Open access peer-reviewed chapter

Perspectives of Phytohormones Application to Enhance Salinity Tolerance in Plants

Written By

Imran Khan, Muhammad Umer Chattha, Rizwan Maqbool, Muqarrab Ali, Muhammad Asif, Muhammad Umair Hassan and Muhammad Talha Aslam

Submitted: 03 October 2023 Reviewed: 07 October 2023 Published: 02 February 2024

DOI: 10.5772/intechopen.1003714

From the Edited Volume

New Insights Into Phytohormones

Basharat Ali and Javed Iqbal

Chapter metrics overview

48 Chapter Downloads

View Full Metrics

Abstract

Plants undergo a wide range of morphological, cellular, anatomical, and physiological changes in response to salinity stress. However, plants produce some signaling molecules, usually known as phytohormones, to combat stress conditions. Salinity tolerance is a complex mechanism, whereas phytohormones have a central role in it. Phytohormone-mediated plant responses improve nutrient uptake, the source-sink relationship, and plant growth and development. Phytohormones triggers the specific gene expressions which are responsible for the modification of various plant mechanisms under salinity stress. This review summarized the most recent research findings about plant responses to salinity stress at physiological and molecular levels and discussed the probable function of several (abscisic acid, indole acetic acid, cytokinins, gibberellic acid, salicylic acid, brassinosteroids, ethylene, and triazoles) phytohormones and their interaction in modulating salinity stress. Further, the understanding of specific genes involved in phytohormonal regulation toward salinity tolerance is a key to developing breeding and transgenic approaches for meeting food demand under sustainable crop production.

Keywords

  • salt stress
  • plant growth
  • phytohormones
  • salinity tolerance
  • future perspectives
  • abiotic stress
  • hormonal regulation

1. Introduction

The world’s population is continuously increasing, and it is expected to reach 9.1 billion in 2050 that would require 70% more food to feed to world [1, 2, 3, 4]. The changing climate is endangering the productivity and sustainability of agricultural production systems [5]. It would be the biggest challenge to meet the food requirement of growing population in near future [5]. Climate change exerts a range of severe abiotic stressors including high salt, drought, cold, and heat stress, which significantly impede crop yields [6]. Among these, salt stress is the most widely spread abiotic stress that covered the 1/3rd area of arable land globally [7] and is increasing by 10% annually due to global climate change [8]. Salt stress lowers seed germination, seedling establishment, enzymatic production, and protein synthesis in plants [9]. Salt stress is detrimental to crop growth, yield, and overall agricultural productivity [10]. Salinity develops osmotic stress and ion toxicity that affect the metabolic processes of plants [11]. The antagonism in essential nutrient uptake by plant roots from saline soil leads to the deficiency of Ca2+, K+, Fe++, and Zn++ ions and water scarcity in plant tissues [12]. Paliwal et al. [13] noticed shrinkage of leaf area, less stomatal conductance, and chlorophyll concentration, alongside the ROS development, which lowers the seed germination, plant height, and biomass [13]. Due to this, the average agricultural yield was reduced by 50% [2, 14].

However, crop plants exhibit complex alterations to mitigate the negative effect of salt stress for survival [15]. These alterations may include both morphological and developmental patterns (growth plasticity) and physiological and biochemical adaptations [16]. Further, plants build up bioactive endogenous compounds, also known as phytohormones, which accelerate the plant growth and development under stress conditions [17]. Phytohormones produce various organic solutes such as sugars, polyols, betaines, and proline to protect the cell structure, ionic balance, reactive oxygen species (ROS) scavenging, protein regulation, and expression of related genes to enhance crop growth and developmental processes under stress conditions [17, 18]. Further, enhancing stress tolerance has emerged as a significant area of focus in the agricultural sector that improves the crop performance under stress, environmental sustainability, horticulture, and economic viability.

The primary emphasis of this chapter is to develop a comprehensive understanding of phytohormones application in modulating the biochemical and physiological processes of plants exposed to salinity-induced stress. Several studies highlighted various detrimental effects of salinity on plants [19]. Contrarily, the salinity tolerance can be regulated through various kinds of phytohormones including auxin, cytokinin, gibberellic acid, abscisic acid, and ethylene [20]. Further, salicylic acid, brassinosteroids, and triazoles are also found imperative to enhance salinity stress tolerance in plants [20]. Plants produce different proteins under stress condition that act as phytohormones to regulate the plant’s normal functions [21]. However, poor hormonal regulation can lead to reduced germination and emergence and limited plant growth and development under salinity stress [22]. Moreover, the exogenous application of phytohormones found a promising practical strategy to deal with salt stress. The exogenous application of phytohormones exhibited a rapid increase in plant growth under stress conditions [17, 23, 24, 25]. Therefore, understanding the function of specific phytohormones will provide essential insights of plant mechanism to adapt salinity tolerance.

Advertisement

2. Soil salinization: a threat to global food production

Soil salinization is the most significant global abiotic stress that reduced the soil fertility of approximately 1100 Mha (7% of total Earth land) area [26]. Soil salinization is increased with natural geochemical processes (primary salinity) and different human activities (secondary salinity) [27]. Primary salinization is caused by various atmospheric depositions, sea-level rise, saltwater intrusion into freshwater aquifers, and rising temperatures, while excessive fertilizer use, poorly managed practices, and intensified agriculture activities increase the secondary salinization that lowered the productivity of 30% cultivated land [28, 29]. The limited availability of freshwater resources for irrigation purposes, coupled with the continuous deterioration of agricultural fields due to salt stress, leads to significant reductions in agricultural productivity of arid and semiarid regions.

Soil salinization began from hundreds to thousands of years ago that degraded the arable land globally [30]. About 30% of irrigated land and 6% of the total land is degraded by soil salinity that caused an estimated $12 billion loss to agricultural production [15, 31]. The rising salinity in arable areas highlights the necessity for the understanding of the plant-salinity-tolerance mechanisms to maintain crop productivity. A significant reduction in plant growth and development, photosynthesis, respiration, and protein synthesis is reported due to salinity stress [32, 33]. The production of reactive oxygen species (ROS), including superoxide anion (O2•–), hydrogen peroxide (H2O2), and hydroxyl radicals (HO) injuring the chloroplasts and mitochondria, is a significant indicator of oxidative damage in salinity [34]. These ROS disrupt the membranous structure and its cell permeability; thus, nutrient availability is reduced under salinity stress [35]. Therefore, plants use antioxidant enzymes as a defensive mechanism to safeguard nucleic acids, proteins, and membrane lipids to avoid the detrimental impacts of ROS [36].

2.1 Causes of soil salinization

The buildup of water-soluble salts (Na+, K+, Cl, and SO42−) inside the root zone raises the salinity. This accumulation leads to osmotic fluctuations that impede the capacity of plant roots to uptake water [37, 38, 39]. The presence of high salt ions develops hyper-ionic salt stress. Soluble sodium and chloride ions are the primary ions responsible for developing soil salinity. In addition, the high Na+ ions develop a sodicity problem in soil [40]. Ancient cells were built to survive salinity since the early evolution of life originated in primeval oceans with similar or even more salt than contemporary oceans [41]. Therefore, numerous terrestrial plants have the ability to endure low to moderate levels of salinity. In contrast, naturally existing salt-tolerant plants (halophytes) exhibit a stringent salinity tolerance mechanism. However, a majority of growing crop species belong to the glycophyte group and therefore exhibit a poorly adapted mechanism to thrive in saline soils. Rice and tomato cultivars are highly susceptible to developing phytotoxic ions due to osmotic damage in saline stress conditions [42]. Salt-induced osmotic stress is observed during the early phase of salt exposure as a result of the gradual absorption of salts and consequent decrease in water potential in the vicinity of the root zone. This reduction in water conductivity within plant cells predominantly hampers plant growth [43, 44]. However, prolonged salinity stress leads to the accumulation of Na+, Cl, and SO42− ions, which cause ion toxicity and hinder nutrient absorption, hence being detrimental to the growing plant cells and tissues [45]. The adverse effects of salt stress on plants can be observed in their morphology, physiology, and biochemical properties. Morphologically, plants experience stunted growth, chlorosis, and impaired seed germination. Physiologically, salt stress inhibits photosynthesis and disrupts nutrient balance. Further, plants undergo various oxidative stresses, electrolyte leakage, and membrane disorganization during biochemical processes [46, 47]. Moreover, salinity negatively affected the reproductive stage of the crop [48]. Therefore, a comprehensive understanding of salinity-induced injuries should be known, highlighting the need for salinity tolerance in plants for sustainable food production around the world.

2.2 Effects of salinity on plants: a challenge to agriculture

2.2.1 Plant growth and development

Salt stress decreases plant growth, and the rate of the decreasing trend depends on the plant’s growth stage and the intensity of the stress [49]. Researchers discovered that the stunted growth of plants acts as an adaptive mechanism for their survival against salt stress [50]. Salinity stress inhibits gene expression (cyclin and cyclin-dependent kinase) necessary for cell growth and number in the plant meristem, nutrient and water absorption, and plant stability. Numerous plants react quickly to stress and consequently stop their growth. Contrarily, some plants face salinity stress while maintaining their normal growth and start dying [51]. In moderate salinity, plants with quick responses activate the dehydration that results in cell shrinkage that recovers later [52, 53, 54]. However, salinity caused cell injury and decreased cell elongation and division, which resulted in less root and shoot growth in plants [55]. Under salinity, soil water component alteration in the rhizosphere disturbs the cell-water relationship of plants that develop osmotic injury [56]. Further, the osmotic injury lowers water uptake, causing inhibition or reduction in photosynthesis and respiration and carbohydrate and protein synthesis, leading to decreased plant growth and development [57, 58].

