Open access peer-reviewed chapter

Salinity Stress in Plants: Challenges in View of Physiological Aspects

Written By

Parastoo Majidian and Hamidreza Ghorbani

Submitted: 29 December 2023 Reviewed: 28 February 2024 Published: 17 July 2024

DOI: 10.5772/intechopen.114385

From the Edited Volume

Abiotic Stress in Crop Plants - Ecophysiological Responses and Molecular Approaches

Edited by Mirza Hasanuzzaman and Kamrun Nahar

Chapter metrics overview

40 Chapter Downloads

View Full Metrics

Abstract

Increasing the worldwide population, the food supply has become a global crisis due to the existence of various environmental stresses. Salinity after drought is one of the devastating environmental stresses that affects about 50% of the world’s agricultural lands. It is considered as one of the important abiotic stresses that cause plant growth restriction in different stages such as seed germination, photosynthesis, hormonal regulation, nutrient uptake, and seed quality and quantity. Under salinity conditions, plants undergo numerous changes as morphological (early flowering, prevention of lateral shoot development, and root adaptations), physiological (Na+/K+ discrimination, osmotic adjustment, ion homeostasis, and stomatal responses), and biochemical (accumulation of polyamines, antioxidant activity, proline, and change the hormone level). With the ever-increasing expansion of saline lands and highly costs spending for their rehabilitation, the preparation of high-yielding lines/genotypes tolerant to salinity will be of particular importance. Being aware of various pathways involved in plant resistance to salinity stress can be an effective tool to increase crop production and cultivated area in different parts of the world.

Keywords

  • antioxidant
  • climate variation
  • hormone
  • limiting factors
  • plant adaptability
  • saline lands

1. Introduction

Abiotic stress is one of the dreadful environmental stresses that affect plant growth and crop yield, even in irrigated areas throughout the world. Salinity stress is one of the major abiotic stresses that limits crop production and agricultural sustainability in many areas of the world because of the increased use of low-quality water for irrigation. An important issue for agricultural science and food production in the world is the study of salt stress mechanisms in crops. The plant response to high NaCl concentrations is complex and comprehensive; it contains numerous processes that should be correctly coordinated. The extra salt concentrations affect plants, leading to ionic imbalance and osmotic stress because of the accumulation of ions like Na+ and Cl. Moreover, salt stress leads to mineral homeostasis of K+ and Ca2+. Recently, a new gene family involved in salt stress response in crops has been reported based on achievements in genomics, molecular biology, and transcriptomic procedures [1].

Soil salinity is considered as a destructive environmental factor that has severe negative effects on seed germination, crop growth, and productivity [2]. Of the 230 million hectares of farmland in use around the world, 20% is affected by salt followed by their increase annually as a result of various factors including (1) inappropriate crop irrigation practices, (2) excessive fertilization and plowing, and (3) natural phenomena such as salt intrusion into coastal zones resulting from rising sea levels [3, 4]. For instance, around 100 million hectares of farmland in China suffer salinity stress [5]. In arid and semi-arid regions, salinity is becoming a more and more dreadful threat due to limited rainfall, high evapotranspiration, and extreme temperatures coupled with poor water and soil management [6, 7]. Globally, salinity is an important stress that affects a large proportion of crop productivity and agricultural soil [8]. Since, demand for food production is growing by enhancing world population, making use of salty lands to compensate for food shortages is of great importance [9]. Thus, scientists are attempting to increase food production by 70% to avoid the risk of food shortages for 9.3 billion people by 2050 [10].

The fact that the continuously growing population and meeting its demand has attracted agricultural politicians to focus on crop productivity enhancement. In this route, environmental stresses could be recognized as major obstacles to crop production [11]. Among all abiotic constraints, salinity stress is one of the major limitations for crop production, which leads to ionic imbalances, osmotic stress, and secondary stresses for glycophytes [12]. Currently, climate change and environmental stresses negatively influence arable lands due to irrigation malpractices, which results in the accumulation of sodium chloride (NaCl) in soil [13]. Besides inaccurate cultivation practices, arable lands may confront some natural phenomena including salty water on the coasts and contamination resulting from parent rock and oceanic salts, which threaten the productivity of arable land under cultivation. Although Na+ causes toxicity in soil under salinity conditions, Cl anion can play an important role in causing some plant species to be sensitive [14].

Salinity has a negative effect on growth and development processes in plants, followed by their seed/biomass yield [15]. Salt stress negatively influences plant growth by germination reduction, decrease of aerial organ growth and dry matter production, as well as disturbance of nutritional imbalances, the complex interaction of hormones, osmotic adjustment, and specific ion toxicity [16, 17]. Increasing salinity decreases the leaf area, followed by hindering leaf development and a decrease in light absorption and photosynthesis rate [18]. If the new leaves appear, they may be weak and show burning symptoms when they turn to mature. In total, there is a positive correlation between leaf number/leaf area and photosynthetic rate, which means the less leaf number/area, the less photosynthetic rate following the decline of dry weight matter [19, 20]. Salinity disturbs ions transport in different plant’s organs, which causes a decline in vegetative and reproductive quality.

Potassium is an essential nutrient for plant growth, which plays the main role in creation of osmotic pressure and protein production of cells [21, 22]. Salinity (sodium chloride) interferes with the active absorption of this element, since it prevents the selective absorption of cells. Besides, the competition of sodium with potassium in cells brings about a decrease in inactive absorption of potassium due to more abundance of sodium than potassium (Figure 1) [23, 24].

Figure 1.

A general sight of physiological, biochemical, and morphological effects of salt stress in plants.

Since crops play an important role in the economy of agricultural societies and salt stress has strong and diverse effects on overall crop growth, the objective of this chapter is to accurately detect and awareness of the biological processes involved in salinity stress to prevent severe damage in consideration of various aspects of morphological, physiological, and biochemical mechanisms. In addition, the introduction of salt-tolerant cultivars could be the other management strategy that leads to decreased damages resulting from salt stress.

Advertisement

2. Influence of salt stress

About 50% of agronomical lands are affected by different ranges of salt stress worldwide [25]. Among different climatic areas, it should be noted that arid and semi-arid zones are extremely sensitive to salt stress, which means salinity is considered the most significant obstacle to agronomical and horticultural production. Salt leads to decreased plant growth in two ways in both soil and water. In one way, the existence of salt in soil solution declines plants’ capability in relation to water absorption, which is known as an osmotic effect or water shortage. In another way, extra salt entrance by transpiration results in damage to leaf cells [26].

Most morphological traits in plants such as leaf area, biomass, seed weight, root system, plant height, spikelet per panicle, and tillers per plant reduce under high salt concentration, leading to a reduction in grain yield and harvest index [27, 28]. Salt stress affects physiological and biochemical processes in plants at all developmental stages, from germination to senescence [29]. Salt stress has both osmotic and ionic or ion-toxicity effects on plants, ultimately causing oxidative stress and nutrient depletion in plant cells [30]. Continuous salt stress reduces plant cell turgor pressure, which in turn reduces cell growth; therefore, plants must osmotically adjust to maintain cell expansion and growth [30, 31, 32]. Osmotic stress (caused by the lower water potential of the external solution) is rapidly sensed by plants soon after exposure to saline conditions and leads to plant water and solute deficits [9, 33]. Osmotic stress also results in rapid stomatal closure, which reduces the plant’s ability to assimilate CO2 and inhibits photosynthesis [34]. Aging and death of plants are caused by the accumulation of sodium (Na+) and chlorine (Cl) in plant cells due to ionic stress [35]. The most common explanation for Na+ toxicity is that it has an inhibitory effect on enzyme activities, affecting metabolism negatively, including the Calvin cycle and other pathways [36, 37]. Excess sodium in the cytoplasm also interferes with the uptake and transport of potassium and other macro and micronutrients, including nitrogen, phosphorus, potassium, calcium, and zinc [38]. Various stresses, such as osmotic stress and salinity, damage the cells and vital molecules in plant cells due to the accumulation of reactive oxygen species [23]. Under salinity tension, plants should regulate biochemical and physiological pathways related to the regulation of osmotic and ionic homeostasis, nutrient balance, and oxidative stress [39].

