Next Article in Journal
Phytochemical Analysis, Antioxidant, Analgesic, Anti-Inflammatory, Hemagglutinin and Hemolytic Activities of Chemically Characterized Extracts from Origanum grosii (L.) and Thymus pallidus (L.)
Next Article in Special Issue
NaCl Modifies Biochemical Traits in Bacterial Endophytes Isolated from Halophytes: Towards Salinity Stress Mitigation Using Consortia
Previous Article in Journal
Melatonin Modulates Tomato Root Morphology by Regulating Key Genes and Endogenous Hormones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Microbes in Alleviating Crop Drought Stress: A Review

1
Engineering and Technical Center for Modern Horticulture, Jiangsu Vocational College of Agriculture and Forestry, Jurong 212400, China
2
Department of Agronomy and Horticulture, Jiangsu Vocational College of Agriculture and Forestry, Jurong 212400, China
3
Nanjing Institute of Agricultural Sciences in Jiangsu Hilly Area, Nanjing 210014, China
*
Author to whom correspondence should be addressed.
Plants 2024, 13(3), 384; https://doi.org/10.3390/plants13030384
Submission received: 29 November 2023 / Revised: 15 December 2023 / Accepted: 18 December 2023 / Published: 27 January 2024
(This article belongs to the Special Issue Role of Microbes in Alleviating Abiotic Stress in Plants)

Abstract

:
Drought stress is an annual global phenomenon that has devastating effects on crop production, so numerous studies have been conducted to improve crop drought resistance. Plant-associated microbiota play a crucial role in crop health and growth; however, we have a limited understanding of the key processes involved in microbiome-induced crop adaptation to drought stress. In this review, we summarize the adverse effects of drought stress on crop growth in terms of germination, photosynthesis, nutrient uptake, biomass, and yield, with a focus on the response of soil microbial communities to drought stress and plant-microbe interactions under drought stress. Moreover, we review the morpho-physiological, biochemical, and molecular mechanisms underlying the mitigation effect of microbes on crop drought stress. Finally, we highlight future research directions, including the characterization of specific rhizosphere microbiome species with corresponding root exudates and the efficiency of rhizobacteria inoculants under drought conditions. Such research will advance our understanding of the complex interactions between crops and microbes and improve crop resistance to drought stress through the application of beneficial drought-adaptive microbes.

1. Introduction

Drought is currently one of the most serious global issues threatening crop growth and productivity, which occurs every year and has devastating effects on crop yields. A meta-analysis found that crops subjected to drought stress suffered a 48% reduction in yield and concluded that global crop losses from drought stress likely exceeded those from all other abiotic stresses combined [1,2]. Furthermore, a continuous increase in global drought severity is predicted for the future, generating serious plant growth problems for over 50% of arable land by 2050 [3,4].
Water shortages have attracted substantial concern and stimulated growing research on plant drought resistance. Under drought stress, plant morphological, physiological, and biochemical processes are adversely affected by inhibited photosynthesis, cell elongation, cell division, and cell turgor pressure. Plants generally adopt three strategies to overcome drought stress: (1) escape, for example, by accelerating or slowing the switch from vegetative to reproductive growth for avoiding complete abortion under the condition of severe drought stress; (2) avoidance via modulation of the root system architecture, adaptation of leaf traits, and the alteration of photosynthesis; and (3) tolerance, including osmotic adjustment, activation of antioxidant defense systems, and responses to phytohormones and chlorophyll content [4,5].
Plant-associated microbiota play an important role in crop growth and have the potential to alleviate the adverse impacts of drought stress on crops in a sustainable manner [6]. In recent years, rapid theoretical and technological advances in soil microbiology have elucidated the effects of the soil microbiome on plant drought resistance. However, our understanding of the key processes underlying microbiome-induced crop adaptations under drought stress remains limited because of the complexities inherent in the rhizosphere and plant cellular systems. Therefore, in this review, we highlight recent research advances related to the following aspects: (1) the adverse impacts of drought stress on crop growth, such as germination, photosynthesis, nutrient uptake, biomass, and yield; (2) the response of soil microbial communities to drought stress; (3) plant–microbe interactions under drought stress; and (4) the morpho-physiological, biochemical, and molecular mechanisms underlying the microbial mitigation of drought stress. Finally, we discuss future research directions to advance our understanding of the complex interactions between soil, crops, and microbes, which is essential for developing microbial biofertilizers as a biologically potent strategy for microbe-mediated drought tolerance.

2. Adverse Impacts of Drought Stress on Crop Growth

Under conditions of water stress, crops typically exhibit suboptimal growth, reduced leaf area, increased root-to-shoot ratios, and premature senescence. Prolonged drought adversely affects numerous plant cellular functions, including nutrient uptake and assimilation, photosynthesis, protein synthesis, and the accumulation of reactive oxygen species (ROS). An overview of the various detrimental morphological and physiological effects of drought stress on crop growth is depicted in Figure 1.

2.1. Germination

Successful seed germination is crucial for seedling establishment and subsequent growth in the field. Seed germination and early seedling growth are the most sensitive stages to water scarcity in many crops, with drought stress negatively impacting parameters such as seed germination index, energy allocation, and seed vigor index across various crop species. The germination of non-dormant seeds unfolds in three distinct phases when a desiccated seed is immersed in water. Phase I of the germination process involves imbibition and water absorption. However, under drought stress, the water potential gradient between the inside and outside of the seed diminishes, impeding water infiltration into the seed coat and subsequent water absorption, ultimately resulting in reduced seedling vigor and inhibited germination [7]. In Phase II, which is known as the lag phase, water uptake ceases, and reserve mobilization commences. Drought stress decreases the activity of hydrolytic enzymes such as amylase, which catalyzes the breakdown of starch into simpler sugars, thereby prolonging the time required for carbohydrate hydrolysis and resulting in failed seed germination [8,9,10,11]. Plant hormones, such as gibberellic acid (GA) and abscisic acid (ABA), play pivotal roles in seed germination; GA activates hydrolytic enzymes, while ABA exerts an opposing influence. Water deficiency triggers an increase in ABA levels but a decrease in GA levels, thus failing to stimulate the synthesis of hydrolytic enzymes and impeding reserve mobilization [12,13,14]. Beyond hydrolytic enzyme activity, the activities of other enzymes such as α-amylase, β-amylase, and proteases are also adversely affected by water deficiency, further hindering seed reserve mobilization [15,16]. Phase III is marked by radicle emergence from the seed coat [8]. In the case of common beans, radicle emergence remains unaffected when drought stress is solely applied during Phase II, whereas drought stress applied after Phase II delays radicle emergence [17].
Drought stress results in the excessive generation of ROS, including superoxide anions (·O2), hydrogen peroxide (H2O2), and hydroxyl radicals (OH·), the accumulation of which leads to DNA or RNA damage, enzyme inactivation, ionic imbalances, and protein denaturation, and ultimately manifests as various physiological and biochemical disruptions in plants [18,19]. An efficient antioxidative defense system in plants regulates the excessive production of ROS through the action of several antioxidant enzymes, including superoxide dismutase, catalase, and peroxidase. However, under drought stress, ROS production surpasses the capacity of the antioxidant defense system owing to impaired enzymatic activity, rendering it incapable of scavenging the elevated ROS levels [20,21]. Consequently, impaired metabolic activity, reduced respiration, and diminished ATP production culminate in the failure of seed germination [22].

2.2. Nutrient Metabolism

Nutrient availability and equilibrium are indispensable for optimizing crop growth and yields. Typically, the negative effects of drought stress on crop growth arise from disruptions in nutrient absorption and utilization. Reduced nutrient absorption under drought stress is partially attributed to a decreased nutrient supply via mineralization and an inhibited nutrient movement through diffusion and mass flow, resulting in stunted crop growth, decreased leaf areas, reduced transpiration rates, and a diminished nutrient uptake [23,24,25]. Roots play a vital role in nutrient uptake in crops; under soil water deficits, plants dynamically adapt by modifying their root system architecture through adjustments in root growth patterns [26,27]. While some plants exhibit robust root growth in the early stages of drought stress to access deep soil water, severe soil water deficits can reduce root elongation, branching, and root tip development [5]. Thus, restricted root growth under severe drought stress offers another explanation for reduced nutrient absorption [7]. Additionally, alterations in hormonal concentrations, such as an increase in jasmonic acid, can also suppress nutrient uptake [28].
Limited nutrient translocation under drought stress results from reduced transpiration rates, inhibited active transportation, and decreased membrane permeability [29,30,31]. The expression of various nutrient transporters and channels under drought stress, including nitrate transporter 1.5 (NRT1.5), the ammonium transporter (AMT) family, the phosphate transporter/phosphate transporter family (PHT/PHO/PT) family, and stelar K+ outward rectifying (SKOR), represents a critical parameter reflecting the degree of drought tolerance in crops [32]. Furthermore, drought stress compromises the nutrient-assimilation-related enzymes activities [33,34], and the reduced energy availability for nutrient assimilation, such as NO3, NH4+, PO43−, and SO42−, also leads to a suboptimal nutrient status under drought stress [35].