2.2.2 Photosynthesis

Photosynthesis is one of the most fundamental physiological attributes of plants, wherein solar energy is transformed into chemical energy. There may be a drop in photosynthesis caused by salt because of poor biosynthesis of chlorophyll [59], changes in enzyme activity [60], the closing of stomata [61], less carbon dioxide supplementation [62], and a damaged photosystem [63]. A significant decrease in chlorophyll levels in plants exposed to high salt concentrations could be due to the enhanced oxidation and degradation of chlorophyll owing to high ROS accumulation [63]. Further, ROS production under salinity stress inhibits the electron transport chain (ETC) that produces pseudocyclic electron transport chains [64]. As a result, modifications in photosynthetic proteins and disruptions in the assembly of photosystems occur [65]. Furthermore, the swelling of the thylakoid membrane under high salinity stress was observed to destroy the chloroplast ultrastructure [66]. Consequently, substantial photosynthesis resulted in less plant growth, which led to a decrease in plant production of Jatropha curcas, O. sativa, P. oleracea, and Solanum melongena in saline conditions [47, 67, 68, 69].

2.2.3 Nutrient balance

Plant growth and development in optimum soil conditions is commonly represented by the “generalized dose-response curve” [70]. Either nutrient-induced deficiency or nutrient toxicity can hinder plant growth due to suboptimal nutrient uptake [26]. The primary cause of less mineral acquisition under salt stress is the interactive effect of Na+ and Cl with calcium (Ca2+), potassium (K+), and magnesium (Mg2+) ions [71]. Many researchers have reported a complex interaction between salt ions and vital mineral nutrients like potassium, phosphorus, and nitrogen [72, 73, 74]. Nitrogen is a vital mineral element that serves as a fundamental part of various cellular components within plants. Due to Cl/NO3 antagonism, greater Cl uptake and accumulation under saline conditions reduce total nitrogen uptake [75]. Phosphorous is needed for photosynthesis, storage, and energy transfer, whereas its uptake reduces in the presence of high Cl and SO42− concentrations and due to the low solubility of the Ca ± P minerals [76]. Though potassium is essential for protein synthesis and water balance, high Na+ ions reduce K+ availability in saline environments [77]. Thus, the maintenance of cellular equilibrium between sodium and potassium is crucial for the survival of plants in saline soil. Although sodium cannot replace potassium in cellular functions, they both share a chemical resemblance that leads to potassium substitution by sodium. Na+ and K+ compete for root uptake sites and lower the K+ and Ca2+ intake during salinity stress [78]. Similarly, a significant decrease in K+ absorption by plants is due to high Na+ ions, regardless of whether Na+, chloride, or sulfate ions predominate in soil solutions [71]. Thus, salinity creates nutritional imbalances and lowers crop productivity in a saline environment [79, 80].

2.2.4 Water relations

The rapid uptake of ions results in the buildup of ions within the cellular structures of plants, hence exerting a detrimental impact on plant water relations [81]. The osmotic potential of plant cells decreased under salinity conditions that created an osmotic gradient; consequently, water moved out, causing a loss in plant cell turgidity [82]. Less water intake and transpiration rate of C. olitorius is noticed from saline soil [83, 84]. Further, these aforementioned findings are supported by recent investigations [85, 86]. However, the maintenance of turgor pressure at a constant level in plants is achieved by decreasing their osmotic potential relative to the overall water potential during salt stress [87]. Moreover, the hydrostatic potential gradient has immense importance as it controls the water transport during transpiration from soil through root xylem cells in an apoplastic pathway. However, the saline stress alters the water intake primarily through the cell-to-cell pathway, thus affecting the water relations more significantly [88, 89].

2.2.5 Yield

Salinity affects energy metabolism, cell signaling, and the synthesis of proteins necessary for plant growth, development, and achieving high yields. Hence, it ultimately impedes agricultural production by impeding plant growth and plant adaptation to stress responses, resulting in an overall reduction in yield [90]. The rate of salt absorption and salt-induced osmotic stress showed cellular and membrane injury leads to a significant loss in crop biomass [91, 92]. Different researchers reported the yield loss of different field crops is affected by salt stress [93]. They suggested that a thorough comprehension of the salinity of plants has the potential to help further enhance agricultural crop productivity in salt-affected areas [26].

Advertisement

3. Role of phytohomones to alleviate salt stress on plants

Plants produce various kinds of phytohormones through biosynthetic pathways that can act locally or be transported to any part of the plant to maintain normal functioning in normal or stressful conditions [94]. These phytohormones, including auxins (IAA), cytokinins, gibberellins (GA), ethylene, abscisic acid (ABA), and brassinosteroids (BRs), regulate numerous physiological and biochemical processes to improve the growth and development of plants in saline conditions [19]. Moreover, strigolactone, nitric oxide (NO), and polyamines also act as phytohormones and regulate responses to environmental stimuli; therefore, phytohormones play a crucial role in salinity stress [95]. Salt stress-induced plants’ survival and avoidance mechanisms hinder their growth and normal functioning [96]. However, phytohormones regulate morphological, physiological, and other biochemical processes under stressful conditions [97]. Eyidogan et al. [98] observed continued changes in the synthesis, distribution, and signal transduction of phytohormones used for plant defensive mechanisms. Further, the salt stress signal transduction triggers the release of different phytohormones that act as baseline transducers [99]. Moreover, the details of the different phytohormones acting against salt stress are given below:

3.1 Abscisic acid (ABA)

ABA is a primitively important phytohormone that regulates the plant stress response through the expression of salt-responsive genes under salinity stress [100]. ABA showed promising results at different plant developmental stages of the crops under stress conditions [101]. Therefore, it is also known as stress hormone as internal ABA signaling activates under stress conditions to survive the plant through various adaptations [102]. Many scientists observed ABA production under various abiotic stresses [103]. ABA production regulates the water potential that tends to decrease in salinity stress [104]. Plants under salinity stress caused ABA generation that influenced the movement of guard cell and leaf water potential and thus considered as an important growth hormone under stress conditions [105]. However, a different response of ABA was observed for plant roots that increase their growth under stress conditions [106]. A contributory effect of soil pH was observed for the redistribution of ABA under stress conditions that control the stomatal opening and closure during salinity stress [107].

ABA and other phytohormones regulate root growth, stress-induced gene and protein expression (dehydrins and late embryogenesis abundant proteins), and compatible solute accumulation [15]. Salinity accelerates ABA buildup, which protects plants [108] and activates salt-induced genes in roots [109]. Another study found that salt-induced ABA helps leaves to restrict Na+ and Cl accumulation [110]. The ABA regulates stomatal closure to reduce transpiration and water loss [111]. Salinity causes stomatal closure due to ABA-induced cytoplasmic Ca++ increases. Further, succeeding plasmalemma ion channel activation and guard cell turgor losses are also linked to ABA-induced H2O2 generation, an intermediate signal of ABA in stomatal closure [112].

The synthesis and accumulation of osmoprotectants (proline) and dehydrins in response to ROS formation during salt-stress-induced dehydration depend on phytohormone [19]. ABA-mediated H2O2 buildup generates NO, which activates MAPK and upregulates ROS-scavenging antioxidant enzyme genes [113]. ABA reduces Na+ accumulation and is translocation in shoot to augment salinity adaptation [114]. ABA increased barley (Hordeum vulgare L.) root vacuolar Na+ while inhibiting xylem transport and plasmalemma influx [115]. Exogenous ABA prevented harmful Cl−1 ions from accumulating in citrus leaves, decreasing ethylene release and leaf abscission under salt stress [116]. Exogenously applied ABA increased K+/Na+ ratio that enhanced the salinity tolerance in rice [117]. Similarly, [118] found that exogenous ABA prevented Na+ and Cl ions and high Na+/K+ ratio in rice grains. Thus, ABA boosted rice grain production via boosting proline, soluble sugars, and K+ and Ca++ homeostasis [118]. Moreover, [109] noticed high accumulation of Ca2+ ions and membrane stability in plants due to ABA concentration under salt stress conditions. High ABA contents were recorded from salt-resistant maize hybrid leaves [119]. Further, a significant amount of ABA contents were noticed from tomato genotypes showing salinity tolerance [24].

In addition, the exogenously applied ABA at 100 μM increased the expression of OsP5CS1 responsible for proline accumulation that augment 20% more survival ratio under salt stress [120]. Mahajan et al. [103] indicated that multiple transcription factors regulate ABA-responsive gene expression. Salinity stress increases ABA and induces salt and osmotic relief genes [121]. ABA and salt stress regulate AtNHX1 expression and tissue distribution as ABA influences the HVP1 and HVP10 for vacuolar H+ inorganic pyrophosphatase and HvVHA-A for the catalytic subunit (subunit A) of ATPase [122]. They quantified the transcript levels and identified that ABA is responsible for salt-stress-adaptable gene expression in barley. Keskin et al. [102] found that ABA treatment induced TIP1 and GLP1 genes expression faster in wheat [123].

3.2 Auxin (IAA)

IAA is commonly known as the first plant hormone based on its discovery [2]. IAA controls cell elongation, vascular tissue development, and apical dominance [121]. Although IAA has been widely recognized for its effects on plant growth and development, it can regulate stress response or coordinate growth under stress [98]. Many researchers found a contributory role of IAA in plants during salinity stress [19, 124]. They proposed that a membrane-bound transcription factor (NTM2) includes auxin signal and modulates seed germination under salt stress. Further, NTM2 IAA30 gene overexpression mediates salt signaling pathway [124]. However, the mechanism of IAA to control salt stress is still unclear [123]. Further, salt alters the IAA metabolism and distribution, homeostasis, and its response against ROS production [125]. Many studies showed that IAA contents varied similar to ABA under salinity [126, 127]. High production of IAA lowers plant growth suggested the imbalance in stress-induced hormonal regulation plants [128]. However, the exogenous application of auxin augments plant growth and development.

IAA as a seed-priming agent also reduced the deleterious effects of salt stress on salt-sensitive wheat cultivar via regulating ionic homeostasis and auxin-induced leaf salicylic acid production [129, 130]. Auxin-response genes increase transcription of many genes in soybean, Arabidopsis, and rice [131]. Auxin/indoleacetic acid (Aux/IAA), GH3, and SAUR gene families contain these sensitive genes. Auxin reduces rice tiller bud growth by decreasing node OsIPT expression and cytokinin production [132]. However, the discovery of novel salt stress genes allows researchers to develop genetic engineering tools for stress tolerance [133]. From the perspectives of phytohormones, it is interesting to know that salinity considerably lowered the IAA production level in maize, while salicylic acid application increased IAA production. These findings show that hormonal homeostasis and cross talk are crucial for stress response signal perception, transduction, and mediation [126].