The continuous transfer of salt into leaves during a long period will eventually lead to high sodium and chlorine accumulation, which causes devastating effects on old leaves, since the death rate of leaves is necessary for plant survival [40]. If the newly created leaves are more than dead leaves, the plant has enough photosynthesizing leaves to produce flowers and seeds, no matter how many of them have been decreased. However, if there are more dead leaves than produced leaves, then plants cannot survive to produce seeds [41].

The plant growth response to salinity stress includes two-phase reactions over time. The first stage of growth reduction, known as “initial growth decrease” or “early phase,” is due to the osmotic effect of salt outside the root zone appearing quickly. Then, the second reduction stage, “late phase or ionic phase,” occurs, which comes from internal damage and continues for a longer period of time. What creates a differentiation between salt-sensitive plants and salt-tolerant plants is the inability to prevent salt from reaching toxic levels in transpiration leaves over time [21].

In most crops, the critical growth stage is seed germination, which exhibits toxic effects on seed embryo water uptake inhibition under salinity. However, legumes, seedlings, and developmental stages are more sensitive than the germination stage. It is observed that salinity results in limitations such as a decrease in the water potential of tissue, photosynthesis, and different stages of growth. For example, the stomatal closure caused by less availability of water to cells brings about less photosynthesis and growth in soybeans under salinity stress [42]. In another study, it was found that the early stages of growth and development in lentils are sensitive to salt stress [43].

Ionic imbalance is the other indicator of salinity conditions that cause disturbance in competitive nutrient uptake, accumulation, and transport in plants. Indeed, the concentrations of cations including Ca2+, K+, and Mg2+ which are vital for photosynthetic activity, decrease by different levels of salt [44]. For instance, the survey of the Na+/K+ ratio in Lotus creticus showed Na+ contents increase in aerial and root organs under salinity stress [45]. Moreover, the competitive uptake of Na+ and K+ ion flux caused an astonishing reduction of the Na+/K+ ratio in mungbean and chickpea, indicating a significant decrease in K+ and yield [46]. In legumes, it was demonstrated that different species had various responses to saline conditions, since it was determined that soybean, mungbean, cowpea, and common bean indicated salt sensitivity as Na+ toxicity, Cl toxicity, Na+, and Cl toxicity, respectively [47]. Lopez-Aguilar et al. reported that the uptake and transport of Ca2+ and Mg2+ were significantly lower in cowpea and frijollilo plants under salt stress; however, these rates in Tepary bean plants showed no significant response to salinity treatments [48]. It was reported that the Na+/K+ ratio increased in the roots and shoots of Cheongcheong and IR28 cultivars of rice under salinity stress [49]. In addition, it was found that there was a significant correlation between grain dry matter and the K+/Na+ ratio under salt stress in wheat, which had an effect on the grain filling rate and its duration [21].

Advertisement

3. Different mechanisms of tolerance to salt stress

Plants show different salinity tolerance reactions, which are reflected in their different growth responses. To increment crop yield under salt stress, it is of great importance to recognize tolerance mechanisms consisting of ion homeostasis, compatible solute accumulation and osmotic protection, antioxidant regulation, and hormonal regulation.

The salt tolerance mechanisms in plants are divided into three groups: osmotic stress tolerance, Na+ expulsion from leaf, and tissue tolerance [50]. With respect to osmotic stress tolerance, this type of stress decreases cell development on the tips of roots and young leaves. A decrease in response to osmotic stress leads to an increase in leaf growth and stomata conduction, while the increment of leaf area would be profitable for plants whose soil has enough water availability. The second group is related to repulsing Na+ from leaves, which is assured that Na+ ions with toxic concentration do not accumulate within leaves, since the deficiency of excretion of Na+ toxic effects causes premature death of older leaves. The third mechanism is tissue tolerance, which appears to be tolerant of some tissues to Na+ and Cl accumulation [51].

Advertisement

4. Salt tolerance strategies

Salt tolerance mechanisms include processes that allow plants to resist stress, take up nutrients from the soil, and complete their life cycle under high salt concentrations in the soil. Plants that are able to grow on higher levels of salt in the rhizosphere are termed as halophytes. To enhance salt tolerance, the following strategies will be briefly reviewed, aiming at crop salt tolerance under saline conditions.

In general, the physiological responses of plants to salinity stress are divided into two main phases. The first phase is ion-independent growth reduction. It occurs within a few minutes to a few days, resulting in closing stomata and cell expansion inhibition [52]. A second phase, which pertains to the build-up of cytotoxic ion levels, happens for a relatively long period, resulting in slowing down metabolic processes, reducing leaf water potential, premature senescence, and ultimately cell death [53, 54].

4.1 Ion homeostasis and its detention in tissue

Under salt stress, plants apply an ion homeostasis mechanism to maintain a high concentration of K+ and a low concentration of Na+ in the cytosol [55]. Ion homeostasis is enough for normal growth and a necessary process for growth during salt stress [56]. Regardless of the nature of halophytes and glycophytes, neither can tolerate high salt concentrations in their cytoplasm; thus, the focus of most projects is on the Na+ ion transfer mechanism and its distribution in various parts of plants. After entering Na+ into cytoplasm, it is transferred to vacuole through antiporter [57]. There are two types of H+ pumps in vacuole membrane vacuolar H+-ATPase (V-ATPase) and Vacuolar H+-pyro-phosphatase [58]. The electrochemical proton gradient created by these two transferring enzymes in the vacuole provides the necessary power for ion transport. Under stress-free conditions, H+ pump plays a significant role in preserving mineral homeostasis, secondary transferring energy, and facilitation of vesicle fusion. Under salt stress, the plant survival depends on the V-ATPase pump, which is dominant in the H+ pump.

In another view, it is considered the vital role of salt overly sensitive gene network (SOS) to salinity for Na+ ion regulation and salt tolerance [59]. The SOS signaling network consisted of three main proteins SOS1, SOS2, and SOS3. SOS1 gene codes anti-porter H+/Na+ of plasma membrane, which facilitates the transport of sodium from root tissue to stem. The SOS2 gene codes serine/threonine kinase which is activated by Ca2+ signals. The third protein involved in SOS signaling is SOS3, which is Ca2+ domain binding to N terminal of SOS2. By increasing Na+ concentration, a severe increment of Ca2+ level in the intracellular area occurs, which in turn facilitates its binding to SOS3 protein. SOS3 activates SOS2 by releasing internal inhibitors. The combination of SOS2-SOS3 on plasma membrane is loaded where SOS1 phosphorylates, which causes an increase in sodium flow and a reduction in its toxicity. In legumes, the ion distribution, specifically the ratio of K+/Na+ in the cytosol, has been reported in various ways among their different species and cultivars.

For example, the response of various legumes to salt stress is determined by Cl toxicity (mungbean), Na+ toxicity (soybean), and Cl and Na+ toxicity (common bean and cowpea) [47]. In one study, it was reported that there were exceptional K+/Na+ transporters in the mash bean cultivar, which conserve lower levels of intracellular Na+ [60]. However, it was observed that the uptake of K+ resulted in the developing Na+/K+ ratio, which was regulated by a decrease of Na+/Ca2+ ratio in pigeon peas under salinity [61]. According to Zhang et al., soybean salt tolerance was obtained by plasma membrane Na+/H+ exchanger GmSOS1, which maintains Na+ homeostasis in soybean [62].

