Next Article in Journal
Evolution of Proliferative Model Protocells Highly Responsive to the Environment
Next Article in Special Issue
Effects of Rhizobium leguminosarum Thy2 on the Growth and Tolerance to Cadmium Stress of Wheat Plants
Previous Article in Journal / Special Issue
PKS5 Confers Cold Tolerance by Controlling Stomatal Movement and Regulating Cold-Responsive Genes in Arabidopsis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular and Physiological Mechanisms to Mitigate Abiotic Stress Conditions in Plants

1
Department of Microbiology, CCS Haryana Agricultural University, Hisar 125004, India
2
Department of Biotechnology, Ch. Devi Lal University, Sirsa 125055, India
3
ICAR-National Research Centre on Camel, Bikaner 334001, India
4
Academy of Biology and Biotechnology, Southern Federal University, 344090 Rostov-on-Don, Russia
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work and both are considered as first author.
Life 2022, 12(10), 1634; https://doi.org/10.3390/life12101634
Submission received: 29 September 2022 / Revised: 14 October 2022 / Accepted: 15 October 2022 / Published: 19 October 2022
(This article belongs to the Special Issue Plant Biotic and Abiotic Stresses)

Abstract

:
Agriculture production faces many abiotic stresses, mainly drought, salinity, low and high temperature. These abiotic stresses inhibit plants’ genetic potential, which is the cause of huge reduction in crop productivity, decrease potent yields for important crop plants by more than 50% and imbalance agriculture’s sustainability. They lead to changes in the physio-morphological, molecular, and biochemical nature of the plants and change plants’ regular metabolism, which makes them a leading cause of losses in crop productivity. These changes in plant systems also help to mitigate abiotic stress conditions. To initiate the signal during stress conditions, sensor molecules of the plant perceive the stress signal from the outside and commence a signaling cascade to send a message and stimulate nuclear transcription factors to provoke specific gene expression. To mitigate the abiotic stress, plants contain several methods of avoidance, adaption, and acclimation. In addition to these, to manage stress conditions, plants possess several tolerance mechanisms which involve ion transporters, osmoprotectants, proteins, and other factors associated with transcriptional control, and signaling cascades are stimulated to offset abiotic stress-associated biochemical and molecular changes. Plant growth and survival depends on the ability to respond to the stress stimulus, produce the signal, and start suitable biochemical and physiological changes. Various important factors, such as the biochemical, physiological, and molecular mechanisms of plants, including the use of microbiomes and nanotechnology to combat abiotic stresses, are highlighted in this article.

1. Introduction

Stress is defined as a stimulus that inhibits the growth of plants and their metabolism and development at the time of both abiotic and biotic stress [1]. Abiotic stresses, such as higher or insufficient water supply, low and high temperature, heavy metals, ultraviolet radiation and salinity, are damaging to plant development and growth, and cause considerable losses in agricultural productivity worldwide [2,3]. When the stress threshold is exceeded, the plant is stressed, followed by activation of physiological, biochemical, morphological, and molecular-level mechanisms. The activation of these mechanisms can show the development of a fresh physiological state and the restoration of homeostasis in plants [4]. Due to abiotic stress, it has been estimated that crop production yield decreases by up to 70% in several commercially significant crops and they execute at just 30% of their genetic makeup in terms of yield [5].
Plants survive in environments that are repeatedly changing and frequently not suitable for plant growth as well as development. These harsh situations for the growth and development of plants arise mainly due to abiotic stress [6]. Moreover, abiotic stresses are anticipated to intensify and occur more frequently in the near future due to climate change, which may cause severe salinization of more than 50% of soil of the arable domain by 2050 [7,8]. Therefore, productive agricultural land and crop yields may gradually decline due to increasing temperature and recurrent flooding over several decades, particularly in the mid-latitudes [7,8].
Along with these factors, anthropogenic activities may increase the amounts of contaminants in the water, soil and air, with which plants must contend. It has been estimated that abiotic stress factors have a greater than 90% impact on plants and crop growth during their growing season in rural areas [9]. On the other hand, the rapid growth of the population has increased the demand for food and other necessary resources. Therefore, to develop stress-resistant cultivars that can endure abiotic stress and feed the expanding population, knowing plants’ stress responses is crucial. When plants faces variety of stresses, activates the stress signal and respond accordingly [9]. Plants with abiotic stress have primary signals for ion toxicity detection, low proline and chlorophyll content, low CO2 assimilation, and osmotic effects, etc., in the cells (Figure 1). Secondary consequences of these abiotic stresses are complicated and comprise oxidative stresses that harm various cellular components such as nucleic acids, proteins present in membranes and lipids, and metabolite malfunction. Thus, different abiotic stresses have distinctive and overlapping signals [10].
Drought and salt stress affect water potential homeostasis and distribution of ions at cellular as well as molecular levels. Alterations in water and ion homeostasis can cause growth inhibition, molecular damage, and even death [10]. Primary stress signals trigger some cellular responses; however, the rest are triggered by secondary stress signals. One of the vital features of these signals is the hyperosmotic signal, which increases phytochrome and abscisic acid in plants and provides a protective role during different abiotic stresses such as drought and salt stress [10]. Plants facing cold or chilling stress first indicate a change in cell membrane structure that affects the plant development, then disrupts protein or protein complex stability and lowers the ROS scavenging enzyme activity. These mechanisms cause photo-inhibition, decreased photosynthesis, and considerable membrane damage [10,11,12]. In addition, stress triggers protein synthesis and gene expression, as it triggers RNA secondary structure formations [13]. All these components of plant activity are critical for stress tolerance for minimizing the internal damage in the new stress environment so that homeostatic conditions must be reestablished and growth must be restored, though at a slower rate [5].
Considerable achievements have been made towards the knowledge of plant cellular and molecular mechanisms under different stress conditions [14]. Recognition of a stress environment causes changes in the expression of genes, resulting in changes in the composition of the plant proteomes, transcriptomes, and metabolomes. Plants’ response to different abiotic stress is not a simple pathway; however, it is an intricate integrated circuit comprising several pathways and precise tissue and cellular compartments and their interactions with additional cofactors, as well as signaling molecules for management of a particular response to a current stimulus. Thiourea (TU), a synthetic plant growth regulator with a composition of 36% nitrogen and 42% sulfur, has received much interest for its participation in plant stress tolerance. Thiourea helps to modulate some of the pathways associated with plants’ resistance to abiotic stress. Understanding the processes during TU-induced tolerance may help improve crop production under stress situations [15].
E3-ubiquitin ligases control abiotic stress responses either positively or negatively. Additionally, the target protein and the outcome of the changes in UPS-mediated breakdown, activity regulation, or translocation depend on the involvement of ubiquitin ligases-E3 enzymes in plants’ abiotic stress behavior. Consequently, recognizing and depicting ubiquitin ligases aims is vital during any stress response investigation [16]. The latest developments in understanding the molecular mechanisms underlying plant behaviors to abiotic stress highlight the complexity of these mechanisms, which involve a variety of processes including sensing, signal activation, transcription with transcript processing, and translation and posttranslational changes for combating the abiotic stress situations in plants (Table 1). The improved knowledge and use of various approaches, including genetic, chemical, and microbial techniques, enhance crop production and agricultural sustainability [17].

2. Abiotic Stresses and Crop Plants

Generally, various stresses act simultaneously, such as combined water, heat, salt, heavy metals, and other light stresses. As a result, these changes interfere with the regular plant metabolism function and the source–sink interaction, which lowers plant growth, metabolism, and production [55]. Moreover, these stresses alter the expression sequence of several plant genes and have huge effects on crop production worldwide, reducing the usual yields of important crops such as wheat and rice [16,56].

2.1. Salt Stress

Soil salinity is a huge risk for agriculture in areas where water shortage and poor drainage systems of irrigated farms lower the productivity of crops significantly [57]. According to the Food and Agricultural Organization (FAO 2016) [58], more than 6% of total global land and 19.5% of total irrigated land is already affected by salt conditions. Both human and natural factors can cause soil salinity. Of a total 932.2 Mha salt-affected soils globally, 76.6 Mha soil salinization has been caused by human beings [59]. The salt-affected lands having higher amounts of either exchangeable sodium or soluble salts, both perhaps due to insufficient leaching of cations that forms the base. The chief soluble salts that act as anions are sulfate (SO2−4SO4), carbonate (CO2−3CO3), chloride (Cl) and nitrate (NO−3NO3) salts and the cations are potassium, calcium, magnesium, and sodium. Many metals, such as selenium, lithium, boron, strontium, silica, fluorine, rubidium, manganese, aluminum, and molybdenum are present in hyper-saline soil water, some of which can be harmful to human, animal, and plant health [6,57].
Salt accumulation slows down the growth of plants and lowers plants’ absorption capacity for nutrients and water, a consequence of osmotic stress. The level of resistance to salt stress changes from species to species. Cereal crops such as barley can resist up to 250 mM NaCl; moderately salt-resistant crops are bread wheat, maize, sorghum, while wheat is less resistant to salinity [60]. Plant development is reduced by subsequent salt exposure during two phases: ionic toxicity and osmotic stress [61].

2.2. Drought Stress

Rainfall distribution is unequal due to climate change, which is responsible for the major stress reported: drought, which is the most prevalent abiotic stress globally, reducing grain output drastically. It has a devastating impact on the ability to fulfill the food requirements of the increasing worldwide population. Drought stress is linked to a lack of water and cellular dehydration. A decrease in water potential, stomatal closure, and turgor pressure results in poor plant growth and development [62]. Low-water stress affects biochemical and physiological functions, including ion acquisition and chlorophyll synthesis, photosynthesis, respiration, and carbohydrate and nutrient metabolism, resulting in reduced plant growth [63]. Plant adaptation to low-water stress is a process that involves physiological and biochemical alterations in the plant system. Under drought conditions, plants limit their shoot development and metabolic demands. In maize, yield reduction is observed up to 40%, and in wheat, 21%, with around 40% water shortage or reduction [64]. A yield decline ranging from 34 to 68% was reported in cowpea during drought stress [65].

2.3. Cold Stress

This stress has been identified as the main abiotic stress, which reduces agricultural crop productivity by decreasing crop quality and postharvest life. Cold stress consists of chilling from 0–15 °C and freezing at 0 °C, negatively affecting plant growth and agriculture production [66,67]. During the comparison of both freezing and chilling stress, it has been found that freezing stress is far more harmful to plants. Typically, freezing’s harmful effects start with the formation of a nucleation of ice within the cells, then progressively grows and forms ice crystals, causing water leakage and cell dehydration [68,69]. Several major crops are still unable to cope with cold acclimation. For example, tomato (Solanum lycopersicum), maize (Zea mays), rice (Oryza sativa), cotton (Gossypium hirusutum), and soybean (Glycine max) cannot withstand lower temperatures and have the capacity to grow and survive only in tropical and subtropical areas [70]. Hence, cold stress has a negative impact on plant growth along with plant metabolism and development, resulting in the reduction of crop yields globally [69,71].

2.4. Heat Stress

During heat stress, plants are highly sensitive; at extremely high temperatures, plants die. Generally, plants perform better at their optimum temperature; below and above the optimum temperature plant growth and development are severely affected [72]. Most biochemical and enzymatic reactions double at temperatures from 20 °C to 30 °C and change with every 10 °C. Temperatures above and below the optimum range reduce the reaction rate because enzymes are denatured and inactivated progressively. One or two degrees of temperature change have a huge impact on plant growth and development, particularly reproduction, causing a negative impact on early stages of male gametophyte formation in various crops, including wheat, rice, sorghum, barley, maize, and chickpea [5,73]. Heat stress causes male sterility and spikelet production abnormalities in rice and wheat [5]. In wheat and rice, cold and heat stress results in tapetum breakdown, and changes in the callose walls of microspores, exine formation and metabolism of carbohydrate, ultimately ensuing in male sterility [5,74]. However, temperature stress shows no adverse effects on female gametophyte development [75].

2.5. Heavy Metals Stress

Metal poisoning is one of the main environmental risks that impairs plants’ ability to operate and engage in normal metabolic activities. Heavy metals (HMs) are a group of non-biodegradable, persistent inorganic substances having an atomic weight of more than 20 and a density of more than 5 g cm−3, which affect and pollute the food chains, irrigation, soils, aquifers, and nearby atmosphere before having mutagenic, cytotoxic, and genotoxic effects on human, plant, and animal health [76,77].
Toxic metals are accumulated in the agricultural soils due to excessive use of chemical fertilizers along with increasing industrialization, showing harmful consequences to the soil–plant interaction system [78]. These metals concentrate and enter the plant system at a slow rate via water and air and enter the food chains over a certain time [79]. This poses a considerable threat to the natural food web and biogeochemical cycle [80]. The unprecedented in vivo heavy metals accumulation and bioaccumulation in the environment presents a dilemma for all plants and organisms. Toxic concentrations of HM can interact with various important cellular molecules, including nucleoproteins and DNA, causing excessive production of ROS [6,81]. This will result in serious plant changes, e.g., proteolysis, shoot chlorosis, and lipid peroxidation [80,82]. Under abiotic stress, it was thought that using osmolytes, nanoparticles, mineral nutrients, hydrogels, antioxidants, protectants, potassium, and plant growth hormones such as uniconazole and salicylic acid would boost plant production [83,84]. Additionally, plants may adapt to the negative impacts of droughts by applying plant hormones such as brassinolide (BR), gibberellic acid (GA), auxins, ABA, cytokinins, JA and ethylene, which govern several beneficial reactions in plants [83,84].

3. Sensing and Responding Mechanisms of Plants during Abiotic Stress Conditions

Biological molecules that act as sensors detect undesirable environmental changes and elicit quick stimuli to abiotic stress by signal molecules that activate the system. Abiotic stress causes additional Ca2+ to enter the cytoplasm from apoplastic sources. Ca2+ entrance passages are one kind of sensor for detecting stress signals [85,86,87,88]. The Ca+, nitric oxide, and reactive oxygen species (ROS) also work as messenger substances that activate plant responses during cold stress. Reactive oxygen species such as hydrogen peroxide, hydroxyl radicals, and superoxides are formed in response to various kinds of stress [89]. Receptor-like kinases have an intracellular and extracellular domain at which ligand binding occurs and protein-with-protein binding will occur. When a sensor protein attaches to the extracellular domain, the histidine residues found in the intracellular domain self-phosphorylate and are activated. The activated ligand or sensor proteins may then induce signal-specific cellular response via a MAPKs cascade. Intracellular signaling, i.e., protein dephosphorylation and phosphorylation, regulate a broad range of cellular functions: enzyme activation, macromolecule assemble, protein localization, and degradation [90,91]. When plants detect abiotic stress, signaling cascades are activated, followed by activation of kinase cascades, the assembly of ROS and plant hormones accumulation, leading to the induction of precise set of genes responsible for combating the plant abiotic stress, as shown in Figure 1.
The metabolism of cytokinins, ABA, and ethylene are all impacted by stress, and these molecules then interact with certain kinases to control various biological functions, from controlling shoot development under stress to stomata movement [92]. Abiotic stress in plants, e.g., low water conditions (drought stress), low temperature, and excessive salinity stimulate the expression of the huge range of genes present in plants. With gene expression, different proteins are formed in different plant parts that prevent damage to the cell and activate the large number of genes essential for several abiotic resistance processes in plants. Various kinds of proteins are produced, such as chaperones and late embryogenesis abundant proteins (LEA proteins), which are primarily involved in the development of tolerance. At the same time, stress-associated genes are all concerned with generating the stress response [93]. Plant genes responsible for stress are regulated at three stages, i.e., transcriptional, posttranscriptional, and posttranslational.

