloti. Journal of Bacteriology. 1996;178(15):4540-4547. DOI: 10.1128/jb.178.15.4540-4547.1996
  80. 80. Colombo C, Palumbo G, He J-Z, Pinton R, Cesco S. Review on iron availability in soil: Interaction of Fe minerals, plants, and microbes. JSS. 2013;14(3):538-548. DOI: 10.1007/s11368-013-0814-z
  81. 81. Lopez-Millan AF, Grusak MA, Abadia A, Abadia J. Iron deficiency in plants: An insight from proteomic approaches. Frontiers in Plant Science. 2013;4:254. DOI: 10.3389/fpls.2013.00254
  82. 82. Tang CK, Chin J, Harford JB, Klausner RD, Rouault TA. Iron regulates the activity of the iron-responsive element binding protein without changing its rate of synthesis or degradation. JBC. 1992;267(34):24466-24470. DOI: 10.1016/S0021-9258(18)35789-2
  83. 83. Burton JW, Harlow C, Theil EC. Evidence for reutilization of nodule iron in soybean seed development. Journal of Plant Nutrition. 2008;21(5):913-927. DOI: 10.1080/01904169809365453
  84. 84. Orozco-Mosqueda Mdel C, Macías-Rodríguez LI, Santoyo G, Farías-Rodríguez R, Valencia-Cantero E. Medicago truncatula increases its iron-uptake mechanisms in response to volatile organic compounds produced by Sinorhizobium meliloti. Folia Microbiologia (Praha). 2013;58(6):579-585. DOI: 10.1007/s12223-013-0243-9
  85. 85. Orozco-Mosqueda M. d C, Rocha-Granados M d C, Glick B R, and Santoyo G: Microbiome engineering to improve biocontrol and plant growth-promoting mechanisms. Microbiological Research. 2018;208:25-31
  86. 86. Sankari S, Babu VMP, Bian K, Alhhazmi A, Andorfer MC, Avalos DM, et al. A haem-sequestering plant peptide promotes iron uptake in symbiotic bacteria. Nature Microbiology. 2022;7(9):1453-1465. DOI: 10.1038/s41564-022-01192-y
  87. 87. Datta B, Chakrabartty PK. Siderophore biosynthesis genes of rhizobium sp. isolated from Cicer arietinum L. 3. Biotech. 2014;4(4):391-401. DOI: 10.1007/s13205-013-0164-y
  88. 88. LeVier K, Guerinot ML. The Bradyrhizobium japonicum fegA gene encodes an iron-regulated outer membrane protein with similarity to hydroxamate-type siderophore receptors. Journal of Bacteriology. 1996;178(24):7265-7275. DOI: 10.1128/jb.178.24.7265-7275.1996
  89. 89. Khan A, Archana G, Desai A. Engineering heterologous iron siderophore complex utilization in rhizobia: Effect on growth of peanut and pigeon pea plants. Applied Soil Ecology. 2012;53. DOI: 10.1016/j.apsoil.2011.10.015
  90. 90. Jach ME, Sajnaga E, Ziaja M. Utilization of legume-nodule bacterial symbiosis in phytoremediation of heavy metal-contaminated soils. Biology (Basel). 2022;11(5):676. DOI: 10.3390/biology11050676
  91. 91. Wani PA, Khan MS, Zaidi A. Effect of metal-tolerant plant growth-promoting rhizobium on the performance of pea grown in metal-amended soil. Archives of Environmental Contamination and Toxicology. 2008;55(1):33-42. DOI: 10.1007/s00244-007-9097-y
  92. 92. Mandal SM, Gouri SS, De D, Das BK, Mondal KC, Pati BR. Effect of arsenic on nodulation and nitrogen fixation of blackgram (Vigna mungo). Indian Journal of Microbiology. 2011;51(1):44-47. DOI: 10.1007/s12088-011-0080-y
  93. 93. Reichman SM. The potential use of the legume–rhizobium symbiosis for the remediation of arsenic contaminated sites. Soil Biology and Biochemistry. 2007;39(10):2587-2593. DOI: 10.1016/j.soilbio.2007.04.030