The mechanisms in relation to cytoplasmic Na+ reduction include restriction of Na+ uptake, increase of Na+ efflux, and compartmentalization of Na+ in the vacuole [63]. The rice plasma membrane Na+/H+ anti-porter (OsSOS1) excludes Na+ from the shoot, promoting a lower cellular Na+/K+ ratio and increasing salt tolerance [64, 65]. The vacuolar Na+/H+ anti-porters OsNHX1, OsNHX2, OsNHX3, OsNHX4, OsNHX5, and OsARP/OsCTP play important roles in the vacuolar compartmentalization of Na+ and K+ that accumulate in the cytoplasm and thereby determine rice salt tolerance [66]. OsHKT1;1, OsHKT2;1, OsHKT2;3/OsHKT3, OsKAT1, OsKAT2, OsHAK5, and OsHAK21/qSE3 transport Na+ or both Na+ and K+, helping to maintain Na+/K+ homeostasis in the cytoplasm and regulate rice response to salt stress [67].

Some genes that include transcription factors affect the salt tolerance of rice by regulating the expression of CLC, NHX, and HKT genes, activating the expression of the K+/Na+ transporter, and reducing sodium absorption into the cells [68].

4.2 Compatible solute accumulation and osmotic protection

The compatible solute (compatible osmolytes or osmoprotectants) are not toxic even at higher concentrations. Their effective role is to protect osmotic structure and balance within cells through continuous infiltration flow of water and water flow reduction. Related to their prominent feature, it can be referred to as being hydrophilic compounds with low molecular weight and no net charge at physiological pH. Under salt stress, there are various types of solute accumulation varying among different plant species including amino acids (proline), carbohydrates (sucrose and fructose), quaternary amines (glycine and betaine), ions (potassium), and organic acids (malate and nitrate) [69]. El-Sabagh et al. reported on the effect of exogenous osmoprotectants such as proline on soybeans under salinity stress in order to increase salinity resistance [70]. The accumulation of proline was observed in common bean (Phaseolus vulgaris L.) subjected to salt stress, followed by its increased accumulation in salt-stressed plants treated with effective microorganisms [71].

The other non-toxic osmolyte is beta-glycine, which raises osmorylation during stress conditions and plays an important role in osmotic adjustment [72], protein fixation [73], photosynthetic system protection [74], and reactive oxygen species (ROS) reduction [75]. Another research has demonstrated the existence of less protein and high protein and amino acids content in salt-tolerant than salt-sensitive beans [76]. In addition, it was shown that osmotic regulation in soybeans occurs by regulating the concentrations of trigonelline, proline, and potassium [70]. Also, the osmotic adjustment caused by accumulation of amino acids, reducing sugars, and ascorbic acid has been reported in pea plants subjected to salt stress [77]. Finally, osmoregulation can be controlled by various osmolytes, which are key factors to decrease osmotic stress in grain legumes.

Salt stress also causes osmotic stress and promotes the biosynthesis and accumulation of compatible osmolytes, which decrease cellular osmotic potential and stabilize cellular proteins and structures [78]. OsP5CS1 and OsP5CR proline synthesis genes increase the accumulation of proline and improve the tolerance of rice to salt stress [79]. OsGMST1, which encodes a monosaccharide transporter, increases monosaccharide accumulation and improves salt tolerance in plants [80]. It is reported that sugars will eventually be exported to transporters (SWEETs) that can regulate sucrose transport/distribution and maintain sugar homeostasis in rice under drought and salinity stresses [81]. In addition, rice glycine betaine is the other osmolyte synthesized by choline monooxygenase OsCMO, and beating aldehyde dehydrogenase OsBADH1 increases the tolerance of rice to salt stress by incrementing the accumulation of glycine betaine [82].

4.3 Regulation of antioxidants in order to tolerance to salt stress

Antioxidant metabolisms like enzymatic and non-enzymatic compounds play a critical role in preventing reactive oxygen species (ROS) resulting from salinity stress [83]. These lethal compounds contain hydrogen peroxide (H2O2), hydroxylradical (OH), superoxide-radical (O−2), and singlet-oxygen (O2), which damage cells, along with different proteins, nucleic acids, and membrane lipids, which lead to oxidative stress [84]. Salt tolerance has a positive correlation with antioxidant enzymes such as glutathione peroxidase (GPX), glutathione reductase (GR), superoxide dismutase (SOD), ascorbate peroxidase (APX), and catalase (CAT) as well as the accumulation of non-enzymatic antioxidant compounds. Under salt stress conditions, the activity of these enzymes increased in the tolerant variety [85]. In the previous research, APX and CAT activities were enhanced in bean salt-tolerant cultivars subjected to salt stress [86]. In conclusion, the enhanced activities of antioxidants in legumes perform as a protective process over oxidative stress.

Superoxide dismutase (SOD) can evacuate the surplus superoxide anions in the cells and hence act as the first line of defense. It is classified into three groups in higher plants. Based on the metal ions present on auxiliary sites, they are iron SOD (Fe-SOD), copper/zinc SOD (Cu/Zn-SOD), and manganese SOD (Mn-SOD). The SOD can dismutase the superoxide anion into the O2 and H2O2 and abolish the superoxide toxicity [87].

Ascorbate peroxidase (APX) is one of the important enzymes that eradicate H2O2. The adverse effects of salt stress in different plant species are mitigated by exogenous application of ascorbate, and it endorses plant recovery from salt stress [88, 89].

Catalase (CAT) is another important enzyme that reduces H2O2 and is present in peroxisomes. It differs from APX in that it has a high affinity for hydrogen peroxide and requires a reducing substrate. Hydrogen peroxide from the light reaction is mainly removed by CAT. The effects of NaCl on H2O2 content and CAT activity were studied in various groups of plants, all of which show high tolerance to NaCl. In all cases, treatment of plants with high salinity levels increased H2O2 concentration, and as a result, CAT enzyme activity also increased [90].

A lower accumulation of O2 and H2O2 was observed in overexpressing OsMn-SOD1 and OsCu/Zn-SOD lines under salt stress [91]. APX gene family in Oryza sativa increase APX activity, decrease H2O2 and malondialdehyde levels, reduce oxidative stress damage, and increase rice tolerance to salinity stress [92]. Glutathione responsive rice glyoxalase II in salt stress conditions with increased photosynthetic efficiency [93], the GR gene family in rice [94] and rice GDP-mannose pyrophosphorylase and dehydroascorbate reductase play an important role in increasing plant tolerance [95].

4.4 Polyamine role in salt tolerance

Polyamines are relatively polycationic low molecular weight compounds that are extremely widespread in all organisms from unicellular to multicellular including putrescine (Put), spermidine (Spd), and spermine (Spm). They are applicable at different development stages including cell division, nucleic acids stability, embryogenesis, dormancy termination, aging regulation, plant growth, and stress resistance [96]. In case of salinity, it was reported that the oxidative damage was decreased by the accumulation of putrescine in soybean [97], chickpea [98], and faba bean [99]. According to spermidine, it was exhibited that spermidine was applicable to alleviate the destructive effects of salinity condition and promote salinity tolerance in various legume species such as faba bean [100], soybean [101], and mungbean [102]. Thus, legumes make use of polyamines for their adaptation and defense mechanisms to cope with the lethal consequences of salinity.

4.5 Role of nitric oxide in salt tolerance

Nitric oxide (NO) is a volatile and small gas molecule that cause plant growth regulation, root growth development, transpiration, stomata closure, flowering, seed germination, and response to salt stress as a signaling molecule [103]. The NO effect in salt tolerance relies on the regulation of H+-ATPase in cell membrane and Na+/K+. Under salinity stress, nitric acid is capable of ROS accumulation followed by prevention of thiobarbituric acid and other aldehyde production to decline disturbing effect of salinity stress on cell membrane [104]. In soybean, it has been reported that the use of NO increases the activity of ascorbate peroxidase enzyme, and also, in Vigna angularis, it increases the activity of oxidant enzymes and non-enzymatic compounds [105, 106].