2.3. Photosynthesis

Chlorophyll, as the primary photosynthetic pigment, plays a pivotal role in determining photosynthetic capacity and consequently crop yield. A decreased chlorophyll content is commonly associated with drought exposure in crops such as wheat [36], quinoa [37], soybean [38], and canola [39]. Genotypes with a higher chlorophyll content consistently exhibit superior photosynthetic abilities and, consequently, higher crop yields [40]. A reduced chlorophyll content is partly attributable to the reduced accessibility to essential nutrients such as N, Fe, and Mg, which are critical for chlorophyll synthesis [41]. The accumulation of ROS under drought stress triggers the breakdown of chloroplasts, resulting in the breakdown of chlorophyll precursors and increased chlorophyllase activity [42]. Furthermore, an imbalance between light capture and utilization, resulting from changes in the chlorophyll-to-carotenoid ratio, adversely affects photoreaction rates [18,43]. Varied responses of carotenoids to water deficiency have been reported in the literature, with some studies noting a decreased carotenoid content in plants such as wheat, sorghum, and sunflower [44,45,46,47,48,49], with others indicating enhanced carotenoid contents [38,50,51]. These divergent findings likely arise from differences in species, drought stress duration, and intensity. Carotenoid contents are likely to reduce and increase under moderate and severe drought stress, respectively, where they play a crucial role in the plant antioxidant defense system by mitigating oxidative damages caused by drought stress.
Water stress affects both the light and dark reactions, consequently reducing the accumulation of photosynthetic products. The reduced stomatal conductance under water stress, driven by an elevated ABA content in guard cells, has been extensively studied [52,53,54,55,56]. ABA-mediated stomatal closure represents the primary response of plants to drought stress, serving as a mechanism to reduce water loss through transpiration, thereby preserving hydraulic function but limiting CO2 intake, ultimately leading to decreased photosynthesis rates [4,56,57,58,59].
Crop photosynthesis is also constrained by mesophyll conductance. Non-stomatal limitations of photosynthesis (mesophyll conductance or Rubisco carboxylation capability) are substantially more pronounced under water stress conditions relative to stomatal limitations [60]. Mesophyll conductance primarily depends on leaf anatomical characteristics such as leaf and cell wall thickness, and chloroplast morphology [61]. A reduced mesophyll conductance and, consequently, lower photosynthesis rates largely depend on crop species, stress duration, and intensity [62]. Roig-Oliver et al. [63] concluded that the variations in cell wall thickness significantly correlate with mesophyll conductance under both short- and long-term water deficit stress, whereas other research suggests that a reduced chloroplast surface area exposed to leaf intercellular air spaces per leaf area (Sc/S), influenced by chloroplast size, number, and arrangement, contributes to a diminished mesophyll conductance under drought stress in rice, wheat, and cotton [64,65,66,67,68]. Furthermore, reduced aquaporin and carbonic anhydrase activities play a role in reduced CO2 diffusion within the mesophyll under drought stress [67,69,70]. Insufficient chloroplastic CO2 concentrations during water scarcity lead to the accumulation of reduced photosynthetic electron transport components, potentially influencing molecular oxygen generation and resulting in ROS production [71]. ROS overaccumulation induces the oxidation of chloroplast proteins, DNA, and lipids [7], along with the reduction in enzymes essential for carbon assimilation and utilization [72,73]. For example, the activity and abundance of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) dramatically decline under water deficiency, negatively impacting photosynthesis [49].

2.4. Biomass and Yield

Drought represents a significant factor contributing to reduced yields and food shortages [74]. Insufficient water during both the early vegetative and reproductive stages substantially hampers crop biomass and yield. Limited water and nutrient uptake and reduced photosynthetic rates impede cell expansion, resulting in a diminished leaf area and restricted internode elongation and plant height [43]. Furthermore, during stem elongation, vegetative growth predominantly determines the final yield, and reduced carbon assimilation leads to impaired flower and grain development [75]. Drought exerts a more pronounced detrimental effect during the reproductive phase, wherein a water deficit not only hinders flower initiation, and ovary and embryo development, but also impacts pollen development, flowering, fertilization, and grain development [76]. Key crop yield components are significantly and negatively affected by water scarcity in crops, such as rice, wheat, cotton, potato, maize, soybean, and sugar beet [18,77,78,79,80,81]. By a meta-analysis, Zhang et al. [82] reported that wheat and rice yields declined by 27.5% and 25.4% under drought stress, respectively.

3. Plant Microbiome Response to Drought Stress

The microbiome, which is considered as the second genome of the plant, is critical for plant growth, especially in unfavorable environments. Drought-triggered changes in crop fungal communities have been studied extensively. For example, large shifts in root-associated fungal communities have been reported under drought conditions in castor bean, with Fusarium, Chytridiomycota, Ascomycota, and Basidiomycota dominating the rhizosphere and bulk soil [83]. Compared with bulk soil and the rhizosphere, the grapevine root endosphere compartment shows the greatest divergence in fungal composition under severe drought stress, with the Chao1 and Shannon diversity of the fungal communities in the root being significantly lower than those in the absence of water deficiency. A significant enrichment in arbuscular mycorrhizal fungus (AMF) Funneliformis was also identified within the roots during drought, which is predominantly attributed to decreased P availability in drought soil [84,85]. Moreover, specific fungal species have been identified in wheat roots under drought stress; for example, Trichoderma longibrachiatum and T. velutinum are only identified under drought stress, whereas Zopfiella sp., M. hedericola, A. verrucaria, G. radicicola, and A. salicis are observed specifically in irrigated plant groups [86]. Changes in the composition of root-associated fungal communities and increased fungal biodiversity in rice plants have also been reported, with the majority of identified OTUs belonging to the Pezizomycotina subphylum [87]. The response of phyllosphere fungal communities to water stress has been studied in tomatoes, whereby drought stress increases the evenness and therefore the Shannon diversity index of fungal communities, and the relative abundance of dominant fungal taxa is also reduced by water deficiency [88]. These results verify that drought stress restricts compartment-specific fungal communities. Using sorghum as a research system, Gao et al. [89] concluded that, under drought stress, the stochastic force was the major force affecting fungal community assembly in leaves and roots. Moreover, they demonstrated that the host compartment was the strongest driving factor influencing fungal assembly, followed by the timing of plant development, and finally the plant genotype. Azarbad et al. [90] concluded that the effect of water stress on fungal diversity depends on the genotype and soil history, as shown by increased fungal diversity in the rhizosphere of wheat growing in soils with a history of water stress; in contrast, no significant changes in rhizosphere fungal α-diversity were found in soils, which never suffers from water stress.
Wheat grown in soils under water stress harbors more fungi and fewer bacteria [91]. Furthermore, bacterial biodiversity is decreased in the soil but increased in the root endosphere [92]. In rice plants, Si et al. [92] found that drought exerts a negligible effect on the alpha diversity of rhizosphere bacterial communities, but substantially enriches Actinobacteria and decreases Firmicutes [93]. Similarly, Santos-Medellin et al. [94] concluded that drought alters bacterial composition in the endosphere and rhizosphere compartments of rice, characterized by an enrichment of Actinobacteria and Chloroflexi but a depletion of Acidobacteria and Deltaproteobacteria. Specifically, the relative abundances of Actinobacteria and Acidobacteria were increased in peanut seedlings and podding stages under drought stress, whereas the relative abundances of Cyanobacteria and Gemmatimonadetes were increased in the flowering stage [95]. Furthermore, in millet plants, Simmons et al. [96] proved that drought intensity is correlated with the enrichment level of Actinobacteria. In conclusion, Actinobacteria enrichment within drought-stressed root microbiomes is strongly conserved among evolutionarily diverse plant species. Moreover, decreases in the phyla Proteobacteria and Verrucomicrobia, as well as increases in the ratio of Gram-positive to Gram-negative bacteria, are also frequently observed under drought conditions [93,97,98]. Drought significantly restructures the sorghum root microbiome composition and functionality during its early stage, which coincides with an increased abundance and activity of monodermal bacteria [99]. The taxonomic patterns in the phyllosphere are totally different from those detected in soil and rhizosphere studies. In the tomato phyllosphere, water stress reduces both leaf bacterial richness and the Shannon index, while simultaneously enriching several Proteobacteria and Bacteroidetes families and depleting several Actinobacteria families [88].

4. Plant–Microbe Interaction under Drought Stress

Changes in crop morphology and metabolism under drought conditions govern the responses of microbial communities [100]. Drought stress induces an increase in root depth and density and therefore an increase in the root/shoot ratio, which affects the composition and distribution of microbiomes and vice versa [19]. Considering the inhibition of photosynthetic activity due to stomatal closure, a previous study found that the abundance of 163 metabolites changed significantly during water stress in soybean roots [101]. Plants regulate their rhizosphere composition to promote the growth of microorganisms and improve plant adaptability in a given ecosystem, with root exudates being the ultimate designers of the rhizosphere microbiome [102,103]. In general, root exudates, including soluble sugars, amino acids, and organic acids, increase with drought stress intensity [104]. Altered carbon exudation through rhizodeposition disrupts microbial processes and shapes root-associated microbial communities [105,106]. Preece and Peñuelas [105] concluded that the effects on rhizodeposition are strongly species-dependent; therefore, the impacts on soil communities also vary. Among the various classes of compounds, the glycolysis intermediate glycerol-3-phosphate, which is a precursor for peptidoglycan biosynthesis and cell wall formation in bacteria [107], accumulates significantly under water deficit stress in sorghum [99]. Strong associations between rhizosphere bacterial communities and root exudates have also been confirmed in rice plants subjected to drought stress. Specifically, Streptomyces exhibits a positive correlation with ABA and a negative correlation with jasmonic acid, whereas some members of Actinobacteria (Conexibacter, Gaiella, Marmoricola, and Nocardioides) are positively associated with L-threonine, L-threonine, L-valine, and L-tryptophan [108].
Moreover, the dominant role of the host plant genotype in plant–microbe interactions under drought stress is associated with the root exudates. The host root zone, where enriched compounds are secreted by both plants and microorganisms, plays a key role in maintaining plant–microbe interactions [109]. For example, the abundance of OTUs and more saprotrophs enriched in the rhizoplane and rhizosphere of drought-tolerant species is far greater than that in the sensitive species in sugarcane, which contributes to enhanced drought resistance by altering the ABA pathways and, therefore, the root exudates [110]. In a drought-tolerant rice plant genotype, an increase in the relative abundance of Bacillus was found to depend on upregulated amino acid exudation [108].