3.3 Cytokinins

Plants produce N6-substituted adenine derivatives commonly known as cytokinins that have either an aromatic or an isoprenoid side chain. Cytokinin affects cell division, chloroplast production, apical dominance, leaf senescence, vascular differentiation, nutrient mobilization, shoot differentiation, anthocyanin production, and photomorphogenic development in plants. Cytokinin is important to reduce the negative effects of salt stress on plant growth [134]. Cytokinin seed priming increased salinity stress tolerance by lowering ABA production in wheat [135]. Cytokinin also counteracts water stress-induced leaf and fruit abscission and seed dormancy under stress conditions. They observed cytokinin as ABA antagonists and IAA antagonists/synergists in plant processes [19]. The increased regulation in plant growth as salt stress adaption was noticed due to the cytokinin production [136].

Ref. [137] found that exogenous cytokinin boosted proline content and salt resistance in egg plants. Cytokinin is an intermediary in the protective action of epibrassinolide and methyl jasmonate in wheat under salt stress [109]. Cytokinin production increased the K-shuttle under salinity. Interestingly, cytokinin production decreased in the salt-resistant cultivar of barley under 65 mM NaCl. However, salinity negatively affected the growth of salt-sensitive cultivars [138]. Salinity lowers the contents of zeatin, zeatin riboside, isopentenyl adenine, and isopentenyl adenine in root and shoot of salt-sensitive barley genotype [139]. Further, the benzyl adenine production during salt stress hinders the growth parameters of barley, while the high cytokinin production improved growth rate and shoot/root ratio [139]. Kinetin acts as a direct free radical scavenger or part of the antioxidant process that protects purine degradation in stress conditions [140].

Moreover, functional investigations of cytokinin receptor mutants showed that all three Arabidopsis cytokinin receptors negatively regulate ABA signaling and osmotic stress responses due to cytokinin-dependent CRE1/AHK4 [141]. Cytokinin receptor genes are regulated by osmotic conditions, suggesting that their function in the osmotic stress response may be similar but poorly known [2].

3.4 Gibberellic acid (GA)

Gibberellic acid concentration increases under salinity stress that helps plants to regulate their mechanism through sugar production and other antioxidant enzyme metabolism [19]. GA substantially affects seed germination, leaf expansion, stem elongation, flower and trichome initiation, and fruit development [142]. The release of photosynthetic enzymes to improve plant photosynthetic efficiency by enlarging the leaf-area index and light interception [143]. GA increases photosynthate source potential and redistribution that helps in food storage [143]. GA3 reverses the morphological and stress-protective effects of triazoles (TR), demonstrating a close link between GA3 and plant stress protection [144]. Gibberellic acid improves plant water interactions and water usage efficiency (WUE) under salinity stress [142]. Maggio et al. [145] found that GA reduced stomatal resistance and improved WUE in tomatoes. The exogenously applied GA restores metabolic activity of plants [19].

Further, GA also regulates other phytohormones that helps in soybean development and higher yield production under salinity stress [146]. This improvement is due to the increased level of bioactive GA1 and GA4 that decreased the ABA, and SA; RNA maintenance and higher protein synthesis in soybean and mustard crops [146147]. Salt stress reduced the enzymatic activity, whereas GA signaling improves source-sink interaction under adverse environmental conditions [19]. GA enlarged the leaf area, root growth by increasing the nitrogen and magnesium uptake under salinity stress [148]. Moreover, GA3 boosted reducing sugars, antioxidant production, and protein synthesis and lowered ribonuclease and polyphenol oxidase in salt-stressed mung bean seedlings [149].

GA3-priming reduces ions (Na+, Cl) uptake and their partitioning in plant root and shoot under salinity stress that improved wheat germination [150]. Salt stress disturbs the hormonal balance that impairs plant development, whereas the use of phytohormone provides an interesting stress-reduction strategy in barley cultivar [151]. Achard et al. [152] noticed the function of DELLA protein for plant survival in salt toxicity. GA production is responsible for salt-inducible DDF1 (dwarf and delayed flowering 1) gene that increases seed germination and growth responses under high saline environments. Further, GA production is stimulated by the influence of IAA in different plant species [153]. Additionally, GA increases ABA catabolism and enhances ethylene that affects its signaling mechanism, suggesting a cross talk between these phytohormones to improve salinity tolerance in crops [154].

3.5 Salicylic acid (SA)

Salicylic acid regulates the plant growth, development, and defense mechanisms against abiotic stresses [101]. SA improves stomatal conductance, transpiration, photosynthetic rate, fruit production, glycolysis, and ion uptake and transport in plants [155].

Though, SA is well-known for its response to biotic stresses, new studies suggested a significant effect of SA to regulate plant functions in salt stress conditions [2]. SA boosts photosynthetic capability by increasing Rubisco activity and pigments [156]. Moreover, SA treatment improved soybean pigments, photosynthesis, and glucose metabolism, thus improving salt tolerance in plants [157].

SA enhanced IAA and lowered ABA in maize that resulted in higher root growth and decreased antioxidant enzyme production while under SS conditions [158]. Bastam et al. [155] found that exogenous SA application enhanced salt tolerance in pistachio seedlings, beans [159], wheat [160], barley [161], mung bean, and mustard [162, 163]. The soil application of SA reduced the salt ion accumulation in maize and mustard crops [164]. SA improves the proline and glycine betaine accumulation in plants, which produce more antioxidants that helps to tolerate the salt stress [164]. Further, the decreased lipid peroxidation and membrane permeability improves during salinity stress in plants [165]. Nazar et al. [163] recorded the antioxidant metabolism and ATP sulfurylase and nitrate reductase activity with SA treatment and improved photosynthesis in mustard growing under saline soils. Additionally, SA pre-treatment of Arabidopsis prevented salt-induced membrane depolarization and guard cell outward rectifying channel K+ loss [166].

The use of SA as a stress management hormone provides a significant improvement in salinity tolerance to agricultural crops and resulted in higher yields. However, SA interaction with other phytohormones is important to open new avenues in the field of crop science [167].

3.6 Brassinosteroids (BRs)

Brassinosteroids are the novel phytohormone that promote plant growth under stress conditions [168]. Plants modify their response in the presence of BRs during stress [169]. The study showed that Arabidopsis’ CPD gene encodes CYP90, which was a steroid hydroxylase-like cytochrome P450 protein [123]. BRs regulate stress response by activating or suppressing important enzymatic pathways, inducing protein synthesis, and producing other hormones [170]. BRs increased the ear number, length, kernel number, and weight of maize crop [171]. Exogenously applied BRs augment the pod formation and seed yield in legumes [172]. Moreover, cotton and rapeseed growth and seed yield also increased with BR application [173]. Additionally, BRs restored chlorophyll and enhanced nitrate reductase that is essential for nitrogen uptake; thus, seed germination and seedling growth of rice (Oryza sativa L.) improved under salinity stress [170]. The experiment showed that 0.5 M NaCl damages the cellular nuclei and chloroplast, whereas the use of BRs protects the cellular structure of barley [174]. For rice, the addition of BRs in salt solution (150 nM NaCl) reduced the Na+ and Cl ion uptake, which promotes seed germination that might be due to the high soluble protein synthesis [175]. The previous studies suggested the positive response of BRs to cope salinity, however, salt stress tolerance mechanism still demands a further study.

3.7 Ethylene

Unlike other phytohormones, ethylene is found in a gaseous form and controls plant growth and development [176, 177]. Researchers called ethylene as a stress hormone that is generated by several stimuli during stress conditions [178]. Plants with high salinity tolerance produce less ethylene level [179]. Contrarily, [180] reported the high ethylene production as a marker of salt tolerance in rice. Pierik et al. [181] observed larger crop growth stimulated with ethylene production. Salinity tolerance in Arabidopsis is due to the ethylene production [152]. In addition, Cao et al. [182] reported ethylene is vital for salt tolerance as the disruption in ethylene receptor activity caused salt sensitivity in plants.

The interaction and/or balance between the receptor and ethylene determines how a plant responds to salt stress. The plant exhibits large rosette with late flowering when receptor signaling is prominent in absence of ethylene signaling, whereas the abundant ethylene signaling triggers the early flowering of tiny rosette. In this, the plant faces two different extreme conditions and needs an adjustment [182]. Fine-tuning at different levels with active homeostasis can aid in plant survival and normal growth against stress conditions [15].

The mutant analysis of triple responses of etiolated seedlings treated with ethylene revealed an Arabidopsis ethylene signal transduction pathway involved in ethylene receptors, CTR1, EIN2, and EIN3, and other components [183]. Based on the structure, the five receptor genes of Arabidopsis are divided into two subfamilies as ETR1 and ERS1 in subfamily I and ETR2, EIN4, and ERS2 in subfamily II. Ethylene binds to all these receptors, and ETR1 kinases ethylene [184]. Both the ethylene receptors and signaling regulate plant development and stress responses. In Arabidopsis, salt stress lowered the ethylene receptor ETR1 expression, which mediates plant responses to abiotic stressors. Further, Khan et al. [20] observed a complex relationship between ethylene genes with GB-mediated salinity tolerance in T. aestivum.

3.8 Triazoles (TRs)

Triazoles are the plant growth regulators that can also be used as a fungicides [15]. Fletcher et al. [185] reported a noteworthy role of TRs against biotic and abiotic stressors in plants. Uniconazole was the most efficient TR that helps to boost the salinity tolerance, but its residual action in plant tissues and soil limits its usage in agriculture to prevent salinity [15]. TR increased net photosynthesis, intercellular CO2 concentration, and dry biomass of radish [185] and pigeon pea growing in saline soils [186]. The seed priming of wheat with paclobutrazol showed less Na+ accumulation and more water-soluble carbohydrates and reducing sugars under salt stress [187]. TR application increased nitrate reductase, protease, POD, SOD, and polyphenol oxidase activity under salt stress [186]. Propiconazole (TR-containing chemical) improved root development (length and biomass) and enzymatic antioxidant activity in Madagascar periwinkle plants under NaCl stress [188]. Though TR has a significant effect on plants, few studies have examined its role in relieving salt stress in diverse crops [185].

