Open access peer-reviewed chapter

Phytohormones-Assisted Management of Salinity Impacts in Plants

Written By

Naser A. Anjum, Asim Masood, Faisal Rasheed, Palaniswamy Thangavel and Nafees A. Khan

Submitted: 09 October 2023 Reviewed: 13 October 2023 Published: 05 November 2023

DOI: 10.5772/intechopen.113734

From the Edited Volume

Making Plant Life Easier and Productive Under Salinity - Updates and Prospects

Edited by Naser A. Anjum, Asim Masood, Palaniswamy Thangavel and Nafees A. Khan

Chapter metrics overview

58 Chapter Downloads

View Full Metrics

Abstract

The salinity of soils has been significantly limiting crop production in most arid and semi-arid regions of the world. Plant hormones (phytohormones), small molecules with versatile roles in plants can be a sustainable approach for minimizing the major salinity-impacts in plants. Most phytohormones are reported to regulate various signaling cascades interrelated with plant development and stress-resilience and -coping mechanisms. In addition to regulating photosynthesis and related variables, phytohormones also modulate nutrient homeostasis, source-sink capacity, osmoregulation, and antioxidant defense systems in plants under abiotic stresses including soil salinity. Molecular studies have confirmed the coordination between phytohormones and signaling networks, which in turn also maintains ionic homeostasis and plant-salinity tolerance. This chapter aims to appraise the literature available on the role of 10 well-characterized stress response hormones (abscisic acid, ABA; ethylene; salicylic acid, SA; jasmonic acid, JA; and nitric oxide, NO) and also other growth-promoting hormones (such as auxins, gibberellins, GA; cytokinins, CKs; brassinosteroids, BRs; and strigolactones, SLs) in the management of salinity impacts in plants. The discussion outcomes may help in devising and furthering the strategies aimed at sustainably strengthening plant-salinity tolerance.

Keywords

  • abiotic stress
  • soil salinity
  • plant health
  • phytohormones
  • stress tolerance

1. Introduction

Human’s major three requirements, namely food, clothing, and shelter are mainly provided by agricultural activity. Unfortunately, the health and productivity of most agricultural crops are impacted by both biotic (insect pests and disease pathogens) and abiotic (temperature, drought, flooding, salinity, heavy metals, radiation, nutrient deficiency, and excess) stress factors. In turn, these stress factors, in isolation and/or combination, are significantly threatening global food security. In particular, abiotic stresses are responsible for about 51–82% of annual loss in yield of the major food crops in the world agriculture. Interestingly, abiotic stress impacts on agricultural crops and food security are further aggravated by the deteriorating agro-climatic conditions [1, 2, 3, 4]. According to the United Nations Population Division, the world population will reach 9 billion in 2037 and 10 billion in 2058 [5]. It would be far easier to feed the projected human population because most grain crops exhibit about 1% annual yield (much < world’s population growth rate), and direct feeding of people shares only 55% of the world’s crop calories [6, 7]. Among the major environmental challenges, the salinization of soils (soil salinity) has been inducing most land degradation, constituting a primary limit on crop health and productivity, and thus, is threatening agriculture across the world [8].

Advertisement

2. Soil salinity: causes, status, and major impacts

2.1 Causes and status

Soils exhibiting the saturation paste extracts (ECe) in the root zone as electrical conductivity (EC) > 4.0 dS m−1 (≈40 mM NaCl) at 25°C and 15% exchangeable Na+ ion are considered saline. There can be five major classes of soil salinity: non-saline (EC = 0–2 dS m−1; low saline (EC = 2–4 dS m−1; moderately saline (EC = 4–8 dS m−1; highly saline (EC = 8–16 dS m−1; and extremely saline (EC = ≥ 16 dS m−1) (Figure 1) [9, 10, 11, 12]. Most saline soils exhibit Na+ as an anion and Cl as a cation [13]. However, Na+, Ca2+, and Mg2+ are the major cations component of total soluble salts in soils, whereas Cl, SO42−, and carbonates (CO32− and HCO3) are the major anions in total soluble salts in soils. Despite the aforementioned fact, most studies aimed at exploring plant-salinity responses and tolerance have considered Na+ and Cl; and have largely ignored other cations (Ca2+ and Mg2+) and/or anions (SO42−, CO32−, and HCO3) [14]. Both Na+ and Cl ions are the most widespread causes of soil salinity, where Cl is more dangerous than Na+ since it is responsible for many physiological disorders in plants [15].