5. Mechanisms Underlying the Mitigation Effect of Microbes on Crop Drought Stress

Under water stress conditions, crop systems actively maintain their physiological water balance either by increasing water uptake or reducing water loss through closing the stomata. Moreover, osmotic regulation is also one of the key measures for plants to cope with drought [4]. Activated stress response pathways regulated by microbes are closely associated with changes in crop morpho-physiology, biochemical processes, and molecular responses (Table 1, Figure 2).

5.1. Morpho-Physiological Mechanisms

Longer roots are profitable for plants to obtain nutrients and water from more soil, and roots respond to changes in soil water content at both the cellular scale and with the entire root system architecture [4]. Under severe drought stress, inoculation with osmotolerant endophytic bacteria increases the root volume, area, and length in pearl millet [112]. Mycorrhizal fungal colonization also significantly increases the volume, number, and length of lateral roots, as well as the root hair density, length, and diameter under drought stress conditions in tea plants [111]. G. diazotrophicus ameliorates root growth under water-restricted conditions by increasing the root length and number of lateral roots in rice plants [116]. According to a meta-analysis, AM-inoculated plants have a 49% greater root dry weight than non-mycorrhizal plants under drought stress, which is the result of increased root length (37%), root surface (31%), and root volume (65%) [135]. Moreover, the coleoptiles of wheat seedlings inoculated with Azospirillum show wider xylem vessels and a less negative water potential under water stress, which explains their improved water status after inoculation [132].
Physiologically, plant-growth-promoting bacteria (PGPR) or AMF inoculation significantly ameliorates the negative effects of drought by increasing nutrient (N and P) uptake, use efficiency, and photosynthetic characteristics such as chlorophyll synthesis, the potential quantum efficiency of PSII (Fv/Fm), non-photochemical efficiency, photochemical efficiency, actual quantum yield, and photosynthesis rate in maize [113,117], chickpea [118], soybean [121], sorghum [123], and grapevines [114,135]. An increased content of compatible solutes, such as proline, sugars, and free amino acids, is also associated with preservation of the relative water content and, therefore, drought stress resistance [113,119,120,136].

5.2. Biochemical Mechanisms

An increase in the synthesis of ethylene from its immediate precursor 1-aminocyclopropane-1-carboxylate (ACC) [137], secreted by plants as a root exudate, has been observed in almost all plants growing under stress conditions. Inoculation with V. paradoxus 5C-2 improves growth, yield, and water-use efficiency during drought by amplifying the xylem ABA concentration and attenuating the xylem ACC concentration [43,134]. The combined effect of the three ACC-deaminase-producing rhizobacteria in Vigna mungo L. and Pisum sativum L results in decreased ACC accumulation and downregulated ACC-oxidase gene expression [122]. Bacillus subtilis provides tolerance to drought stress in wheat by increasing the indole-3-acetic acid content and counteracting the increased ACC content [124]. The application of endophytes or AMF in the alleviation of drought stress, either by producing ACC deaminase or increasing the content of hormones, such as auxin, brassinosteroids, ABA, indole-3-acetic acid, and GA, has also been clarified in soybean [115], mung bean [126], Capsicum annuum [125], pearl millet [112], Alhagi sparsifolia [138], tea [111], and wheat [133]. In the presence of Bacillus safensis, enhanced antioxidant responses occur because of the elevated activities of enzymes, including catalase, peroxidase, ascorbate peroxidase, superoxide dismutase, and glutathione reductase in wheat plants, which contributes to their outstanding resistance to water stress [130]. Activated antioxidant enzyme activity after AMF inoculation under drought stress has also been observed in Alhagi sparsifolia [138], wheat [129], and mung bean [131]. Moreover, reduced emissions of stress volatiles, including benzaldehyde, β-pinene, and geranyl acetone, are likely candidates responsible for drought stress mitigation in the early stress stages [129].

5.3. Molecular Mechanisms

After 1 h of drought stress induction, 194 genes with modified expression in B. phytofirmans PsJN colonizing non-stressed and drought-stressed potato plants were identified. These genes were upregulated in the transcriptome mostly in the following functional categories: cellular homeostasis, homeostatic processes, and cell redox homeostasis [128]. Moreover, the positive impact of PGPR inoculation on drought resistance is associated with the modulated expression of a regulatory component (CTR1) of the ethylene signaling pathway and the DREB2 transcription factor [124,131]. Similarly, the dehydration-responsive binding protein Gmdreb1a and the galactinol synthase Gmgols were strongly expressed in response to drought stress in the presence of Bacillus cereus in soybeans [121]. Except for DREB2A, the expression levels of genes involved in phytohormone biosynthesis (SbNCED, SbGA20oX, and SbYUC) and those coding for drought-responsive transcription factors (SbAP2 and SbSNAC1) are significantly higher in endophyte-inoculated plants than in uninoculated pearl millet plants subjected to severe drought stress [112]. Furthermore, the differential expression of genes involved in ROS scavenging (CAT, APX, GST), ethylene biosynthesis (ACO and ACS), salicylic acid (PR1), and jasmonate (MYC2) signaling has also been observed under drought conditions in PGPR-inoculated plants, in contrast to uninoculated control plants [131]. Colonization with T. harzianum can alter the expression of aquaporin and dehydrin genes in rice genotypes under drought stress, thereby contributing to the advancement in sustained crop productivity under drought stress [127]. Additionally, reverse transcription PCR analysis of rice roots revealed significant differences in the expression of root-development-modulating genes during the water restriction period, with the relative gene expression increasing 10–50-fold compared to uninoculated field-grown plants [116].

6. Future Perspectives

In this review, we have addressed the primary impacts of drought stress on crop growth and the intricate interplay between crops and microorganisms in drought-prone environments. We have systematically elucidated the underlying mechanisms governing how microorganisms can ameliorate the adverse effects of drought stress on crops. The body of evidence presented herein unequivocally highlights the profound influence of drought stress on crop responses, particularly in shaping the composition of root microbial communities, with root exudates emerging as pivotal contributors to this dynamic process. Nevertheless, a substantial research gap exists in the characterization of root exudates across diverse crop species. The complex milieu of hundreds of chemically complex primary and secondary metabolites exuded from roots, compounded by simultaneous releases by microbes, renders the identification of specific root exudates a formidable task.
Given the pivotal role of root exudates in modulating root microbial communities, the application of exudates holds significant promise for mitigating the adverse effects of drought stress on crop growth. Hence, sustained efforts are required to characterize and quantify root exudates under drought stress conditions.
Furthermore, microbial inoculants are the best possible candidates for promoting healthier and more sustainable food production while bolstering environmental friendliness and agricultural sustainability. A deeper exploration of distinct rhizobacterial microbes could enhance our understanding of their interconnectedness and their potential to fortify crops against drought stress. However, many questions remain unanswered. First, specific rhizosphere microbiome species designed using the corresponding root exudates are difficult to discriminate. Second, research into the efficiency of rhizobacteria inoculants under drought conditions is limited. Nevertheless, the application of beneficial drought-adaptive microbes to crops could be a sustainable solution enabling plants to withstand stress conditions, paving the way for improved economic yields, soil health, and fertility. Thus, considerable future research is required to identify optimal rhizobacterial strains and potential field candidates, as well as to develop more effective PGPR strains with longer shelf-life to ensure sustainable crop production in arid areas.