3.1. Gene Regulation at the Transcriptional Level

Transcriptional regulation, consisting of chromatin modification and remodeling enhancers and promoters, has regular binding sites positioned downstream and upstream of the coding area known as cis-regulatory and trans-regulatory elements, typically transcription factors. Various abiotic stresses cause alterations in the methylation pattern of histone proteins and DNA, which repress or promote gene transcription. Promoters are unique sequences that have a regulatory role, binding RNA polymerase and other transcription factors to initiate transcription [94]. C-repeat binding factors (CBF), dehydration responsive element binding (DREB), MYB, zinc finger families, and leucine zipper (bZIP) are examples of the regulatory elements concerned in plant defense systems along with genes responsible for stress during binding of the responsive gene promoter’s cis-element [95]. Drought tolerance was improved by Oryza sativa WRKY 11 (trans-regulatory element) overexpression under heat shock protein 101 (HSP 101) promoter control [96,97]. A significant discovery is new cis-acting elements, C-repeat and DRE, which respond to low temperature, drought (low water stress), and excess salt stress [98]. C-repeat binding factor proteins have been isolated progressively since their discovery by identification of DNA-associated proteins which attach the DRE and CRT motifs [99]. C-repeat binding factors 1–3 are cold-induced CBF genes found in Arabidopsis and are located on chromosome IV in tandem; CBF1-3 include Apetala2 or ethylene responsive type transcription factors, which directly bind with the CRT/DRE-conserved motifs present in the promoters of CBF regulons, also known as COR genes, and trigger gene expression during a low-temperature environment [99,100]. Transgenic Arabidopsis with CBF1 overexpression has more COR expression and is more resistant to freezing [101]. Orthologous expression of CBFs has been reported in various plants types such as tomato, wheat, rice, maize, and barley, along with heterologous expression of CBFs in Arabidopsis, which also improves their freezing tolerance mechanism [102], and suggests that it shows tolerance to cold only in those plants with CBF genes from tomato only. Hence, it is reported that CBF1-3 play an essential role in modulating cold tolerance and it is not more conserved in every species, but also it is species-specific [102].

3.2. Gene Regulation at Posttranscriptional Level

Posttranscriptional gene regulation refers to the regulation that happens between the stages of pre-mRNA and mRNA translation. It involves four levels: (A) pre-messenger RNA processing via capping as well as splicing along with polyadenylation; (B) nucleoplasmic trafficking of mRNA; (C) mRNA turnover and stability; (D) translocation of mRNA [103]. Alternative splicing is another strategy that plays an essential role in gene regulation during heat and cold stress, e.g., a gene stabilized 1 (STAT1) encoding a nuclear pre-mRNA provides cold resistance in A. thaliana [18,104]. Cold responsive (COR) gene regulation is essential for posttranscriptional regulation in plants. A dead box gene expression regulator (RCF1) plays an important role in the correct pre-mRNA splicing of various COR genes during low-temperature stress [105]. It has been investigated that alternate splicing pathways alter gene expression to cope with temperature stress [106]. Small RNA segments consisting of 20 to 25 nucleotides are formed from non-coding dsRNA precursors via dicer-like (DCL) RNases, generating several posttranscriptional gene silencing processes. One of these processes, mediated by 21 nucleotide microRNAs, cleaves mRNAs or blocks their translation [107].

3.3. Gene Regulation at the Posttranslational Level

Protein phosphorylation, ubiquitination, and sumoylation are the posttranscriptional activities that play an essential role in modifying plant behavior to many abiotic stresses. SNF1-related protein kinase (SnRKs) and MAPKs known as mitogen-activated protein kinases are key players in various signal transduction cascades initiated by osmotic stress and dehydration via phosphorylation of particular residues [108]. They include XERICO and SnRK2 gene codes for an H2-type zinc finger, and E3-ubiquitin ligases, which are associated in ABA-dependent response during water stress, such as stomata closure [109].
Posttranslational histone modifications and DNA methylation are linked to gene expression changes in response to chilling or cold stress. Histone protein acetylation and deacetylation are due to histone acetyl transferase (HAT) and histone deacetylases (HDAs) being involved in the plant during cold stress [110]. HAD 6 of Arabidopsis is overexpressed due to low temperature and positively enhances the freezing resistance [111]. During low-temperature stress, HDAs come into sight directly with maize DREB1 gene activation and histone hyper-acetylation. It has been reported that histone acetylation of ZmDREB1Aas well as ZmCOR413 in maize and OsDREB1 gene in rice histone is activated by lower temperature [112,113]. In the case of Arabidopsis, the RNA-mediated methylation 4 (RMP4) protein was found to play an important role in RNA-directed DNA methylation by combining with RNA polymerase Pol II and Pol V [114]. During cold stress, gene RDM4 is essential for Pol II possession at CBF2 and CBF3 gene promoters to fight abiotic stress in plants [115].

4. Crucial Signal Transduction Mechanism for Abiotic Stress

Abiotic stresses, which induce signal transduction pathways, such as drought, cold, light, heat and salt, are classified into three types. The first one is MAPK modules, which are used in osmotic/oxidative stress signaling to create antioxidant substances, ROS scavenging enzymes, and osmolytes; the second is calcium-dependent signaling, which promotes the triggers of LEA-type genes; followed by calcium-dependent SOS (salt overlay sensitive), which maintains ion homeostasis, as shown in Figure 2.

4.1. Oxidative or Osmotic Stress Signaling in Plants

All stresses involving salt, heat, drought, oxidative, and cold stress cause the formation of ROS species and cause significant damage to plants [116]. A greater level of ROS functions as a signal and one of the preventive plant responses is the production of ROS scavengers. Osmotic stress triggers various protein kinases, such as MAPKs, in restoring osmotic homeostasis. As a result, osmotic stress activates sensor or receptor proteins such as G protein-coupled receptor proteins, tyrosine and histidine kinases, which activate the MAPK network and signaling cascade, and are associated with the production of more osmolytes, required during osmotic stress. The primary function of osmolytes in cell turgor pressure maintenance is to act as a driving force for water uptake. Well-suited solutes such as proline, glycine betaine, mannitol, and trehalose will serve as ROS and chemical chaperones by stabilizing membrane proteins [117].
The MAP kinase network modulates and transmits the signals from the cell surface to the nucleus. Three kinases are triggered consecutively by the upstream kinase in the core MAPK cascades. The MAP kinase phosphorylates on threonine and serine residues after the activation of MAPKKK, and once activated the MAPK either transfers to the nucleus directly and triggers the transcription factor, stimulating other signal substances to regulate gene expression to cope with stress [118].

4.2. Ca2+-Dependent Activation of LEA Genes

Abiotic stress increases with calcium ions’ entry into the cytoplasm. Calcium ion entry channels serve as sensors for abiotic stress detection. Calcium ions trigger CDPKs (calcium-dependent protein kinases) and are represented by multigene families; their expression stages are modulated both geographically and temporarily during complete development. The CDPK pathway plays an important role in the production of large numbers of anti-desiccation proteins via the activation of LEA-type genes, indicating the damage and repair processes, which are distinct from the pathways that regulate osmolyte synthesis [119,120]. Seeds are naturally desiccated during maturation to minimize desiccation shock at the time of germination by accumulating high levels of LEA proteins [121]. Water deficiency, low temperature, excessive osmolarity, and low-temperature stress cause crop plants to accumulate LEA protein. These proteins are utilized to protect proteins from denaturation or renaturation, maintain membrane integrity and protein structure, and sequester ions in affected tissues. Many scientific publications state that chaperones and LEA proteins protect macromolecules against dehydration, such as lipids, enzymes, and mRNA [122,123]. LEA proteins specialize in membrane desiccation defense, while antioxidant enzymes and compounds have a role in desiccation tolerance. Both LEA proteins and osmolytes work in association with membrane structure and protein stabilization by conferring favored hydration during moderate desiccation conditions and changing water levels to protect against abiotic stress in plants [124].

4.3. Calcium Ion-Dependent SOS Signaling

Plants response to high salt concentration, change the ion transporter channel, helps in the regulation of ion homeostasis during salinity. Higher intracellular or extracellular sodium ions function during the SOS pathway, primarily increasing a cytoplasmic calcium ions signal, which affects the expression and activity of different ion transporters, including K+, sodium ions, and H+. A shift in turgor is used as the input for osmotic stress signaling. Salt stress signaling pathways including osmotic and ionic homeostasis signaling routes, detoxification pathways, and growth maintenance pathways are reported in Arabidopsis [125,126]. During abiotic stress signaling, evidence shows that CDPKs and SOS3 of calcium ion sensors have a main role in coupling an inorganic signal with a specified protein phosphorylation pathway and seem important for plant salinity resistance [127].

5. Functions of the Microbiome in Abiotic Stress Management

Plants are not isolated entities and are not able to live alone. Instead, they coexist with various microbes, including bacteria, fungi, protists, viruses, and other microorganisms [128]. These microorganisms coexist in various plant tissues, forming the plant’s microbiome, which lives in three different places: the phyllosphere, endosphere, and rhizosphere [127]. It is becoming increasingly proven that mycorrhizal fungi and useful soil bacteria, including PGPB and PGPR, have a significant role in sustainable farming and agriculture by stimulating plant growth and increasing plant tolerance to abiotic stresses [127]. Many microbes are essential in plant development, metabolism, and growth under abiotic stress situations (Table 2; Figure 3).

5.1. Arbuscular Mycorrhizal Fungi (AMF)

The development and health of plants may be impacted by AMF, which are symbiotic soil-borne fungi [128]. Agroecosystem services related to AMF still have significant knowledge gaps that need to be filled to be optimized, even though they are currently thought to have significant potential stability and sustainability in the agriculture system, and huge progress has been reported in the understanding of the arbuscular mycorrhizal symbiosis.
It has been found that AMF plays an essential role during abiotic stress, when various stresses are combined [129]. Recently, the role of AMF for protection aligned with abiotic stress in tomatoes has been reported [130]. In conjunction with halophytes and glycophytes, AMF can protect against salt stress [131]. A researcher [132] demonstrated a contrary higher dependence of glycophytes versus AMF compared to halophytes during salt stress. The advantages of AMF under salt stress were the subject of meta-analyses of various genes required for stress tolerance [130,133] (Auge et al. 2014, Chandrasekaran et al. 2021). The osmolyte, carbohydrate, and antioxidant systems can all be improved by AMF [134]. A high K+/Na+ level is maintained when they are in symbiosis with plants, preventing the absorption or transfer of harmful Na+ [135].
AMF can also affect how efficiently carbon is used by maintaining larger grain yields, faster rates of stomatal conductance, net photosynthesis, and poorer internal water use efficiency during salinity stress [136]. AMF-associated mechanisms have recently been discussed in [137], ranging from the uptake of nutrients to increased water use efficiency, from osmoprotectant and ionic homeostasis to enhanced photosynthetic efficiency, from cell structure protection to strengthening functions along with triggering antioxidant metabolism, till phytochrome profile modulation. AM-colonized plants show a modification in plant metabolism, specifically, an enhancement in proline amount with greater H2O2 and isoprene emission in contrast to not-inoculated plants [135]. In the context of water stress conditions, AMF also has shown a beneficial outcome in tomato plants; however, the effects varied depending on the features and fungi species taken into account [138,139]. When three AMFs from various genera were examined to see how they affected tomato resistance to salt or drought stress, it became clear that all studied AMFs shared certain responses; however, others were unique to individual isolates [140].
Table 2. Microbiome-mediated abiotic stress resistance and mechanism in plants.
Table 2. Microbiome-mediated abiotic stress resistance and mechanism in plants.
Type of MicrobesAbiotic Stress TypePlant and Tolerance MechanismReferences
Pseudomonas putida P45DroughtSunflower (Helianthus annuus) showed EPS production and enhanced soil aggregation[141]
PseudomonasDroughtPea (Pisum sativum) plants reduced the production of ethylene[142]
Azospirillium sp.DroughtWheat (Triticum aestivum) has better water relations[143]
AM fungiDrought and salinitySorghum (Sorghum bicolor) showed better water relations[144]
ScytonemaCoastal salinityRice (Oryza sativa) with extracellular products and gibberellic acid[145]
Burkholderia phytofirmansColdGrapevine (Vitis vinifera) with ACC-deaminase synthesis[146]
Burkholderia sp. and Methylobacterium oryzaeCd and Ni toxicityTomato (Solanum lycopersicum) with lower uptake along with translocation of heavy metals[147]
Pseudomonas fluorescencesSalinityGroundnut (Arachis hypogaea) with ACC-deaminase synthesis[148]
Rhizobium tropici; P. polymyxaDroughtCommon bean (Phaseolus vulgaris) with change in hormonal composition and stomatal conductance[149]
Glomus intraradices and Pseudomonas mendocinaDroughtLettuce (Lactuca sativa) has antioxidant status improved[150]
Pseudomonas strain AMK-P6HeatSorghum (Sorghum bicolor) with better biochemical status due to activation of heat shock proteins[151]
Glomus sp. and Bacillus megateriumDroughtTrifolium (Trifolium repens) with proline and IAA production[152]
Paraphaeosphaeria quadriseptataDroughtArabidopsis (Arabidopsis thaliana) has HSP-heat shock protein induction [153]
Bacillus subtilisSalinityArabidopsis (Arabidopsis thaliana) has reduced root Na+ import by reduced transcriptional expression of AtHTK1 (a high-affinity KC transporter) genes[153]
Pseudomonas putidaSalinityCotton (Gossypium hirusutum) stopped the salinity-associated accumulation of ABA in seedlings[154]
Glomus etunicatum and Glomus clarumSalinityWheat (Triticum aestivum), Chilli (Capsicum annum) and mung bean (Vigna radiata) have increased KC concentration in root and reduced NaC in shoots and root[155]
PGPRsHeatClover (Trifolium repens) plants with greater nitrogen fixation[156]
Bacillus licheformisDroughtCapsicum annum with expression and activation of stress-related proteins and genes[157]
Bacillus thuringiensisDroughtWheat (Triticum aestivum) showed the organic compound production[158]
Pantoea dispersa and Azospirillium brailenseSalinityCapsicum annuum has an increase in photosynthesis rates well as stomatal conductance[159]
Burkholderia phytofirmans and Enterobacter sp.DroughtMaize (Zea mays) showed an increased rate of shoot and root biomass[160]
Pseudomonas koreensis strainSalinitySoyabean (Glycine max) has increase KC level and decreased Na+ level[161]
Enterobacter intermediusZn toxicityWhite mustard (Sinapis alba) with ACC deaminase, IAA IAA, hydrocyanic acid, and solubilization of phosphate[162]
Serratia sp. and Bacillus cereusDroughtCucumber (Cucumis sativa) showed the production of genes responsible for the synthesis of proline, an antioxidant enzyme, and monodehydroascorbate[163] Wang et al.; 2012
Photobacterium spp.Mercury toxicityCommon reed (Phragmites australis) showed activity of IAA and mercury reductase [164]
Rhizobium leguminosarum and Pseudominas brassicacerumZinc toxicityMustard (Brassica juncea) with metal chelating molecules[165]
RhizobiumSalinityAsian rice (Oryza sativa) with RAB 18 salt stress-associated gene expression[166]
PB 50 strain of B. megaterium DroughtRice (Oryza sativa) showed better plant growth under osmotic stress, plants protected via stomatal closure with enhanced soluble sugar, carotenoid content and protein content[167]
Bacillus albus and Bacillus cereusDroughtMaize (Zea mays) seeds have a higher germination rate and increased seedling length with reduced toxic effects[168]
Gluconacetobacter diazotrophics (Pal5)DroughtRice (Oryza sativa L.) shows gor, P5CR, BADH and cat genes expression with increase glycine betaine and proline content [169]
Penicillium sp. and CalcoaceticusDroughtFoxtail millet (Setaria italica) with increased glycine betaine and proline content, sugars and chlorophyll a and b with the decrease in lipid oxidation[170]
Streptomyces pactum and actinomycesDroughtWheat (T. aestivum) reduces stress via an enhancement in sugar levels and antioxidant enzymes[171]
B. amyloliquefaciens; Pseudomonas putidaDrought Chick pea (Cicer arietinum) with better photosynthesis, chlorophyll content, biomass and osmolyte content[172]
PGPR consortiumSalinitCommon bean (Phaseolus vulgaris) with available iron content in the soil increased[173,174]
B. gladioli; P. aeruginosaCd toxicityTomato (Solanum lycopersicum) with higher expression of metal transporter genes[175]
Mesorhizobium; RhizobiumSalinityChickpea (Cicer arietinum) with enhanced nitrogen fixation [176]
Bacillus megateriumOsmoticMaize (Zea mays) with higher expression of 2 ZmPIP isoforms in roots[177]
Additionally, in drought circumstances, AMF has valuable effects on maize, promoting plant development and photosynthesis by considerably increasing mineral absorption, assimilation and chlorophyll content, compatible solute concentration, and triggering the antioxidant defensive system [129]. The structures and functions of the photosystem PSII and PSI are less likely to be harmed by water stress due to AMF, as has also been shown [178,179,180]. The latest research discovered that the stimulation of fungal genes’ encoding for SOD scavenging enzymes and non-enzymatic defenses such as glutaredoxin, metallothioein, etc., had a protective function against ROS burst [181]. The contrary impacts of AMF on aquaporin expression and plant hormone levels rely on the type of fungus, aquaporin, and applied stress [182]. Since dryness and salt cause generalized stressors, it stands to reason that AMF, which enables plants to survive under high-salinity environments, makes plants resistant to drought [182]. Transcriptomics has helped understand how several crop species, including rice [182,183], tomatoes [184]), grapevines [185] (Balestrini et al. 2017), as well as wheat [186], regulate expression of fungal as well as plant genes in AM interactions.
Transcriptomic analysis has recently been used in tomato roots colonized by AM to confirm the contribution of the AM symbiosis to tomato plant response, nematode function, and water stress, and revealed unique information regarding the response to AM symbiosis. A function for arbuscular mycorrhizal colonization in initiating defensive behavior against root-knot nematodes was also suggested by alterations in the tomato gene expression associated with nematode attack during AM symbiosis [187].