Figure 1.

Schematic representation of the major classes of soil salinity [9, 10]. dSm−1, decisiemens per meter.

Notably, the soil salinity may be developed as natural (geological, hydrological, and pedological processes) or induced by human activity (human-caused factors). The long-term natural accumulation of salts (including Cl of Na+, Ca2+, and Mg2+and sometimes SO42− and CO32−) in the soil or surface water contributes to the primary or natural salinity. On the other hand, the disruption of the hydrologic balance of the soil between water applied (irrigation or rainfall) and water used by crops (transpiration), as a result of anthropogenic activities, causes secondary soil salinity [16]. In fact, the secondary salinization (rising NaCl levels in groundwater) impacts irrigated land and eventually leads to the loss of agricultural soils [17]. About 1125 million hectares worldwide have already been impacted by the soil salinization [18]. Huge annual global loss in crop production (≈ US$27.3 billion) has been reported in saline soils in irrigated areas, representing about 20% of the total salinity-affected soils, mainly in North America, Oceania, and the Middle East. Interestingly, salinity may impact half of the world’s irrigated land by 2050 [19, 20, 21, 22].

2.2 Major impacts

In terms of the tolerance to salinity levels, plants are grouped into two categories, namely halophytes (salinity-tolerant and adapted to salinized environments) and glycophytes (salinity-sensitive plants). Unfortunately, most agricultural crop plants are glycophytes (salt-sensitive), where soil salinization adversely inhibits growth, metabolism, development, and productivity (yield). Thus, the salinity-sensitivity of most crop plants, increasing rate of land salinization, and salinity-caused serious loss in crop productivity are challenging food security. Soil salinity significantly limits the growth, metabolism, seed germination, flowering, fruiting, and productivity (crop yields) in most crop plants, mainly as a result of osmotic stress and ionic stress [3, 23].

2.2.1 Osmotic stress and ionic stress

High soil salinity is bound to cause osmotic stress, followed by ionic stress. These two high soil salinity-caused direct stresses (ionic and osmotic) induce secondary stresses such as oxidative stress. Increased Na+ ions in the soil mainly cause osmotic stress. Equilibrium in ion homeostasis is severely disturbed, which includes an increase in the Na+ levels, K+ efflux, K+ leakage, K+ deficiency in the cytosol, replacement of Ca2+ with Na+, and eventual impaired Na+/K+ ratio, nutritional disbalance, and disrupted enzyme activity [11, 12, 15]. Significantly reduced water absorption, decreased osmotic potential, closure of the stomata, inhibited CO2-influx, and impaired downstream processes were reported as a result of osmotic stress in plants [22, 23, 24].

2.2.2 Oxidative stress and antioxidant metabolism

Elevated salinity-induced ionic and osmotic stresses are bound to cause oxidative stress, which is a physiological condition of elevated cellular generation of reactive oxygen species (ROS) and or diminished scavenging (antioxidant-mediated) of ROS leading to impaired cellular redox homeostasis. In turn, elevation in the generation of ROS and impaired ROS metabolism/scavenging has been widely reported in plants under soil salinization. The list of major ROS includes O2•−, •OH, H2O2, and 1O2 are produced in different including chloroplast, mitochondria, apoplast, and peroxisome, under normal conditions of stress such as salinity. If not metabolized and/or scavenged, most elevated cellular ROS may lead to cytotoxicity and severe damage to biological molecules such as proteins, lipids, and nucleic acids. In order to avoid the harmful effects of oxidative stress, plants have triggered an antioxidant defense system based on two types of machinery [12, 25]. Fortunately, plants are endowed with an antioxidant defense system comprising (enzymes and non-enzymes/antioxidant metabolites), which in isolation and/or combination tend to scavenge/metabolize most ROS and thereby avert elevated ROS-caused consequences [26, 27]. Major enzymatic antioxidants comprise superoxide dismutase (SOD), catalase (CAT), guaiacol peroxidase (GPX), glutathione sulfotransferase (GST), ascorbate peroxidase (APX), monodehydroascorbate reductase (MDHAR), dehydroascorbate reductase (DHAR) and glutathione reductase (GR). On the other hand, ascorbate (AsA), glutathione (GSH), carotenoids, tocopherols, and phenolics are among the major non-enzymatic antioxidants in the plant stress defense system [27, 28, 29, 30, 31, 32, 33, 34].