Author Contributions

Conceptualization, Z.G. and L.G.; methodology, C.H.; investigation, Y.G.; writing—original draft preparation, Z.G., J.Z., and G.T.; writing—review and editing, Z.G. and L.G.; supervision, L.G.; project administration, G.T.; funding acquisition, Z.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Jiangsu Vocational and Technical College of Agriculture and Forestry Youth Support Project (2021kj33), Basic Science (Natural Science) Research Project of Universities of Jiangsu Province (21KJB210009), and Yafu Technology and Service Innovation Major project (2023kj02).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cohen, I.; Zandalinas, S.I.; Huck, C.; Fritschi, F.B.; Mittler, R. Meta-analysis of drought and heat stress combination impact on crop yield and yield components. Physiol. Plant. 2021, 171, 66–76. [Google Scholar] [CrossRef] [PubMed]
  2. Alqudah, A.M.; Samarah, N.H.; Mullen, R.E. Drought Stress Effect on Crop Pollination, Seed Set, Yield and Quality. In Alternative Farming Systems, Biotechnology, Drought Stress and Ecological Fertilisation; Springer: Berlin/Heidelberg, Germany, 2011; pp. 193–213. [Google Scholar]
  3. Kasim, W.A.; Osman, M.E.; Omar, M.N.; Abd El-Daim, I.A.; Bejai, S.; Meijer, J. Control of Drought Stress in Wheat Using Plant-Growth-Promoting Bacteria. J. Plant Growth Regul. 2012, 32, 122–130. [Google Scholar] [CrossRef]
  4. Gupta, A.; Rico-Medina, A.; Caño-Delgado, A.I. The physiology of plant responses to drought. Science 2020, 368, 266–269. [Google Scholar] [CrossRef] [PubMed]
  5. Fang, Y.; Xiong, L. General mechanisms of drought response and their application in drought resistance improvement in plants. Cell. Mol. Life Sci. 2015, 72, 673–689. [Google Scholar] [CrossRef] [PubMed]
  6. Haichar, F.E.Z.; Cernava, T.; Liu, J.; Timm, C.M. Editorial: Novel Insights into the Response of the Plant Microbiome to Abiotic Factors. Front Plant Sci 2021, 12, 607874. [Google Scholar] [CrossRef] [PubMed]
  7. Batool, M.; El-Badri, A.M.; Hassan, M.U.; Yang, H.; Wang, C.; Yan, Z.; Kuai, J.; Wang, B.; Zhou, G. Drought Stress in Brassica napus: Effects, Tolerance Mechanisms, and Management Strategies. J. Plant Growth Regul. 2023, 42, 21–45. [Google Scholar] [CrossRef]
  8. Khan, M.N.; Fu, C.; Li, J.; Tao, Y.; Li, Y.; Hu, J.; Chen, L.; Khan, Z.; Wu, H.; Li, Z. Seed nanopriming: How do nanomaterials improve seed tolerance to salinity and drought? Chemosphere 2023, 310, 136911. [Google Scholar] [CrossRef]
  9. Damaris, R.N.; Lin, Z.; Yang, P.; He, D. The Rice Alpha-Amylase, Conserved Regulator of Seed Maturation and Germination. Int. J. Mol. Sci. 2019, 20, 450. [Google Scholar] [CrossRef]
  10. Yang, N.; Guo, X.; Wu, Y.; Hu, X.; Ma, Y.; Zhang, Y.; Wang, H.; Tang, Z. The inhibited seed germination by ABA and MeJA is associated with the disturbance of reserve utilizations in Astragalus membranaceus. J. Plant Interact. 2018, 13, 388–397. [Google Scholar] [CrossRef]
  11. Biju, S.; Fuentes, S.; Gupta, D. Silicon improves seed germination and alleviates drought stress in lentil crops by regulating osmolytes, hydrolytic enzymes and antioxidant defense system. Plant Physiol. Biochem. 2017, 119, 250–264. [Google Scholar] [CrossRef]
  12. Hassan, M.J.; Geng, W.; Zeng, W.; Raza, M.A.; Khan, I.; Iqbal, M.Z.; Peng, Y.; Zhu, Y.; Li, Z. Diethyl Aminoethyl Hexanoate Priming Ameliorates Seed Germination via Involvement in Hormonal Changes, Osmotic Adjustment, and Dehydrins Accumulation in White Clover under Drought Stress. Front. Plant Sci. 2021, 12, 709187. [Google Scholar] [CrossRef] [PubMed]
  13. El Moukhtari, A.; Ksiaa, M.; Zorrig, W.; Cabassa, C.; Abdelly, C.; Farissi, M.; Savoure, A. How Silicon Alleviates the Effect of Abiotic Stresses during Seed Germination: A Review. J. Plant Growth Regul. 2022, 42, 3323–3341. [Google Scholar] [CrossRef]
  14. Liu, J.; Hasanuzzaman, M.; Wen, H.; Zhang, J.; Peng, T.; Sun, H.; Zhao, Q. High temperature and drought stress cause abscisic acid and reactive oxygen species accumulation and suppress seed germination growth in rice. Protoplasma 2019, 256, 1217–1227. [Google Scholar] [CrossRef] [PubMed]
  15. Hameed, A.; Farooq, T.; Hameed, A.; Sheikh, M.A. Sodium nitroprusside mediated priming memory invokes water-deficit stress acclimation in wheat plants through physio-biochemical alterations. Plant Physiol. Biochem. 2021, 160, 329–340. [Google Scholar] [CrossRef] [PubMed]
  16. Martinez, M.; Gomez-Cabellos, S.; Gimenez, M.J.; Barro, F.; Diaz, I.; Diaz-Mendoza, M. Plant Proteases: From Key Enzymes in Germination to Allies for Fighting Human Gluten-Related Disorders. Front. Plant Sci. 2019, 10, 721. [Google Scholar] [CrossRef] [PubMed]
  17. Stephan, C.; Laval-Martin, D.L. Changes in NAD+ kinase activity during germination of Phaseolus vulgaris and P. acutifolius, and effects of drought stress. J. Plant Physiol. 2000, 157, 65–73. [Google Scholar] [CrossRef]
  18. Hussain, H.A.; Hussain, S.; Khaliq, A.; Ashraf, U.; Anjum, S.A.; Men, S.; Wang, L. Chilling and Drought Stresses in Crop Plants: Implications, Cross Talk, and Potential Management Opportunities. Front. Plant Sci. 2018, 9, 393. [Google Scholar] [CrossRef]
  19. Kang, J.; Peng, Y.; Xu, W. Crop Root Responses to Drought Stress: Molecular Mechanisms, Nutrient Regulations, and Interactions with Microorganisms in the Rhizosphere. Int. J. Mol. Sci. 2022, 23, 9310. [Google Scholar] [CrossRef]
  20. Khan, M.N.; Khan, Z.; Luo, T.; Liu, J.; Rizwan, M.; Zhang, J.; Xu, Z.; Wu, H.; Hu, L. Seed priming with gibberellic acid and melatonin in rapeseed: Consequences for improving yield and seed quality under drought and non-stress conditions. Ind. Crops Prod. 2020, 156, 112850. [Google Scholar] [CrossRef]
  21. Gill, S.S.; Tuteja, N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant Physiol. Biochem. 2010, 48, 909–930. [Google Scholar] [CrossRef]
  22. Wei, J.; Xu, L.; Shi, Y.; Cheng, T.; Tan, W.; Zhao, Y.; Li, C.; Yang, X.; Ouyang, L.; Wei, M.; et al. Transcriptome profile analysis of Indian mustard (Brassica juncea L.) during seed germination reveals the drought stress-induced genes associated with energy, hormone, and phenylpropanoid pathways. Plant Physiol. Biochem. 2023, 200, 107750. [Google Scholar] [CrossRef] [PubMed]
  23. Islam, M.; Sandhi, A. Heavy Metal and Drought Stress in Plants: The Role of Microbes—A Review. Gesunde Pflanz. 2022, 75, 695–708. [Google Scholar] [CrossRef]
  24. Rasheed, A.; Hassan, M.U.; Aamer, M.; Batool, M.; Fang, S.; Wu, Z.; Li, H. A critical review on the improvement of drought stress tolerance in rice (Oryza sativa L.). Not. Bot. Horti Agrobot. Cluj-Napoca 2020, 48, 1756–1788. [Google Scholar] [CrossRef]
  25. He, M.; Dijkstra, F.A. Drought effect on plant nitrogen and phosphorus: A meta-analysis. New Phytol. 2014, 204, 924–931. [Google Scholar] [CrossRef] [PubMed]
  26. Smith, S.; De Smet, I. Root system architecture: Insights from Arabidopsis and cereal crops. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2012, 367, 1441–1452. [Google Scholar] [CrossRef] [PubMed]
  27. Malamy, J.E. Intrinsic and environmental response pathways that regulate root system architecture. Plant Cell Environ. 2005, 28, 67–77. [Google Scholar] [CrossRef] [PubMed]
  28. Iqbal, S.; Iqbal, M.A.; Li, C.; Iqbal, A.; Abbas, R.N. Overviewing Drought and Heat Stress Amelioration—From Plant Responses to Microbe-Mediated Mitigation. Sustainability 2023, 15, 1671. [Google Scholar] [CrossRef]
  29. Shafiq, B.A.; Nawaz, F.; Majeed, S.; Aurangzaib, M.; Al Mamun, A.; Ahsan, M.; Ahmad, K.S.; Shehzad, M.A.; Ali, M.; Hashim, S.; et al. Sulfate-Based Fertilizers Regulate Nutrient Uptake, Photosynthetic Gas Exchange, and Enzymatic Antioxidants to Increase Sunflower Growth and Yield under Drought Stress. J. Soil Sci. Plant Nutr. 2021, 21, 2229–2241. [Google Scholar] [CrossRef]