4.6 Hormones role in salt tolerance

It is proved that crops utilize adaptive responses against salt stress (downstream of the initial salt signaling phase), which alters plant hormone levels and the salt-induced signaling cascade. Phytohormones, endogenous chemical signals, are one of the adaptive responses that coordinate plant growth and development under optimal conditions and environmental challenges [28]. Plant hormones are chemical compounds that exist in very low concentrations and play important key roles in controlling and regulating all aspects of growth and development. Hormones are classified into two main groups containing (1) growth promoters (gibberellins, cytokinins, and auxins) and (2) growth inhibitors (ABA and ethylene) that individually affect salt stress. Among all phytohormones, ABA has a significant role in saline conditions. It is considered as the central regulator of abiotic stress tolerance in plants by up-regulation of PYR, PYL, RCAR proteins complex, protein phosphatases 2C (PP2C), ABF transcription factors, and SnRK2 protein kinases, which can decrease the preventive saline effect on growth and photosynthesis. In fact, ABA is responsible for metabolic variations, stomatal closure, and stress--responsive gene regulation under salinity condition in crops, even at high concentrations of ABA, which are more helpful in osmotic adjustment using more stress protein production and salinity adaptation. For example, Shu et al. suggested that ABA could be a potential plant growth regulator in soybeans, which could improve soybean germination under salt stress [107]. In addition, it was reported that the stomata closure occurred in response to ABA in Lupinus albus L. [108]. Thus, ABA also plays a major role in mediating physiological responses to salt stress through alteration in osmotic adjustment and photosynthetic and plant developmental growth. The ABA biosynthesis genes OsNCED3 and OsNCED5, encoding 9-cis-epoxycarotenoid dioxygenases, increase ABA levels and enhance salt stress tolerance [109]. The rice ABA receptor gene OsPYL/RCAR5 is a positive regulator of the ABA signal transduction pathway in abiotic stress tolerance [110]. OsbZIP23 gene directly affects the gene responsible for ABA synthesis and the component gene of ABA signaling and can regulate ABA signaling and biosynthesis by regulating stress response pathways [111].

In one study, it was noticed that salinity tolerance incremented by promotion of ethylene biosynthesis and signal transduction in rice, while, their inhibition increased salinity sensitivity in Arabidopsis and soybean [112]. However, in tomato, the direct ethylene precursor 1-aminocyclopropane-1 carboxylic acid (ACC) causes Na+ accumulation and oxidative stress in leaves and promotes leaf senescence under salinity stress [113, 114]. Meanwhile, protein salt intolerance 1 (OsSIT1) in rice increases sensitivity to salinity stress by increasing ethylene production in the plant [115]. DOF transcription factors and ethylene-responsive element binding protein (EREBP) transcription factor OsBIERF3 cause positive changes in plant root length by limiting ethylene synthesis [116].

Gibberellic acid (GA) is another important hormone that regulates plant growth and is associated with the regulation of growth under abiotic stress. Gibberellin 2-oxidase 5 (OsGA2ox5) and GA catabolic pathway genes reduce GA accumulation and increase plant salt tolerance with delayed growth [117].

In case of cytokinins (CKs), its function is to control plant adaptation to environmental stresses. In Arabidopsis, by the reduction of various levels of CKs, the salt tolerance increased [118], while it was exhibited that the exogenous CK application caused salt tolerance enhancement in Solanum melongena. Indeed, the enzyme results in CK degradation pathway called as cytokinin oxidase OsCKX2, which negatively regulates salt stress tolerance by controlling CK levels [119].

Jasmonic acid (JA) is the other hormone that reduces root length and activates antioxidant enzymes under salt stress conditions [120]. It was reported that JA accumulates endogenously in rice roots and exogenously increases salt stress tolerance in rice and wheat [121, 122].

Salicylic acid (SA) is also known as adaptive response in plants under salinity stress, which plays an important role in plant salt tolerance. Exogenous SA application in plants can control salinity conditions through (1) enhancement of antioxidant system, (2) increases in the synthesis of osmolytes, and (3) promotes photosynthesis and nitrogen fixation [123].

Advertisement

5. Genetic engineering for salinity stress tolerance

Salt tolerance is generally controlled by large numbers of genes and different physiological mechanisms. Recently, researchers have focused on analyzing genes controlling salt tolerance by introduction of hypothetical premise for augmentation of the stress signal system and upgrading plant tolerance to stress [124]. Genetic transformation technology allows genetic scientists to transfer genes in a precise and predictable way. Based on this, manipulation of the biosynthetic pathways involved in osmotic protection and physiological traits of plants could be efficient tools to (1) raise the level of tolerance to salinity, (2) upgrade the accumulation of molecules that remove ROS, (3) reduce the concentration of lipid peroxidation, (4) maintain the three-dimensional and functional structure of proteins, etc. [125]. To produce salt-resistant plants, plant engineering methods can be used to overexpress resistance genes. Gene expression studies using constitutive promoters provide limited biological information compared to using inducible promoters or cell type-specific promoters. In this direction, the selection of strong promoters inducing abiotic stress is very important in plant breeding programs. In addition, a wide range of stress-inducible promoters are available that vary in their ability for regulation of gene expression patterns to enhance salinity tolerance in crop plants. The selection of these promoters can significantly affect the results of a transgenic manipulation. For instance, salt tolerance is fundamentally enhanced by vacuolar Na+/H+ anti-porter gene AtNHX1 and plasma membrane Na+/H+ anti-porter gene SOS1 in transgenic A. thaliana [126]. A substantial number of genes, HKT and NHX, encoding K+ transporters and channels have been distinguished and cloned in different plant species. The class 1 HKT transporters, distinguished in Arabidopsis, shield the plant from the antagonistic impacts of salinity by anticipating overabundance aggregation of Na+ in leaves [127]. Similar results were observed in an experiment with rice, where the class 1 HKT transporter removes excess sodium from the xylem, thus protecting photosynthetic leaf tissues against the toxic agent Na+ [128]. Transgenic genes such as Ta-UnP, TaZNF, TaSST, TaDUF1, and TaSP can significantly increase salt tolerance in wheat [129]. Studies conducted by Sun et al. revealed that an overexpressing salt-responsive gene, ZFP179, encoding a Cys2/His2 zinc finger protein, enhanced salt tolerance in rice [130]. It has also been reported that overexpression of genes for signaling proteins such as kinases and phosphatases are involved in salt tolerance in various plants. One of these proteins is mitogen-activated protein kinases (MAPKs), which are conserved signaling proteins and play an essential role in improving plant tolerance to various stresses of crops. It was reported that different transgenic plants engineered with MAPK cascade are tolerant to salt stress [131].

Advertisement

6. Conclusions

Totally, various studies have been demonstrated that salt tolerance consists of different cellular, metabolical, molecular, and physiological responses in the whole plant. During evolution, plants have developed the ability to detoxify excessive levels of Na+ and use multiple strategies to mitigate damages resulting from salinity stress. There are various ways to decline the effects of salt stress on plant growth, which require a thorough conception of which defensive responses to environmental stresses could be more appropriate. Thus, it is indispensable to apply integrated approaches such as molecular, biochemical, and physiological techniques in order to develop salt-tolerant cultivars. A comprehensive understanding of molecular mechanisms in crop salt tolerance has been gained through the development of molecular salt tolerance markers, complete sequencing of plant genomes, and the widespread use of microarray analysis and genome editing technology. In order to create salt-tolerant crops using genetic engineering methods, various tools and techniques could be distinguished, including an extensive number of salt-responsive genes, methods of gene transfer, and their promoter sequences. These powerful techniques can offer advantages and solutions to the complex and fascinating traits involved in salt tolerance plants.