  94. 94. Ye L, Yang P, Zeng Y, Li C, Jian N, Wang R, et al. Rhizobium symbiosis modulates the accumulation of arsenic in Medicago truncatula via nitrogen and NRT3.1-like genes regulated by ABA and linalool. Journal of Hazardous Materials. 2021;415:125611. DOI: 10.1016/j.jhazmat.2021.125611
  95. 95. Quiñones MA, Ruiz-Díez B, Fajardo S, López-Berdonces MA, Higueras PL, Fernández-Pascual M. Lupinus albus plants acquire mercury tolerance when inoculated with an Hg-resistant Bradyrhizobium strain. Plant Physiology and Biochemistry. 2013;73:168-175. DOI: 10.1016/j.plaphy.2013.09.015
  96. 96. Zribi K, Djébali N, Mrabet M, Khayat N, Smaoui A, Mlayah A, et al. Physiological responses to cadmium, copper, lead, and zinc of Sinorhizobium sp. strains nodulating Medicago sativa grown in Tunisian mining soils. Annales de Microbiologie. 2012;62(3):1181-1188. DOI: 10.1007/s13213-011-0358-7
  97. 97. Wani PA, Khan MS, Zaidi A. Chromium-reducing and plant growth-promoting Mesorhizobium improves chickpea growth in chromium-amended soil. Biotechnology Letters. 2008;30(1):159-163. DOI: 10.1007/s10529-007-9515-2
  98. 98. Duan C, Mei Y, Wang Q , Wang Y, Li Q , Hong M, et al. Rhizobium inoculation enhances the resistance of alfalfa and microbial characteristics in copper-contaminated soil. Frontiers in Microbiology. 2021;12:781831. DOI: 10.3389/fmicb.2021.781831
  99. 99. Wekesa C, Muoma JO, Reichelt M, Asudi GO, Furch ACU, Oelmüller R. The cell membrane of a novel rhizobium phaseoli strain is the crucial target for aluminum toxicity and tolerance. Cell. 2022;11(5):873. DOI: 10.3390/cells11050873
  100. 100. Rother JA, Millbank JW, Thornton I. Nitrogen fixation by white clover (Trifolium repens) in grasslands on soils contaminated with cadmium, lead and zinc. Journal of Soil Science. 1983;34(1):127-136. DOI: 10.1111/j.1365-2389.1983.tb00819.x
  101. 101. Fan M, Liu Z, Nan L, Wang E, Chen W, Lin Y, et al. Isolation, characterization, and selection of heavy metal-resistant and plant growth-promoting endophytic bacteria from root nodules of Robinia pseudoacacia in a Pb/Zn mining area. Microbiological Research. 2018;217:51-59. DOI: 10.1016/j.micres.2018.09.002
  102. 102. Arregui G, Hipólito P, Pallol B, Lara-Dampier V, García-Rodríguez D, Varela HP, et al. Coba de la Peña T, Lucas M M, and Pueyo J J: Mercury-tolerant Ensifer medicae strains display high mercuric reductase activity and a protective effect on nitrogen fixation in Medicago truncatula nodules under mercury stress. Frontiers in Plant Science. 2020;11:560768. DOI: 10.3389/fpls.2020.560768
  103. 103. Elbagory M, El-Nahrawy S, Omara AE. Synergistic interaction between symbiotic N2 fixing bacteria and bacillus strains to improve growth, physiological parameters, antioxidant enzymes and ni accumulation in faba bean plants (Vicia faba) under nickel stress. Plants (Basel). 2022;11(14):1812. DOI: 10.3390/plants11141812
  104. 104. Jebara SH, Abdelkerim S, Fatnassi IC, Chiboub M, Saadani O, Jebara M. Identification of effective Pb resistant bacteria isolated from Lens culinaris growing in lead contaminated soils. Journal of Basic Microbiology. 2015;55(3):346-353. DOI: 10.1002/jobm.201300874
  105. 105. Armendariz AL, Talano MA, Olmos Nicotra MF, Escudero L, Breser ML, Porporatto C, et al. Impact of double inoculation with Bradyrhizobium japonicum E109 and Azospirillum brasilense Az39 on soybean plants grown under arsenic stress. Plant Physiology and Biochemistry. 2019;138:26-35. DOI: 10.1016/j.plaphy.2019.02.018