5.2. Actinomycetes

Several studies have shown Actinomycetes to thrive in various stressful environments, including drought, high temperatures, and high salt. They significantly reduce the harmful effects of abiotic stresses while encouraging plant development [188]. Streptomyces play a synergistic role in wheat crops under severe abiotic conditions and promote plant development and their capacity to withstand such conditions [189]. Furthermore, because of their unusual shape, they mix with the soil particles in the rhizosphere and create a solid link with the plants. This approach assists the plant in strained soil and enables more effective utilization of water and nutrients in the rhizosphere. These bacteria increase their capacity to withstand abiotic stresses by an array of mechanisms, comprising changes in root and cell wall morphology, 1-ACC deaminase activity, and the ability to protect from oxidative damage with the production of proline and glycine betaine, which aid in osmoregulation processes [190]. It has been found that wheat seeds soaked with streptomyces inoculum showed greatly enhanced shoot length, root depth, and high root and shoot biomass, and appreciably boosted the germination of roots under high-saline conditions [191]. Tomato plants inoculated with Streptomyces sp. PGPA39, during salt stress, showed a substantial reduction in leaf proline concentration and enhancement in plant biomass compared to the non-treated plants [192].
Treating maize plants with Actinomycetes during low-water conditions produced plants with higher growth, survival rate, shoot and root dry weight, and chlorophyll content than untreated plants [193]. Furthermore, Hasegawa and colleagues [194] demonstrated that endophytic Actinomycetes II improves drought resistance in Kalmia latifolia L. (mountain laurel) by causing callose buildup and lignification in the cell wall. According to next-generation sequencing methods used for microbial communities’ identification, it has been investigated that Actinobacteria have regularly been discovered as one of the bacteria most found in soils [195]. However, although actinomycetes have played an essential role in reducing abiotic plant stress, slight knowledge regarding the dynamics of their interactions with plants limits the potential for using these microorganisms in agriculture.

5.3. PGPR/Plant Growth-Promoting Bacteria

The plant rhizosphere is a highly delicate and dynamic ecosystem, with different housing types of microorganisms that play various roles in plant development and survival [196]. Numerous distinctive and illustrative PGPR bacteria, comprising the Bacillus, Pseudomonas, Enterobacter, Agrobacterium, Klebsiella, Azotobacter, Ervinia, Bradyrhizobium, Burkholderia, and Serratia, have been isolated and described [197,198]. Conversely, several variables, including plant growth stage, species, cultivation method, soil ecology, and testing settings may affect how effective PGPR treatment is at enhancing plant growth. Rhizobacteria that promote plant development can lessen the drought stress effects and increase yield in treated plants [199,199]. It has been reported that Rhizobacterium induces drought endurance and resilience, known as RIDER, which results in biochemical alterations, and allows PGPR to decrease the drought stress effects [200,201]. Phytohormones development, bacterial exopolysaccharides formation, cyclic metabolic pathway conventions, and optimization of the antioxidant defense system are associated with the deposition of various carbon-based substances, including amino acids; polyamines, sugars, and heat shock protein production are examples of RIDER mechanisms [202,203].
Recently, it has been investigated that PGPB, along with other plant microbiomes, acts as a biological technique for decreasing plant salt stress [173]. Rhizobacteria that encourage plant development can successfully lessen the effects of drought on Zea mays and wheat grass. Two bacterial strains, namely Enterobacter sp. 16i and Bacillus sp. 12D6, were employed to counteract drought stress in plants and it was reported that Bacillus sp. 12D6 performed better under drought stress [204]. Azospirillum (GQ255950)-treated maize plants perform better and have more root and shoot fresh weight biomass along with proline, soluble sugars, amino acids, and osmotic levels compared to non-treated maize plants [205]. A significant biotic stress element that reduces agricultural output is drought. In prior research, Pseudomonas libanensis EU-LWNA-33 was used to treat wheat plants, and measurements of growth parameters showed that the root length and biomass were enhanced after treatment. Additionally, biochemical analysis revealed that at 75% stress, levels of proline increased up to two times while levels of glycine betaine were enhanced 1.2-fold [206].
Plant survival under various types of stress is challenging and depends specifically on the plant’s root microbiome. It has been demonstrated that bacteria isolated from harsh environmental circumstances contain traits that make them resistant to salt stress. P. fluorescens was isolated from Saharan rhizosphere soil and showed a PGPB property in maize under salt stress. Numerous microorganisms living on plants with high salt content exhibit adaptations to salinity stress and do well in these environments [207]. One study was conducted in a greenhouse where the soil was treated with three isolated bacteria from rice fields. The soil showed how drought stress might change microbial interactions, and soil microbial interactions with plants were changed with water limitation, e.g., Actinobacteria and Chloroflexi were abundant in the changing patterns, while Acidobacteria and Deltaproteobacteria were lost. Plant survival under drought circumstances results from compartment-specific reorganization [208].
Several methods of PGPB might respond to drought stress effects and promote phytohormones as well as solute production, chlorophyll synthesis, and activation of 1-aminocyclopropane-1-carboxylate deaminase, exo-polysaccharides, and mineral solubilization [209]. Bacillus amyloliquefaciens GB03 and Pseudomonas flurescencs WCS417R- treated plants showed that activating the antioxidant defense system considerably decreased the effects of drought stress [210].
Rhizobacteria that encourage plant development, including Enterobacter, Bacillus, Pseudomonas and Moraxella, have been identified. Under drought stress, wheat plants treated with Bacillus species showed enhanced auxin production up to 25.9 g mL−1 along with an increase in field capability of up to 10% as well as increased crop output of 34% [211]. In Uttar Pradesh, India, PGPR strains including Pseudomonas putida and B. amyloliquefaciens have been identified in alkaline soils, and the combined impact was studied in chickpea plants, which produced plants with boosted biomass, photosynthesis, osmolyte, and chlorophyll content, and reduced abiotic stress, as well as being eco-friendly to the environment [172]. Wheat seedlings treated with Azospirillum brasilense Sp245 had enhanced water control, greater mineral content, higher amounts of K, Mg, and Ca, and a 12.4% overall improvement in grain yield when compared to untreated plants [143].
Foxtail millet crops were used in an experiment, treated with Penicillium and Acinetobacter, which revealed increased physiological growth and a reduction in the harmful effects of drought stress [170]. During this study, buildup of proline, osmolytes, and glycine betaine with elevated levels of chlorophyll content was reported that helped reduce drought stress [170].

6. Nanoparticles’ Application in Combating Abiotic Stress in Plants

Numerous plant responses are triggered by abiotic stress, ranging from growth and morphological changes to crop output and yield [212]. Nanotechnology is one of the newest and most promising methods for treating abiotic stress conditions in plants, such as HMs stress, heat stress, drought stress, and salinity stress. It is also an environmentally friendly technique. With the increasing cellular antioxidants, nutrient uptake, photosynthetic efficiency, and molecular as well as biochemical pathways, NPs dramatically increase the ability of plants to withstand abiotic stress [213]. Modernizing the agricultural sector with prospective applications for improving plant growth as well as development under stress conditions, nanotechnology has recently made strides [213].
Several studies have been reported on the possible use of NPs in the remediation of HM-contaminated soil [214,215,216]. A study of FeO NPs found they can relieve drought stress [217,218] and reduce Cd toxicity in wheat plants by enhancing biomass, chlorophyll levels, and antioxidant enzymes [216]. Si NPs have been reported to reduce HM-induced phytotoxicity in wheat as well as rice and pea [219,220]. To lessen the detrimental effects of HMs on plant growth and development, new nanoremediation techniques must be developed.
It has been found that Si NP treatment enhances plants’ ability to withstand drought stress [221]. By modifying the morpho-physiological characteristics of barley plants, Si NPs showed promising drought stress recovery [222]. It has been revealed that in saline- and water-deficient circumstances, Si NPs improved cucumber growth and yield [223]. Wheat plants under drought stress exhibited higher relative water content, CAT, SOD activity, yield, and biomass after application of chitosan [224]. Silver nanoparticles were used to lessen the harmful effects of drought in lentil (Lens culinaris Medic.) plants [225]. Abscisic acid administration aided by Si NPs was described as an efficient management technique to increase drought resistance in Arabidopsis thaliana [226].
Wheat plants’ growth, chlorophyll concentrations, and antioxidant enzymes were improved by treatment with FeO NPs, which reduced the effects of salt stress [216]. It has been found that manganese NP seed-priming controls salt stress by altering the molecular reactions in pepper (C. annuum L.) plants [227]. To better understand how NPs work to increase plants’ resistance to salinity and other abiotic stress in plants, more molecular and physiological research is required. Numerous studies have demonstrated the possible use of NPs to increase the ability of plants to withstand heat stress [228,229]. Recently it has been reported that the use of organically produced Se NPs (100 g/mL) enhanced wheat development by enhancing plant tolerance to heat stress [230]. Researchers have discovered that silver nanoparticles considerably improved the morphological characteristics of wheat plants under heat stress conditions [231]. Overall, using metallic NPs as nanofertilizers can help plants be more resilient to heat stress, which is important for sustainable agriculture.

7. Conclusions

Evidence from various studies regarding molecular, biochemical, and physio-morphological characteristics, plant responses to diverse types of abiotic stresses, and the microbe interaction mitigation methods in plants have been examined. These studies have improved our knowledge of the processes behind gene cascades, microbial interactions, and metabolic pathways, along with augmentation of different proteins, enzymes, metabolites, and the down- and upregulation of numerous genes. This paper provides dynamic information about plants’ response to different types of abiotic stress. This paper gives innovative ideas for improving the current methods to mitigate abiotic stress by employing various gene regulation-based and microbial-mediated plant interactions, facilitating improved germination, increased ability to withstand and mitigate adverse environment conditions, and better yield in plants. By supplying vital nutrients, nanotechnology can be a cost effective and promising way to increase plants’ ability to withstand abiotic stress. However, because of their widespread use, there are some possible worries regarding their adverse impacts on the environment.
Plants have evolved built-in adaptation mechanisms to deal with various complex abiotic challenges. With the aid of science and technological advancements, it is now feasible to comprehend gene function, establish gene-manipulation strategies, and generate plant characteristics to combat abiotic stress. Signaling pathways must be seen as intricate networks. Hence, abiotic stress signal-transduction cascades are better understood by molecular investigations of the signaling molecules. Abiotic stresses have a detrimental effect on plants that is exacerbated by ongoing climate change, reducing agricultural production. Abiotic stress tolerance is a multigenic response that involves stress-responsive gene expression, signal transduction, and sensing. As a result of these coordinated efforts, agricultural production, productivity, and food security will improve.

8. Future Perspectives

Genetic modification and recombinant techniques should be integrated with traditional and marker-assisted breeding practices in the future to produce plants with modified characteristics that cope with adverse conditions. In addition to this, nanotechnology will be the most potent tool for combating the abiotic stresses in plants. Moreover, finding new adaptable germplasm is crucial for directing breeding programs to identify plants for a changing climate and future research should focus on developing NPs that are inexpensive, nontoxic, self-degradable, and eco-friendly using green methods. By addressing the effects of climate change, particularly stress caused by drought, heat and cold, these combined initiatives will significantly advance the cause of food security through increased agricultural output and productivity.