Given the rising demands on crop yield for keeping up with the burgeoning human population, strategies for sustainably maintaining an optimum crop plant health under increasing soil salinization and climate change and the development of salt-tolerant crop varieties are being exhaustively explored.

Advertisement

3. Phytohormones and plant-salinity tolerance

Phytohormones stand second to none among the regulatory compounds and chemical messengers in terms of their importance in almost every aspect of plant life. Although phytohormones are small molecules, they are widely known to play key roles in plant growth, development, and various plant physiological processes. These also regulate internal and external stimuli, stress-involved signal transduction pathways, oxidative stress-scavenging, and stress tolerance in plants [35, 36, 37]. Plants under salinity stress usually tend to accumulate a range of osmolytes (osmoprotectants/compatible solutes), namely proline, glycine betaine, polyamines, and sugars. Interestingly, the cellular levels of most of these osmolytes are significantly modulated by varied phytohormones [38].

Employing the protocols, based mainly on the use of various phytohormones, may help sustainably improve optimum growth, metabolism, photosynthesis and productivity (yield) under salinity-affected soils, thereby minimizing increasing strain on global food security. Notably, the list of well-characterized stress response hormones includes abscisic acid (ABA), ethylene, salicylic acid (SA), and jasmonic acid (JA). In contrast, phytohormones classified as growth promotion hormones are auxin, gibberellin (GA), cytokinins (CKs), brassinosteroids (BRs), nitric oxide (NO), and strigolactones (SLs) [39, 40].

Apart from presenting a brief overview, the following sections attempt to enlighten the major roles (and the basic mechanisms involved) of ABA, auxins, BRs, CKs, ethylene, GAs, JA, NO, SA, and SLs in plant-salinity tolerance (Figure 2).

Figure 2.

Schematic representation of the typical structures of phytohormones discussed in the chapter.

3.1 Abscisic acid

Chemically, a sesquiterpenoid, 15-C compound, abscisic acid (ABA) is a naturally occurring and enigmatic stress phytohormone widely known to play key roles in plant growth and development. Important processes including leaf abscission, seed dormancy, embryo morphogenesis, stomatal opening and cell turgor maintenance are known to varyingly involve ABA [37, 41, 42, 43]. ABA has also been argued as a modulator of plant’s adaptive stress responses via integrating various stress signals, controlling downstream responses, biosynthesizing dehydrins, osmolytes, and protective proteins, and regulating protein-encoding genes [43, 44, 45]. ABA-supply helped pepper (Capsicum annuum) seeds to exhibit high germination percentage, radicle emergence, and cotyledon expansion of seeds under NaCl stress mainly as a result of low expression in seeds of ABA signaling components such as CaABI, CaPYL2, CaPYL4, CaSnRK2.3, and CaSnRK2.6 [46]. ABA-nitrogen coordination alleviated salinity-inhibited photosynthetic potential in mustard (Brassica juncea) by improving proline accumulation and antioxidant activity [47]. Major physiological mechanisms underlying ABA-induced salinity tolerance in plants may also involve enhanced activity of antioxidant enzymes (CAT, APX, peroxidase, and POD) and the contents of antioxidant non-enzyme/metabolites (AsA and GSH) [48]. Additionally, significant reductions in Na+ content, increased contents of K+, Mg2+, and Ca2+; and that of hormones such as 1-aminocyclopropane carboxylic acid, trans-zeatin, N6-isopentenyladenosine, indole-3-acetic acid (IAA), and ABA were also observed in ABA-supplied plants under salinity stress [48]. Involvement of ABA signaling in reduction of transpiration flow, regulation of Na+ ion homeostasis and antioxidant enzyme activities was reported to induce salinity tolerance in wheat (Triticum aestivum) seedlings [49]. Earlier, ABA-mediated improved Indica rice (Oryza sativa)-tolerance to salinity stress involved the calmodulin signaling cascade and the ABA-mediated induction of OsP5CR gene expression in osmolyte (proline) accumulation [50].