  106. 106. Fatnassi IC, Chiboub M, Saadani O, Jebara M, Jebara SH. Impact of dual inoculation with rhizobium and PGPR on growth and antioxidant status of Vicia faba L. under copper stress. Comptes Rendus Biologies. 2015;338(4):241-254. DOI: 10.1016/j.crvi.2015.02.001
  107. 107. Ju W, Liu L, Fang L, Cui Y, Duan C, Wu H. Impact of co-inoculation with plant-growth-promoting rhizobacteria and rhizobium on the biochemical responses of alfalfa-soil system in copper contaminated soil. Ecotoxicology and Environmental Safety. 2019;167:218-226. DOI: 10.1016/j.ecoenv.2018.10.016
  108. 108. Han X, Wang J, Zhang Y, Kong Y, Dong H, Feng X, et al. Changes in the m6A RNA methylome accompany the promotion of soybean root growth by rhizobia under cadmium stress. Journal of Hazardous Materials. 2022;441:129843. DOI: 10.1016/j.jhazmat.2022.129843
  109. 109. Zhang J, Xu Y, Cao T, Chen J, Rosen BP, Zhao FJ. Arsenic methylation by a genetically engineered rhizobium-legume symbiont. Plant and Soil. 2017;416(1-2):259-269. DOI: 10.1007/s11104-017-3207-z
  110. 110. Cruz de Carvalho MH. Drought stress and reactive oxygen species: Production, scavenging and signaling. Plant Signaling & Behavior. 2008;3(3):156-165. DOI: 10.4161/psb.3.3.5536
  111. 111. Sarker U, Oba S. Drought stress effects on growth, ros markers, compatible solutes, phenolics, flavonoids, and antioxidant activity in Amaranthus tricolor. Applied Biochemistry and Biotechnology. 2018;186(4):999-1016. DOI: 10.1007/s12010-018-2784-5
  112. 112. You J, Chan Z. ROS regulation during abiotic stress responses in crop plants. Frontiers in Plant Science. 2015;6:1092. DOI: 10.3389/fpls.2015.01092
  113. 113. Zahran HH, Sprent JI. Effects of sodium chloride and polyethylene glycol on root-hair infection and nodulation of Vicia faba L. plants by rhizobium leguminosarum. Planta. 1986;167(3):303-309. DOI: 10.1007/bf00391332
  114. 114. Tu JC. Effect of salinity on rhizobium-root-hair interaction, nodulation and growth of soybean. Canadian Journal of Plant Science. 1981;61(2):231-239. DOI: 10.4141/cjps81-035
  115. 115. Borucki W, Sujkowska M. The effects of sodium chloride-salinity upon growth, nodulation, and root nodule structure of pea (Pisum sativum L.) plants. Acta Physiologiae Plantarum. 2007;30(3):293-301. DOI: 10.1007/s11738-007-0120-8
  116. 116. L'Taief B, Sifi B, Zaman-Allah M, Drevon JJ, Lachaâl M. Effect of salinity on root-nodule conductance to the oxygen diffusion in the Cicer arietinum-Mesorhizobium ciceri symbiosis. Journal of Plant Physiology. 2007;164(8):1028-1036. DOI: 10.1016/j.jplph.2006.05.016
  117. 117. Chien CT, Maundu J, Cavaness J, Dandurand LM, Orser CS. Characterization of salt-tolerant and salt-sensitive mutants of rhizobium leguminosarum biovar viciae strain C1204b. FEMS Microbiology Letters. 1992;69(2):135-140. DOI: 10.1016/0378-1097(92)90617-w
  118. 118. Dong R, Zhang J, Huan H, Bai C, Chen Z, Liu G. High salt tolerance of a Bradyrhizobium strain and its promotion of the growth of Stylosanthes guianensis. International Journal of Molecular Sciences. 2017;18(8):1625. DOI: 10.3390/ijms18081625
  119. 119. Franzini VI, Azcón R, Ruiz-Lozano JM, Aroca R. Rhizobial symbiosis modifies root hydraulic properties in bean plants under non-stressed and salinity-stressed conditions. Planta. 2019;249(4):1207-1215. DOI: 10.1007/s00425-018-03076-0