References

  1. 1. Isayenkov SV. Physiological and molecular aspects of salt stress in plants. Cytology and Genetic. 2012;46:302-318. DOI: 10.3103/S0095452712050040
  2. 2. Yang Y, Guo Y. Elucidating the molecular mechanisms mediating plant salt-stress responses. New Phytologist. 2018;217(2):523-539
  3. 3. Shabala S. Learning from halophytes: Physiological basis and strategies to improve abiotic stress tolerance in crops. Annals of Botany. 2013;112(7):1209-1221
  4. 4. Kamran M, Parveen A, Ahmar S, Malik Z, Hussain S, Chattha MS, et al. An overview of hazardous impacts of soil salinity in crops, tolerance mechanisms, and amelioration through selenium supplementation. International Journal of Molecular Sciences. 2019;21(1):148
  5. 5. Naveed SA, Zhang F, Zhang J, Zheng TQ , Meng LJ, Pang YL, et al. Identification of QTN and candidate genes for salinity tolerance at the germination and seedling stages in rice by genome-wide association analyses. Scientific Reports. 2018;8(1):6505
  6. 6. de Azevedo Neto AD, Prisco JT, Enéas-Filho J, de Abreu CEB, 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. Environmental and Experimental Botany. 2006;56(1):87-94
  7. 7. Lodeyro AF, Carrillo N. Salt stress in higher plants: Mechanisms of toxicity and defensive responses. Stress Responses in Plants: Mechanisms of Toxicity and Tolerance. 2015:1-33
  8. 8. Munns R. Comparative physiology of salt and water stress. Plant, Cell and Environment. 2002;25(2):239-250
  9. 9. Qin H, Huang R. The phytohormonal regulation of Na+/K+ and reactive oxygen species homeostasis in rice salt response. Molecular Breeding. 2020;40(5):47
  10. 10. Shabala S, Cuin TA. Potassium transport and plant salt tolerance. Physiologia Plantarum. 2008;133(4):651-669
  11. 11. Vinocur B, Altman A. Recent advances in engineering plant tolerance to abiotic stress: Achievements and limitations. Current Opinion in Biotechnology. 2005;16(2):123-132
  12. 12. Ahmad R, Hussain S, Anjum MA, Khalid MF, Saqib M, Zakir I, et al. Oxidative stress and antioxidant defense mechanisms in plants under salt stress. In: Plant Abiotic Stress Tolerance. USA: Springer International Publishing; 2019. pp. 191-205
  13. 13. Carillo P, Annunziata MG, Pontecorvo G, Fuggi A, Woodrow P. Salinity stress and salt tolerance. Abiotic Stress in Plants-mechanisms and Adaptations. 2011;1:21-38
  14. 14. Islam F, Yasmeen T, Arif MS, Ali S, Ali B, Hameed S, et al. Plant growth promoting bacteria confer salt tolerance in Vigna radiata by up-regulating antioxidant defense and biological soil fertility. Plant Growth Regulation. 2016;80(1):23-36
  15. 15. Ghassemi-Golezani K, Taifeh-Noori M, Oustan S, Moghaddam M. Response of soybean cultivars to salinity stress. Journal of Food Agricultural and Environment. 2009;7(2):401-404
  16. 16. Nadeem M, Li J, Yahya M, Wang M, Ali A, Cheng A, et al. Grain legumes and fear of salt stress: Focus on mechanisms and management strategies. International Journal of Molecular Science. 2019;20(4):799. DOI: 10.3390/ijms20040799
  17. 17. Tlahig S, Bellani L, Karmous I, Barbieri F, Loumerem M, Muccifora S. Response to salinity in legume species: An insight on the effects of salt stress during seed germination and seedling growth. Chemistry and Biodiversity. 2021;18(4):e2000917
  18. 18. Manchanda G, Garg N. Salinity and its effects on the functional biology of legumes. Acta Physiologiae Plantarum. 2008;30:595-618
  19. 19. Bybordi A. The influence of salt stress on seed germination, growth and yield of canola cultivars. Notulae Botanicae Horti Agrobotanici Cluj-Napoca. 2010;38(1):128-133
  20. 20. Dadkhah A. Effect of salinity on growth and leaf photosynthesis of two sugar beet (Beta vulgaris L.) cultivars. Journal of Agricultural Science and Technology. 2011;13(7):1001-1012
  21. 21. Poustini K, Siosemardeh A. Ion distribution in wheat cultivars in response to salinity stress. Field Crops Research. 2004;85(2-3):125-133
  22. 22. Fan M, Bie Z, Krumbein A, Schwarz D. Salinity stress in tomatoes can be alleviated by grafting and potassium depending on the rootstock and K-concentration employed. Scientia Horticulturae. 2011;130(3):615-623
  23. 23. Ahanger MA, Tomar NS, Tittal M, Argal S, Agarwal R. Plant growth under water/salt stress: ROS production; antioxidants and significance of added potassium under such conditions. Physiology and Molecular Biology of Plants. 2017;23(4):731-744
  24. 24. Wu H, Li Z. The importance of Cl exclusion and vacuolar Cl− sequestration: Revisiting the role of Cl transport in plant salt tolerance. Frontiers in Plant Science. 2019;10:1418
  25. 25. Jouyban Z. The effects of salt stress on plant growth. Technical Journal of Engineering and Applied Sciences. 2012;2(1):7-10
  26. 26. Ondrasek G, Rengel Z, Veres S. Soil salinization and salt stress in crop production. Abiotic Stress in Plants-Mechanisms and Adaptations. 2011;1:171-190
  27. 27. Razzaq A, Ali A, Safdar LB, Zafar MM, Rui Y, Shakeel A, et al. Salt stress induces physiochemical alterations in rice grain composition and quality. Journal of Food Science. 2020;85(1):14-20
  28. 28. Van Zelm E, Zhang Y, Testerink C. Salt tolerance mechanisms of plants. Annual Review of Plant Biology. 2020;71:403-433
  29. 29. Muranaka S, Shimizu K, Kato M. Ionic and osmotic effects of salinity on single-leaf photosynthesis in two wheat cultivars with different drought tolerance. Photosynthetica. 2002;40:201-207
  30. 30. Zhu JK. Abiotic stress signaling and responses in plants. Cell. 2016;167(2):313-324
  31. 31. Hakim MA, Juraimi AS, Hanafi MM, Ismail MR, Selamat A, Rafii MY, et al. Biochemical and anatomical changes and yield reduction in rice (Oryza sativa L.) under varied salinity regimes. BioMed Research International. 2014;2014:1-11
  32. 32. Munns R, Guo J, Passioura JB, Cramer GR. Leaf water status controls day-time but not daily rates of leaf expansion in salt-treated barley. Functional Plant Biology. 2000;27(10):949-957
  33. 33. Zhou J, Li Z, Xiao G, Zhai M, Pan X, Huang R, et al. CYP71D8L is a key regulator involved in growth and stress responses by mediating gibberellin homeostasis in rice. Journal of Experimental Botany. 2020;71(3):1160-1170
  34. 34. Wegner LH, Stefano G, Shabala L, Rossi M, Mancuso S, Shabala S. Sequential depolarization of root cortical and stelar cells induced by an acute salt shock–implications for Na+ and K+ transport into xylem vessels. Plant Cell Environment. 2011;34(5):859-869
  35. 35. Yang Y, Guo Y. Unraveling salt stress signaling in plants. Journal of Integrative Plant Biology. 2018;60(9):796-804
  36. 36. Wu TM, Lin WR, Kao CH, Hong CY. Gene knockout of glutathione reductase 3 results in increased sensitivity to salt stress in rice. Plant Molecular Biology. 2015;87:555-564