Author Contributions

Conceptualization, B.S.S. and B.B.; validation, J.S.D., V.D.R., R.K. and S.M.; formal analysis, B.S.S.; data curation, B.B. and S.M.; writing—original draft preparation, B.B., V.D.R., J.S.D. and T.M.; writing—review and editing, J.S.D., R.K.; visualization, S.M.; supervision, B.S.S.; project administration B.S.S. and B.B. equally contributed to MS and are considered both first authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors are highly grateful to the Department of Microbiology, CCS HAU, Hisar, for providing infrastructural support for preparing the manuscript. V.D.R. and T.M. acknowledge support from the laboratory «Soil Health» of the Southern Federal University with the financial support of the Ministry of Science and Higher Education of the Russian Federation, agreement No. 075-15-2022-1122.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Umar, O.B.; Ranti, L.A.; Abdulbaki, A.S.; Bola, A.L.; Abdulhamid, A.K.; Biola, M.R.; Victor, K.O. Stresses in Plants: Biotic and Abiotic. In Current Trends in Wheat Research; Ansari, M., Ed.; IntechOpen: London, UK, 2021. [Google Scholar]
  2. Wania, S.H.; Kumar, V.; Shriram, V.; Sah, S.K. Phytohormones and their metabolic engineering for abiotic stress tolerance in crop plants. Crop J. 2016, 4, 162–176. [Google Scholar] [CrossRef] [Green Version]
  3. Mei, H.; Cheng-Qiang, H.; Nai-Zheng, D. Abiotic Stresses: General defenses of land plants and chances for engineering multistress tolerance. Front. Plant Sci. 2018, 7, 1771. [Google Scholar] [CrossRef] [Green Version]
  4. Jogaiah, S.; Govind, S.R.; Tran, L.S.P. Systems biology-based approaches toward understanding drought tolerance in food crops. Crit. Rev. Biotechnol. 2013, 33, 23–29. [Google Scholar] [CrossRef] [PubMed]
  5. 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] [Green Version]
  6. Sihag, S.; Brar, B.; Joshi, U.N. Salicylic acid induces amelioration of chromium toxicity and affects antioxidant enzyme activity in Sorghum bicolor L. Int. J. Phytoremediation 2019, 21, 293–304. [Google Scholar] [CrossRef]
  7. Fedoron, N.V.; Battisti, D.S.; Beachy, R.N.; Cooper, P.J.; Fischhon, D.A.; Hodges, C.N.; Knauf, V.C.; Lobell, D.; Mazur, B.J.; Molden, D.; et al. Radically rethinking agriculture for the 21st century. Science 2010, 327, 833–834. [Google Scholar] [CrossRef] [Green Version]
  8. Kundzewicz, Z.; Ulbrich, U.; Brucher, T.; Graczyk, D.; Kruger, A.; Leckebusch, G.C. Summer floods in Central Europe—Climate change track? Nat. Hazards 2005, 36, 165–189. [Google Scholar] [CrossRef]
  9. Cramer, G.R.; Urano, K.; Delrot, S.; Pezzotti, M.; Shinozaki, K. Effects of abiotic stress on plants: A systems biology perspective. BMC Plant Biol. 2011, 11, 163–176. [Google Scholar] [CrossRef] [Green Version]
  10. Mathivanan, S. Abiotic Stress-Induced Molecular and Physiological Changes and Adaptive Mechanisms in Plants. In Abiotic Stress in Plants; Fahad, S., Saud, S., Chen, Y., Wu, C., Wang, D., Eds.; IntechOpen: London UK, 2021. [Google Scholar] [CrossRef]
  11. Ruelland, E.; Vaultier, M.N.; Zachowski, A.; Hurry, V. Cold signalling and cold acclimation in plants. Adv. Bot. Res. 2009, 49, 35–150. [Google Scholar]
  12. Siddiqui, K.S.; Cavicchioli, R. Cold-adapted enzymes. Annu. Rev. Biochem. 2006, 75, 403–433. [Google Scholar] [CrossRef]
  13. Rajkowitsch, L.; Chen, D.; Stampfl, S.; Semrad, K.; Waldsich, C.; Mayer, O.; Jantsch, M.F.; Konrat, R.; Bläsi, U.; Schroeder, R. RNA chaperones, RNA annealers and RNA helicases. RNA Biol. 2007, 4, 118–130. [Google Scholar] [CrossRef] [Green Version]
  14. Hadiarto, T.; Tran, L.S.P. Progress studies of drought-responsive genes in rice. Plant Cell Rep. 2011, 30, 297–310. [Google Scholar] [CrossRef]
  15. Waqas, M.A.; Kaya, C.; Riaz, A.; Farooq, M.; Nawaz, I.; Wilkes, A.; Li, Y. Potential mechanisms of abiotic stress tolerance in crop plants induced by thiourea. Front. Plant Sci. 2019, 10, 1336. [Google Scholar] [CrossRef]
  16. Melo, F.V.; Oliveira, M.M.; Saibo, N.J.; Lourenço, T.F. Modulation of abiotic stress responses in rice by E3-ubiquitin ligases: A promising way to develop stress-tolerant crops. Front. Plant Sci. 2021, 12, 368. [Google Scholar] [CrossRef]
  17. Zhang, H.; Zhu, J.; Gong, Z.; Zhu, J.K. Abiotic stress responses in plants. Nat. Rev. Genet. 2022, 23, 104–119. [Google Scholar] [CrossRef]
  18. Hwarari, D.; Guan, Y.; Ahmad, B.; Movahedi, A.; Min, T.; Hao, Z.; Lu, Y.; Chen, J.; Yang, L. ICE-CBF-COR signaling cascade and its regulation in plants responding to cold stress. Int. J. Mol. Sci. 2022, 23, 1549. [Google Scholar] [CrossRef]
  19. Eremina, M.; Rozhon, W.; Poppenberger, B. Hormonal control of cold stress responses in plants. Cell. Mol. Life Sci. 2016, 73, 797–810. [Google Scholar] [CrossRef]
  20. Rizhsky, L.; Liang, H.; Shuman, J.; Shulaev, V.; Davletova, S.; Mittler, R. When Defense Pathways Collide. The Response of Arabidopsis to a Combination of Drought and Heat Stress. Plant Physiol. 2004, 134, 1683–1696. [Google Scholar] [CrossRef] [Green Version]
  21. Zhang, Y.; Min, H.; Shi, C.; Xia, G.; Lai, Z. Transcriptome analysis of the role of autophagy in plant response to heat stress. PLoS ONE 2021, 16, e0247783. [Google Scholar] [CrossRef]
  22. Pan, T.; Sun, X.; Liu, Y.; Li, H.; Deng, G.; Lin, H.; Wang, S. Heat stress alters genome-wide profiles of circular RNAs in Arabidopsis. Plant Mol. Biol. 2018, 96, 217–229. [Google Scholar] [CrossRef]
  23. Sperdouli, I.; Moustakas, M. Spatio-temporal heterogeneity in Arabidopsis thaliana leaves under drought stress. Plant Biol. 2012, 14, 118–128. [Google Scholar] [CrossRef]
  24. Dubey, A.K.; Kumar, N.; Kumar, A.; Ansari, M.A.; Ranjan, R.; Gautam, A.; Sahu, N.; Pandey, V.; Behera, S.K.; Mallick, S.; et al. Over-expression of CarMT gene modulates the physiological performance and antioxidant defense system to provide tolerance against drought stress in Arabidopsis thaliana L. Ecotoxicol. Environ. Saf. 2019, 171, 54–65. [Google Scholar] [CrossRef]
  25. Shen, J.; Xing, T.; Yuan, H.; Liu, Z.; Jin, Z.; Zhang, L.; Pei, Y. Hydrogen sulfide improves drought tolerance in Arabidopsis thaliana by microRNA expressions. PLoS ONE 2013, 8, e77047. [Google Scholar] [CrossRef]
  26. Jin, Z.; Wang, Z.; Ma, Q.; Sun, L.; Zhang, L.; Liu, Z.; Liu, D.; Hao, X.; Pei, Y. Hydrogen sulfide mediates ion fluxes inducing stomatal closure in response to drought stress in Arabidopsis thaliana. Plant Soil 2017, 419, 141–152. [Google Scholar] [CrossRef]
  27. Panta, G.; Rieger, M.; Rowland, L. Effect of cold and drought stress on blueberry dehydrin accumulation. J. Hortic. Sci. Biotechnol. 2001, 76, 549–556. [Google Scholar]
  28. Ergin, S.; Kesici, M.; Gulen, H. Changes in H2O2 and peroxidase activities in strawberry plants under heat stress. Harran TarımveGıdaBilimleri Derg. 2012, 16, 25–35. [Google Scholar]
  29. Gulen, H.; Eris, A. Effect of heat stress on peroxidase activity and total protein content in strawberry plants. Plant Sci. 2004, 166, 739–744. [Google Scholar] [CrossRef]
  30. Kim, J.H.; Lim, S.D.; Jang, C.S. Oryza sativa heat-induced RING finger protein 1 (OsHIRP1) positively regulates plant response to heat stress. Plant Mol. Biol. 2019, 99, 545–559. [Google Scholar] [CrossRef]
  31. Pan, Y.; Wang, W.; Zhao, X.; Zhu, L.; Fu, B.; Li, Z. DNA methylation alterations of rice in response to cold stress. Plant Omics J. 2011, 4, 364–369. [Google Scholar]
  32. De Azevedo Neto, A.D.; Prisco, J.T.; Eneas-Filho, J.; de Abreu, C.E.B.; Gomes-Filho, E. Effect of salt stress on antioxidative enzymes and lipid peroxidation in leaves and roots of salt-tolerant and salt-sensitive maize genotypes. Environ. Exp. Bot. 2006, 56, 87–94. [Google Scholar] [CrossRef]
  33. 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]
  34. Kilasi, N.L.; Singh, J.; Vallejos, C.E.; Ye, C.; Jagadish, S.K.; Kusolwa, P.; Rathinasabapathi, B. Heat stress tolerance in rice (Oryza sativa L.): Identification of quantitative trait loci and candidate genes for seedling growth under heat stress. Front. Plant Sci. 2018, 9, 1578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Paredes, M.; Quiles, M.J. The effects of cold stress on photosynthesis in Hibiscus plants. PLoS ONE 2015, 10, e0137472. [Google Scholar] [CrossRef] [PubMed]
  36. Munoz, R.; Quiles, M.J. Water deficit and heat affect the tolerance to high illumination in hibiscus plants. Int. J. Mol. Sci. Mar. 2013, 7, 5432–5444. [Google Scholar] [CrossRef] [Green Version]
  37. Ni, L.; Wang, Z.; Liu, X.; Wu, S.; Hua, J.; Yin, Y.; Li, H.; Gu, C. Transcriptome Analysis of Salt Stress in Hibiscus hamabo Sieb. et Zucc Based on Pacbio Full-Length Transcriptome Sequencing. Int. J. Mol. Sci. 2022, 23, 138. [Google Scholar] [CrossRef]
  38. Hajihashemi, S.; Noedoost, F.; Geuns, J.; Djalovic, I.; Siddique, K.H. Effect of cold stress on photosynthetic traits, carbohydrates, morphology, and anatomy in nine cultivars of Stevia rebaudiana. Front. Plant Sci. 2018, 9, 1430. [Google Scholar] [CrossRef] [Green Version]
  39. Debnath, M.; Ashwath, N.; Midmore, D.J. Physiological and morphological responses to abiotic stresses in two cultivars of Stevia rebaudiana (Bert.) Bertoni. S. Afr. J. Bot. 2019, 123, 124–132. [Google Scholar] [CrossRef]
  40. Karimi, M.; Ahmadi, A.; Hashemi, J.; Abbasi, A.; Tavarini, S.; Pompeiano, A.; Angelini, L.G. The positive role of steviol glycosides in stevia (Stevia rebaudianaBertoni) under drought stress condition. Plant Biosyst. 2016, 150, 1323–1331. [Google Scholar] [CrossRef]
  41. Pandey, M.; Chikara, S.K. Effect of Salinity and Drought Stress on Growth Parameters, Glycoside Content and Expression Level of Vital Genes in Steviol Glycosides Biosynthesis Pathway of Stevia rebaudiana (Bertoni). Int. J. Genet. 2015, 7, 153–160. [Google Scholar]
  42. Li, X.; Cai, J.; Liu, F.; Dai, T.; Cao, W.; Jiang, D. Cold priming drives the sub-cellular antioxidant systems to protect photosynthetic electron transport against subsequent low temperature stress in winter wheat. Plant Physiol. Biochem. 2014, 82, 34–43. [Google Scholar] [CrossRef]
  43. Ni, Z.; Li, H.; Zhao, Y.; Peng, H.; Hu, Z.; Xin, M.; Sun, Q. Genetic improvement of heat tolerance in wheat: Recent progress in understanding the underlying molecular mechanisms. Crop J. 2018, 6, 32–41. [Google Scholar] [CrossRef]
  44. Schmidt, J.; Tricker, P.J.; Eckermann, P.; Kalambettu, P.; Garcia, M.; Fleury, D. Novel alleles for combined drought and heat stress tolerance in wheat. Front. Plant Sci. 2020, 10, 1800. [Google Scholar] [CrossRef] [Green Version]
  45. Hussain, N.; Ghaffar, A.; Zafar, Z.U.; Javed, M.; Shah, K.H.; Noreen, S.; Manzoor, H.; Iqbal, M.; Hassan, I.F.Z.; Bano, H.; et al. Identification of novel source of salt tolerance in local bread wheat germplasm using morpho-physiological and biochemical attributes. Sci. Rep. 2021, 11, 10854. [Google Scholar] [CrossRef]
  46. Omrani, S.; Arzani, A.; Esmaeilzadeh Moghaddam, M.; Mahlooji, M. Genetic analysis of salinity tolerance in wheat (Triticum aestivum L.). PLoS ONE 2022, 17, e0265520. [Google Scholar] [CrossRef]
  47. Wang, L.; Zhu, W.; Fang, L.; Sun, X.; Su, L.; Liang, Z.; Wang, N.; Londo, J.P.; Li, S.; Xin, H. Genome-wide identification of WRKY family genes and their response to cold stress in Vitis Vinifera. BMC Plant Biol. 2014, 14, 103. [Google Scholar] [CrossRef] [Green Version]
  48. Yu, W.; Wang, L.; Zhao, R.; Sheng, J.; Zhang, S.; Li, R.; Shen, L. Knockout of SlMAPK3 enhances tolerance to heat stress involving ROS homeostasis in tomato plants. BMC Plant Biol. 2019, 19, 1–13. [Google Scholar] [CrossRef] [Green Version]
  49. Niron, H.; Barlas, N.; Salih, B.; Türet, M. Comparative Transcriptome, Metabolome, and Ionome Analysis of Two Contrasting Common Bean Genotypes in Saline Conditions. Front. Plant Sci. 2020, 11, 2007. [Google Scholar] [CrossRef]
  50. Stoeva, N.; Kaymakanova, M. Effect of salt stress on the growth and photosynthesis rate of bean plants (Phaseolus vulgaris L.). J. Cent. Eur. Agric. 2008, 9, 385–391. [Google Scholar]
  51. Singh, M.; Khan, M.M.A.; Uddin, M.; Naeem, M.; Qureshi, M.I. Proliferating effect of radiolytically depolymerized carrageenan on physiological attributes, plant water relation parameters, essential oil production and active constituents of Cymbopogon flexuosus Steud. under drought stress. PLoS ONE 2017, 12, e0180129. [Google Scholar] [CrossRef] [Green Version]