3.2 Auxins

A chemical messenger involved in the light and gravity-stimulated shoot-to-root transport of a ‘growth stimulus’ was argued to be auxin, whose chemical nature was later identified as IAA [51, 52]. Chemically similar to the amino acid tryptophan, IAA is the representative and most studied auxin in plants [51, 52, 53, 54]. Auxins are mainly involved in cell division, cell elongation, and cell differentiation [45]. However, auxins can also regulate plant abiotic stress responses, where its homeostasis, distribution, and metabolism can be modulated by most abiotic stress factors [35, 45, 52]. Auxins can mediate the root growth plasticity in response to salinity stress [55, 56]. Earlier, a significant remodeling of root architecture was reported under high salinity, which was argued due to salinity-led altered auxin-accumulation and -redistribution [57, 58]. Pre- or post-treatment of seeds with IAA significantly alleviated salinity impacts and improved seed germination and early seedling establishments of T. aestivum under salinity stress [59]. Hence, an optimum concentration and timely exogenous application of auxins would be a promising approach for countering the salinity stress impacts in crop plants [36].

3.3 Brassinosteroids

Considered ubiquitous in the plant kingdom, polyhydroxy steroidal phytohormones, namely brassinosteroids (BRs), promote growth, seed germination, rhizogenesis, and senescence in plants, as well as their stress-tolerance capacity. Notably, among so far identified 60 BRs-related compounds, the list of bioactive BRs includes only three: brassinolide (BL), 28-homobrassinolide (28-HomoBL), and 24-epibrassinolide (24-EpiBL) [60, 61, 62]. Extensive reports are available on the role of BRs in salinity-impact control in different test plants [62, 63, 64, 65, 66, 67].

In several salinity-exposed test plants, 28-homoBL detoxified the NaCl-caused stress by elevating the activities of antioxidative enzymes including (SOD, CAT, GR, APX, and GPX) [68, 69]. Seed priming with BL can improve seed germination and seedling growth by significantly increasing POD, SOD, and CAT activity under salt stress [70]. The supply of polyhydroxylated spirostanic brassinosteroid analog (BB-16) can also enhance the activity of CAT, SOD, and GR and thereby mitigate the salinity impacts in plants [71]. 24-EpiBL-mediated improved tolerance of different test plants to varying salinity levels involved a 24-EpiBL-mediated decrease in oxidative stress via induction in the activity of ROS-metabolizing enzymatic antioxidants including APX, CAT, and POD [72, 73, 74, 75, 76, 77]. Significant decreases in the cellular levels of electrolyte leakage, O2•− production, MDA, H2O2, and improved growth, carbonic anhydrase activity, photosynthetic efficiency in epiBL-supplied salinity-treated test plants were corroborated with enhanced activity of SOD, POD, GPX, CAT and APX enzymes and the improved contents of AsA and GSH [78, 79]. Interestingly, BRs have been extensively reported to regulate plant-salinity tolerance via interacting with a number of plant hormones including auxins [80, 81], ethylene [63, 65], ABA [82, 83, 84, 85], and NO [64, 79, 82, 86, 87]. Moreover, BR signaling components can be directly regulated by salt stress signals at both transcriptional and post-translational levels [84, 85, 88, 89, 90].

3.4 Cytokinins

Cytokinins (CKs) are the derivatives of adenine (such as zeatin, kinetin, and N6-benzyladenine, BA) or of phenylurea (such as diphenylurea and thidiazuron) [45, 91]. Notably, the first naturally occurring CK was zeatin, which was identified and purified from immature maize (Zea mays) kernels [92]. Kinetin was the first CK discovered as an adenine (aminopurine) derivative [92]. CKs mainly regulate the major plant growth and developmental processes [93, 94]. However, literature also supports the immense roles (and underlying mechanisms) of CKs in plant abiotic stress tolerance [95, 96, 97]. Notably, the genetic engineering of CKs-metabolism was argued as one of the prospective ways to improve agricultural traits of crop plants [98]. CK signaling-mediated promotion in salt tolerance in Z. mays was argued to involve CK-mediated modulation of shoot Cl exclusion [99]. Soaking Z. mays seeds in zeatin-type cytokinin biostimulators (namely cis-zeatin-type CKs, c-Z-Ck; trans-zeatin, t-Z-Ck isomers) was reported to enhance antioxidant system and photosynthetic efficiency and thereby improve Z. mays salt tolerance [100]. Notable contradictory results yielded in several studies on the functional analyses of CK receptor mutants and the involvement of CK in ABA-mediated stress signaling in plants under osmotic/salinity stress warrant further molecular-genetic clarifications regarding the role of CKs in plant osmotic/salinity stress tolerance [101, 102, 103].