  37. 37. Cheeseman JM. The integration of activity in saline environments: Problems and perspectives. Functional Plant Biology. 2013;40(9):759-774
  38. 38. Iqbal MN, Rasheed R, Ashraf MY, Ashraf MA, Hussain I. Exogenously applied zinc and copper mitigate salinity effect in maize (Zea mays L.) by improving key physiological and biochemical attributes. Environmental Science and Pollution Research. 2018;25:23883-23896
  39. 39. Shahid MA, Sarkhosh A, Khan N, Balal RM, Ali S, Rossi L, et al. Insights into the physiological and biochemical impacts of salt stress on plant growth and development. Agronomy. 2020;10(7):938
  40. 40. Abdelraheem A, Esmaeili N, O’Connell M, Zhang J. Progress and perspective on drought and salt stress tolerance in cotton. Industrial Crops and Products. 2019;130:118-129
  41. 41. Ibrahim EA. Seed priming to alleviate salinity stress in germinating seeds. Journal of Plant Physiology. 2016;1(92):38-46
  42. 42. Do TD, Vuong TD, Dunn D, Smothers S, Patil G, Yungbluth DC, et al. Mapping and confirmation of loci for salt tolerance in a novel soybean germplasm, Fiskeby III. Theoretical Applied Genetics. 2018;131(3):513-524
  43. 43. Sidari M, Santonoceto C, Anastasi U, Preiti G, Muscolo A. Variations in four genotypes of lentil under NaCl-salinity stress. American Journal of Agriculture and Biological Science. 2008;3:410-416
  44. 44. Yang CW, Xu HH, Wang LL, Liu J, Shi DC, Wang DL. Comparative effects of salt-stress and alkali-stress on the growth, photosynthesis, solute accumulation, and ion balance of barley plants. Photosynthetica. 2009;47(1):79-86
  45. 45. Rejili M, Vadel AM, Guetet A, Mahdhi M, Lachiheb B, Ferchichi A, et al. Influence of temperature and salinity on the germination of Lotus creticus (L.) from the arid land of Tunisia. African Journal of Ecology. 2010;48(2):329-337
  46. 46. Nandwal AS, Godara M, Kamboj DV, Kundu BS, Mann A, Kumar B, et al. Nodule functioning in trifoliate and Pentafoliate mungbean genotypes as influenced by salinity. Biologia Plantarum. 2000;43:459-462. DOI: 10.1023/A:1026704107525
  47. 47. Le LTT, Kotula L, Siddique KH, Colmer TD. Na+ and/or Cl− toxicities determine salt sensitivity in soybean (Glycine max (L.) Merr.), mungbean (Vigna radiata (L.) R. Wilczek), cowpea (Vigna unguiculata (L.) Walp.), and common bean (Phaseolus vulgaris L.). International Journal of Molecular Sciences. 2021;22(4):1909
  48. 48. López-Aguilar R, Orduño-Cruz A, Lucero-Arce A, Murillo-Amador B, Troyo-Diéguez E. Response to salinity of three grain legumes for potential cultivation in arid areas. Soil Science and Plant Nutrition. 2003;49(3):329-336
  49. 49. Farooq M, Park JR, Jang YH, Kim EG, Kim KM. Rice cultivars under salt stress show differential expression of genes related to the regulation of Na+/K+ balance. Frontiers in Plant Science. 2021;12:680131
  50. 50. Sairam RK, Tyagi A. Physiology and molecular biology of salinity stress tolerance in plants. Current Science. 2004;86:407-421
  51. 51. Acosta-Motos JR, Ortuño MF, Bernal-Vicente A, Diaz-Vivancos P, Sanchez-Blanco MJ, Hernandez JA. Plant responses to salt stress: Adaptive mechanisms. Agronomy. 2017;7(1):18
  52. 52. Rajendran K, Tester M, Roy SJ. Quantifying the three main components of salinity tolerance in cereals. Plant Cell Environment. 2009;32:237-249
  53. 53. Khan MIR, Khan NA. Ethylene reverses photosynthetic inhibition by nickel and zinc in mustard through changes in PSII activity, photosynthetic nitrogen use efficiency, and antioxidant metabolism. Protoplasma. 2014;251:1007-1019
  54. 54. Roy SJ, Negrão S, Tester M. Salt resistant crop plants. Current Opinion in Biotechnology. 2014;26:115-124
  55. 55. Serrano R, Rodriguez-Navarro A. Ion homeostasis during salt stress in plants. Current Opinion in Cell Biology. 2001;13(4):399-404
  56. 56. Manishankar P, Wang N, Köster P, Alatar AA, Kudla J. Calcium signaling during salt stress and in the regulation of ion homeostasis. Journal of Experimental Botany. 2018;69(17):4215-4226
  57. 57. Basu S, Kumar A, Benazir I, Kumar G. Reassessing the role of ion homeostasis for improving salinity tolerance in crop plants. Physiologia Plantarum. 2021;171(4):502-519
  58. 58. Silva P, Gerós H. Regulation by salt of vacuolar H+-ATPase and H+-pyrophosphatase activities and Na+/H+ exchange. Plant Signaling Behavior. 2009;4(8):718-726
  59. 59. Liu M, Wang TZ, Zhang WH. Sodium extrusion associated with enhanced expression of SOS1 underlies different salt tolerance between Medicago falcata and Medicago truncatula seedlings. Environment Experimental Botany. 2015;110:46-55
  60. 60. Mishra S, Panda SK, Sahoo L. Transgenic asiatic grain legumes for salt tolerance and functional genomics. Reveiws in Agricultural Sciences. 2014;2:21-36
  61. 61. Subbarao GV, Johansen C, Jana MK, Rao JK. Effects of the sodium/calcium ratio in modifying salinity response of pigeonpea (Cajanus cajan). Journal of Plant Physiology. 1990;136(4):439-443
  62. 62. Zhang M, Cao J, Zhang T, Xu T, Yang L, Li X, et al. A putative plasma membrane Na+/H+ antiporter GmSOS1 is critical for salt stress tolerance in Glycine max. Frontier in Plant Sciences. 2022;16(13):870695. DOI: 10.3389/fpls.2022.870695
  63. 63. Tester M, Davenport R. Na+ tolerance and Na+ transport in higher plants. Annals of Botany. 2003;91(5):503-527
  64. 64. El Mahi H, Pérez-Hormaeche J, De Luca A, Villalta I, Espartero J, Gámez-Arjona F, et al. A critical role of sodium flux via the plasma membrane Na+/H+ exchanger SOS1 in the salt tolerance of rice. Plant Physiology. 2019;180(2):1046-1065
  65. 65. Martínez-Atienza J, Jiang X, Garciadeblas B, Mendoza I, Zhu JK, Pardo JM, et al. Conservation of the salt overly sensitive pathway in rice. Plant Physiology. 2007;143(2):1001-1012
  66. 66. Fukuda A, Nakamura A, Hara N, Toki S, Tanaka Y. Molecular and functional analyses of rice NHX-type Na+/H+ antiporter genes. Planta. 2011;233:175-188
  67. 67. Wang R, Jing W, Xiao L, Jin Y, Shen L, Zhang W. The rice high-affinity potassium transporter1; 1 is involved in salt tolerance and regulated by an MYB-type transcription factor. Plant Physiology. 2015;168(3):1076-1090
  68. 68. Lee SK, Kim BG, Kwon TR, Jeong MJ, Park SR, Lee JW, et al. Overexpression of the mitogen-activated protein kinase gene OsMAPK33 enhances sensitivity to salt stress in rice (Oryza sativa L.). Journal of Biosciences. 2011;36:139-151
  69. 69. Rontein D, Basset G, Hanson AD. Metabolic engineering of osmoprotectant accumulation in plants. Metabolic Engineering. 2002;4(1):49-56
  70. 70. El Sabagh A, Sorour S, Ragab A, Islam MS, Barutçular C, Ueda A, et al. Alleviation of adverse effects of salt stress on soybean (Glycine max. L.) by using osmoprotectants and organic nutrients. International Journal of Agricultural and Biosystems Engineering. 2015;9(9):1014-1018