  52. Shalata, A.; Tal, M. The effect of salt stress on lipid peroxidation and antioxidants in the leaf of the cultivated tomato and its wild salt-tolerant relative Lycopersicon pennellii. Physiol. Plant. 1998, 104, 169–174. [Google Scholar] [CrossRef]
  53. Maggio, A.; Raimondi, G.; Martino, A.; De Pascale, S. Salt stress response in tomato beyond the salinity tolerance threshold. Environ. Exp. Bot. 2007, 59, 276–282. [Google Scholar] [CrossRef]
  54. Schubert, S.; Serraj, R.; Plies-Balzer, E.; Mengel, K. Effect of drought stress on growth, sugar concentrations and amino acid accumulation in N2-fixing alfalfa (Medicago sativa). J. Plant Physiol. 1995, 146, 541–546. [Google Scholar] [CrossRef]
  55. Khan, F.; Hakeem, K.R. Cell signaling during drought and salt stress. In Plant Signaling: Understanding the Molecular Crosstalk; Springer: New Delhi, India, 2014; pp. 227–239. [Google Scholar]
  56. Abhinandan, K.; Skori, L.; Stanic, M.; Hickerson, N.M.; Jamshed, M.; Samuel, M.A. Abiotic stress signaling in wheat—An inclusive overview of hormonal interactions during abiotic stress responses in wheat. Front. Plant Sci. 2018, 9, 734. [Google Scholar] [CrossRef] [Green Version]
  57. Kumar, P.; Sharma, P.K. Soil salinity and food security in India. Front. Sustain. Food Syst. 2020, 4, 533781. [Google Scholar] [CrossRef]
  58. FAO. The State of Food and Agriculture—Climate Change, Agriculture and Food Security; Food and Agricultural Organization of the United Nations: Rome, Italy, 2016. [Google Scholar]
  59. Shahid, S.A.; Zaman, M.; Heng, L. Soil salinity: Historical perspectives and a world overview of the problem. In Guideline for Salinity Assessment, Mitigation and Adaptation Using Nuclear and Related Techniques; Springer: Cham, Switzerland, 2018; pp. 43–53. [Google Scholar] [CrossRef]
  60. Maas, E.V.; Honman, G.J. Crop salt tolerance—Current assessment. J. Irrig. Drain. Div. 1977, 103, 115–134. [Google Scholar] [CrossRef]
  61. Yildiz, C.; Acikgoz, H. A kernel extreme learning machine-based neural network to forecast very short-term power output of an on-grid photovoltaic power plant. Energy Sources Part A Recovery Util. Environ. Eff. 2021, 43, 395–412. [Google Scholar] [CrossRef]
  62. 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] [Green Version]
  63. Li, X.N.; Pu, H.C.; Liu, F.L.; Zhou, Q.; Cai, J.; Dai, T.B.; Cao, W.; Jiang, D. Winter wheat photosynthesis and grain yield responses to spring freeze. Agron. J. 2015, 107, 1002–1010. [Google Scholar] [CrossRef]
  64. Daryanto, S.; Wang, L.; Jacinthe, P.A. Global synthesis of drought enacts on maize and wheat production. PLoS ONE 2016, 11, e0156362. [Google Scholar] [CrossRef] [Green Version]
  65. Farooq, M.; Gogoi, N.; Barthakur, S.; Baroowa, B.; Bharadwaj, N.; Alghamdi, S.S. Drought stress in grain legumes during reproduction and grain filling. J. Agron. Crop Sci. 2017, 203, 81–102. [Google Scholar] [CrossRef]
  66. Guo, X.Y.; Liu, D.F.; Chong, K. Cold signaling in plants: Insights into mechanisms and regulation. J. Integr. Plant Biol. 2018, 60, 45–756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Liu, H.; Yu, C.; Li, H.; Ouyang, B.; Wang, T.; Zhang, J.; Wang, X.; Ye, Z. Overexpression of ShDHN, a dehydrin gene from Solanum habrochaites enhances tolerance to multiple abiotic stresses in tomato. Plant Sci. 2015, 231, 198–211. [Google Scholar] [CrossRef] [PubMed]
  68. Dowgert, M.F.; Steponkus, P.L. Behavior of the plasma membrane of isolated protoplasts during a freeze-thaw cycle. Plant Physiol. 1984, 75, 1139–1151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Pearce, R.S. Plant freezing and damage. Ann. Bot. 2001, 87, 417–424. [Google Scholar] [CrossRef]
  70. Chinnusamy, V.; Zhu, J.; Zhu, J.K. Cold stress regulation of gene expression in plants. Trends Plant Sci. 2007, 12, 444–451. [Google Scholar] [CrossRef]
  71. Zhang, Q.; Chen, Q.; Wang, S.; Hong, Y.; Wang, Z. Rice and cold stress: Methods for its evaluation and summary of cold tolerance-related quantitative trait loci. Rice 2014, 7, 1–12. [Google Scholar] [CrossRef] [Green Version]
  72. Liliane, T.N.; Charles, M.S. Factors affecting yield of crops. In Agronomy-Climate Change Food Security; Amanullah, Ed.; Intech Open: London, UK, 2020. [Google Scholar] [CrossRef]
  73. Liu, X.; Zhou, Y.; Xiao, J.; Bao, F. Effects of chilling on the structure, function and development of chloroplasts. Front. Plant Sci. 2018, 22, 1715. [Google Scholar] [CrossRef] [Green Version]
  74. Mamun, E.A.; Alfred, S.; Cantrill, L.C.; Overall, R.L.; Sutton, B.G. Effects of chilling on male gametophyte development in rice. Cell Biol. Int. 2006, 30(7), 583–591. [Google Scholar] [CrossRef]
  75. Fahad, S.; Chen, Y.; Saud, S.; Wang, K.; Xiong, D.; Chen, C.; Wu, C.; Shah, F.; Nie, L.; Huang, J. Ultraviolet radiation effect on photosynthetic pigments, biochemical attributes, antioxidant enzyme activity and hormonal contents of wheat. J. Food Agric. Environ. 2013, 11, 1635–1641. [Google Scholar]
  76. Ali, H.; Khan, E.; Sajad, M.A. Phytoremediation of heavy metals-concepts and applications. Chemosphere 2013, 91, 869–881. [Google Scholar] [CrossRef]
  77. Rascio, N.; Navari-Izzo, F. Heavy metal hyperaccumulating plants: How and why do they do it? And what makes them so interesting? Plant Sci. 2011, 18, 169–181. [Google Scholar] [CrossRef]
  78. Nievola, C.C.; Carvalho, C.P.; Carvalho, V.; Rodrigues, E. Rapid response of plants to temperature changes. Temperature 2017, 4, 371–405. [Google Scholar] [CrossRef]
  79. Rauwane, M.; Ntushelo, K. Understanding biotic stress and hormone signalling in cassava (Manhot esculenta): Potential for using hyphenated analytical techniques. Appl. Sci. 2020, 20, 8152. [Google Scholar] [CrossRef]
  80. Haider, F.U.; Liqun, C.; Coulter, J.A.; Cheema, S.A.; Wu, J.; Zhang, R.; Farooq, M. Cadmium toxicity in plants: Impacts and remediation strategies. Ecotoxicol. Environ. Saf. 2021, 211, 111887. [Google Scholar] [CrossRef]
  81. Emamverdian, A.; Ding, Y.; Mokhberdoran, F.; Xie, Y. Heavy metal stress and some mechanisms of plant defense response. Sci. World J. 2015, 2015, 756120. [Google Scholar] [CrossRef] [Green Version]
  82. Mani, A.; Sankaranarayanan, K. Heavy Metal and Mineral Element-Induced Abiotic Stress in Rice Plant. In Rice Crop—Current Developments; Shah, F., Khan, Z.H., Iqbal, A., Eds.; IntechOpen: London, UK, 2018. [Google Scholar] [CrossRef]
  83. Lamaoui, M.; Jemo, M.; Datla, R.; Bekkaoui, F. Heat and drought stresses in crops and approaches for their mitigation. Front. Chem. 2018, 6, 26. [Google Scholar] [CrossRef]
  84. Hewedy, O.A.; Abdel Lateif, K.S.; Seleiman, M.F.; Shami, A.; Albarakaty, F.M.; M El-Meihy, R. Phylogenetic diversity of Trichoderma strains and their antagonistic potential against soil-borne pathogens under stress conditions. Biology 2020, 9, 189. [Google Scholar] [CrossRef]
  85. Zhu, J.K. Abiotic Stress Signaling and Responses in Plants. Cell 2016, 167, 313–324. [Google Scholar] [CrossRef] [Green Version]
  86. Plieth, C.; Hansen, U.P.; Knight, H.; Knight, M.R. Temperature sensing by plants: The primary characteristics of signal perception and calcium response. Plant J. 1999, 18, 491–497. [Google Scholar] [CrossRef] [Green Version]
  87. Knight, H.; Trewavas, A.J.; Knight, M.R. Cold calcium signaling in Arabidopsis involves two cellular pools and a change in calcium signature after acclimation. Plant Cell 1996, 8, 489–503. [Google Scholar]
  88. Knight, M.R.; Knight, H. Low-temperature perception leading to gene expression and cold tolerance in higher plants. New Phytologist. 2012, 195, 737–751. [Google Scholar] [CrossRef]
  89. Tyystjarvi, E. Photoinhibition of photosystem II. Int. Rev. Cell Mol. Biol. 2013, 300, 243–303. [Google Scholar]
  90. Xiong, L.; Zhu, J.K. Abiotic stress signal transduction in plants: Molecular and genetic perspectives. Physiol. Plant. 2001, 112, 152–166. [Google Scholar] [CrossRef] [Green Version]
  91. Sha Valli Khan, P.S.; Nagamallaiah, G.V.; Dhanunjay Rao, M.; Sergeant, K.; Hausman, J.F. Abioticstress tolerance in plants insights from proteomics. In Emerging Technologies and Management of Crop Stress Tolerance; Ahmad, P., Ed.; Elsevier Academic Press: London, UK, 2014; Volume 2, pp. 23–68. [Google Scholar]
  92. Skalak, J.; Nicolas, K.L.; Vankova, R.; Hejatko, J. Signal integration in plant abiotic stress responses via multistep phosphorelay signaling. Front. Plant Sci. 2021, 12, 644823. [Google Scholar] [CrossRef]
  93. Magwanga, R.O.; Lu, P.; Kirungu, J.N.; Lu, H.; Wang, X.; Cai, X.; Zhou, Z.; Zhang, Z.; Salih, H.; Wang, K.; et al. Characterization of the late embryogenesis abundant (LEA) proteins family and their role in drought stress tolerance in upland cotton. BMC Genet. 2018, 15, 6. [Google Scholar] [CrossRef]
  94. Juven-Gershon, T.; Hsu, J.Y.; Theisen, J.W.; Kadonaga, J.T. The RNA polymerase II core promoter—The gateway to transcription. Curr. Opin. Cell Biol. 2008, 20, 253–259. [Google Scholar] [CrossRef] [Green Version]
  95. Hu, H.; Dai, M.; Yao, J.; Xiao, B.; Li, X.; Zhang, Q.; Xiong, L. Overexpressing a NAM, ATAF, and CUC (NAC) transcription factor enhances drought resistance and salt tolerance in rice. Proc. Natl. Acad. Sci. USA 2006, 103, 12987–12992. [Google Scholar] [CrossRef] [Green Version]
  96. Phukan, U.J.; Jeena, G.S.; Shukla, R.K. WRKY transcription factors: Molecular regulation and stress responses in plants. Front. Plant Sci. 2016, 3, 760. [Google Scholar] [CrossRef] [Green Version]
  97. Wu, X.; Shiroto, Y.; Kishitani, S.; Ito, Y.; Toriyama, K. Enhanced heat and drought tolerance in transgenic rice seedlings overexpressing OsWRKY11 under the control of HSP101 promoter. Plant Cell Rep. 2009, 28, 21–30. [Google Scholar] [CrossRef]
  98. Singh, D.; Laxmi, A. Transcriptional regulation of drought response: A tortuous network of transcriptional factors. Front. Plant Sci. 2015, 6, 895. [Google Scholar] [CrossRef] [Green Version]
  99. Liu, Q.; Kasuga, M.; Sakuma, Y.; Abe, H.; Miura, S.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Two transcription factors, DREB1 and DREB2, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought- and low-temperature-responsive gene expression, respectively. Arabidopsis. Plant Cell 1998, 10, 1391–1406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Medina, J.; Bargues, M.; Terol, J.; Perez-Alonso, M.; Salinas, J. The Arabidopsis CBF gene family is composed of three genes encoding AP2 domain-containing proteins whose expression is regulated by low temperature but not by abscisic acid or dehydration. Plant Physiol. 1999, 119, 463–470. [Google Scholar] [CrossRef] [Green Version]
  101. Jaglo-Ottosen, K.R.; Gilmour, S.J.; Zarka, D.G.; Schabenberger, O.; Thomashow, M.F. Arabidopsis CBF1 overexpression induces COR genes and enhances freezing tolerance. Science 1998, 280, 104–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Shi, Y.; Ding, Y.; Yang, S. Molecular regulation of CBF signaling in cold acclimation. Trends Plant Sci. 2018, 23, 623–637. [Google Scholar] [CrossRef] [PubMed]
  103. Floris, M.; Mahgoub, H.; Lanet, E.; Robaglia, C.; Menand, B. Posttranscriptional regulation of gene expression in plants during abiotic stress. Int. J. Mol. Sci. 2009, 10, 3168–3185. [Google Scholar] [CrossRef] [PubMed]
  104. Lee, B.H.; Kapoor, A.; Zhu, J.; Zhu, J.K. Stabilized1, a stress-upregulated nuclear protein, is required for pre-mRNA splicing, mRNA turnover, and stress tolerance in Arabidopsis. Plant Cell 2006, 18, 1736–1749. [Google Scholar] [PubMed] [Green Version]
  105. Guan, Q.; Wu, J.; Zhang, Y.; Jiang, C.; Liu, R.; Chai, C.; Zhu, J.A. Dead box RNA helicase is critical for pre-mRNA splicing, cold-responsive gene regulation, and cold tolerance in Arabidopsis. Plant Cell 2013, 25, 342–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Calixto, C.P.G.; Guo, W.B.; James, A.B.; Tzioutziou, N.A.; Entizne, J.C.; Panter, P.E.; Knight, H.; Nimmo, H.G.; Zhang, R.; Brown, J.W. Rapid and dynamic alternative splicing impacts the Arabidopsis cold response transcriptome. Plant Cell 2018, 30, 1424–1444. [Google Scholar]
  107. Sunkar, R.; Kapoor, A.; Zhu, J.K. Posttranscriptional induction of two Cu/Zn superoxide dismutase genes in Arabidopsis is mediated by downregulation of miR398 and important for oxidative stress tolerance. Plant Cell 2006, 18, 2051–2065. [Google Scholar]
  108. Zhu, J.K. Salt and drought stress signal transduction in plants. Annu. Rev. Plant Biol. 2002, 53, 247–273. [Google Scholar]