  120. 120. Chakraborty S, Harris JM. At the crossroads of salinity and rhizobium-legume symbiosis. Molecular Plant-Microbe Interactions. 2022;35(7):540-553. DOI: 10.1094/mpmi-09-21-0231-fi
  121. 121. Miransari M, Balakrishnan P, Smith D, Mackenzie AF, Bahrami HA, Malakouti MJ, et al. Overcoming the stressful effect of low pH on soybean root hair curling using lipochitooligosacharides. Communications in Soil Science and Plant Analysis. 2006;37(7-8):1103-1110. DOI: 10.1080/00103620600586391
  122. 122. Richardson AE, Simpson RJ, Djordjevic MA, Rolfe BG. Expression of nodulation genes in rhizobium leguminosarum biovar trifolii is affected by low pH and by Ca and Al ions. Applied and Environmental Microbiology. 1988;54(10):2541-2548. DOI: 10.1128/aem.54.10.2541-2548.1988
  123. 123. Checcucci A, Marchetti M. The rhizosphere talk show: The rhizobia on stage. Frontiers in Agronomy. 2020;2. DOI: 10.3389/fagro.2020.591494
  124. 124. Han Q , Ma Q , Chen Y, Tian B, Xu L, Bai Y, et al. Variation in rhizosphere microbial communities and its association with the symbiotic efficiency of rhizobia in soybean. The ISME Journal. 2020;14(8):1915-1928. DOI: 10.1038/s41396-020-0648-9
  125. 125. Mehboob I, Naveed M, Zahir ZA, Sessitsch A. Potential of rhizosphere bacteria for improving rhizobium-legume symbiosis. In: Arora N, editor. Plant Microbe Symbiosis: Fundamentals and Advances. New Delhi: Springer; 2013:305-349. DOI: 10.1007/978-81-322-1287-4_12. ISBN: 978-81-322-1286-7
  126. 126. Kiss E, Huguet T, Poinsot V, Batut J. The typA gene is required for stress adaptation as well as for symbiosis of Sinorhizobium meliloti 1021 with certain Medicago truncatula lines. Molecular Plant-Microbe Interactions. 2004;17(3):235-244. DOI: 10.1094/mpmi.2004.17.3.235
  127. 127. Martini MC, Vacca C, Torres Tejerizo GA, Draghi WO, Pistorio M, Lozano MJ, et al. ubiF is involved in acid stress tolerance and symbiotic competitiveness in rhizobium favelukesii LPU83. Brazilian Journal of Microbiology. 2022;53(3):1633-1643. DOI: 10.1007/s42770-022-00780-8
  128. 128. Reeve WG, Bräu L, Castelli J, Garau G, Sohlenkamp C, Geiger O, et al. The Sinorhizobium medicae WSM419 lpiA gene is transcriptionally activated by FsrR and required to enhance survival in lethal acid conditions. Microbiology (Reading). 2006;152(Pt 10):3049-3059. DOI: 10.1099/mic.0.28764-0
  129. 129. Tiwari S, Sarangi BK, Thul ST. Identification of arsenic resistant endophytic bacteria from Pteris vittata roots and characterization for arsenic remediation application. Journal of Environmental Management. 2016;180:359-365. DOI: 10.1016/j.jenvman.2016.05.029
  130. 130. Vinuesa P, Neumann-Silkow F, Pacios-Bras C, Spaink HP, Martínez-Romero E, Werner D. Genetic analysis of a pH-regulated operon from rhizobium tropici CIAT899 involved in acid tolerance and nodulation competitiveness. Molecular Plant-Microbe Interactions. 2003;16(2):159-168. DOI: 10.1094/mpmi.2003.16.2.159
  131. 131. Plett JM, Solomon J, Snijders F, Marlow-Conway J, Plett KL, Bithell SL. Order of microbial succession affects rhizobia-mediated biocontrol efforts against Phytophthora root rot. Microbiological Research. 2021;242:126628. DOI: 10.1016/j.micres.2020.126628
  132. 132. Tewari S, Sharma S. Rhizobial-metabolite based biocontrol of fusarium wilt in pigeon pea. Microbial Pathogenesis. 2020;147:104278. DOI: 10.1016/j.micpath.2020.104278