3.5 Ethylene

Ethylene is an important signaling molecule and a gaseous phytohormone. Its coordination with downstream signaling components has been reported to help plants in varyingly tolerating salinity stress [104, 105, 106]. Induction of ethylene generation in salinity-exposed plants is indicative of its significance as a downstream signal and modulation of gene expression [104]. Ethylene-homeostasis and ethylene signaling have been argued as an important factor required for plant-salinity tolerance [107, 108]. The maintenance of cellular ethylene (via endogenous production-induced accumulation and/or by exogenously supplied of ethylene precursor, 1-aminocyclopropane-1-carboxylic acid) has been reported to enhance Na+ and K+ homeostasis and induce downstream signaling for ROS-homeostasis; and eventually to improve plant salt tolerance [109, 110, 111]. Earlier, ethylene-mediated improvement in Arabidopsis salt tolerance mainly involved enhanced retention of K+ in shoots and roots rather than decrease in tissue Na+ content [112]. Ethylene (or its biochemical precursor, 1-aminocyclopropane-1-carboxylic acid) supplies improved plant tolerance to high salinity [110, 113, 114, 115]. Ethylene can trigger plant salt tolerance by modulating polyamine catabolism enzymes associated with H2O2 production [116]. S-nitrosylation of ACO homolog 4 (1-aminocyclopropane-1-carboxylate oxidase homolog 4; ACOh4) improved ethylene synthesis and improved salt tolerance in salinity-exposed tomato plants [117]. The roles of S-adenosylmethionine (SAM, involved in ethylene biosynthesis) and its derivatives in plant salt tolerance have also been recently discussed [118].

Molecular studies have unveiled the ‘MdNAC047-ETHYLENE RESPONSE FACTOR (MdERF3)-ethylene-salt tolerance’ regulatory pathway in apple [108]. Apple MdERF4 was reported to negatively regulate salt tolerance by inhibiting MdERF3 transcription [119]. MdMYB46 enhanced salt (and osmotic) stress tolerance in apple by directly activating stress-responsive signals [120]. Additionally, plant responses to salt stress may also involve EIN3/EIL1-dependent genes and other ROS scavenger-coding genes [110]. The outcomes of crosstalk between miR319 and ethylene contribute to plant-salinity tolerance. To this end, overexpression of Osa-MIR319b and targeting mimicry form of miR319 (MIM319) confirmed the role of miR319-mediated positive regulation of ethylene synthesis, and eventually improved salinity tolerance in switchgrass (Panicum virgatum) [121]. However, negative roles of ethylene have also been reported in salinity-exposed plants, where enhanced ethylene levels did not help plants in counteracting salinity stress impacts [122, 123, 124]. Thus, the reported few ambiguous roles of ethylene in plant-salinity stress responses require further explanations.

3.6 Gibberellins

Gibberellins (gibberellic acid, GA), a large family of tetracyclic di-terpenoid compounds, are classical plant hormones denoted largely by ‘gibberellin numbers’ (GAn) in order of discovery, such as GA1, GA2, …, GAn. In general, GAs are involved in growth and development [125, 126]. However, the literature is full on the involvement of GA in plant tolerance to a number of abiotic stresses [96, 126, 127]. In salinity-exposed barley, exogenous GA3 increased the shoot and root length of germinated barley seeds; significantly reduced ion-leakage, osmolyte (proline) accumulation; and thereby rescued the expression of the HvABI5, HvABA7, and HvKO1 by 3, 10, and 33 fold, respectively [128]. Exogenously applied GA enhanced growth and salinity stress tolerance in Z. mays by modulating the morpho-physiological, biochemical, and molecular attributes [129]. Moreover, GA3-supply improved pigment content, plant growth, and development, reduced Na+ concentration in shoots and roots, increased the water absorption and metabolic activities in seeds, uplifted the seed dormancy, modulated cell division, and cell elongation. In this way, GA3-supply increased the growth of root, shoot, and number of leaves; increased photosynthetic activities and the dry matter production; maintained a fine-tuning among AsA-GSH cycle components; improved the plant height, yield, and yield-related traits, Ca2+ and K+ concentrations, and transpiration rates; and decreased Na+ concentrations in different test plants under salinization [129, 130, 131, 132, 133].

As reported in plants under most abiotic stresses, a higher accumulation of DELLA proteins was reported in salinity-exposed plants, which in turn was argued to restrain growth and enhance stress tolerance through reducing GA signaling activity [134, 135]. Seed priming with GA was reported to induce high salinity tolerance in Pisum sativum, where the applied GA modulated antioxidants, secondary metabolites, and upregulated antiporter genes [136]. Moreover, pre-treating/soaking of seeds with GAs was widely evidenced to improve increased α-amylase; salinity-caused nutritional disorders; decreased Na+ content; enhanced ion uptake, photosynthesis, and redox homeostasis; improved coordination among CAT, APX, and SOD, as an adaptive mechanism to salt stress [137, 138, 139, 140].