Advertisement

4. Conclusion

The global impact of salinity and the trend of a growing population are alarming for food security. Crop yield and productivity is hindered in saline soil due to osmotic and ionic damage. Salt-responsive gene regulation, the antioxidant defense system, nutrient transport, osmolyte synthesis, and salt compartmentalization are badly affected under salinity stress. Plants regulate antioxidant enzymes, calcium-mediated responses, and hormonal signaling against ROS production during salinity stress to maintain homeostasis for optimal growth. However, the production of phytohormones during salinity stress regulates plant growth, seed germination, metabolism, and physiological activities. Plants use phytohormones as phytoprotectants against salinity stress. Exogenously applied phytohormones improved the stomata regulation, gas exchange mechanism that positively affected plant–water relations, and nutrient uptake, and subsequently, plant growth and productivity increased. Moreover, the hormonal priming of seeds that improves the seed germination in saline soil that results in utmost plant population. Plants integrate exogenous and endogenous cues during stressful and non-stressful situations, linking their stress response by hormonal route. Thus, hormones and their interactions with other plant components are vital to develop appropriate management strategies and phytoprotectants against salinity stress.

Advertisement

Acknowledgments

Our heartfelt gratitude goes out to everyone who has contributed to this book chapter, whether via their skill or unwavering encouragement. We hope to leave a lasting impression on the scientific community, and we look forward to continuing this adventure of research and discovery with you all. We are also grateful to this publishing group for their supportive policy to spread knowledge in the form of this book.