  30. Meisser, M.; Vitra, A.; Deléglise, C.; Dubois, S.; Probo, M.; Mosimann, E.; Buttler, A.; Mariotte, P. Nutrient limitations induced by drought affect forage N and P differently in two permanent grasslands. Agric. Ecosyst. Environ. 2019, 280, 85–94. [Google Scholar] [CrossRef]
  31. Wang, Y.; Zhao, L.; Xie, X.; Huang, J.; Li, D.; Chen, W.; Zhu, A. Transcriptomic responses in Neolitsea sericea leaves under acute drought stress. Acta Physiol. Plant. 2018, 40, 2–22. [Google Scholar] [CrossRef]
  32. Barzana, G.; Carvajal, M. Genetic regulation of water and nutrient transport in water stress tolerance in roots. J. Biotechnol. 2020, 324, 134–142. [Google Scholar] [CrossRef] [PubMed]
  33. Ashraf, M.; Iram, A. Drought stress induced changes in some organic substances in nodules and other plant parts of two potential legumes differing in salt tolerance. Flora-Morphol. Distrib. Funct. Ecol. Plants 2005, 200, 535–546. [Google Scholar] [CrossRef]
  34. Gloser, V.; Dvorackova, M.; Mota, D.H.; Petrovic, B.; Gonzalez, P.; Geilfus, C.M. Early Changes in Nitrate Uptake and Assimilation under Drought in Relation to Transpiration. Front. Plant Sci. 2020, 11, 602065. [Google Scholar] [CrossRef] [PubMed]
  35. Farooq, M.; Wahid, A.; Kobayashi, N.; Fujita, D.; Basra, S.M.A. Plant drought stress: Effects, mechanisms and management. Agron. Sustain. Dev. 2009, 29, 185–212. [Google Scholar] [CrossRef]
  36. Wasaya, A.; Manzoor, S.; Yasir, T.A.; Sarwar, N.; Mubeen, K.; Ismail, I.A.; Raza, A.; Rehman, A.; Hossain, A.; El Sabagh, A. Evaluation of Fourteen Bread Wheat (Triticum aestivum L.) Genotypes by Observing Gas Exchange Parameters, Relative Water and Chlorophyll Content, and Yield Attributes under Drought Stress. Sustainability 2021, 13, 4799. [Google Scholar] [CrossRef]
  37. Saddiq, M.S.; Wang, X.; Iqbal, S.; Hafeez, M.B.; Khan, S.; Raza, A.; Iqbal, J.; Maqbool, M.M.; Fiaz, S.; Qazi, M.A.; et al. Effect of Water Stress on Grain Yield and Physiological Characters of Quinoa Genotypes. Agronomy 2021, 11, 1934. [Google Scholar] [CrossRef]
  38. Amani Machiani, M.; Javanmard, A.; Morshedloo, M.R.; Janmohammadi, M.; Maggi, F. Funneliformis mosseae Application Improves the Oil Quantity and Quality and Eco-physiological Characteristics of Soybean (Glycine max L.) under Water Stress Conditions. J. Soil Sci. Plant Nutr. 2021, 21, 3076–3090. [Google Scholar] [CrossRef]
  39. Zhang, X.; Lu, G.; Long, W.; Zou, X.; Li, F.; Nishio, T. Recent progress in drought and salt tolerance studies in Brassica crops. Breed. Sci. 2014, 64, 60–73. [Google Scholar] [CrossRef]
  40. Zhang, J.; Mason, A.S.; Wu, J.; Liu, S.; Zhang, X.; Luo, T.; Redden, R.; Batley, J.; Hu, L.; Yan, G. Identification of Putative Candidate Genes for Water Stress Tolerance in Canola (Brassica napus). Front. Plant Sci. 2015, 6, 1058. [Google Scholar] [CrossRef]
  41. Six, J. Plant nutrition for sustainable development and global health. Plant Soil 2011, 339, 1–2. [Google Scholar] [CrossRef]
  42. Mohammadi, H.; Amirikia, F.; Ghorbanpour, M.; Fatehi, F.; Hashempour, H. Salicylic acid induced changes in physiological traits and essential oil constituents in different ecotypes of Thymus kotschyanus and Thymus vulgaris under well-watered and water stress conditions. Ind. Crops Prod. 2019, 129, 561–574. [Google Scholar] [CrossRef]
  43. Zia, R.; Nawaz, M.S.; Siddique, M.J.; Hakim, S.; Imran, A. Plant survival under drought stress: Implications, adaptive responses, and integrated rhizosphere management strategy for stress mitigation. Microbiol. Res. 2021, 242, 126626. [Google Scholar] [CrossRef] [PubMed]
  44. Yadav, R.K.; Sangwan, R.S.; Sabir, F.; Srivastava, A.K.; Sangwan, N.S. Effect of prolonged water stress on specialized secondary metabolites, peltate glandular trichomes, and pathway gene expression in Artemisia annua L. Plant Physiol. Biochem. 2014, 74, 70–83. [Google Scholar] [CrossRef] [PubMed]
  45. Al Hassan, M.; Lopez-Gresa, M.D.P.; Boscaiu, M.; Vicente, O. Stress tolerance mechanisms in Juncus: Responses to salinity and drought in three Juncus species adapted to different natural environments. Funct. Plant Biol. 2016, 43, 949–960. [Google Scholar] [CrossRef] [PubMed]
  46. Hammad, S.A.R.; Ali, O.A.M. Physiological and biochemical studies on drought tolerance of wheat plants by application of amino acids and yeast extract. Ann. Agric. Sci. 2014, 59, 133–145. [Google Scholar] [CrossRef]
  47. Manivannan, P.; Rabert, G.A.; Rajasekar, M.; Somasundaram, R. Drought stress-induced modification on growth and pigments composition in different genotypes of Helianthus annuus L. Curr. Bot. 2014, 5, 7–13. [Google Scholar]
  48. Arivalagan, M.; Somasundaram, R. Effect of Propiconazole and Salicylic acid on the Growth and Photosynthetic Pigments Variations in Sorghum bicolour L. under Drought Condition. J. Plant Stress Physiol. 2015, 7, 17–23. [Google Scholar] [CrossRef]
  49. Hassan, M.U.; Aamer, M.; Chattha, M.U.; Haiying, T.; Shahzad, B.; Barbanti, L.; Nawaz, M.; Rasheed, A.; Afzal, A.; Liu, Y.; et al. The Critical Role of Zinc in Plants Facing the Drought Stress. Agriculture 2020, 10, 396. [Google Scholar] [CrossRef]
  50. Wijewardana, C.; Alsajri, F.A.; Irby, J.T.; Krutz, L.J.; Golden, B.; Henry, W.B.; Gao, W.; Reddy, K.R. Physiological assessment of water deficit in soybean using midday leaf water potential and spectral features. J. Plant Interact. 2019, 14, 533–543. [Google Scholar] [CrossRef]
  51. Gajanayake, B.; Reddy, K.R. Sweetpotato Responses to Mid- and Late-Season Soil Moisture Deficits. Crop Sci. 2016, 56, 1865–1877. [Google Scholar] [CrossRef]
  52. Elferjani, R.; Soolanayakanahally, R. Canola Responses to Drought, Heat, and Combined Stress: Shared and Specific Effects on Carbon Assimilation, Seed Yield, and Oil Composition. Front. Plant Sci. 2018, 9, 1224. [Google Scholar] [CrossRef] [PubMed]
  53. Correia, B.; Hancock, R.D.; Amaral, J.; Gomez-Cadenas, A.; Valledor, L.; Pinto, G. Combined Drought and Heat Activates Protective Responses in Eucalyptus globulus That Are Not Activated When Subjected to Drought or Heat Stress Alone. Front. Plant Sci. 2018, 9, 819. [Google Scholar] [CrossRef] [PubMed]
  54. Zandalinas, S.I.; Mittler, R.; Balfagon, D.; Arbona, V.; Gomez-Cadenas, A. Plant adaptations to the combination of drought and high temperatures. Physiol. Plant. 2018, 162, 2–12. [Google Scholar] [CrossRef] [PubMed]
  55. Fahad, S.; Bajwa, A.A.; Nazir, U.; Anjum, S.A.; Farooq, A.; Zohaib, A.; Sadia, S.; Nasim, W.; Adkins, S.; Saud, S.; et al. Crop Production under Drought and Heat Stress: Plant Responses and Management Options. Front. Plant Sci. 2017, 8, 1147. [Google Scholar] [CrossRef] [PubMed]
  56. Diksaityte, A.; Virsile, A.; Zaltauskaite, J.; Januskaitiene, I.; Juozapaitiene, G. Growth and photosynthetic responses in Brassica napus differ during stress and recovery periods when exposed to combined heat, drought and elevated CO2. Plant Physiol. Biochem. 2019, 142, 59–72. [Google Scholar] [CrossRef] [PubMed]
  57. Rodrigues, J.; Inze, D.; Nelissen, H.; Saibo, N.J.M. Source-Sink Regulation in Crops under Water Deficit. Trends Plant Sci. 2019, 24, 652–663. [Google Scholar] [CrossRef] [PubMed]
  58. Okamoto, M.; Peterson, F.C.; Defries, A.; Park, S.Y.; Endo, A.; Nambara, E.; Volkman, B.F.; Cutler, S.R. Activation of dimeric ABA receptors elicits guard cell closure, ABA-regulated gene expression, and drought tolerance. Proc. Natl. Acad. Sci. USA 2013, 110, 12132–12137. [Google Scholar] [CrossRef] [PubMed]
  59. Waseem, M.; Nie, Z.F.; Yao, G.Q.; Hasan, M.; Xiang, Y.; Fang, X.W. Dew absorption by leaf trichomes in Caragana korshinskii: An alternative water acquisition strategy for withstanding drought in arid environments. Physiol. Plant. 2021, 172, 528–539. [Google Scholar] [CrossRef]
  60. Li, S.-L.; Tan, T.-T.; Fan, Y.-F.; Raza, M.A.; Wang, Z.-L.; Wang, B.-B.; Zhang, J.-W.; Tan, X.-M.; Chen, P.; Shafiq, I.; et al. Responses of leaf stomatal and mesophyll conductance to abiotic stress factors. J. Integr. Agric. 2022, 21, 2787–2804. [Google Scholar] [CrossRef]