  71. 71. Nounjan N, Theerakulpisut P. Effects of exogenous proline and trehalose on physiological responses in rice seedlings during salt-stress and after recovery. Plant Soil and Environment. 2012;58(7):309-315
  72. 72. Meloni DA, Oliva MA, Ruiz HA, Martinez CA. Contribution of proline and inorganic solutes to osmotic adjustment in cotton under salt stress. Journal of Plant Nutrition. 2001;24(3):599-612
  73. 73. Mäkelä P, Kärkkäinen J, Somersalo SJBP. Effect of glycinebetaine on chloroplast ultrastructure, chlorophyll and protein content, and RuBPCO activities in tomato grown under drought or salinity. Biologia Plantarum. 2000;43(3):471-475
  74. 74. Chou TS, Chao YY, Kao CH. Involvement of hydrogen peroxide in heat shock-and cadmium-induced expression of ascorbate peroxidase and glutathione reductase in leaves of rice seedlings. Journal of Plant Physiology. 2012;169(5):478-486
  75. 75. Saxena M, Saxena J, Nema R, Singh D, Gupta A. Phytochemistry of medicinal plants. Journal of Pharmacognosy and Phytochemistry. 2013;1:168-182
  76. 76. Khan MSA, Karim MA, Haque MM. Genotypic differences in growth and ions accumulation in soybean under NaCl salinity and water stress conditions. Bangladesh Agronomy Journal. 2014;17(1):47-58
  77. 77. Zhu M, Zhou M, Shabala L, Shabala S. Linking osmotic adjustment and stomatal characteristics with salinity stress tolerance in contrasting barley accessions. Functional Plant Biology. 2014;42(3):252-263
  78. 78. Zhao C, Zhang H, Song C, Zhu JK, Shabala S. Mechanisms of plant responses and adaptation to soil salinity. The Innovation. 2020;1:1-41
  79. 79. Sripinyowanich S, Klomsakul P, Boonburapong B, Bangyeekhun T, Asami T, Gu H, et al. Exogenous ABA induces salt tolerance in indica rice (Oryza sativa L.): The role of OsP5CS1 and OsP5CR gene expression during salt stress. Environmental and Experimental Botany. 2013;86:94-105
  80. 80. Cao H, Guo S, Xu Y, Jiang K, Jones AM, Chong K. Reduced expression of a gene encoding a Golgi localized monosaccharide transporter (OsGMST1) confers hypersensitivity to salt in rice (Oryza sativa). Journal of Experimental Botany. 2011;62(13):4595-4604
  81. 81. Mathan J, Singh A, Ranjan A. Sucrose transport in response to drought and salt stress involves ABA-mediated induction of OsSWEET13 and OsSWEET15 in rice. Physiologia Plantarum. 2021;171(4):620-637
  82. 82. Hasthanasombut S, Supaib-ulwatana K, Mii M, Nakamura I. Genetic manipulation of japonica rice using the OsBADH1 gene from Indica rice to improve salinity tolerance. Plant Cell Tissue and Organ Culture (PCTOC). 2011;104:79-89
  83. 83. Irshad A, Rehman RNU, Abrar MM, Saeed Q , Sharif R, Hu T. Contribution of rhizobium–legume symbiosis in salt stress tolerance in Medicago truncatula evaluated through photosynthesis, antioxidant enzymes, and compatible solutes accumulation. Sustainability. 2021;13(6):3369
  84. 84. Rubio MC, Bustos-Sanmamed P, Clemente MR, Becana M. Effects of salt stress on the expression of antioxidant genes and proteins in the model legume Lotus japonicus. New Phytologist. 2009;181(4):851-859
  85. 85. Hernandez JA, Jimenezm A, Mullineaux P, Sevilla F. Tolerance of pea (Pisum sativum L.) to long term salt stress is associated with induction of antioxidant defenses. Plant Cell Environment. 2000;23:853-862
  86. 86. Yasar F, Ellialtioglu S, Yildiz K. Effect of salt stress on antioxidant defense systems, lipid peroxidation, and chlorophyll content in green bean. Russian Journal of Plant Physiology. 2008;55:782-786. DOI: 10.1134/S1021443708060071
  87. 87. Ma L, Zhang H, Sun L. NADPH oxidase AtrbohD and AtrbohF function in ROS-dependent regulation of Na+/K+ homeostasis in Arabidopsis under salt stress. Journal of Experimental Botany. 2012;63:305-317
  88. 88. Agarwal S, Shaheen R. Stimulation of antioxidant system and lipid peroxidation by abiotic stresses in leaves of Momordica charantia. Brazilian Journal of Plant Physiology. 2007;19:149-161
  89. 89. Munir N. Aftab F enhancement of salt tolerance in sugarcane by ascorbic acid pre-treatment. African journal of Biotechnology. 2011;10:18362-18370
  90. 90. Liu C, Mao B, Ou S, Wang W, Liu L, Wu Y, et al. OsbZIP71, a bZIP transcription factor, confers salinity and drought tolerance in rice. Plant Molecular Biology. 2014;84:19-36
  91. 91. Guan Q , Liao X, He M, Li X, Wang Z, Ma H, et al. Tolerance analysis of chloroplast OsCu/Zn-SOD overexpressing rice under NaCl and NaHCO3 stress. PLoS One. 2017;12(10):e0186052
  92. 92. Zhang Z, Zhang Q , Wu J, Zheng X, Zheng S, Sun X, et al. Gene knockout study reveals that cytosolic ascorbate peroxidase 2 (OsAPX2) plays a critical role in growth and reproduction in rice under drought, salt and cold stresses. PLoS One. 2013;8(2):e57472
  93. 93. Ghosh A, Pareek A, Sopory SK, Singla-Pareek SL. A glutathione responsive rice glyoxalase II, Os GLYII-2, functions in salinity adaptation by maintaining better photosynthesis efficiency and anti-oxidant pool. The Plant Journal. 2014;80(1):93-105
  94. 94. Lima-Melo Y, Carvalho FE, Martins MO, Passaia G, Sousa RH, Neto MCL, et al. Mitochondrial GPX1 silencing triggers differential photosynthesis impairment in response to salinity in rice plants. Journal of Integrative Plant Biology. 2016;58(8):737-748
  95. 95. Ushimaru T, Nakagawa T, Fujioka Y, Daicho K, Naito M, Yamauchi Y, et al. Transgenic Arabidopsis plants expressing the rice dehydroascorbate reductase gene are resistant to salt stress. Journal of Plant Physiology. 2006;163(11):1179-1184
  96. 96. Todorova D, Sergiev I, Alexieva V, Karanov E, Smith A, Hall M. Polyamine content in Arabidopsis thaliana (L.) Heynh during recovery after low and high temperature treatments. Plant Growth Regulation. 2007;51:185-191
  97. 97. Zhang GW, Xu SC, Hu QZ, Mao WH, Gong YM. Putrescine plays a positive role in salt-tolerance mechanisms by reducing oxidative damage in roots of vegetable soybean. Journal of Integrative Agriculture. 2014;13:349-357
  98. 98. Sheokand S, Kumari A, Sawhney V. Effect of nitric oxide and putrescine on antioxidative responses under NaCl stress in chickpea plants. Physiology and Molecular Biology of Plants. 2008;14:355-362. DOI: 10.1007/s12298-008-0034-y
  99. 99. Suleiman S. Effects of exogenous application of diamine (putrescine) on growth and mineral elements distribution in Faba bean plants under saline conditions. Tishreen University Journal of Research Science Studies. 2007;30:257-265
  100. 100. Mahdi AHA. Improvement of salt tolerance in Vicia faba (L.) plants by exogenous application of polyamines. Egyptian Journal of Agronomy. 2016;38:1-21
  101. 101. Radhakrishnan R, Lee IJ. Spermine promotes acclimation to osmotic stress by modifying antioxidant, abscisic acid, and jasmonic acid signals in soybean. Journal of Plant Growth Regulation. 2013;32:22-30