References

  1. 1. Food and Agricultural Organization (FAO). Land and Plant Nutrition Management Service. Rome, Italy: FAO; 2009. Available from: www.fao.org/wsfs/forum2050/
  2. 2. Fahad S, Hussain S, Bano A, Saud S, Hassan S, Shan D, et al. Potential role of phytohormones and plant growth-promoting rhizobacteria in abiotic stresses: Consequences for changing environment. Environmental Science and Pollution Research. 2014;3:1-15. DOI: 10.1007/s11356-014-3754-2
  3. 3. He W, Goodkind D, Kowal PR. An aging world. In: International Population Reports. Chapter 1. 2015. Washington DC: U.S. Government Printing Office; 2016. pp. 2-3
  4. 4. Falcon WP, Naylor RL, Shankar ND. Rethinking global food demand for 2050. Population and Development Review. 2022;48:921-957. DOI: 10.1111/padr.12508
  5. 5. Alrteimei HA, Ash’aari ZH, Muharram FM. Last decade assessment of the impacts of regional climate change on crop yield variations in the Mediterranean region. Agriculture. 2022;12:1787. DOI: 10.3390/agriculture12111787
  6. 6. Saud S, Chen Y, Baowen L, Fahad S, Arooj S. The different impact on the growth of cool season turf grass under the various conditions on salinity and drought stress. International Journal of Agricultural Science Research. 2013;3:77-84
  7. 7. Biswas S, Seal P, Majumder B, Biswas AK. Efficacy of seed priming strategies for enhancing salinity tolerance in plants: An overview of the progress and achievements. Plant Stress. 2023;9:100186. DOI: 10.1016/j.stress.2023.100186
  8. 8. Farhat F, Ashaq N, Noman A, Aqeel M, Raja S, Naheed R, et al. Exogenous application of Moringa leaf extract confers salinity tolerance in sunflower by concerted regulation of antioxidants and secondary metabolites. Journal of Soil Science and Plant Nutrition. 2023;23:1-17. DOI: 10.1007/s42729-023-01301-8
  9. 9. Kumar A, Singh S, Gaurav AK, Srivastava S, Verma JP. Plant growth-promoting bacteria: Biological tools for the mitigation of salinity stress in plants. Frontiers in Microbiology. 2020;11:1216. DOI: 10.3389/fmicb.2020.01216
  10. 10. Khan A, Khan AL, Muneer S, Kim YH, Al-Rawahi A, Al-Harrasi A. Silicon and salinity: Crosstalk in crop-mediated stress tolerance mechanisms. Frontiers in Plant Sciences. 2019;10:1429. DOI: 10.3389/fpls.2019.01429
  11. 11. Rahman MM, Mostofa MG, Rahman MA, Islam MR, Keya SS, Das AK, et al. Acetic acid: A cost-effective agent for mitigation of seawater-induced salt toxicity in mung bean. Scientific Reports. 2019;9:15186. DOI: 10.1038/s41598-019-51178-w
  12. 12. Khan WUD, Aziz T, Maqsood MA, Farooq M, Abdullah Y, Ramzani PMA, et al. Silicon nutrition mitigates salinity Stress in maize by modulating ion accumulation, photosynthesis, and antioxidants. Photosynthetica. 2018;56:1047-1057. DOI: 10.3390/w15040739
  13. 13. Paliwal S, Tripathi MK, Tiwari S, Tripathi N, Payasi DK, Tiwari PN, et al. Molecular advances to combat different biotic and abiotic stresses in linseed (Linum usitatissimum L.): A comprehensive review. Genes. 2023;14:1461. DOI: 10.3390/genes14071461
  14. 14. Akram R, Fahad S, Masood N, Rasool A, Ijaz M, Ihsan MZ, et al. Plant growth and morphological changes in rice under abiotic stress. In: Handbook of Advances in Rice Research for Abiotic Stress Tolerance. Sawston, Cambridge, United Kingdom: Woodhead publishing; 2019. pp. 69-85. DOI: 10.1016/B978-0-12-814332-2.00004-6
  15. 15. Fahad S, Hussain S, Matloob A, Khan FA, Khaliq A, Saud S, et al. Phytohormones and plant responses to salinity stress: A review. Plant Growth Regulation. 2015;75:391-404. DOI: 10.1007/s10725-014-0013-y
  16. 16. Saud S, Li X, Chen Y, Zhang L, Fahad S, Hussain S, et al. Silicon application increases drought tolerance of Kentucky bluegrass by improving plant water relations and morphophysiological functions. The Scientific World Journal. 2014;2014:368694. DOI: 10.1155/2014/368694
  17. 17. Muchate NS, Nikalje GC, Rajurkar NS, Suprasanna P, Nikam TD. Plant salt stress: Adaptive responses, tolerance mechanism and bioengineering for salt tolerance. The Botanical Review. 2016;82:371-406. DOI: 10.1007/s12229-016-9173-y
  18. 18. Sharma I, Chin I, Saini S, Bhardwaj R, Pati PK. Exogenous application of brassinosteroid offers tolerance to salinity by altering stress responses in rice variety Pusa Basmati-1. Plant Physiology and Biochemistry. 2013;69:17-26. DOI: 10.1016/j.plaphy.2013.04.013
  19. 19. Iqbal N, Umar S, Khan NA, Khan MIR. A new perspective of phytohormones in salinity tolerance: Regulation of proline metabolism. Environmental and Experimental Botany. 2014;100:34-42. DOI: 10.1016/j.envexpbot.2013.12.006
  20. 20. Khan MIR, Iqbal N, Masood A, Per TS, Khan NA. Salicylic acid alleviates adverse effects of heat stress on photosynthesis through changes in proline production and ethylene formation. Plant Signaling & Behavior. 2013;8:e26374. DOI: 10.4161/psb.26374
  21. 21. Khan M, Ali S, Manghwar H, Saqib S, Ullah F, Ayaz A, et al. Melatonin function and crosstalk with other phytohormones under normal and stressful conditions. Genes. 2022;13:1699. DOI: 10.3390/genes13101699
  22. 22. Chen J, Pang X. Phytohormones unlocking their potential role in tolerance of vegetable crops under drought and salinity stresses. Frontiers in Plant Science. 2023;14:1121780. DOI: 10.3389/fpls.2023.1121780
  23. 23. Iqbal N, Masood A, Khan NA. Phytohormones in salinity tolerance: Ethylene and gibberellins cross talk. In: Khan NA, Nazar R, Iqbal N, Anjum NA, editors. Phytohormones and Abiotic Stress Tolerance in Plants. Berlin: Springer; 2012. pp. 77-98. DOI: 10.1007/978-3-642-25829-9_3
  24. 24. Amjad M, Akhtar J, Anwar-ul-Haq M, Yang A, Akhtar SC, Jacobsen E. Integrating role of ethylene and ABA in tomato plants adaptation to salt stress. Scientia Horticulturae. 2014;172:109-116. DOI: 10.1016/j.scienta.2014.03.024
  25. 25. Kaur Y, Das N. Roles of polyamines in growth and development of the Solanaceous crops under normal and stressful conditions. Journal of Plant Growth Regulation. 2022;2:1-22. DOI: 10.1007/s00344-022-10841-9
  26. 26. Balasubramaniam T, Shen G, Esmaeili N, Zhang H. Plants’ response mechanisms to salinity stress. Plants. 2023;12:2253-2267. DOI: 10.3390/plants12122253
  27. 27. Singh A. Soil salinity: A global threat to sustainable development. Soil Use and Management. 2022;38:39-67. DOI: 10.1111/sum.12772
  28. 28. Machado RMA, Serralheiro RP. Soil salinity: Effect on vegetable crop growth. Management practices to prevent and mitigate soil salinization. Horticulturae. 2017;3:30-37. DOI: 10.3390/horticulturae3020030
  29. 29. Hopmans JW, Qureshi AS, Kisekka I, Munns R, Grattan SR, Rengasamy P, et al. Critical knowledge gaps and research priorities in global soil salinity. Advances in Agronomy. 2021;169:1-191. DOI: 10.1016/bs.agron.2021.03.001
  30. 30. Chang AC, Brawer SD. Salinity and Drainage in San Joaquin Valley, California. Berlin/Heidelberg, Germany: Springer; 2016. pp. 6-21. DOI: 10.1007/978-94-007-6851-2
  31. 31. Shabala S. Learning from halophytes: Physiological basis and strategies to improve abiotic stress tolerance in crops. Annals of Botany. 2013;112:1209-1221. DOI: 10.1093/aob/mct205
  32. 32. Hussain S, Khaliq A, Matloob A, Wahid MA, Afzal I. Germination and growth response of three wheat cultivars to NaCl salinity. Soil and Environment. 2013;32:36-43
  33. 33. Mustafa Z, Pervez MA, Ayyub CM, Matloob A, Khaliq A, Hussain S, et al. Morpho-physiological characterization of chilli genotypes under NaCl salinity. Soil and Environment. 2014;33:133-141
  34. 34. Ul Islam SN, Asgher M, Khan NA. Hydrogen peroxide and its role in abiotic Stress tolerance in plants. In: Gasotransmitters Signaling in Plant Abiotic Stress: Gasotransmitters in Adaptation of Plants to Abiotic Stress. Cham: Springer International Publishing; 2023. pp. 167-195. DOI: 10.1007/978-3-031-30858-1_9
  35. 35. Shakya R. Markers of oxidative Stress in plants. In: Ecophysiology of Tropical Plants. Boca Raton, Florida, United States: CRC Press; 2024. pp. 298-310
  36. 36. Singh A, Rajput VD, Sharma R, Ghazaryan K, Minkina T. Salinity stress and nanoparticles: Insights into antioxidative enzymatic resistance, signaling, and defense mechanisms. Environmental Research. 2023;235:116585. DOI: 10.1016/j.envres.2023.116585
  37. 37. Munns R, Tester M. Mechanisms of salinity tolerance. The Annual Review of Plant Biology. 2008;59:651-681. DOI: 10.1146/annurev.arplant.59.032607.092911
  38. 38. 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;21:148. DOI: 10.3390/ijms21010148
  39. 39. Stavi I, Thevs N, Priori S. Soil salinity and sodicity in drylands: A review of causes, effects, monitoring, and restoration measures. Frontiers in Environmental Science. 2021;330:712831. DOI: 10.3389/fenvs.2021.712831
  40. 40. Foronda DA. Reclamation of a saline-sodic soil with organic amendments and leaching. Environmental Sciences Proceedings. 2022;16:56-67. DOI: 10.3390/environsciproc2022016056
  41. 41. Knauth LP. Salinity history of the Earth’s early ocean. Nature. 1998;395:554-555
  42. 42. Ullah A, Bano A, Khan N. Climate change and salinity effects on crops and chemical communication between plants and plant growth-promoting microorganisms under stress. Frontiers in Sustainable Food Systems. 2021;5:618092. DOI: 10.3389/fsufs.2021.618092
  43. 43. Munns R. Genes and salt tolerance: Bringing them together. New Phytologist. 2005;167:645-663. DOI: 10.1111/j.1469-8137.2005.01487.x
  44. 44. Abbasi H, Jamil M, Haq A, Ali S, Ahmad R, Malik Z, et al. Salt stress manifestation on plants, mechanism of salt tolerance and potassium role in alleviating it: A review. Zemdirbyste-Agriculture. 2016;103:229-238. DOI: 10.13080/z-a.2016.103.030
  45. 45. Isayenkov SV, Maathuis FJ. Plant salinity stress: Many unanswered questions remain. Frontiers in Plant Science. 2019;10:80-88. DOI: 10.3389/fpls.2019.00080
  46. 46. Ji X, Tang J, Zhang J. Effects of salt stress on the morphology growth physiological parameters of Juglans microcarpa L. Seedl. Plants. 2022;11:2381. DOI: 10.3390/plants11182381
  47. 47. Hannachi S, Steppe K, Eloudi M, Mechi L, Bahrini I, Van Labeke MC. Salt stress induced changes in photosynthesis and metabolic profiles of one tolerant (‘Bonica’) and one sensitive (‘black beauty’) eggplant cultivars (Solanum melongena L.). Plants. 2022;11:590-601. DOI: 10.3390/plants11050590