  61. Gao, L.; Lu, Z.; Ding, L.; Xie, K.; Wang, M.; Ling, N.; Guo, S. Anatomically induced changes in rice leaf mesophyll conductance explain the variation in photosynthetic nitrogen use efficiency under contrasting nitrogen supply. BMC Plant Biol. 2020, 20, 527. [Google Scholar] [CrossRef]
  62. Zou, J.; Hu, W.; Li, Y.; Zhu, H.; He, J.; Wang, Y.; Meng, Y.; Chen, B.; Zhao, W.; Wang, S.; et al. Leaf anatomical alterations reduce cotton’s mesophyll conductance under dynamic drought stress conditions. Plant J. 2022, 111, 391–405. [Google Scholar] [CrossRef] [PubMed]
  63. Roig-Oliver, M.; Bresta, P.; Nadal, M.; Liakopoulos, G.; Nikolopoulos, D.; Karabourniotis, G.; Bota, J.; Flexas, J. Cell wall composition and thickness affect mesophyll conductance to CO2 diffusion in Helianthus annuus under water deprivation. J. Exp. Bot. 2020, 71, 7198–7209. [Google Scholar] [CrossRef] [PubMed]
  64. Ouyang, W.; Struik, P.C.; Yin, X.; Yang, J. Stomatal conductance, mesophyll conductance, and transpiration efficiency in relation to leaf anatomy in rice and wheat genotypes under drought. J. Exp. Bot. 2017, 68, 5191–5205. [Google Scholar] [CrossRef] [PubMed]
  65. Terashima, I.; Hanba, Y.T.; Tholen, D.; Niinemets, Ü. Leaf functional anatomy in relation to photosynthesis. Plant Physiol. 2011, 155, 108–116. [Google Scholar] [CrossRef]
  66. Li, Y.; Yang, X.; Ren, B.; Shen, Q.; Guo, S. Why Nitrogen Use Efficiency Decreases under High Nitrogen Supply in Rice (Oryza sativa L.). Seedlings. J. Plant Growth Regul. 2012, 31, 47–52. [Google Scholar] [CrossRef]
  67. Han, J.; Lei, Z.; Zhang, Y.; Yi, X.; Zhang, W.; Zhang, Y. Drought-introduced variability of mesophyll conductance in Gossypium and its relationship with leaf anatomy. Physiol. Plant. 2019, 166, 873–887. [Google Scholar] [CrossRef] [PubMed]
  68. Tholen, D.; Boom, C.; Zhu, X.-G. Opinion: Prospects for improving photosynthesis by altering leaf anatomy. Plant Sci. 2012, 197, 92–101. [Google Scholar] [CrossRef] [PubMed]
  69. Kawase, M.; Hanba, Y.T.; Katsuhara, M. The photosynthetic response of tobacco plants overexpressing ice plant aquaporin McMIPB to a soil water deficit and high vapor pressure deficit. J. Plant Res. 2013, 126, 517–527. [Google Scholar] [CrossRef] [PubMed]
  70. Gao, L.; Lu, Z.; Ding, L.; Guo, J.; Wang, M.; Ling, N.; Guo, S.; Shen, Q. Role of Aquaporins in Determining Carbon and Nitrogen Status in Higher Plants. Int. J. Mol. Sci. 2018, 19, 35. [Google Scholar] [CrossRef]
  71. Basu, S.; Ramegowda, V.; Kumar, A.; Pereira, A. Plant adaptation to drought stress. F1000Research 2016, 5, 1554. [Google Scholar] [CrossRef]
  72. Ramachandra Reddy, A.; Chaitanya, K.V.; Vivekanandan, M. Drought-induced responses of photosynthesis and antioxidant metabolism in higher plants. J. Plant Physiol. 2004, 161, 1189–1202. [Google Scholar] [CrossRef] [PubMed]
  73. Majidia, M.M.; Rashidib, F.; Sharafi, Y. Physiological traits related to drought tolerance in Brassica. Int. J. Plant Prod. 2015, 9, 541–559. [Google Scholar]
  74. Al-Qubati, A.; Zhang, L.; Pyarali, K. Climatic drought impacts on key ecosystem services of a low mountain region in Germany. Environ. Monit. Assess. 2023, 195, 800. [Google Scholar] [CrossRef] [PubMed]
  75. Sinclair, T.R.; Jamieson, P.D. Grain number, wheat yield, and bottling beer: An analysis. Field Crops Res. 2006, 98, 60–67. [Google Scholar] [CrossRef]
  76. Barnabas, B.; Jager, K.; Feher, A. The effect of drought and heat stress on reproductive processes in cereals. Plant Cell Environ. 2008, 31, 11–38. [Google Scholar] [CrossRef] [PubMed]
  77. Palanog, A.D.; Swamy, B.P.M.; Shamsudin, N.A.A.; Dixit, S.; Hernandez, J.E.; Boromeo, T.H.; Cruz, P.C.S.; Kumar, A. Grain yield QTLs with consistent-effect under reproductive-stage drought stress in rice. Field Crops Res. 2014, 161, 46–54. [Google Scholar] [CrossRef]
  78. Kamara, A.Y.; Menkir, A.; Badu-Apraku, B.; Ibikunle, O. The influence of drought stress on growth, yield and yield components of selected maize genotypes. J. Agric. Sci. 2004, 141, 43–50. [Google Scholar] [CrossRef]
  79. Hu, W.; Snider, J.L.; Wang, H.; Zhou, Z.; Chastain, D.R.; Whitaker, J.; Perry, C.D.; Bourland, F.M. Water-induced variation in yield and quality can be explained by altered yield component contributions in field-grown cotton. Field Crops Res. 2018, 224, 139–147. [Google Scholar] [CrossRef]
  80. Diaz-Valencia, P.; Melgarejo, L.M.; Arcila, I.; Mosquera-Vasquez, T. Physiological, Biochemical and Yield-Component Responses of Solanum tuberosum L. Group Phureja Genotypes to a Water Deficit. Plants 2021, 10, 638. [Google Scholar] [CrossRef]
  81. Alavilli, H.; Yolcu, S.; Skorupa, M.; Aciksoz, S.B.; Asif, M. Salt and drought stress-mitigating approaches in sugar beet (Beta vulgaris L.) to improve its performance and yield. Planta 2023, 258, 30. [Google Scholar] [CrossRef]
  82. Zhang, J.; Zhang, S.; Cheng, M.; Jiang, H.; Zhang, X.; Peng, C.; Lu, X.; Zhang, M.; Jin, J. Effect of Drought on Agronomic Traits of Rice and Wheat: A Meta-Analysis. Int. J. Environ. Res. Public Health 2018, 15, 839. [Google Scholar] [CrossRef]
  83. Hereira-Pacheco, S.E.; Estrada-Torres, A.; Dendooven, L.; Navarro-Noya, Y.E. Shifts in root-associated fungal communities under drought conditions in Ricinus communis. Fungal Ecol. 2023, 63, 101225. [Google Scholar] [CrossRef]
  84. Carbone, M.J.; Alaniz, S.; Mondino, P.; Gelabert, M.; Eichmeier, A.; Tekielska, D.; Bujanda, R.; Gramaje, D. Drought Influences Fungal Community Dynamics in the Grapevine Rhizosphere and Root Microbiome. J Fungi 2021, 7, 686. [Google Scholar] [CrossRef]
  85. Augé, R.M. Water relations, drought and vesicular-arbuscular mycorrhizal symbiosis. Mycorrhiza 2001, 11, 3–42. [Google Scholar] [CrossRef]
  86. Salamon, S.; Mikolajczak, K.; Blaszczyk, L.; Ratajczak, K.; Sulewska, H. Changes in root-associated fungal communities in Triticum aestivum ssp. spelta L. and Triticum aestivum ssp. vulgare L. under drought stress and in various soil processing. PLoS ONE 2020, 15, e0240037. [Google Scholar] [CrossRef]
  87. Andreo-Jimenez, B.; Vandenkoornhuyse, P.; Le Van, A.; Heutinck, A.; Duhamel, M.; Kadam, N.; Jagadish, K.; Ruyter-Spira, C.; Bouwmeester, H. Plant host and drought shape the root associated fungal microbiota in rice. PeerJ 2019, 7, e7463. [Google Scholar] [CrossRef]
  88. Debray, R.; Socolar, Y.; Kaulbach, G.; Guzman, A.; Hernandez, C.A.; Curley, R.; Dhond, A.; Bowles, T.; Koskella, B. Water stress and disruption of mycorrhizas induce parallel shifts in phyllosphere microbiome composition. New Phytol. 2022, 234, 2018–2031. [Google Scholar] [CrossRef]
  89. Gao, C.; Montoya, L.; Xu, L.; Madera, M.; Hollingsworth, J.; Purdom, E.; Singan, V.; Vogel, J.; Hutmacher, R.B.; Dahlberg, J.A.; et al. Fungal community assembly in drought-stressed sorghum shows stochasticity, selection, and universal ecological dynamics. Nat. Commun. 2020, 11, 34. [Google Scholar] [CrossRef] [PubMed]
  90. Azarbad, H.; Tremblay, J.; Giard-Laliberte, C.; Bainard, L.D.; Yergeau, E. Four decades of soil water stress history together with host genotype constrain the response of the wheat microbiome to soil moisture. FEMS Microbiol. Ecol. 2020, 96, fiaa098. [Google Scholar] [CrossRef] [PubMed]
  91. Azarbad, H.; Constant, P.; Giard-Laliberté, C.; Bainard, L.D.; Yergeau, E. Water stress history and wheat genotype modulate rhizosphere microbial response to drought. Soil Biol. Biochem. 2018, 126, 228–236. [Google Scholar] [CrossRef]
  92. Si, J.; Froussart, E.; Viaene, T.; Vazquez-Castellanos, J.F.; Hamonts, K.; Tang, L.; Beirinckx, S.; De Keyser, A.; Deckers, T.; Amery, F.; et al. Interactions between soil compositions and the wheat root microbiome under drought stress: From an in silico to in planta perspective. Comput. Struct. Biotechnol. J. 2021, 19, 4235–4247. [Google Scholar] [CrossRef]
  93. Rosa, A.P.; Dias, T.; Mouazen, A.M.; Cruz, C.; Santana, M.M. Finding optimal microorganisms to increase crop productivity and sustainability under drought—A structured reflection. J. Plant Interact. 2023, 18, 2178680. [Google Scholar] [CrossRef]