  109. Yoshida, R.; Umezawa, T.; Mizoguchi, T.; Takahashi, S.; Takahashi, F.; Shinozaki, K. The regulatory domain of SRK2E/OST1/SnRK2.6 interacts with ABI1 and integrates abscisic acid (ABA) and osmotic stress signals controlling stomatal closure in Arabidopsis. J. Biologic. Chem. 2006, 281, 5310–5318. [Google Scholar] [CrossRef] [Green Version]
  110. Kim, J.M.; Sasaki, T.; Ueda, M.; Sako, K.; Seki, M. Chromatin changes in response to drought, salinity, heat, and cold stresses in plants. Front. Plant Sci. 2015, 6, 114. [Google Scholar] [CrossRef] [Green Version]
  111. To, T.K.; Nakaminami, K.; Kim, J.M.; Morosawa, T.; Ishida, J.; Tanaka, M.; Yokoyama, S.; Shinozaki, K.; Seki, M. Arabidopsis HDA6 is required for freezing tolerance. Biochem. Biophys. Res. Commun. 2011, 406, 414–419. [Google Scholar] [CrossRef]
  112. Roy, D.; Paul, A.; Roy, A.; Ghosh, R.; Ganguly, P.; Chaudhuri, S. Deferential acetylation of histone H3 at the regulatory region of OsDREB1b promoter facilitates chromatin remodelling and transcription activation during cold stress. PLoS ONE 2004, 9, e100343. [Google Scholar]
  113. Hu, Y.; Zhang, L.; Zhao, L.; Li, J.; He, S.B.; Zhou, K.; Yang, F.; Huang, M.; Jiang, L.; Li, L. Trichostatin A selectively suppresses the cold-induced transcription of the ZmDREB1 gene in maize. PLoS ONE 2011, 6, e22132. [Google Scholar] [CrossRef]
  114. He, X.J.; Hsu, Y.F.; Zhu, S.H.; Liu, H.L.; Pontes, O.; Zhu, J.H.; Cui, X.; Wang, C.S.; Zhu, J.K. A conserved transcriptional regulator is required for RNA-directed DNA methylation and plant development. Genes Dev. 2009, 23, 2717–2722. [Google Scholar] [CrossRef] [Green Version]
  115. Chan, Z.L.; Wang, Y.P.; Cao, M.J.; Gong, Y.H.; Mu, Z.X.; Wang, H.Q.; Hu, Y.; Deng, X.; He, X.J.; Zhu, J.K. RDM4 modulates cold stress resistance in Arabidopsis partially through the CBF-mediated pathway. New Phytol. 2016, 209, 1527–1539. [Google Scholar] [CrossRef] [Green Version]
  116. Sunkar, R.; Bartels, D.; Kirch, H.H. Overexpression of a stress-inducible dehydrogenase gene from Arabidopsis thaliana in transgenic plants improves stress tolerance. Plant J. 2003, 35, 452–464. [Google Scholar] [CrossRef]
  117. Wang, W.; Vinocur, B.; Altman, A. Plant responses to drought, salinity and extreme temperatures: Towards genetic engineering for stress tolerance. Planta 2003, 218, 1–14. [Google Scholar] [CrossRef]
  118. Xiong, L.; Schumaker, K.S.; Zhu, J.K. Cell signaling during cold, drought, and salt stress. Plant Cell 2002, 14, 165–183. [Google Scholar] [CrossRef] [Green Version]
  119. Serrano, R.; Gaxiola, R.; Rios, G.; Forment, J.; Vicente, O.; Ros, R. Salt stress proteins identified by a functional approach in yeast. Monatshefte Chem./Chem. Mon. 2003, 134, 1445–1464. [Google Scholar] [CrossRef]
  120. Fahad, S.; Nie, L.; Chen, Y.; Wu, C.; Xiong, D.; Saud, S.; Hongyan, L.; Cui, K.; Huang, J. Crop plant hormones and environmental stress. Sustain. Agric. Rev. 2015, 15, 371–400. [Google Scholar]
  121. Garay-Arroyo, A.; Colmenero-Flores, J.M.; Garciarrubio, A.; Covarrubias, A.A. Highly hydrophilic proteins in prokaryotes and eukaryotes are common during conditions of water deficit. J. Biol. Chem. 2000, 275, 5668–5674. [Google Scholar] [CrossRef] [Green Version]
  122. Fahad, S.; Bano, A. Effect of salicylic acid on physiological and biochemical characterization of maize grown in the saline area. Pak. J. Bot. 2012, 44, 1433–1438. [Google Scholar]
  123. Eimer, M. Transgenic drought- and salt-tolerant plants. Genet. Eng. Newsl. 2004, 15, 1–14. [Google Scholar]
  124. Serrano, R.; Montesinos, C. Molecular bases of desiccation tolerance in plant cells and potential applications in food dehydration. Food Sci. Technol. Int. 2003, 9, 157–161. [Google Scholar] [CrossRef]
  125. Quintero, F.J.; Ohta, M.; Shi, H.; Zhu, J.K.; Pardo, J.M. Reconstitution in yeast of the Arabidopsis SOS signaling pathway for Na+ homeostasis. Proc. Natl. Acad. Sci. USA 2002, 25, 9061–9066. [Google Scholar] [CrossRef] [Green Version]
  126. Seifikalhor, M.; Aliniaeifard, S.; Shomali, A.; Azad, N.; Hassani, B.; Lastochkina, O.; Li, T. Calcium signaling and salt tolerance are diversely entwined in plants. Plant Signal Behav. 2019, 14, 1665455. [Google Scholar] [CrossRef]
  127. Sandrini, M.; Nerva, L.; Sillo, F.; Balestrini, R.; Chitarra, W.; Zampieri, E. Abiotic stress and belowground microbiome: The potential of omics approaches. Int. J. Mol. Sci. 2022, 23, 1091. [Google Scholar] [CrossRef]
  128. Antoine, S.; Heriche, M.; Boussageon, R.; Noceto, P.A.; van Tuinen, D.; Wipf, D.; Courty, P.E. A historical perspective on mycorrhizal mutualism emphasizing arbuscular mycorrhizas and their emerging challenges. Mycorrhiza 2021, 31, 637–653. [Google Scholar]
  129. 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] [PubMed] [Green Version]
  130. Chandrasekaran, M.; Boopathi, T.; Manivannan, P. Comprehensive assessment of ameliorative effects of AMF in alleviating abiotic stress in tomato plants. J. Fungi 2021, 7, 303. [Google Scholar] [CrossRef] [PubMed]
  131. Kosova, K.; Vitamvas, P.; Urban, M.O.; Prasil, I.T. Plant proteome responses to salinity Stress-Comparison of glycophytes and halophytes. Funct. Plant Biol. 2013, 40, 775–786. [Google Scholar] [CrossRef] [PubMed]
  132. Pan, J.; Peng, F.; Tedeschi, A.; Xue, X.; Wang, T.; Liao, J.; Zhang, W.; Huang, C. Do halophytes and glycophytes differ in their interactions with arbuscular mycorrhizal fungi under salt stress? A meta-analysis. Bot. Stud. 2020, 61, 13. [Google Scholar] [CrossRef] [Green Version]
  133. Auge, R.M.; Toler, H.D.; Saxton, A.M. Arbuscular mycorrhizal symbiosis and osmotic adjustment in response to NaCl stress: A meta-analysis. Front. Plant Sci. 2014, 5, 562. [Google Scholar]
  134. Evelin, H.; Giri, B.; Kapoor, R. Ultrastructural evidence for AMF mediated salt stress mitigation in Trigonella foenum-graecum. Mycorrhiza 2013, 23, 71–86. [Google Scholar] [CrossRef]
  135. Evelin, H.; Kapoor, R.; Giri, B. Arbuscular mycorrhizal fungi in alleviation of salt stress: A review. Ann. Bot. 2009, 104, 1263–1280. [Google Scholar] [CrossRef] [Green Version]
  136. Eroglu, C.G.; Cabral, C.; Ravnskov, S.; Bak Topbjerg, H.; Wollenweber, B. Arbuscular mycorrhiza influences carbon-use efficiency and grain yield of wheat grown under pre- and post-anthesis salinity stress. Plant Biol. 2020, 22, 863–871. [Google Scholar] [CrossRef]
  137. Sharma, K.; Gupta, S.; Thokchom, S.D.; Jangir, P.; Kapoor, R. Arbuscular mycorrhiza-mediated regulation of polyamines and aquaporins during abiotic stress: Deep insights on the recondite players. Front. Plant Sci. 2021, 12, 1072. [Google Scholar] [CrossRef]
  138. Volpe, V.; Chitarra, W.; Cascone, P.; Volpe, M.G.; Bartolini, P.; Moneti, G.; Pieraccini, G.; Di Serio, C.; Maserti, B.; Guerrieri, E.; et al. The association with two different arbuscular mycorrhizal fungi differently affects water stress tolerance in tomato. Front. Plant Sci. 2018, 9, 1480. [Google Scholar] [CrossRef] [Green Version]
  139. Chitarra, W.; Pagliarani, C.; Maserti, B.; Lumini, E.; Siciliano, I.; Cascone, P.; Schubert, A.; Gambino, G.; Balestrini, R.; Guerrieri, E. Insights on the impact of arbuscular mycorrhizal symbiosis on tomato tolerance to water stress. Plant Physiol. 2016, 171, 1009–1023. [Google Scholar] [CrossRef] [Green Version]
  140. Rivero, J.; Alvarez, D.; Flors, V.; Azcon-Aguilar, C.; Pozo, M.J. Root metabolic plasticity underlies functional diversity in mycorrhiza-enhanced stress tolerance in tomato. New Phytol. 2018, 220, 1322–1336. [Google Scholar] [CrossRef] [Green Version]
  141. Sandhya, V.; Ali, S.K.Z.; Minakshi, G.; Reddy, G.; Venkateswarlu, B. Alleviation of drought stress effects in sunflower seedlings by the exopolysaccharides producing Pseudomonas putida strain GAP-P45. Biol. Fertil. Soils 2009, 46, 17–26. [Google Scholar] [CrossRef]
  142. Arshad, M.; Sharoona, B.; Mahmood, T. Inoculation with Pseudomonas spp. containing ACC deaminase partially eliminate the effects of drought stress on growth, yield and ripening of pea (Pisum sativum L.). Pedosphere 2008, 18, 611–620. [Google Scholar] [CrossRef]
  143. Creus, C.M.; Sueldo, R.J.; Barassi, C.A. Water relations and yield in Azospirillum-inoculated wheat exposed to drought in the field. Can. J. Bot. 2004, 82, 273–281. [Google Scholar] [CrossRef] [Green Version]
  144. Cho, E.K.; Hong, C.B. Over-expression of tobacco NtHSP70-1 contributes to drought-stress tolerance in plants. Plant Cell Rep. 2006, 25, 349–358. [Google Scholar] [CrossRef]
  145. Rodriguez, A.A.; Stella, A.M.; Storni, M.M.; Zulpa, G.; Zaccaro, M.C. Effect of cyanobacterial extracellular products and gibberellic acid on salinity tolerance in Oryza sativa L. Saline Syst. 2006, 2, 7. [Google Scholar] [CrossRef] [Green Version]
  146. Barka, E.A.; Nowak, J.; Clement, C. Enhancement of chilling resistance of inoculated grapevine plantlets with a plant growth-promoting rhizobacterium Burkholderiaphytofirmans strain PsJN. Appl. Environ. Microbiol. 2006, 72, 7246–7252. [Google Scholar] [CrossRef] [Green Version]
  147. Madhaiyan, M.; Poonguzhali, S.; Sa, T. Metal tolerating methylotrophic bacteria reduces nickel and cadmium toxicity and promotes plant growth of tomato (Lycopersicon esculentum L.). Chemosphere 2007, 69, 220–228. [Google Scholar] [CrossRef]
  148. SaravanaMartins, S.J.; Rocha, G.A.; De Melo, H.C.; De Castro Georg, R.; Ulhoa, C.J.; De Campos Dianese, É.; Oshiquiri, L.H.; da Cunha, M.G.; da Rocha, M.R.; de Araújo, L.G.; et al. Plant-associated bacteria mitigate drought stress in soybean. Environ. Sci. Pollut. Res. 2018, 25, 13676–13686. [Google Scholar] [CrossRef]
  149. Figueiredo, M.V.B.; Burity, H.A.; Martinez, C.R.; Chanway, C.P. Alleviation of drought stress in common bean (Phaseolus vulgaris L.) by coinoculation with Paenibacilluspolymyxa and Rhizobium tropici. Appl. Soil Ecol. 2008, 40, 182–188. [Google Scholar] [CrossRef]
  150. Kohler, J.; Hernández, J.A.; Caravaca, F.; Roldán, A. Plant growth promoting rhizobacteria and arbuscular mycorrhizal fungi modify alleviation biochemical mechanisms in water stressed plants. Funct. Plant. Biol. 2008, 35, 141–151. [Google Scholar] [CrossRef]
  151. Ali, S.Z.; Sandhya, V.; Grover, M.; Kishore, N.; Rao, L.V.; Venkateswarlu, B. Pseudomonas sp. strain AKM-P6 enhances tolerance of sorghum seedlings to elevated temperatures. Biol. Fert. Soils 2009, 46, 45–55. [Google Scholar] [CrossRef]
  152. Marulanda, A.; Porcel, R.; Barea, J.M.; Azcon, R. Drought tolerance and antioxidant activities in lavender plants colonized by native drought tolerant or drought sensitive Glomus species. Microb. Ecol. 2007, 54, 543–552. [Google Scholar] [CrossRef]
  153. McLellan, C.A.; Turbyville, T.J.; Wijeratne, K.; Kerschen, A.; Vierling, E.; Queitsch, C.; Whiteshell, L.; Gunatilaka, A.A.L. A rhizosphere fungus enhances Arabidopsis thermotolerance through production of an HSP90 inhibitor. Plant Physiol. 2007, 145, 174–182. [Google Scholar] [CrossRef]
  154. Zhang, H.; Kim, M.S.; Sun, Y.; Dowd, S.E.; Shi, H.; Paré, P.W. Soil bacteria confer plant salt tolerance by tissue-specific regulation of the sodium transporter HKT1. Mol. Plant Microbe Interact. 2008, 21, 737–744. [Google Scholar] [CrossRef] [Green Version]
  155. Yao, L.X.; Wu, Z.S.; Zheng, Y.Y.; Kaleem, I.; Li, C. Growth promotion and protection against salt stress by Pseudomonas putida Rs-198 on cotton. Eur. J. Soil Biol. 2010, 46, 49–54. [Google Scholar] [CrossRef]
  156. Daei, G.; Ardekani, M.R.; Rejali, F.; Teimuri, S.; Miransari, M. Alleviation of salinity stress on wheat yield, yield components, and nutrient uptake using arbuscular mycorrhizal fungi under field conditions. J. Plant Physiol. 2009, 66, 617–625. [Google Scholar] [CrossRef]
  157. Lazzarotto, P.; Calanca, P.; Semenov, M.; Fuhrer, J. Transient responses to increasing CO2 and climate change in an unfertilized grass clover sward. Clim. Res. 2010, 41, 221–232. [Google Scholar] [CrossRef]
  158. Lim, J.H.; Kim, S.D. Induction of drought stress resistance by multifunctional PGPR Bacillus licheniformis K11 in pepper. Plant Pathol. J. 2013, 29, 201–208. [Google Scholar] [CrossRef]
  159. Timmusk, S.; El-Daim, I.A.A.; Copolovici, L.; Tanilas, T.; Kannaste, A.; Behers, L.; Nevo, E.; Seisenbaeva, G.; Stenström, E.; Niinemets, Ü. 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] [PubMed] [Green Version]