  48. 48. Rafaliarivony S, Ranarijaona HLT, Rasoafalimanana M, Radanielina T, Wissuwa M. Evaluation of salinity tolerance of lowland rice genotypes at the reproductive stage. bioRxiv. 2022;2:1-25. DOI: 10.1101/2022.08.22.504861
  49. 49. Yadav SP, Bharadwaj R, Nayak H, Mahto R, Singh RK, Prasad SK. Impact of salt stress on growth, productivity and physicochemical properties of plants: A review. International Journal of Chemical Studies. 2019;7:1793-1798
  50. 50. Munns R. Comparative physiology of salt and water stress. Plant, Cell and Environment. 2002;25:239-250. DOI: 10.1046/j.0016-8025.2001.00808.x
  51. 51. Chinnusamy V, Zhu JK. Plant salt tolerance. In: Plant Responses to Abiotic Stress. Berlin/Heidelberg, Germany: Springer; 2003. pp. 241-270. DOI: 10.1007/978-3-540-39402-0_10
  52. 52. Budagovskaya NV. Rapid response reactions of buckwheat plant shoots on changes in sodium chloride concentration at the root zone and blockage of calcium channels. The European Journal of Plant Science and Biotechnology. 2010;4:128-130
  53. 53. Ma Y, Dias MC, Freitas H. Drought and salinity stress responses and microbe-induced tolerance in plants. Frontiers in Plant Science. 2020;11:591911. DOI: 10.3389/fpls.2020.591911
  54. 54. Tufa KN. Review on effects, mechanisms and managements of plants water stress. Irrigation and Drainage Systems. 2022;11:2-7. DOI: 10.37421/2168-9768.2022.11.357
  55. 55. Ansari ZG, Rao GR, Aparna K. Physiological and molecular aspects of salinity stress tolerance in crop plants. Plant Stress Biology. 2019;3:47-59
  56. 56. Alkharabsheh HM, Seleiman MF, Hewedy OA, Battaglia ML, Jalal RS, Alhammad BA, et al. Field crop responses and management strategies to mitigate soil salinity in modern agriculture: A review. Agronomy. 2021;11:2299. DOI: 10.3390/agronomy11112299
  57. 57. Petretto GL, Urgeghe PP, Massa D, Melito S. Effect of salinity (NaCl) on plant growth, nutrient content, and glucosinolate hydrolysis products trends in rocket genotypes. Plant Physiology and Biochemistry. 2019;141:30-39. DOI: 10.1016/j.plaphy.2019.05.012
  58. 58. Denaxa NK, Nomikou A, Malamos N, Liveri E, Roussos PA, Papasotiropoulos V. Salinity effect on plant growth parameters and fruit bioactive compounds of two strawberry cultivars, coupled with environmental conditions monitoring. Agronomy. 2022;12:2279-2288. DOI: 10.3390/agronomy12102279
  59. 59. Qin C, Ahanger MA, Zhou J, Ahmed N, Wei C, Yuan S, et al. Beneficial role of acetylcholine in chlorophyll metabolism and photosynthetic gas exchange in Nicotiana benthamiana seedlings under salinity stress. Plant Biology. 2020;22:357-365. DOI: 10.1111/plb.13079
  60. 60. Al Hinai MS, Ullah A, Al-Rajhi RS, Farooq M. Proline accumulation, ion homeostasis and antioxidant defence system alleviate salt stress and protect carbon assimilation in bread wheat genotypes of Omani origin. Environmental and Experimental Botany. 2022;193:104687. DOI: 10.1016/j.envexpbot.2021.104687
  61. 61. Orzechowska A, Trtílek M, Tokarz KM, Szymanska R, Niewiadomska E, Rozpadek P, et al. Thermal analysis of stomatal response under salinity and high light. International Journal of Molecular Sciences. 2021;22:4663. DOI: 10.3390/ijms22094663
  62. 62. Zahra N, Al Hinai MS, Hafeez MB, Rehman A, Wahid A, Siddique KH, et al. Regulation of photosynthesis under salt stress and associated tolerance mechanisms. Plant Physiology and Biochemistry. 2022;178:55-69. DOI: 10.1016/j.plaphy.2022.03.003
  63. 63. Seleiman MF, Aslam MT, Alhammad BA, Hassan MU, Maqbool R, Chattha MU, et al. Salinity stress in wheat: Effects, mechanisms and management strategies. Phyton. 2022;91:667-694. DOI: 10.32604/phyton.2022.017365
  64. 64. Zahra N, Wahid A, Shaukat K, Hafeez MB, Batool A, Hasanuzzaman M. Oxidative stress tolerance potential of milk thistle ecotypes after supplementation of different plant growth-promoting agents under salinity. Plant Physiology and Biochemistry. 2021;166:53-65. DOI: 10.1016/j.plaphy.2021.05.042
  65. 65. Huihui Z, Yue W, Xin L, Guoqiang H, Yanhui C, Zhiyuan T, et al. Chlorophyll synthesis and the photoprotective mechanism in leaves of mulberry (Morus alba L.) seedlings under NaCl and NaHCO3 stress revealed by TMT-based proteomics analyses. Ecotoxicology and Environmental Safety. 2020;190:110164. DOI: 10.1016/j.ecoenv.2020.110164
  66. 66. Goussi R, Manaa A, Derbali W, Cantamessa S, Abdelly C, Barbato R. Comparative analysis of salt stress, duration and intensity, on the chloroplast ultrastructure and photosynthetic apparatus in Thellungiella salsuginea. Journal of Photochemistry and Photobiology B: Biology. 2018;183:275-287. DOI: 10.1016/j.jphotobiol.2018.04.047
  67. 67. Hnilickova H, Kraus K, Vachova P, Hnilicka F. Salinity stress affects photosynthesis, malondialdehyde formation, and proline content in Portulaca oleracea L. Plants. 2021;10:845-852. DOI: 10.3390/plants10050845
  68. 68. Wang X, Wang W, Huang J, Peng S, Xiong D. Diffusional conductance to CO2 is the key limitation to photosynthesis in salt-stressed leaves of rice (Oryza sativa). Physiologia Plantarum. 2018;163:45-58. DOI: 10.1111/ppl.12653
  69. 69. Silva END, Ribeiro RV, Ferreira-Silva SL, Viégas RA, Silveira JAG. Salt stress induced damages on the photosynthesis of physic nut young plants. Scientia Agricola. 2011;68:62-68. DOI: 10.1590/S0103-90162011000100010
  70. 70. Berry W, Wallace A. Toxicity: The concept and relationship to the dose response curve. Journal of Plant Nutrition. 1981;3:13-19. DOI: 10.1080/01904168109362814
  71. 71. Grattan SR, Grieve CM. Mineral element acquisition and growth response of plants grown in saline environments. Agriculture, Ecosystems & Environment. 1992;38:275-300. DOI: 10.1016/0167-8809(92)90151-Z
  72. 72. Kumar A, Kumar A, Bihari B, Qasmi M. Soil fertility and mineral nutrition of plants. Current Research in Soil Fertility. 2020;65:23-35
  73. 73. Aslam MT, Imran K, Chattha MU, Maqbool R, Ziaulhaq M, Lihong W, et al. The critical role of nitrogen in plants facing the salinity stress: Review and future prospective. Notulae Botanicae Horti Agrobotanici Cluj-Napoca. 2023;51:13347-13347. DOI: 10.15835/nbha51313347
  74. 74. Janaki D. Nutrient transformation in salt affected soil. In: Current Research in Soil Fertility. Vol. 24. AkiNik Publications; 2022. pp. 83-100
  75. 75. Munns R, Termaat A. Whole-plant responses to salinity. Functional Plant Biology. 1986;13:143-160. DOI: 10.1071/PP9860143
  76. 76. Hao D, Li X, Kong W, Chen R, Liu J, Guo H, et al. Phosphorylation regulation of nitrogen, phosphorus, and potassium uptake systems in plants. The Crop Journal. 2023;3:45-56. DOI: 10.1016/j.cj.2023.06.003
  77. 77. Turcios AE, Papenbrock J, Tränkner M. Potassium, an important element to improve water use efficiency and growth parameters in quinoa (Chenopodium quinoa) under saline conditions. Journal of Agronomy and Crop Science. 2021;207:618-630. DOI: 10.1111/jac.12477
  78. 78. Ran X, Huang X, Wang X, Liang H, Wang Y, Li J, et al. Ion absorption, distribution and salt tolerance threshold of three willow species under salt stress. Frontiers in Plant Science. 2022;13:969896. DOI: 10.3389/fpls.2022.969896
  79. 79. Hasana R, Miyake H. Salinity stress alters nutrient uptake and causes the damage of root and leaf anatomy in maize. KnE Life Sciences. 2017;3:219-225. DOI: 10.18502/kls.v3i4.708
  80. 80. Cruz JL, Coelho EF, Coelho Filho MA, Santos AAD. Salinity reduces nutrients absorption and efficiency of their utilization in cassava plants. Ciência Rural. 2018;48:e20180351. DOI: 10.1590/0103-8478cr20180351
  81. 81. Hailu B, Mehari H. Impacts of soil salinity/sodicity on soil-water relations and plant growth in dry land areas: A review. Journal of Natural Sciences Research. 2021;12:1-10. DOI: 10.7176/JNSR/12-3-01
  82. 82. Betzen BM, Smart CM, Maricle KL, MariCle BR. Effects of increasing salinity on photosynthesis and plant water potential in Kansas salt marsh species. Transactions of the Kansas Academy of Science. 2019;122:49-58. DOI: 10.1660/062.122.0105
  83. 83. Taneenah A, Nulit R, Yusof UK, Janaydeh M. Tolerance of Molokhia (Corchorus olitorius L.) seed with dead sea water, sea water, and NaCl: Germination and anatomical approach. Advances in Environmental Biology. 2015;9:106-116
  84. 84. Chaudhuri K, Choudhuri MA. Effects of short-term NaCl stress on water relations and gas exchange of two jute species. Journal of Agricultural Research. 1997;40:373-380. DOI: 10.1023/A:1001013913773
  85. 85. Álvarez S, Sánchez-Blanco MJ. Long-term effect of salinity on plant quality, water relations, photosynthetic parameters and ion distribution in Callistemon citrinus. Plant Biology. 2014;16:757-764. DOI: 10.1111/plb.12106
  86. 86. Sheldon AR, Dalal RC, Kirchhof G, Kopittke PM, Menzies NW. The effect of salinity on plant-available water. Plant and Soil. 2017;418:477-491. DOI: 10.1007/s11104-017-3309-7
  87. 87. Rajasekaran LR, Aspinall D, Jones GP, Paleg LG, Stress metabolism. IX. Effect of salt stress on trigonelline accumulation in tomato. Canadian Journal of Plant Science. 2001;81:487-498. DOI: 10.4141/P00-079
  88. 88. Shaheen S, Naseer S, Ashraf M, Akram NA. Salt stress affects water relations, photosynthesis, and oxidative defense mechanisms in Solanum melongena L. Journal of Plant Interaction. 2013;8:85-96. DOI: 10.1080/17429145.2012.718376
  89. 89. Mirfattahi Z, Karimi S, Roozban MR. Salinity induced changes in water relations, oxidative damage and morpho-physiological adaptations of pistachio genotypes in soilless culture. Acta Agriculturae Slovenica. 2017;109:291-302
  90. 90. Munns R, Gilliham M. Salinity tolerance of crops–what is the cost? New Phytologist. 2015;208:668-673. DOI: 10.1111/nph.13519
  91. 91. Maas EV, Hoffman GJ. Crop salt tolerance-current assessment. Journal of the Irrigation and Drainage Division. 1977;103:115-134. DOI: 10.1061/JRCEA4.0001137
  92. 92. Volkmar KM, Hu Y, Steppuhn H. Physiological responses of plants to salinity: A review. Canadian Journal of Plant Science. 1998;78:19-27. DOI: 10.4141/P97-020
  93. 93. Amombo E, Ashilenje D, Hirich A, Kouisni L, Oukarroum A, Ghoulam C, et al. Exploring the correlation between salt tolerance and yield: Research advances and perspectives for salt-tolerant forage sorghum selection and genetic improvement. Planta. 2022;255:71-88. DOI: 10.1007/s00425-022-03847-w
  94. 94. Peleg Z, Blumwald E. Hormone balance and abiotic stress tolerance in crop plants. Current Opinion in Plant Biology. 2011;14:290-295. DOI: 10.1016/j.pbi.2011.02.001