  133. 133. Parveen G, Noreen R, Shafique HA, Sultana V, Ehteshamul-Haque S, Athar M. Role of rhizobia in suppressing the root diseases of soybean under soil amendment. Planta Daninha. 2019;37. DOI: 10.1590/s0100-83582019370100038
  134. 134. Paau AS. Improvement of rhizobium inoculants by mutation, genetic engineering and formulation. Biotechnology Advances. 1991;9(2):173-184. DOI: 10.1016/0734-9750(91)90002-D
  135. 135. Geetha SJ, Joshi SJ. Engineering rhizobial bioinoculants: A strategy to improve iron nutrition. Scientific World Journal. 2013;2013:315890. DOI: 10.1155/2013/315890
  136. 136. Kibido T, Kunert K, Makgopa M, Greve M, Vorster J. Improvement of rhizobium-soybean symbiosis and nitrogen fixation under drought. Food and Energy Security. 2020;9(1):e177. DOI: 10.1002/fes3.177
  137. 137. Pii Y, Crimi M, Cremonese G, Spena A, Pandolfini T. Auxin and nitric oxide control indeterminate nodule formation. BMC Plant Biology. 2007;7:21. DOI: 10.1186/1471-2229-7-21
  138. 138. Ali S, Kim WC. Plant growth promotion under water: Decrease of waterlogging-induced acc and ethylene levels by acc deaminase-producing bacteria. Frontiers in Microbiology. 2018;9:1096. DOI: 10.3389/fmicb.2018.01096
  139. 139. Bosworth AH, Williams MK, Albrecht KA, Kwiatkowski R, Beynon J, Hankinson TR, et al. Alfalfa yield response to inoculation with recombinant strains of rhizobium meliloti with an extra copy of dctABD and/or modified nifA expression. Applied and Environmental Microbiology. 1994;60(10):3815-3832. DOI: 10.1128/aem.60.10.3815-3832.1994
  140. 140. Ghosh P, Adolphsen KN, Yurgel SN, Kahn ML. Sinorhizobium medicae WSM419 genes that improve Symbiosis between Sinorhizobium meliloti Rm1021 and Medicago truncatula Jemalong A17 and in other symbiosis systems. Applied and Environmental Microbiology. 2021;87(15):e0300420. DOI: 10.1128/aem.03004-20
  141. 141. Nascimento F, Brígido C, Alho L, Glick BR, Oliveira S. Enhanced chickpea growth-promotion ability of a Mesorhizobium strain expressing an exogenous ACC deaminase gene. Plant and Soil. 2011;353(1-2):221-230. DOI: 10.1007/s11104-011-1025-2
  142. 142. Saad MM, Michalet S, Fossou R, Putnik-Delić M, Crèvecoeur M, Meyer J, et al. Loss of nifq leads to accumulation of porphyrins and altered metal-homeostasis in nitrogen-fixing symbioses. Molecular Plant-Microbe Interactions. 2019;32(2):208-216. DOI: 10.1094/mpmi-07-18-0188-r
  143. 143. Hirsch PR, Spokes JD. Survival and dispersion of genetically modified rhizobia in the field and genetic interactions with native strains. FEMS Microbiology Ecology. 1994;15(1-2):147-159. DOI: 10.1111/j.1574-6941.1994.tb00239.x
  144. 144. Andrews M, De Meyer S, James EK, Stępkowski T, Hodge S, Simon MF, et al. Horizontal transfer of Symbiosis genes within and between Rhizobial genera: Occurrence and importance. Genes (Basel). 2018;9(7):321. DOI: 10.3390/genes9070321
  145. 145. Epstein B, Tiffin P. Comparative genomics reveals high rates of horizontal transfer and strong purifying selection on rhizobial symbiosis genes. Proceedings of the Biological Sciences. 1942;2021(288):20201804. DOI: 10.1098/rspb.2020.1804

Written By

Jemaneh Z. Habtewold and Ravinder K. Goyal

Submitted: 07 December 2022 Reviewed: 12 December 2022 Published: 01 February 2023