  160. del Amor, F.; Cuadra-Crespo, P. Plant growth-promoting bacteria as a tool to improve salinity tolerance in sweet pepper. Funct. Plant Biol. 2012, 39, 82–90. [Google Scholar] [CrossRef] [PubMed]
  161. Naveed, M.; Mitter, B.; Reichenauer, T.G.; Wieczorek, K.; Sessitsch, A. Increased drought stress resilience of maize through endophytic colonization by Burkholderiaphytofirmans PsJN and Enterobacter sp FD17. Environ. Exp. Bot. 2014, 97, 30–39. [Google Scholar] [CrossRef]
  162. Kasotia, A.; Varma, A.; Choudhary, D.K. Pseudomonas-mediated mitigation of salt stress and growth promotion in Glycine max. Agric. Res. 2015, 4, 31–41. [Google Scholar] [CrossRef]
  163. Plociniczak, T.; Sinkkonen, A.; Romantschuk, M.; Piotrowska-seget, Z. Characterization of Enterobacter intermedius MH8b and its use for the enhancement of heavy metal uptake by Sinapsis alba L. Appl. Soil Ecol. 2013, 63, 1–7. [Google Scholar] [CrossRef]
  164. Wang, C.; Yang, W.; Wang, C.; Gu, C.; Niu, D.; Liu, H.X.; Wang, Y.P.; Guo, J.H. Induction of drought tolerance in cucumber plants by a consortium of three plant growth promoting rhizobacterium strains. PLoS ONE 2012, 7, e52565. [Google Scholar] [CrossRef] [Green Version]
  165. Mathew, D.C.; Ho, Y.N.; Gicana, R.G.; Mathew, G.M.; Chien, M.C.; Huang, C.C. A rhizosphere-associated symbiont, Photobacterium spp. strain MELD1, and its targeted synergistic activity for phytoprotection against mercury. PLoS ONE 2015, 10, e0121178. [Google Scholar] [CrossRef]
  166. Adediran, G.A.; Ngwenya, B.T.; Mosselmans, J.F.W.; Heal, K.V. Bacteria–zinc co-localization implicates enhanced synthesis of cysteine-rich peptides in zinc detoxification when Brassica juncea is inoculated with accumulation in N2-fixing alfalfa (Medicago sativa). J. Plant. Physiol. 2016, 146, 541–546. [Google Scholar]
  167. Jha, Y.; Sablok, G.; Subbarao, N.; Sudhakar, R.; Fazil, M.H.U.T.; Subramanian, R.B.; Squartini, A.; Kumar, S. Bacterial-induced expression of RAB18 protein in Orzya sativa salinity stress and insights into molecular interaction with GTP ligand. J. Mol. Recognit. 2014, 27, 521–527. [Google Scholar] [CrossRef]
  168. Arun, K.D.; Sabarinathan, K.G.; Gomathy, M.; Kannan, R.; Balachandar, D. Mitigation of drought stress in rice crop with plant growth-promoting abiotic stress-tolerant rice phyllosphere bacteria. J. Basic Microbiol. 2020, 60, 768–786. [Google Scholar] [CrossRef]
  169. Ashry, N.M.; Alaidaroos, B.A.; Mohamed, S.A.; Badr, O.A.; El-Saadony, M.T.; Esmael, A. Utilization of drought-tolerant bacterial strains isolated from harsh soils as a plant growth-promoting rhizobacteria (PGPR). Saudi J. Biol. Sci. 2022, 29, 1760–1769. [Google Scholar] [CrossRef]
  170. Filgueiras, L.; Silva, R.; Almeida, I.; Vidal, M.; Baldani, J.I.; Meneses, C.H.S.G. Gluconacetobacterdiazotrophicus mitigates drought stress in Oryza sativa L. Plant Soil 2020, 451, 57–73. [Google Scholar] [CrossRef]
  171. Kour, D.; Rana, K.L.; Yadav, A.N.; Sheikh, I.; Kumar, V.; Dhaliwal, H.S.; Saxena, A.K. Amelioration of drought stress in Foxtail millet (Setariaitalica L.) by P-solubilizing drought-tolerant microbes with multifarious plant growth promoting attributes. Environ. Sustain. 2020, 3, 23–34. [Google Scholar] [CrossRef]
  172. Li, H.; Guo, Q.; Jing, Y.; Liu, Z.; Zheng, Z.; Sun, Y.; Xue, Q.; Lai, H. Application of Streptomyces pactum Act12 enhances drought resistance in wheat. J. Plant Growth Regul. 2020, 39, 122–132. [Google Scholar] [CrossRef]
  173. Kumar, M.; Mishra, S.; Dixit, V.; Kumar, M.; Agarwal, L.; Chauhan, P.S.; Nautiyal, C.S. Synergistic effect of Pseudomonas putida and Bacillus amyloliquefaciens ameliorates drought stress in chickpea (Cicer arietinum L.). Plant Signal. Behav. 2016, 11, e1071004. [Google Scholar] [CrossRef]
  174. Kumar, A.; Singh, S.; Gaurav, A.K.; Srivastava, S.; Verma, J.P. Plant growth-promoting bacteria: Biological tools for the mitigation of salinity stress in plants. Front. Microbiol. 2020, 11, 1216. [Google Scholar] [CrossRef]
  175. Ferreira, M.J.; Silva, H.; Cunha, A. Siderophore-producing rhizobacteria as a promising tool for empowering plants to cope with iron limitation in saline soils: A review. Pedosphere 2019, 29, 409–420. [Google Scholar] [CrossRef]
  176. Khanna, K.; Jamwal, V.L.; Gandhi, S.G.; Ohri, P.; Bhardwaj, R. Metal resistant PGPR lowered Cd uptake and expression of metal transporter genes with improved growth and photosynthetic pigments in Lycopersicon esculentum under metal toxicity. Sci. Rep. 2019, 9, 5855. [Google Scholar] [CrossRef] [Green Version]
  177. Chaudhary, D.; Sindhu, S.S. Amelioration of salt stress in chickpea (Cicer arietinum L.) by coinculation of ACC deaminase-containing rhizospheric bacteria with Mesorhizobium strains. Legume Res. 2017, 40, 80–86. [Google Scholar] [CrossRef] [Green Version]
  178. Marulanda, A.; Azcón, R.; Chaumont, F.; Ruiz-Lozano, J.M.; Aroca, R. Regulation of plasma membrane aquaporins by inoculation with a Bacillus megaterium strain in maize (Zea mays L.) plants under unstressed and salt-stressed conditions. Planta 2010, 232, 533–543. [Google Scholar] [CrossRef]
  179. Balestrini, R.; Brunetti, C.; Chitarra, W.; Nerva, L. Photosynthetic traits and nitrogen uptake in crops: Which is the role of arbuscular mycorrhizal fungi? Plants 2020, 9, 1105. [Google Scholar] [CrossRef]
  180. Mathur, S.; Tomar, R.S.; Jajoo, A. Arbuscular mycorrhizal fungi (AMF) protects photosynthetic apparatus of wheat under drought stress. Photosynth. Res. 2019, 139, 227–238. [Google Scholar] [CrossRef]
  181. Zou, Y.N.; Wu, Q.S.; Kuca, K. Unravelling the role of arbuscular mycorrhizal fungi in mitigating the oxidative burst of plants under drought stress. Plant Biol. 2021, 23, 50–57. [Google Scholar] [CrossRef]
  182. Qin, Y.; Druzhinina, I.S.; Pan, X.; Yuan, Z. Microbially mediated plant salt tolerance and microbiome-based solutions for saline agriculture. Biotechnol. Adv. 2016, 34, 1245–1259. [Google Scholar] [CrossRef]
  183. Fiorilli, V.; Vallino, M.; Biselli, C.; Faccio, A.; Bagnaresi, P.; Bonfante, P. Host and non-host roots in rice: Cellular and molecular approaches reveal differential responses to arbuscular mycorrhizal fungi. Front. Plant Sci. 2015, 6, 636. [Google Scholar] [CrossRef]
  184. Fiorilli, V.; Catoni, M.; Miozzi, L.; Novero, M.; Accotto, G.P.; Lanfranco, L. Global and cell-type gene expression profiles in tomato plants colonized by an arbuscular mycorrhizal fungus. New Phytol. 2009, 184, 975–987. [Google Scholar] [CrossRef]
  185. Balestrini, R.; Salvioli, A.; Dal Molin, A.; Novero, M.; Gabelli, G.; Papparelli, E.; Marroni, F.; Bonfante, P. Impact of an arbuscular mycorrhizal fungus versus a mixed microbial inoculum on the transcriptome reprogramming of grapevine roots. Mycorrhiza 2017, 7, 417–430. [Google Scholar] [CrossRef]
  186. Fiorilli, V.; Vannini, C.; Ortolani, F.; Garcia-Seco, D.; Chiapello, M.; Novero, M.; Domingo, G.; Terzi, V.; Morcia, C.; Bagnaresi, P.; et al. Omics approaches revealed how arbuscular mycorrhizal symbiosis enhances yield and resistance to leaf pathogen in wheat. Sci. Rep. 2018, 8, 9625. [Google Scholar] [CrossRef] [Green Version]
  187. Balestrini, R.; Rosso, L.C.; Veronico, P.; Melillo, M.T.; De Luca, F.; Fanelli, E.; Colagiero, M.; Salvioli, A.; Ciancio, A.; Pentimone, I. Transcriptomic responses to water deficit and nematode infection in mycorrhizal tomato roots. Front. Microbiol. 2019, 10, 1807. [Google Scholar] [CrossRef] [Green Version]
  188. Grover, M.; Bodhankar, S.; Maheswari, M.; Srinivasarao, C. 6 Actinomycetes as mitigators of climate change and abiotic stress. In Plant Growth Promoting Actinobacteria; Subramaniam, G., Arumugam, S., Rajendran, V., Eds.; Springer: Singapore, 2016; pp. 203–212. [Google Scholar]
  189. Yandigeri, M.S.; Meena, K.K.; Singh, D.; Malviya, N.; Singh, D.P.; Solanki, M.K.; Yadav, A.K.; Arora, D.K. Drought-tolerant endophytic Actinobacteria promote growth of wheat (Triticum aestivum) under water stress conditions. Plant Growth Regul. 2012, 68, 411–420. [Google Scholar] [CrossRef]
  190. Singh, R.P.; Pandey, D.M.; Jha, P.N.; Ma, Y. ACC deaminase producing rhizobacterium Enterobacter cloacae ZNP-4 enhance abiotic stress tolerance in wheat plant. PLoS ONE 2022, 17, e0267127. [Google Scholar] [CrossRef] [PubMed]
  191. Aly, M.M.; El-Sabbagh, S.M.; El-Shouny, W.A.; Ebrahim, M.K.H. Physiological response of Zea mays to NaCl stress with respect to Azotobacter chroococcum and Streptomyces niveus. Pak. J. Biol. Sci. 2003, 6, 2073–2080. [Google Scholar] [CrossRef]
  192. Palaniyandi, S.A.; Damodharan, K.; Yang, S.H.; Suh, J.W. Streptomyces sp. Strain PGPA39 alleviates salt stress and promotes growth of ‘Micro Tom’ tomato plants. J. Appl. Microbiol. 2014, 117, 766–773. [Google Scholar] [CrossRef] [PubMed]
  193. Ghavami, N.; Alikhani, H.A.; Pourbabaei, A.A.; Besharati, H. Effects of two new siderophore-producing rhizobacteria on growth and iron content of maize and canola plants. J. Plant Nutr. 2017, 40, 736–746. [Google Scholar] [CrossRef]
  194. Hasegawa, S.; Meguro, A.; Toyoda, K.; Nishimura, T.; Kunoh, H. Drought tolerance of tissue-cultured seedlings of mountain laurel (Kalmia latifolia L.) induced by an endophytic actinomycete II. Acceleration of callose accumulation and lignification. Actinomycetologica 2005, 19, 13–17. [Google Scholar] [CrossRef]
  195. Johnston-Monje, D.; Lundberg, D.S.; Lazarovits, G.; Reis, V.M.; Raizada, M.N. Bacterial populations in juvenile maize rhizospheres originate from both seed and soil. Plant Soil 2016, 405, 337–355. [Google Scholar] [CrossRef] [Green Version]
  196. Genitsaris, S.; Stefanidou, N.; Leontidou, K.; Matsi, T.; Karamanoli, K.; Mellidou, I. Bacterial communities in the rhizosphere and phyllosphere of halophytes and drought-tolerant plants in mediterranean ecosystems. Microorganisms 2020, 8, 1708. [Google Scholar] [CrossRef]
  197. Ullah, M.A.; Mahmood, I.A.; Ali, A.; Nawaz, Q.; Sultan, T.; Zaman, B.U. Effect of inoculation methods of biozote-max (plant growth promoting rhizobacteria-pgpr) on growth and yield of rice under naturally salt-affected soil. Res. Plant Biol. 2017, 7, 24–26. [Google Scholar] [CrossRef]
  198. Grover, M.; Bodhankar, S.; Sharma, A.; Sharma, P.; Singh, J.; Nain, L. PGPR mediated alterations in root traits: Way toward sustainable crop production. Front. Sustain. Food Syst. 2021, 4, 618230. [Google Scholar] [CrossRef]
  199. Vanamala, P.; Sultana, U.; Sindhura, P.; Gul, M.Z. Plant Growth-Promoting Rhizobacteria (PGPR): A Unique Strategy for Sustainable Agriculture. In Handbook of Research on Microbial Remediation and Microbial Biotechnology for Sustainable Soil; IGI Global: Hershey, PA, USA, 2021; pp. 332–357. [Google Scholar]
  200. Ansari, F.A.; Ahmad, I. Alleviating drought stress of crops through PGPR: Mechanism and application. In Microbial Interventions in Agriculture and Environment; Singh, D.P., Gupta, V.K., Prabha, R., Eds.; Springer: Singapore, 2019; pp. 341–358. [Google Scholar] [CrossRef]
  201. Khan, N.; Bano, A.; Shahid, M.A.; Nasim, W.; Babar, A. Interaction between PGPR and PGR for water conservation and plant growth attributes under drought condition. Biologia 2018, 73, 1083–1098. [Google Scholar] [CrossRef]
  202. Sinha, D.; Mukherjee, S.; Mahapatra, D. Multifaceted potential of Plant Growth Promoting Rhizobacteria (PGPR): An overview. In Handbook of Research on Microbial Remediation and Microbial Biotechnology for Sustainable Soil; Ahmad Malik, J., Ed.; IGI Global: Hershey, PA, USA, 2021; pp. 205–268. [Google Scholar] [CrossRef]
  203. Monteiro, G.; Nogueira, G.; Neto, C.; Nascimento, V.; Freitas, J. Promotion of Nitrogen Assimilation by Plant Growth-Promoting Rhizobacteria. In Nitrogen in Agriculture-Physiological, Agricultural and Ecological Aspects; Intech Open: London, UK, 2021. [Google Scholar]
  204. Jochum, M.D.; Mcwilliams, K.L.; Borrego, E.J.; Kolomiets, M.V.; Niu, G.; Pierson, E.A.; Jo, Y.K. Bioprospecting plant growth-promoting rhizobacteria that mitigate drought stress in grasses. Front. Microbiol. 2019, 10, 2106. [Google Scholar] [CrossRef]
  205. Bano, Q.; Ilyas, N.; Bano, A.; Zafar, N.; Akram, A.; Hassan, F. Effect of Azospirillum inoculation on maize (Zea mays L.) under drought stress. Pak. J. Bot. 2013, 45, 13–20. [Google Scholar]