  102. 102. Ghosh S, Mitra S, Paul A. Physiochemical studies of sodium chloride on mungbean (Vigna radiata L. Wilczek) and its possible recovery with spermine and gibberellic acid. The Scientific World Journal. 2015;2015:1-8. DOI: 10.1155/2015/858016
  103. 103. Dadasoglu E, Ekinci M, Kul R, Shams M, Turan M, Yildirim E. Nitric oxide enhances salt tolerance through regulating antioxidant enzyme activity and nutrient uptake in pea. Legume Research. 2021;44(1):41-45
  104. 104. Salahuddin M, Nawaz F, Shahbaz M, Naeem M, Zulfiqar B, Shabbir RN, et al. Effect of exogenous nitric oxide (NO) supply on germination and seedling growth of mungbean (cv. Nm-54) under salinity stress. Seed. 2017;40:1-7
  105. 105. Ahanger MA, Aziz U, Alsahli AA, Alyemeni MN, Ahmad P. Influence of exogenous salicylic acid and nitric oxide on growth, photosynthesis, and ascorbate-glutathione cycle in salt stressed Vigna angularis. Biomolecules. 2019;10(1):42
  106. 106. Sofo A, Scopa A, Nuzzaci M, Vitti A. Ascorbate peroxidase and catalase activities and their genetic regulation in plants subjected to drought and salinity stresses. International Journal of Molecular Sciences. 2015;16(6):13561-13578
  107. 107. Shu K, Qi Y, Chen F, Meng Y, Luo X, Shuai H, et al. Salt stress represses soybean seed germination by negatively regulating GA biosynthesis while positively mediating ABA biosynthesis. Frontier in Plant Sciences. 2017;10(8):1372. DOI: 10.3389/fpls.2017.01372
  108. 108. Wolf O, Jeschke WD. Long distance transport of abscisic acid in NaCl-treated intact plants of Lupinus albus. Journal of Experimental Botany. 1990;41:593-600
  109. 109. Huang Y, Jiao Y, Xie N, Guo Y, Zhang F, Xiang Z, et al. OsNCED5, a 9-cis-epoxycarotenoid dioxygenase gene, regulates salt and water stress tolerance and leaf senescence in rice. Plant Science. 2019;287:110188
  110. 110. Kim H, Lee K, Hwang H, Bhatnagar N, Kim DY, Yoon IS, et al. Overexpression of PYL5 in rice enhances drought tolerance, inhibits growth, and modulates gene expression. Journal of Experimental Botany. 2014;65(2):453-464
  111. 111. Zong W, Tang N, Yang J, Peng L, Ma S, Xu Y, et al. Feedback regulation of ABA signaling and biosynthesis by a bZIP transcription factor targets drought-resistance-related genes. Plant Physiology. 2016;171(4):2810-2825
  112. 112. Ellouzi H, Hamed KB, Hernández I, Cela J, Müller M, Magné C, et al. Comparative study of the early osmotic, ionic, redox and hormonal signaling response in leaves and roots of two halophytes and a glycophyte to salinity. Planta. 2014;240:1299-1317
  113. 113. Ghanem ME, Albacete A, Martínez-Andújar C, Acosta M, Romero-Aranda R, Dodd IC, et al. Hormonal changes during salinity-induced leaf senescence in tomato (Solanum lycopersicum L.). Journal of Experimental Botany. 2008;59(11):3039-3050
  114. 114. Zhang M, Smith JAC, Harberd NP, Jiang C. The regulatory roles of ethylene and reactive oxygen species (ROS) in plant salt stress responses. Plant Molecular Biology. 2016;91:651-659
  115. 115. Li CH, Wang G, Zhao JL, Zhang LQ , Ai LF, Han YF, et al. The receptor-like kinase SIT1 mediates salt sensitivity by activating MAPK3/6 and regulating ethylene homeostasis in rice. The Plant Cell. 2014;26(6):2538-2553
  116. 116. Qin H, Wang J, Chen X, Wang F, Peng P, Zhou Y, et al. Rice OsDOF15 contributes to ethylene-inhibited primary root elongation under salt stress. New Phytologist. 2019;223(2):798-813
  117. 117. Shan C, Mei Z, Duan J, Chen H, Feng H, Cai W. OsGA2ox5, a gibberellin metabolism enzyme, is involved in plant growth, the root gravity response and salt stress. PLoS One. 2014;9(1):e87110
  118. 118. Nishiyama R, Watanabe Y, Fujita Y, Le DT, Kojima M, Werner T, et al. Analysis of cytokinin mutants and regulation of cytokinin metabolic genes reveals important regulatory roles of cytokinins in drought, salt and abscisic acid responses, and abscisic acid biosynthesis. The Plant Cell. 2011;23(6):2169-2183
  119. 119. Wu X, He J, Chen J, Yang S, Zha D. Alleviation of exogenous 6-benzyladenine on two genotypes of eggplant (Solanum melongena mill.) growth under salt stress. Protoplasma. 2014;251:169-176
  120. 120. Toda Y, Tanaka M, Ogawa D, Kurata K, Kurotani KI, Habu Y, et al. Rice salt sensitive forms a ternary complex with JAZ and class-C bHLH factors and regulates jasmonate-induced gene expression and root cell elongation. The Plant Cell. 2013;25(5):1709-1725
  121. 121. Qiu Z, Guo J, Zhu A, Zhang L, Zhang M. Exogenous jasmonic acid can enhance tolerance of wheat seedlings to salt stress. Ecotoxicology and Environmental Safety. 2014;104:202-208
  122. 122. Kang DJ, Seo YJ, Lee JD, Ishii R, Kim KU, Shin DH, et al. Jasmonic acid differentially affects growth, ion uptake and abscisic acid concentration in salt-tolerant and salt-sensitive rice cultivars. Journal of Agronomy and Crop Science. 2005;191(4):273-282
  123. 123. Filgueiras CC, Martins AD, Pereira RV, Willett DS. The ecology of salicylic acid signaling: Primary, secondary and tertiary effects with applications in agriculture. International Journal of Molecular Sciences. 2019;20:1-19
  124. 124. Liang W, Ma X, Wan P, Liu L. Plant salt-tolerance mechanism: A review. Biochemical and Biophysical Research Communications. 2018;495:286-291
  125. 125. Abogadallah GM. Antioxidative defense under salt stress. Plant Signal Behavior. 2010;5:369-374
  126. 126. He AL, Niu SQ , Zhao Q , Li YS, Gou JY, Gao HJ, et al. Induced salt tolerance of perennial ryegrass by a novel bacterium strain from the rhizosphere of a desert shrub Haloxylon ammodendron. International Journal of Molecular Sciences. 2018;19(2):69
  127. 127. An BY, Luo Y, Li JR, Qiao WH, Zhang XS, Gao XQ. Expression of a vacuolar Na+/H+ antiporter gene of alfalfa enhances salinity tolerance in transgenic Arabidopsis. Acta Agronomical Sinica. 2008;34:557-564
  128. 128. Schroeder JI, Delhaize E, Frommer WB. Using membrane transporters to improve crops for sustainable food production. Nature. 2013;497:60-66
  129. 129. Liang W, Cui W, Ma X. Function of wheat Ta-UnP gene in enhancing salt tolerance in transgenic Arabidopsis and rice. Biochemical and Biophysical Research Communications. 2014;450:794-801
  130. 130. Sun SJ, Guo SQ , Yang X, Bao YM, Tang HJ, Sun H. Functional analysis of a novel Cys2/His2 type zinc finger protein involved in salt tolerance in rice. Journal of Experimental Botany. 2010;61:2807-2818
  131. 131. Zhang L, Xi D, Li S, Gao Z, Zhao S. Shi J a cotton group CMAP kinase gene, GhMPK2, positively regulates salt and drought tolerance in tobacco. Plant Molecular Biology. 2011;77:17-31

Written By

Parastoo Majidian and Hamidreza Ghorbani

Submitted: 29 December 2023 Reviewed: 28 February 2024 Published: 17 July 2024