  94. Santos-Medellin, C.; Edwards, J.; Liechty, Z.; Nguyen, B.; Sundaresan, V. Drought Stress Results in a Compartment-Specific Restructuring of the Rice Root-Associated Microbiomes. mBio 2017, 8, e00764-17. [Google Scholar] [CrossRef]
  95. Dai, L.; Zhang, G.; Yu, Z.; Ding, H.; Xu, Y.; Zhang, Z. Effect of Drought Stress and Developmental Stages on Microbial Community Structure and Diversity in Peanut Rhizosphere Soil. Int. J. Mol. Sci. 2019, 20, 2265. [Google Scholar] [CrossRef]
  96. Simmons, T.; Styer, A.B.; Pierroz, G.; Goncalves, A.P.; Pasricha, R.; Hazra, A.B.; Bubner, P.; Coleman-Derr, D. Drought Drives Spatial Variation in the Millet Root Microbiome. Front. Plant Sci. 2020, 11, 599. [Google Scholar] [CrossRef]
  97. Acosta-Martínez, V.; Cotton, J.; Gardner, T.; Moore-Kucera, J.; Zak, J.; Wester, D.; Cox, S. Predominant bacterial and fungal assemblages in agricultural soils during a record drought/heat wave and linkages to enzyme activities of biogeochemical cycling. Appl. Soil Ecol. 2014, 84, 69–82. [Google Scholar] [CrossRef]
  98. Naylor, D.; Coleman-Derr, D. Drought Stress and Root-Associated Bacterial Communities. Front. Plant Sci. 2017, 8, 2223. [Google Scholar] [CrossRef]
  99. Xu, L.; Naylor, D.; Dong, Z.; Simmons, T.; Pierroz, G.; Hixson, K.K.; Kim, Y.M.; Zink, E.M.; Engbrecht, K.M.; Wang, Y.; et al. Drought delays development of the sorghum root microbiome and enriches for monoderm bacteria. Proc. Natl. Acad. Sci. USA 2018, 115, E4284–E4293. [Google Scholar] [CrossRef]
  100. Hartman, K.; Tringe, S.G. Interactions between plants and soil shaping the root microbiome under abiotic stress. Biochem. J. 2019, 476, 2705–2724. [Google Scholar] [CrossRef] [PubMed]
  101. Tripathi, P.; Rabara, R.C.; Reese, R.N.; Miller, M.A.; Rohila, J.S.; Subramanian, S.; Shen, Q.J.; Morandi, D.; Bucking, H.; Shulaev, V.; et al. A toolbox of genes, proteins, metabolites and promoters for improving drought tolerance in soybean includes the metabolite coumestrol and stomatal development genes. BMC Genom. 2016, 17, 102. [Google Scholar] [CrossRef] [PubMed]
  102. Zhalnina, K.; Louie, K.B.; Hao, Z.; Mansoori, N.; da Rocha, U.N.; Shi, S.; Cho, H.; Karaoz, U.; Loque, D.; Bowen, B.P.; et al. Dynamic root exudate chemistry and microbial substrate preferences drive patterns in rhizosphere microbial community assembly. Nat. Microbiol. 2018, 3, 470–480. [Google Scholar] [CrossRef]
  103. Williams, A.; de Vries, F.T. Plant root exudation under drought: Implications for ecosystem functioning. New Phytol. 2020, 225, 1899–1905. [Google Scholar] [CrossRef]
  104. Chen, Y.; Yao, Z.; Sun, Y.; Wang, E.; Tian, C.; Sun, Y.; Liu, J.; Sun, C.; Tian, L. Current Studies of the Effects of Drought Stress on Root Exudates and Rhizosphere Microbiomes of Crop Plant Species. Int. J. Mol. Sci. 2022, 23, 2374. [Google Scholar] [CrossRef]
  105. Preece, C.; Peñuelas, J. Rhizodeposition under drought and consequences for soil communities and ecosystem resilience. Plant Soil 2016, 409, 1–17. [Google Scholar] [CrossRef]
  106. Ren, C.; Chen, J.; Lu, X.; Doughty, R.; Zhao, F.; Zhong, Z.; Han, X.; Yang, G.; Feng, Y.; Ren, G. Responses of soil total microbial biomass and community compositions to rainfall reductions. Soil Biol. Biochem. 2018, 116, 4–10. [Google Scholar] [CrossRef]
  107. Brown, S.; Santa Maria, J.P., Jr.; Walker, S. Wall teichoic acids of gram-positive bacteria. Annu. Rev. Microbiol. 2013, 67, 313–336. [Google Scholar] [CrossRef]
  108. Li, G.; Wang, K.; Qin, Q.; Li, Q.; Mo, F.; Nangia, V.; Liu, Y. Integrated Microbiome and Metabolomic Analysis Reveal Responses of Rhizosphere Bacterial Communities and Root exudate Composition to Drought and Genotype in Rice (Oryza sativa L.). Rice 2023, 16, 19. [Google Scholar] [CrossRef]
  109. Gaete, A.; Pulgar, R.; Hodar, C.; Maldonado, J.; Pavez, L.; Zamorano, D.; Pastenes, C.; Gonzalez, M.; Franck, N.; Mandakovic, D. Tomato Cultivars with Variable Tolerances to Water Deficit Differentially Modulate the Composition and Interaction Patterns of Their Rhizosphere Microbial Communities. Front. Plant Sci. 2021, 12, 688533. [Google Scholar] [CrossRef] [PubMed]
  110. Zhao, X.; Liu, Q.; Xie, S.; Jiang, Y.; Yang, H.; Wang, Z.; Zhang, M. Response of Soil Fungal Community to Drought-Resistant Ea-DREB2B Transgenic Sugarcane. Front. Microbiol. 2020, 11, 562775. [Google Scholar] [CrossRef] [PubMed]
  111. Liu, C.-Y.; Hao, Y.; Wu, X.-L.; Dai, F.-J.; Abd-Allah, E.F.; Wu, Q.-S.; Liu, S.-R. Arbuscular mycorrhizal fungi improve drought tolerance of tea plants via modulating root architecture and hormones. Plant Growth Regul. 2023. [Google Scholar] [CrossRef]
  112. Manjunatha, B.S.; Nivetha, N.; Krishna, G.K.; Elangovan, A.; Pushkar, S.; Chandrashekar, N.; Aggarwal, C.; Asha, A.D.; Chinnusamy, V.; Raipuria, R.K.; et al. Plant growth-promoting rhizobacteria Shewanella putrefaciens and Cronobacter dublinensis enhance drought tolerance of pearl millet by modulating hormones and stress-responsive genes. Physiol. Plant. 2022, 174, e13676. [Google Scholar] [CrossRef] [PubMed]
  113. Hagaggi, N.S.A.; Abdul-Raouf, U.M. Drought-tolerant Sphingobacterium changzhouense Alv associated with Aloe vera mediates drought tolerance in maize (Zea mays). World J. Microbiol. Biotechnol. 2022, 38, 248. [Google Scholar] [CrossRef] [PubMed]
  114. Ye, Q.; Wang, H.; Li, H. Arbuscular Mycorrhizal Fungi Improve Growth, Photosynthetic Activity, and Chlorophyll Fluorescence of Vitis vinifera L. cv. Ecolly under Drought Stress. Agronomy 2022, 12, 1563. [Google Scholar] [CrossRef]
  115. Dubey, A.; Saiyam, D.; Kumar, A.; Hashem, A.; Abd_Allah, E.F.; Khan, M.L. Bacterial Root Endophytes: Characterization of Their Competence and Plant Growth Promotion in Soybean (Glycine max (L.) Merr.) under Drought Stress. Int. J. Environ. Res. Public Health 2021, 18, 931. [Google Scholar] [CrossRef]
  116. Silva, R.; Filgueiras, L.; Santos, B.; Coelho, M.; Silva, M.; Estrada-Bonilla, G.; Vidal, M.; Baldani, J.I.; Meneses, C. Gluconacetobacter diazotrophicus Changes The Molecular Mechanisms of Root Development in Oryza sativa L. Growing under Water Stress. Int. J. Mol. Sci. 2020, 21, 333. [Google Scholar] [CrossRef] [PubMed]
  117. Pereira, S.I.A.; Abreu, D.; Moreira, H.; Vega, A.; Castro, P.M.L. Plant growth-promoting rhizobacteria (PGPR) improve the growth and nutrient use efficiency in maize (Zea mays L.) under water deficit conditions. Heliyon 2020, 6, e05106. [Google Scholar] [CrossRef]
  118. Hashem, A.; Kumar, A.; Al-Dbass, A.M.; Alqarawi, A.A.; Al-Arjani, A.-B.F.; Singh, G.; Farooq, M.; Abd Allah, E.F. Arbuscular mycorrhizal fungi and biochar improves drought tolerance in chickpea. Saudi J. Biol. Sci. 2019, 26, 614–624. [Google Scholar] [CrossRef]
  119. Begum, N.; Ahanger, M.A.; Su, Y.; Lei, Y.; Mustafa, N.S.A.; Ahmad, P.; Zhang, L. Improved Drought Tolerance by AMF Inoculation in Maize (Zea mays) Involves Physiological and Biochemical Implications. Plants 2019, 8, 579. [Google Scholar] [CrossRef]
  120. Qiang, X.; Ding, J.; Lin, W.; Li, Q.; Xu, C.; Zheng, Q.; Li, Y. Alleviation of the detrimental effect of water deficit on wheat (Triticum aestivum L.) growth by an indole acetic acid-producing endophytic fungus. Plant Soil 2019, 439, 373–391. [Google Scholar] [CrossRef]
  121. Martins, S.J.; Rocha, G.A.; de Melo, H.C.; Georg, R.d.C.; Ulhoa, C.J.; Dianese, E.d.C.; Oshiquiri, L.H.; da Cunha, M.G.; da Rocha, M.R.; de Araujo, L.G.; et al. Plant-associated bacteria mitigate drought stress in soybean. Environ. Sci. Pollut. Res. 2018, 25, 13676–13686. [Google Scholar] [CrossRef]
  122. Saikia, J.; Sarma, R.K.; Dhandia, R.; Yadav, A.; Bharali, R.; Gupta, V.K.; Saikia, R. Alleviation of drought stress in pulse crops with ACC deaminase producing rhizobacteria isolated from acidic soil of Northeast India. Sci. Rep. 2018, 8, 3560. [Google Scholar] [CrossRef] [PubMed]