  206. Kour, D.; Rana, K.L.; Sheikh, I.; Kumar, V.; Yadav, A.N.; Dhaliwal, H.S.; Saxena, A.K. Alleviation of drought stress and plant growth promotion by Pseudomonas libanensis EU-LWNA-33, a drought-adaptive phosphorus-solubilizing bacterium. Proc. Natl. Acad. Sci. India B Biol. Sci. 2020, 90, 785–795. [Google Scholar] [CrossRef]
  207. Zerrouk, I.Z.; Benchabane, M.; Khelifi, L.; Yokawa, K.; Ludwig-Muller, J.; Baluska, F. Pseudomonas strain isolated from date-palm rhizospheres improves root growth and promotes root formation in maize exposed to salt and aluminum stress. J. Plant Physiol. 2016, 191, 111–119. [Google Scholar] [CrossRef]
  208. 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] [Green Version]
  209. Priyanka, J.P.; Goral, R.T.; Rupal, K.S.; Saraf, M. Rhizospheric microflora: A natural alleviator of drought stress in agricultural crops. In Plant Growth Promoting Rhizobacteria for Sustainable Stress Management; Sayyed, R.Z., Ed.; Springer: Singapore, 2019; pp. 103–115. [Google Scholar] [CrossRef]
  210. Chiappero, J.; Del Rosario Cappellari, L.; Alderete, L.G.S.; Palermo, T.B.; Banchio, E. Plant growth promoting rhizobacteria improve the antioxidant status in Mentha piperita grown under drought stress leading to an enhancement of plant growth and total phenolic content. Industr. Crops Prod. 2019, 139, 111553. [Google Scholar] [CrossRef]
  211. Raheem, A.; Shaposhnikov, A.; Belimov, A.A.; Dodd, I.C.; Ali, B. Auxin production by rhizobacteria was associated with improved yield of wheat (Triticum aestivum L.) under drought stress. Arch. Agron. Soil Sci. 2018, 64, 574–587. [Google Scholar] [CrossRef]
  212. Hanif, S.; Saleem, M.F.; Sarwar, M.; Irshad, M.; Shakoor, A.; Wahid, M.A.; Khan, H.Z. Biochemically triggered heat and drought stress tolerance in rice by proline application. J Plant Growth Regul. 2021, 40, 305–312. [Google Scholar] [CrossRef]
  213. Rana, R.A.; Siddiqui, M.N.; Skalicky, M.; Brestic, M.; Hossain, A.; Kayesh, E.; Popov, M.; Hejnak, V.; Gupta, D.R.; Mahmud, N.U.; et al. Prospects of nanotechnology in improving the productivity and quality of horticultural crops. Horticulturae 2021, 7, 332. [Google Scholar] [CrossRef]
  214. Tripathi, D.K.; Singh, V.P.; Prasad, S.M.; Chauhan, D.K.; Dubey, N.K. Silicon nanoparticles (SiNp) alleviate chromium (VI) phytotoxicity in Pisum sativum (L.) seedlings. Plant Physiol. Biochem. 2015, 96, 189–198. [Google Scholar] [CrossRef]
  215. Hussain, A.; Ali, S.; Rizwan, M.; Zia ur Rehman, M.; Javed, M.R.; Imran, M.; Chatha, S.A.S.; Nazir, R. Zinc oxide nanoparticles alter the wheat physiological response and reduce the cadmium uptake by plants. Environ. Pollut. 2018, 242, 1518–1526. [Google Scholar] [CrossRef]
  216. Manzoor, N.; Ahmed, T.; Noman, M.; Shahid, M.; Nazir, M.M.; Ali, L.; Alnusaire, T.S.; Li, B.; Schulin, R.; Wang, G. Iron oxide nanoparticles ameliorated the cadmium and salinity stresses in wheat plants, facilitating photosynthetic pigments and restricting cadmium uptake. Sci. Total Environ. 2021, 769, 145221. [Google Scholar] [CrossRef]
  217. Hussain, A.; Ali, S.; Rizwan, M.; Rehman, M.Z.; ur Qayyum, M.F.; Wang, H.; Rinklebe, J. Responses of wheat (Triticum aestivum) plants grown in a Cd contaminated soil to the application of iron oxide nanoparticles. Ecotoxicol. Environ. Saf. 2019, 173, 156–164. [Google Scholar] [CrossRef]
  218. Adrees, M.; Khan, Z.S.; Ali, S.; Hafeez, M.; Khalid, S.; ur Rehman, M.Z.; Hussain, A.; Hussain, K.; Chatha, S.A.S.; Rizwan, M. Simultaneous mitigation of cadmium and drought stress in wheat by soil application of iron nanoparticles. Chemosphere 2020, 238, 124681. [Google Scholar] [CrossRef]
  219. Gao, M.; Zhou, J.; Liu, H.; Zhang, W.; Hu, Y.; Liang, J.; Zhou, J. Foliar spraying with silicon and selenium reduces cadmium uptake and mitigates cadmium toxicity in rice. Sci. Total Environ. 2018, 631–632, 1100–1108. [Google Scholar] [CrossRef]
  220. Rehman, M.Z.; Rizwan, M.; Hussain, A.; Saqib, M.; Ali, S.; Sohail, M.I.; Shafiq, M.; Hafeez, F. Alleviation of cadmium (cd) toxicity and minimizing its uptake in wheat (Triticum aestivum) by using organic carbon sources in cd-spiked soil. Environ. Pollut. 2018, 241, 557–565. [Google Scholar] [CrossRef]
  221. Ashkavand, P.; Tabari, M.; Zarafshar, M.; Tomaskova, I.; Struve, D. Effect of SiO2 nanoparticles on drought resistance in hawthorn seedlings. For. Res. Pap. 2018, 76, 350–359. [Google Scholar] [CrossRef] [Green Version]
  222. Ghorbanpour, M.; Mohammadi, H.; Kariman, K. Nanosilicon-based recovery of barley (Hordeum vulgare) plants subjected to drought stress. Environ. Sci. Nano 2020, 7, 443–461. [Google Scholar] [CrossRef]
  223. Alsaeedi, A.; El-Ramady, H.; Alshaal, T.; El-Garawany, M.; Elhawat, N.; Al-Otaibi, A. Silica nanoparticles boost growth and productivity of cucumber under water deficit and salinity stresses by balancing nutrients uptake. Plant Physiol. Biochem. 2019, 139, 1–10. [Google Scholar] [CrossRef]
  224. Behboudi, F.; Tahmasebi-Sarvestani, Z.; Kassaee, M.Z.; Modarres-Sanavy, S.A.M.; Sorooshzadeh, A.; Mokhtassi-Bidgoli, A. Evaluation of chitosan nanoparticles effects with two application methods on wheat under drought stress. J. Plant Nutr. 2019, 42, 1439–1451. [Google Scholar] [CrossRef]
  225. Das, A.; Das, B. Nanotechnology a potential tool to mitigate abiotic stress in crop plants. In Abiotic and Biotic Stress in Plants; De Oliveira, A., Ed.; IntechOpen: London, UK, 2019. [Google Scholar]
  226. Sun, D.; Hussain, H.I.; Yi, Z.; Rookes, J.E.; Kong, L.; Cahill, D.M. Delivery of abscisic acid to plants using glutathione responsive mesoporous silica nanoparticles. J. Nanosci. Nanotechnol. 2017, 18, 1615–1625. [Google Scholar] [CrossRef] [PubMed]
  227. Ye, Y.; Cota-Ruiz, K.; Hernández-Viezcas, J.A.; Valdés, C.; Medina-Velo, I.A.; Turley, R.S.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Manganese nanoparticles control salinity-modulated molecular responses in Capsicum annuum L. through priming: A sustainable approach for agriculture. ACS Sustain. Chem. Eng. 2020, 8, 1427–1436. [Google Scholar] [CrossRef]
  228. Wu, J.; Wang, T. Synergistic effect of zinc oxide nanoparticles and heat stress on the alleviation of transcriptional gene silencing in Arabidopsis thaliana. Bull. Environ. Contam. Toxicol. 2020, 104, 49–56. [Google Scholar] [CrossRef] [PubMed]
  229. Thakur, S.; Asthir, B.; Kaur, G.; Kalia, A.; Sharma, A. Zinc oxide and titanium dioxide nanoparticles influence heat stress tolerance mediated by antioxidant defense system in wheat. Cereal Res. Commun. 2021, 49, 1–12. [Google Scholar] [CrossRef]
  230. El-Saadony, M.T.; Saad, A.M.; Najjar, A.A.; Alzahrani, S.O.; Alkhatib, F.M.; Shafi, M.E.; Selem, E.; Desoky, E.S.M.; Fouda, S.E.; El-Tahan, A.M.; et al. The use of biological selenium nanoparticles to suppress Triticum aestivum L. crown and root rot diseases induced by Fusarium species and improve yield under drought and heat stress. Saudi J. Biol. Sci. 2021, 28, 4461–4471. [Google Scholar] [CrossRef]
  231. Iqbal, M.; Raja, N.I.; Mashwani ZU, R.; Hussain, M.; Ejaz, M.; Yasmeen, F. Effect of silver nanoparticles on growth of wheat under heat stress. Iran. J. Sci. Technol. Trans. A Sci. 2019, 43, 387–395. [Google Scholar] [CrossRef]
Figure 1. Plant behaviors during abiotic stress conditions.
Figure 1. Plant behaviors during abiotic stress conditions.
Life 12 01634 g001
Figure 2. Signaling mechanisms for abiotic stress tolerance in plants.
Figure 2. Signaling mechanisms for abiotic stress tolerance in plants.
Life 12 01634 g002
Figure 3. Functions of different microbes in combating abiotic stress conditions.
Figure 3. Functions of different microbes in combating abiotic stress conditions.
Life 12 01634 g003
Table 1. Different studies concerning abiotic stress tolerance in plants.
Table 1. Different studies concerning abiotic stress tolerance in plants.
Abiotic Stress TypeMechanism and Key Parameters StudiedPlants/CropsReferences
Cold stressThe CBF (C repeat binding factor) transcriptional cascade, CBF expression and CBF-independent regulons mediate the transcriptional regulation and pre-mRNA processing, export, and degradation involved in post-regulatory mechanismsArabidopsis[18]
Low-temperature stressAlter hormonal expression[19]
Heat and drought stressEnhancing the accumulation of carbohydrates[20]
Heat stressAutophagy plays a vital role in cellular homeostasis, metabolism, and other processes[21]
Heat stressceRNA networks are mediated by the differentially expressed circRNA, by the influence of various important genes, and participate in response to hydrogen peroxide, heat stress, and phytochrome signaling pathway[22]
Water stressLipid peroxidation decreases with scavenging reactive oxygen species and higher excitation energy dissolution due to photochemical quenching with reduced excitation pressure[23]
Drought stressThe physiological activities and antioxidant protective systems modulate CarMT gene overexpression[24]
Drought stressH2S endogenous production rate increases and a noteworthy transcriptional reorganization of pertinent miRNAs[25]
Drought stressA transmembrane potassium ions efflux as well as calcium and chloride ions influxes are induced due to endogenous hydrogen sulfides[26]
Cold and drought stress Dehydrins concentrated in roots and stemsBlueberry[27]
Heat stressLower accumulation of H2O2 and damage to cellsStrawberry [28,29]
Salinity stressPlant response is positively regulated due to OsH1RP1-ring finger protein 1 Maize[30]
Cold stressChanges in DNA methylationRice[31]
Heat stressLipid peroxidation as well as antioxidant enzymes in roots and leaves[32]
Drought stressE3-ubiquitin breakdown[16,33]
Heat stressCandidate genes as well as quantitative trait loci [34]
Cold stressLinear electron transport chain is downregulated and PSII is repressed, as represented by the lowering in the PSII photochemistry efficiency along with electron transport efficiencyHibiscus[35]
Water deficit and heat stress Contribution of ferredoxin-mediated cyclic pathway and chlororespiration[36]
Salinity stressSimultaneous expression of variable expressed genes [37]
Cold stressEnhances epidermal cell density, stomatal density and index, width of xylem vessel and phloem tissue and sclerenchymaCandyleaf[38]
Salt stressAccumulation of biomass, ions concentration in tissue and steviolglycosides[39]
Drought stressThe use of steviol glycosides enhances the harvest index [40]
Salinity and drought stressSodium chloride serves as an activator, and mannitol works for the downregulation of genes involved in the steviol glycosides synthesis pathways that alter the steviol glycosides production[41]
Cold stressPhotosynthetic electron transport chain protection by the sub-cellular antioxidant systemWheat[42]
Heat and drought stressSignaling of phytohormone and epigenetic control[43,44]
Salinity stressMaintenance of osmoprotectants, photosynthetic activity and sodium/potassium ions ratio[45,46]
Cold stressWRKY gene expressionGrapevine[47]
Heat stressHSPs genes, along with antioxidant enzyme expressionTomato[48]
Salinity stressReduced the accumulation of different ions such as sodium, magnesium, and zinc in leaves and rootsCommonbean[49]
Drought stressEnhances plant metabolism with water relation parameters, antioxidant enzyme water relation parameters, activities of antioxidant enzymes and yield per plant increasesLemon grass[50,51]
Salinity stressAntioxidants in leaves and lipid peroxidationTomato[52]
Salinity stressBiomass production as well as stomatal conductance[53]
Drought stressSugar and amino acid content accumulationAlfalfa [54]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Saharan, B.S.; Brar, B.; Duhan, J.S.; Kumar, R.; Marwaha, S.; Rajput, V.D.; Minkina, T. Molecular and Physiological Mechanisms to Mitigate Abiotic Stress Conditions in Plants. Life 2022, 12, 1634. https://doi.org/10.3390/life12101634

AMA Style

Saharan BS, Brar B, Duhan JS, Kumar R, Marwaha S, Rajput VD, Minkina T. Molecular and Physiological Mechanisms to Mitigate Abiotic Stress Conditions in Plants. Life. 2022; 12(10):1634. https://doi.org/10.3390/life12101634

Chicago/Turabian Style

Saharan, Baljeet Singh, Basanti Brar, Joginder Singh Duhan, Ravinder Kumar, Sumnil Marwaha, Vishnu D. Rajput, and Tatiana Minkina. 2022. "Molecular and Physiological Mechanisms to Mitigate Abiotic Stress Conditions in Plants" Life 12, no. 10: 1634. https://doi.org/10.3390/life12101634

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