  95. 95. Davies PJ. Plant Hormones: Biosynthesis, Signal Transduction, Action. Kluwer, Dordrecht: Springer Science & Business Media; 2004
  96. 96. Skirycz A, Inze´ D. More from less: Plant growth under limited water. Current Opinion in Biotechnology. 2010;21:197-203. DOI: 10.1016/j.copbio.2010.03.002
  97. 97. Kosakivska IV, Vedenicheva NP, Babenko LM, Voytenko LV, Romanenko KO, Vasyuk VA. Exogenous phytohormones in the regulation of growth and development of cereals under abiotic stresses. Molecular Biology Reports. 2022;49:617-628. DOI: 10.1007/s11033-021-06802-2
  98. 98. Eyidogan F, Oz MT, Yucel M, Oktem HA. Signal transduction of phytohormones under abiotic stresses. In: Khan NA, Nazar R, Iqbal N, Anjum NA, editors. Phytohormones and Abiotic Stress Tolerance in Plants. Berlin: Springer; 2012. pp. 1-48. DOI: 10.1007/978-3-642-25829-9_1
  99. 99. Harrison MA. Cross-talk between phytohormone signaling pathways under both optimal and stressful environmental conditions. In: Khan NA, Nazar R, Iqbal N, Anjum NA, editors. Phytohormones and Abiotic Stress Tolerance in Plants. Berlin: Springer; 2012. pp. 49-76. DOI: 10.1007/978-3-642-25829-9_2
  100. 100. Sharma N, Abrams SR, Waterer DR. Uptake, movement, activity, and persistence of an abscisic acid analog (80 acetylene ABA methyl ester) in marigold and tomato. Journal of Plant Growth Regulation. 2005;24:28-35. DOI: 10.1007/s00344-004-0438-z
  101. 101. Devinar G, Llanes A, Masciarelli O, Luna V. Different relative humidity conditions combined with chloride and sulfate salinity treatments modify abscisic acid and salicylic acid levels in the halophyte Prosopis strombulifera. Plant Growth Regulation. 2013;70:247-256. DOI: 10.1007/s10725-013-9796-5
  102. 102. Keskin BC, Sarikaya AT, Yuksel B, Memon AR. Abscisic acid regulated gene expression in bread wheat. Australian Journal of Crop Science. 2010;4:617-625. DOI: 10.3316/INFORMIT.857732547077080
  103. 103. Mahajan S, Tuteja N. Cold, salinity and drought stresses: An overview. Archives of Biochemistry and Biophysics. 2005;444:139-158. DOI: 10.1016/j.abb.2005.10.018
  104. 104. Yildiz M, Poyraz İ, Çavdar A, Özgen Y, Beyaz R. Plant responses to salt stress. In: Plant Breeding-Current and Future Views. London, UK: IntechOpen; 2020. DOI: 10.5772/intechopen.93920
  105. 105. Zhang J, Jia W, Yang J, Ismail AM. Role of ABA in integrating plant responses to drought and salt stresses. Field Crops Research. 2006;97:111-119. DOI: 10.1016/j.fcr.2005.08.018
  106. 106. Jia W, Wang Y, Zhang S, Zhang J. Salt -stress-induced ABA accumulation is more sensitively triggered in roots than in shoots. Journal of Experimental Botany. 2002;53:2201-2206. DOI: 10.1093/jxb/erf079
  107. 107. Hartung W, Sauter A, Hose E. Abscisic acid in the xylem: Where does it come from, where does it go to? Journal of Experimental Botany. 2002;53:27-32. DOI: 10.1093/jexbot/53.366.27
  108. 108. Shakirova FM, Avalbaev AM, Bezrukova MV, Kudoyarova GR. Role of endogenous hormonal system in the realization of the antistress action of plant growth regulators on plants. Plant Stress. 2010;4:32-38
  109. 109. Parida AK, Das AB. Salt tolerance and salinity effects on plants: A review. Ecotoxicology & Environmental Safety. 2005;60:324-349. DOI: 10.1016/j.ecoenv.2004.06.010
  110. 110. Cabot C, Sibole JV, Barcelo J, Poschenrieder C. Abscisic acid decreases leaf Na+ exclusion in salt treated Phaseolus vulgaris L. Journal of Plant Growth Regulation. 2009;28:187-192. DOI: 10.1007/s00344-009-9088-5
  111. 111. Wilkinson S, Davies WJ. Drought, ozone, ABA and ethylene: New insights from cell to plant to community. Plant Cell and Environment. 2010;33:510-525. DOI: 10.1111/j.1365-3040.2009.02052.x
  112. 112. Kim TW, Wang ZY. Brassinosteroid signal transduction from receptor kinases to transcription factors. Annual Review of Plant Biology. 2010;61:681-704. DOI: 10.1146/annurev.arplant.043008.092057
  113. 113. Lee BR, La VH, Park SH, Mamun MA, Bae DW, Kim TH. H2O2-responsive hormonal status involves oxidative burst signaling and proline metabolism in rapeseed leaves. Antioxidants. 2022;11:566. DOI: 10.3390/antiox11030566
  114. 114. Khadri M, Tejera NA, Lluch C. Sodium chloride-ABA interaction in two common bean (Phaseolus vulgaris) cultivars differing in salinity tolerance. Environmental and Experimental Botany. 2007;60:211-218. DOI: 10.1016/j.envexpbot.2006.10.008
  115. 115. Behl R, Jeschke WD. Influence of abscisic acid on unidirectional fluxes and intracellular compartmentation of K+ and Na+ in excised barley root segments. Physiologia Plantarum. 1981;53:95-100. DOI: 10.1111/j.1399-3054.1981.tb04116.x
  116. 116. Gomez CA, Arbona V, Jacas J, Primomillo E, Talon M. Abscisic acid reduces leaf abscission and increases salt tolerance in citrus plants. The Journal of Plant Growth Regulation. 2002;21:234-240. DOI: 10.1007/s00344-002-0013-4
  117. 117. Bohra JS, Dorffling H, Dorffling K. Salinity tolerance of rice (Oryza sativa L.) with reference to endogenous and exogenous abscisic acid. Journal of Agronomy and Crop Science. 1995;174:79-86. DOI: 10.1111/j.1439-037X.1995.tb00197.x
  118. 118. Gurmani AR, Bano A, Ullah N, Khan H, J-ahangir M, Flowers TJ. Exogenous abscisic acid (ABA) and silicon (Si) promote salinity tolerance by reducing sodium (Na+) transport and bypass flow in rice (Oryza sativa indica). Australian Journal of Crop Science. 2013;7:1219-1226. DOI: 10.3316/INFORMIT.619232089838636
  119. 119. Zorb C, Geilfusb CM, Muhlingb KH, Ludwig-Muller J. The influence of salt stress on ABA and auxin concentrations in two maize cultivars differing in salt resistance. Journal of Plant Physiology. 2014;170:220-224. DOI: 10.1016/j.jplph.2012.09.012
  120. 120. 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. DOI: 10.1016/j.envexpbot.2010.01.009
  121. 121. Wang Y, Mopper S, Hasentein KH. Effects of salinity on endogenous ABA, IAA, JA, and SA in Iris hexagona. Journal of Chemical Ecology. 2001;27:327-342. DOI: 10.1023/A:1005632506230
  122. 122. Shi HZ, Zhu JK. Regulation of expression of the vacuolar Na+/H+ antiporter gene AtNHX1 by salt stress and abscisic acid. Plant Molecular Biology. 2002;50:543-550. DOI: 10.1023/A:1019859319617
  123. 123. Javid MG, Sorooshzadeh A, Moradi F, Sanavy SAMM, Allahdadi I. The role of phytohormones in alleviating salt stress in crop plants. Australian Journal of Crop Science. 2011;5:726-734
  124. 124. Jung J, Park C. Auxin modulation of salt stress signaling in arabidopsis seed germination. Plant Signaling and Behavior. 2011;6:1198-1200. DOI: 10.4161/psb.6.8.15792
  125. 125. Schopfer P, Liszkay A, Bechtold M, Frahry G, Wagner A. Evidence that hydroxyl radicals mediate auxin-induced extension growth. Planta. 2002;214:821-828. DOI: 10.1007/s00425-001-0699-8
  126. 126. Fahad S, Saud S, Chen Y, Wu C, Wang D. Abiotic Stress in Plants. London, UK: BoD–Books on Demand; 2021. DOI: 10.5772/intechopen.91549
  127. 127. El Sabagh A, Islam MS, Hossain A, Iqbal MA, Mubeen M, Waleed M, et al. Phytohormones as growth regulators during abiotic stress tolerance in plants. Frontiers in Agronomy. 2022;4:765068. DOI: 10.3389/fagro.2022.765068
  128. 128. Ribaut JM, Pilet PE. Water stress and indole-3ylacetic acid content of maize roots. Planta. 1994;193:502-507. DOI: 10.1007/BF02411554
  129. 129. Iqbal M, Ashraf M. Seed treatment with auxins modulates growth and ion partitioning in salt-stressed wheat plants. Journal of Integrated Plant Biology. 2007;49:1003-1015. DOI: 10.1111/j.1672-9072.2007.00488.x
  130. 130. Iqbal M, Ashraf M. Salt tolerance and regulation of gas exchange and hormonal homeostasis by auxin-priming in wheat. Pesquisa Agropecuaria Brasileira. 2013;48:1210-1219. DOI: 10.1590/S0100-204X2013000900004
  131. 131. Hagen G, Guilfoyle T. Auxin-responsive gene expression: Genes, promoters and regulatory factors. Plant Molecular Biology. 2002;49:373-385. DOI: 10.1023/A:1015207114117
  132. 132. Liu Y, Xu J, Ding Y, Wang Q , Li G, Wang S. Auxin inhibits the outgrowth of tiller buds in rice (Oryza sativa L.) by downregulating OsIPT expression and cytokinin biosynthesis in nodes. Australian Journal of Crop Science. 2011;5:169-174
  133. 133. Zhu JK. Salt and drought stress signal transduction in plants. Annual Review of Plant Biology. 2002;53:247-273. DOI: 10.1146/annurev.arplant.53.091401.143329
  134. 134. Barciszewski J, Siboska G, Rattan SIS, Clark BFC. Occurrence, biosynthesis and properties of kinetin (N6-furfuryladenine). Plant Growth Regulation. 2000;32:257-265. DOI: 10.1023/A:1010772421545
  135. 135. Iqbal M, Ashraf M, Jamil A. Seed enhancement with cytokinins: Changes in growth and grain yield in salt stressed wheat plants. Plant Growth Regulation. 2006;50:29-39. DOI: 10.1007/s10725-006-9123-5
  136. 136. Hadiarto T, Tran LS. Progress studies of drought-responsive genes in rice. Plant Cell Reports. 2011;30:297-310. DOI: 10.1007/s00299-010-0956-z
  137. 137. Wu X, He J, Chen J, Yang S, Zha D. Alleviation of exogenous 6-benzyladenine on two genotypes of egg-plant (Solanum melongena mill.) growth under salt stress. Protoplasma. 2013;251:169-176. DOI: 10.1007/s00709-013-0535-6
  138. 138. Kuiper D, Kuiper PJC, Lambers H, Schuit J, Staal M. Cytokinin concentration in relation to mineral nutrition and benzyladenine treatment in Plantago major ssp. Pleiosperma. Physiologia Plantarum. 1989;75:511-517. DOI: 10.1111/j.1399-3054.1989.tb05617.x
  139. 139. Kuiper D, Schuit J, Kuiper PJC. Actual cytokinin concentrations in plant tissue as an indicator for salt resistance in cereals. Plant and Soil. 1990;123:243-250. DOI: 10.1007/BF00011276
  140. 140. Chakrabarti N, Mukherji S. Alleviation of NaCl stress by pretreatment with phytohormones in Vigna radiata. Plant Biology. 2003;46:589-594. DOI: 10.1023/A:1024827931134
  141. 141. Tran LSP, Urao T, Qin F, Maruyama K, Kakimoto T, Shinozaki K, et al. Functional analysis of AHK1/ATHK1 and cytokinin receptor histidine kinases in response to abscisic acid, drought, and salt stress in Arabidopsis. National Academy of Sciences of the United States of America. 2007;104:20623-20628. DOI: 10.1073/pnas.0706547105
  142. 142. Yamaguchi S. Gibberellin metabolism and its regulation. Annual Review of Plant Biology. 2008;59:225-251. DOI: 10.1146/annurev.arplant.59.032607.092804
  143. 143. Khan NA, Singh S, Nazar R, Lone PM. The source–sink relationship in mustard. Asian and Australasian Journal of Plant Science and Biotechnology. 2007;1:10-18