  123. Zhang, Q.; Gong, M.; Yuan, J.; Hou, Y.; Zhang, H.; Wang, Y.; Hou, X. Dark Septate Endophyte Improves Drought Tolerance in Sorghum. Int. J. Agric. Biol. 2017, 19, 53–60. [Google Scholar] [CrossRef]
  124. Barnawal, D.; Bharti, N.; Pandey, S.S.; Pandey, A.; Chanotiya, C.S.; Kalra, A. Plant growth-promoting rhizobacteria enhance wheat salt and drought stress tolerance by altering endogenous phytohormone levels and TaCTR1/TaDREB2 expression. Physiol. Plant. 2017, 161, 502–514. [Google Scholar] [CrossRef]
  125. Maxton, A.; Singh, P.; Masih, S.A. ACC deaminase-producing bacteria mediated drought and salt tolerance in Capsicum annuum. J. Plant Nutr. 2017, 41, 574–583. [Google Scholar] [CrossRef]
  126. Kumari, S.; Vaishnav, A.; Jain, S.; Varma, A.; Choudhary, D.K. Induced drought tolerance through wild and mutant bacterial strain Pseudomonas simiae in mung bean (Vigna radiata L.). World J. Microbiol. Biotechnol. 2016, 32, 4. [Google Scholar] [CrossRef] [PubMed]
  127. Pandey, V.; Ansari, M.W.; Tula, S.; Yadav, S.; Sahoo, R.K.; Shukla, N.; Bains, G.; Badal, S.; Chandra, S.; Gaur, A.K.; et al. Dose-dependent response of Trichoderma harzianum in improving drought tolerance in rice genotypes. Planta 2016, 243, 1251–1264. [Google Scholar] [CrossRef] [PubMed]
  128. Sheibani-Tezerji, R.; Rattei, T.; Sessitsch, A.; Trognitz, F.; Mitter, B. Transcriptome Profiling of the Endophyte Burkholderia phytofirmans PsJN Indicates Sensing of the Plant Environment and Drought Stress. MBio 2015, 6, e00621-15. [Google Scholar] [CrossRef] [PubMed]
  129. Timmusk, S.; Abd El-Daim, I.A.; Copolovici, L.; Tanilas, T.; Kannaste, A.; Behers, L.; Nevo, E.; Seisenbaeva, G.; Stenstrom, E.; Niinemets, U. Drought-Tolerance of Wheat Improved by Rhizosphere Bacteria from Harsh Environments: Enhanced Biomass Production and Reduced Emissions of Stress Volatiles. PLoS ONE 2014, 9, e96086. [Google Scholar] [CrossRef]
  130. Chakraborty, U.; Chakraborty, B.N.; Chakraborty, A.P.; Dey, P.L. Water stress amelioration and plant growth promotion in wheat plants by osmotic stress tolerant bacteria. World J. Microbiol. Biotechnol. 2013, 29, 789–803. [Google Scholar] [CrossRef]
  131. Sarma, R.K.; Saikia, R. Alleviation of drought stress in mung bean by strain Pseudomonas aeruginosa GGRJ21. Plant Soil 2013, 377, 111–126. [Google Scholar] [CrossRef]
  132. Pereyra, M.A.; García, P.; Colabelli, M.N.; Barassi, C.A.; Creus, C.M. A better water status in wheat seedlings induced by Azospirillum under osmotic stress is related to morphological changes in xylem vessels of the coleoptile. Appl. Soil Ecol. 2012, 53, 94–97. [Google Scholar] [CrossRef]
  133. Arzanesh, M.H.; Alikhani, H.A.; Khavazi, K.; Rahimian, H.A.; Miransari, M. Wheat (Triticum aestivum L.) growth enhancement by Azospirillum sp. under drought stress. World J. Microbiol. Biotechnol. 2010, 27, 197–205. [Google Scholar] [CrossRef]
  134. Belimov, A.A.; Dodd, I.C.; Hontzeas, N.; Theobald, J.C.; Safronova, V.I.; Davies, W.J. Rhizosphere bacteria containing 1-aminocyclopropane-1-carboxylate deaminase increase yield of plants grown in drying soil via both local and systemic hormone signalling. New Phytol. 2009, 181, 413–423. [Google Scholar] [CrossRef] [PubMed]
  135. Chandrasekaran, M. Arbuscular Mycorrhizal Fungi Mediated Enhanced Biomass, Root Morphological Traits and Nutrient Uptake under Drought Stress: A Meta-Analysis. J. Fungi 2022, 8, 660. [Google Scholar] [CrossRef] [PubMed]
  136. Kour, D.; Yadav, A.N. Bacterial Mitigation of Drought Stress in Plants: Current Perspectives and Future Challenges. Curr. Microbiol. 2022, 79, 248. [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. Chen, X.; Aili, Y.; Ma, X.; Wang, H.; Dawuti, M. Mycorrhizal fungal colonization promotes apparent growth and physiology of Alhagi sparsifolia seedlings under salt or drought stress at vulnerable developmental stage. Plant Growth Regul. 2023. [Google Scholar] [CrossRef]
Figure 1. Effect of drought stress on crop growth.
Figure 1. Effect of drought stress on crop growth.
Plants 13 00384 g001
Figure 2. Mechanisms underlying the effect of microbes on crop drought resistance.
Figure 2. Mechanisms underlying the effect of microbes on crop drought resistance.
Plants 13 00384 g002
Table 1. Effects of microbes on crop drought resistance and related mechanisms.
Table 1. Effects of microbes on crop drought resistance and related mechanisms.
CropsMicrobesRelated MechanismsReferencesDOI
MorphologicalPhysiologicalBiochemicalMolecular
TeaClaroideoglomus etunicatum Liu et al. [111]https://doi.org/10.1007/s10725-023-00972-8
Pearl milletShewanella putrefaciens, Cronobacter dublinensis Manjunatha et al. [112]https://doi.org/10.1111/ppl.13676
MaizeSphingobacterium changzhouense Hagaggi and Abdul-Raouf [113]https://doi.org/10.1007/s11274-022-03441-y
GrapeAMF Ye et al. [114]https://doi.org/10.3390/agronomy12071563
SoybeanBacillus cereus, Pseudomonas otitidis, Pseudomonas sp. Dubey et al. [115]https://doi.org/10.3390/ijerph18030931
RiceGluconacetobacter diazotrophicus Silva et al. [116]https://doi.org/10.3390/ijms21010333
MaizePGPR Cupriavidus necator (B1), Pseudomonas fluorescens S3X Pereira et al. [117]https://doi.org/10.1016/j.heliyon.2020.e05106
ChickpeaAMF Hashem et al. [118]https://doi.org/10.1016/j.sjbs.2018.11.005
MaizeAMF Begum et al. [119]https://doi.org/10.3390/plants8120579
WheatAlternaria alternata Qiang et al. [120]https://doi.org/10.1007/s11104-019-04028-7
SoybeanBacillus subtilis, Bacillus thuringiensis, Bacillus cereus Martins et al. [121]https://doi.org/10.1007/s11356-018-1610-5
Vigna mungo L.Ochrobactrum pseudogrignonense RJ12, Pseudomonas sp. RJ15 and Bacillus subtilis RJ46 Saikia et al. [122]https://doi.org/10.1038/s41598-018-25174-5
Pisum sativum L.Ochrobactrum pseudogrignonense RJ12, Pseudomonas sp. RJ15 and Bacillus subtilis RJ46 Saikia et al. [122]https://doi.org/10.1038/s41598-018-25174-5
SorghumExophiala pisciphila GM25 Zhang et al. [123]https://doi.org/10.17957/IJAB/15.0241
WheatBacillus subtilis Barnawal et al. [124]https://doi.org/10.1111/ppl.12614
Capsicum annuumBulkhorderia cepacian, Citrobacter feurendii, Serratia marcescens Maxton et al. [125]https://doi.org/10.1080/01904167.2017.1392574
Mung beanPseudomonas simiae Kumari et al. [126]https://doi.org/10.1007/s11274-015-1974-3
RiceTrichoderma harzianum Pandey et al. [127]https://doi.org/10.1007/s00425-016-2482-x
PotatoBurkholderia phytofirmans Sheibani-Tezerji et al. [128]https://doi.org/10.1128/mBio.00621-15
WheatBacillus thuringiensis AZP2, Paenibacillus polymyxa B Timmusk et al. [129]https://doi.org/10.1371/journal.pone.0096086
WheatBacillus safensis, Ochrobactrum pseudogregnonense Chakraborty et al. [130]https://doi.org/10.1007/s11274-012-1234-8
Mung beanPseudomonas aeruginosa GGRJ21 Sarma and Saikia [131]https://doi.org/10.1007/s11104-013-1981-9
WheatAzospirillum brasilense Sp245 Pereyra et al. [132]https://doi.org/10.1016/j.apsoil.2011.11.007
WheatAzospirillum lipoferum Arzanesh et al. [133]https://doi.org/10.1007/s11274-010-0444-1
PeaVariovorax paradoxus 5C-2 Belimov et al. [134]https://doi.org/10.1111/j.1469-8137.2008.02657.x
Note: Green, blue, yellow and pink box represented morphological, physiological, biochemical, and molecular mechanisms were involved in the corresponding references, respectively.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gu, Z.; Hu, C.; Gan, Y.; Zhou, J.; Tian, G.; Gao, L. Role of Microbes in Alleviating Crop Drought Stress: A Review. Plants 2024, 13, 384. https://doi.org/10.3390/plants13030384

AMA Style

Gu Z, Hu C, Gan Y, Zhou J, Tian G, Gao L. Role of Microbes in Alleviating Crop Drought Stress: A Review. Plants. 2024; 13(3):384. https://doi.org/10.3390/plants13030384

Chicago/Turabian Style

Gu, Zechen, Chengji Hu, Yuxin Gan, Jinyan Zhou, Guangli Tian, and Limin Gao. 2024. "Role of Microbes in Alleviating Crop Drought Stress: A Review" Plants 13, no. 3: 384. https://doi.org/10.3390/plants13030384

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