  144. 144. Gilley A, Flecher RA. Gibberellin antagonizes paclobutrazole induced stress protection in wheat seedlings. Journal of Plant Physiology. 2007;103:200-207. DOI: 10.1016/S0176-1617(98)80066-7
  145. 145. Maggio A, Barbieri G, Raimondi G, de Pascale S. Contrasting effects of GA3 treatments on tomato plants exposed to increasing salinity. Journal of Plant Growth Regulation. 2010;29:63-72. DOI: 0.1007/s00344-009-9114-7
  146. 146. Hamayun M, Khan SA, Khan AL, Shin JH, Ahmad B, Shin DH, et al. Exogenous gibberellic acid reprograms soybean to higher growth and salt stress tolerance. Journal of Agricultural and Food Chemistry. 2010;58:7226-7232. DOI: 10.1021/jf101221t
  147. 147. Siddiqui MH, Khan MN, Mohammad F, Khan MMA. Role of nitrogen and gibberellic acid (GA3) in the regulation of enzyme activities and in osmoprotectant accumulation in Brassica juncea L. under salt stress. Journal of Agronomy and Crop Science. 2008;194:214-224. DOI: 10.1111/j.1439-037X.2008.00308.x
  148. 148. Tuna AL, Kaya C, Dikilitas M, Higgs D. The combined effects of gibberellic acid and salinity on some antioxidant enzyme activities, plant growth parameters and nutritional status in maize plants. Environmental and Experimental Botany. 2008;62:1-9. DOI: 10.1016/j.envexpbot.2007.06.007
  149. 149. Mohammed AHMA. Physiological aspects of mungbean plant (Vigna radiata L. Wilczek) in response to salt stress and gibberellic acid treatment. Research Journal of Agriculture and Biological Sciences. 2007;3:200-213
  150. 150. Radi AF, Shaddad MAK, El-Enany AE, Omran FM. Interactive effects of plant hormones (GA3 or ABA) and salinity on growth and some metabolites of wheat seedlings. In: Plant Nutrition: Food Security and Sustainability of Agro-Ecosystems Through Basic and Applied Research. Vol. 92. Springer Science & Business Media; 2006. pp. 436-437. DOI: 10.1007/0-306-47624-X_211
  151. 151. Vettakkorumakankav NA. Crucial role for gibberellin in stress protecting of plants. Plant and Cell Physiology. 1999;40:542-548. DOI: 10.1093/oxfordjournals.pcp.a029575
  152. 152. Achard P, Cheng H, De Grauwe L, Decat J, Schoutteten H, Moritz T, et al. Integration of plant responses to environmentally activated phytohormonal signals. Science. 2006;311:91-94. DOI: 10.1126/science.111864
  153. 153. Wolbang CM, Chandler PM, Smith JJ, Ross JJ. Auxin from the developing inflorescence is required for the biosynthesis of active gibberellins in barley stems. Plant Physiology. 2004;134:769-776. DOI: 10.1104/pp.103.030460
  154. 154. Gonai T, Kawahara S, Tougou M, Satoh S, Hashiba T, Hirai N, et al. Abscisic acid in the thermoinhibition of lettuce seed germination and enhancement of its catabolism by gibberellin. Journal of Experimental Botany. 2004;55:111-118. DOI: 10.1093/jxb/erh023
  155. 155. Bastam N, Baninasab B, Ghobadi C. Improving salt tolerance by exogenous application of salicylic acid in seedlings of pistachio. Plant Growth Regulation. 2013;69:275-284. DOI: 10.1007/s10725-012-9770-7
  156. 156. Mimouni H, Wasti S, Manaa A, Gharbi E, Chalh A, Vandoorne B, et al. Does salicylic acid (SA) improve tolerance to salt stress in plants? A study of SA effects on tomato plant growth, water dynamics, photosynthesis, and biochemical parameters. Omics: a Journal of Integrative Biology. 2016;20:180-190. DOI: 10.1089/omi.2015.0161
  157. 157. Khodary SEA. Effect of salicylic acid on growth, photosynthesis and carbohydrate metabolism in salt stressed maize plants. International Journal of Agriculture and Biology. 2004;6:5-8
  158. 158. Fahad S, Bano A. Effect of salicylic acid on physiological and biochemical characterization of maize grown in saline area. Pakistan Journal of Botany. 2012;44:1433-1438
  159. 159. Azooz MM. Salt stress mitigation by seed priming with salicylic acid in two faba bean genotypes differing in salt tolerance. International Journal of Agriculture and Biology. 2009;11:343-350
  160. 160. Sakhabutdinova AR, Fatkhutdinova DR, Bezrukova MV, Shakirova FM. Salicylic acid prevents the damaging action of stress factors on wheat plants. Bulgarian Journal of Plant Physiology. 2003;29:314-319
  161. 161. El-Tayeb MA. Response of barley grains to the interactive effect of salinity and salicylic acid. Plant Growth Regulation. 2005;45:215-225. DOI: 10.1007/s10725-005-4928-1
  162. 162. Khan NA, Syeed S, Masood A, Nazar R, Iqbal N. Application of salicylic acid increases contents of nutrients and antioxidative metabolism in mungbean and alleviates adverse effects of salinity stress. International Journal of Plant Biology. 2010;1:11-21. DOI: 10.4081/pb.2010.e1
  163. 163. Nazar R, Iqbal N, Syeed S, Khan NA. Salicylic acid alleviates decreases in photosynthesis under salt stress by enhancing nitrogen and sulfur assimilation and antioxidant metabolism differentially in two mungbean cultivars. Journal of Plant Physiology. 2011;168:807-815. DOI: 10.1016/j.jplph.2010.11.001
  164. 164. Gunes A, Inal A, Alpaslan M, Eraslan F, Bagci EG, Cicek N. Salicylic acid induced changes on some physiological parameters symptomatic for oxidative stress and mineral nutrition in maize (Zea mays L.) grown under salinity. Journal of Plant Physiology. 2007;164:728-736. DOI: 10.1016/j.jplph.2005.12.009
  165. 165. Horvath E, Szalai G, Janda T. Induction of abiotic stress tolerance by salicylic acid signaling. Journal of Plant Growth Regulation. 2007;26:290-300. DOI: 10.1007/s00344-007-9017-4
  166. 166. Jayakannan M, Bose J, Babourina O, Rengel Z, Shabala S. Salicylic acid improves salinity tolerance in Arabidopsis by restoring membrane potential and preventing salt-induced K+ loss via a GORK channel. Journal of Experimental Botany. 2013;64:2255-2268. DOI: 10.1093/jxb/ert085
  167. 167. Pieterse CM, Leon-Reyes A, Van Der Ent S, Van Wees SCM. Networking by small molecule hormones in plant immunity. Nature Chemical Biology. 2009;5:308-316
  168. 168. Bhardwaj R, Arora HK, Nagar PK, Thukral AK. Brassinosteroids-a novel group of plant hormones. In: Trivedi PC, editor. Plant Molecular Physiology-Current Scenario and Future Projections. Jaipur, India: Aavishkar Publishers; 2006. pp. 58-84
  169. 169. Yang J, Miao W, Chen J. Roles of jasmonates and brassinosteroids in rice responses to high temperature stress–a review. The Crop Journal. 2021;9:977-985. DOI: 10.1016/j.cj.2021.02.007
  170. 170. Bajguz A, Hayat S. Effects of brassinosteroids on the plant responses to environmental stresses. Plant Physiology and Biochemistry. 2009;47:1-8. DOI: 10.1016/j.plaphy.2008.10.002
  171. 171. Ali Q , Athar HR, Ashraf M. Modulation of growth, photosynthetic capacity and water relations in salt stressed wheat plants by exogenously applied 24-epibrassinolide. Plant Growth Regulation. 2008;56:107-116. DOI: 10.1007/s10725-008-9290-7
  172. 172. Rao AAR, Vardhini BV, Sujatha E, Anuradha S. Brassinosteroids a new class of phytohormones. Current Science. 2002;82:1239-1245
  173. 173. Hayat Q , Hayat S, Irfan M, Ahmad A. Effect of exogenous salicylic acid under changing environment: A review. Environmental and Experimental Botany. 2010;68:14-25. DOI: 10.1016/j.envexpbot.2009.08.005
  174. 174. Krishna P. Brassinosteroid-mediated stress responses. Journal of Plant Growth Regulation. 2003;22:289-297. DOI: 10.1007/s00344-003-0058-z
  175. 175. Anuradha S, Rao SSR. Effect of brassinosteroids on salinity stress induced inhibition of seed germination and seedling growth of rice (Oryza sativa L.). Plant Growth Regulation. 2001;33:151-153. DOI: 10.1023/A:1017590108484
  176. 176. Bakshi A, Shemansky JM, Chang C, Binder BM. History of research on the plant hormone ethylene. Journal of Plant Growth Regulation. 2015;34:809-827. DOI: 10.1007/s00344-015-9522-9
  177. 177. Kwon TH, Park CG, Lee BH, Jeong IH, Lee SE. A new approach: Ethyl formate fumigation to control Bemisia tabaci (Hemiptera: Aleyrodidae) in a yellow melon vinyl house. Applied Sciences. 2022;12:5173. DOI: 10.3390/app12105173
  178. 178. Wang B, Zhang J, Xia X, Zhang WH. Ameliorative effect of brassinosteroid and ethylene on germination of cucumber seeds in the presence of sodium chloride. Plant Growth Regulation. 2011;65:407-413. DOI: 10.1007/s10725-011-9595-9
  179. 179. El-Iklil Y, Karrou M, Benichou M. Salt stress effect on epinasty in relation to ethylene production and water relations in tomato. Agronomie. 2000;20:399-406. DOI: 10.1051/agro:2000136
  180. 180. Khan AA, Akbar M, Seshu DV. Ethylene as an indicator of salt tolerance in rice. Crop Science. 1987;27:1242-1248. DOI: 10.2135/cropsci1987.0011183X002700060031x
  181. 181. Pierik R, Tholen D, Poorter H, Visser EJW, Voesenek LACJ. The janus face of ethylene: Growth inhibition and stimulation. Trends in Plant Science. 2006;11:176-183. DOI: 10.1016/j.tplants.2006.02.006
  182. 182. Cao WH, Liu J, He XJ, Mu RL, Zhou HL, Chen SY, et al. Modulation of ethylene responses affects plant salt-stress responses. Plant Physiology. 2007;143:707-719. DOI: 10.1104/pp.106.094292
  183. 183. Guo H, Ecker JR. The ethylene signaling pathway: New insights. Current Opinion in Plant Biology. 2004;7:40-49. DOI: 10.1016/j.pbi.2003.11.011
  184. 184. Hall AE, Findell JL, Schaller GE, Sisler EC, Bleecker AB. Ethylene perception by the ERS1 protein in arabidopsis. Plant Physiology. 2000;123:1449-1457. DOI: 10.1104/pp.123.4.1449
  185. 185. Fletcher RA, Gill A, Davis TD, Sankhla N. Triazoles as plant growth regulators and stress protectants. Horticultural Reviews. 2000;24:55-138
  186. 186. Karikalan L, Rajan SN, Gopi R, Sujatha BM, Pannersevalam R. Induction of salt tolerance by triadimefon in pigeon pea (Cajanus cajan L.) mill sp. Indian Journal of Experimental Biology. 1999;37:825-829
  187. 187. Hajihashemi S, Kiarostami K, Enteshari S, Saboora A. Effect of paclobutrazol on wheat salt tolerance at pollination stage. Russian Journal of Plant Physiology. 2009;56:251-257. DOI: 10.1134/S1021443709020149
  188. 188. Jaleel CA, Gopi R, Manivannan P, Gomathinayagam M, Murali PV, Panneerselvam R. Soil applied propiconazole alleviates the impact of salinity on Catharanthus roseus by improving antioxidant status. Pesticide Biochemistry and Physiology. 2008;90:135-139. DOI: 10.1016/j.pestbp.2007.11.003

Written By

Imran Khan, Muhammad Umer Chattha, Rizwan Maqbool, Muqarrab Ali, Muhammad Asif, Muhammad Umair Hassan and Muhammad Talha Aslam

Submitted: 03 October 2023 Reviewed: 07 October 2023 Published: 02 February 2024