3.7 Jasmonic acid

Important critical signaling molecule jasmonic acid (JA) is among the most abundant members of the jasmonate class of plant hormones. Derived from linolenic acid (as cyclopentanone) and lipids, JA is known to regulate plant growth, development, and stress responses [45, 141]. Extensive reports are available on the role (and underlying mechanisms) of JA-mediated plant-salinity tolerance. Methyl jasmonic acid supply shifted the endogenous fatty acid levels and supported O. sativa growth in saline soil [142]. Transcriptomic analysis has revealed methyl jasmonate-mediated salt tolerance in alfalfa (Medicago sativa) as a result of antioxidant activity regulation and ion homeostasis [99]. JA-supply mediated mitigation of the inhibitory effect of salt stress in T. aestivum by increasing the endogenous levels of CK and IAA, reducing ABA contents, increasing α-tocopherol, phenolics, and flavonoids levels, and triggering SOD and APX activity [143]. In salinity-exposed Anchusa italica, methyl jasmonate improved test plant-salinity tolerance by enhancing contents of photosynthetic pigment, soluble sugars, K+ and Ca2+, declining Na+ content, and eventually improving the major growth attributes [144]. Salinity stress-mediated induction in endogenous JA is also known in plants [145, 146]. Interaction outcomes of JA with ABA can also improve plant-salinity tolerance [141, 147]. JA-mediated saline stress tolerance in O. sativa involved autophagy and programmed cell death as critical pathways [148]. Jasmonate biosynthesis gene OsOPR7 was involved in the mitigation of salinity-induced mitochondrial oxidative stress [149]. Conferment of a greater cell elongation under salt stress was achieved with mutations of JA-receptor CORONATINE INSENSITIVE1 (COI1) and MYC2/3/4 along with the stabilized JASMONATE ZIM mutant jaz3-1 [150].

3.8 Nitric oxide

A highly versatile gaseous, free-radical, redox-signaling molecule, nitric oxide (NO), has been widely reported to perform diverse functions in plants [15, 151]. In a wide range of studies on salinity-exposed plants, exogenous supply of NO (or sodium nitroprusside, SNP; a NO donor) improved seed vigor, germination, and plant health and productivity through alleviating oxidative damage as a result of decreased levels of electrolyte leakage, MDA, and H2O2, improving antioxidant defense mechanism, decreasing methylglyoxal toxicity, and upregulating the glyoxalase system, adjusting the levels of osmolytes, and maintaining ionic balance [16, 152, 153, 154]. NO-supply can also reverse the glucose-mediated photosynthetic repression in plants under salinity exposure [154]. In NO (0.1 mM SNP)-mediated T. aestivum seed priming (for 20 h) helped improve germination rate, the weight of radical and coleoptile, and K+ ion and Na+ ion homeostasis [155, 156]. Exogenous application of NO (or its donor, SNP) can increase leaf area, plant dry mass, and the lengths of shoot and root in NaCl-stressed plants [157, 158]. Additionally, the maintenance of ion homeostasis (via enhanced K+ uptake and reduced Na+ uptake) and the modulation of the Na+/H+ antiporter enzyme were also reported in salinity-exposed and NO (or its donor, SNP)-supplemented test plants [159]. NO-mediated high salt tolerance in plants may also involve a NO-accrued increase in H+-ATPase activity and eventual reduced leakage-mediated maintenance of high cytosolic K+/Na+ ratio [156, 160]. NO-mediated improved defense against salt-induced stress also involves NO’s interaction with signaling molecules [15]. In earlier studies, NO (or its donor, SNP)-supply mediated improvements in the photosynthetic capacity involved NO-mediated protection of photosynthetic pigments; maintenance of normal shape of thylakoids and increase in chloroplast size; enhancement in the quenching of additional energy and quantum-yield of photosystem II; increase in ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO) activity; induction in the influx and efflux of Ca2+; regulation of stomatal behavior and guard cells-ABA concentration; and efficient energy dissipation [158, 161, 162, 163].

The asA-GSH cycle is a central modulator of the plant stress responses and defense, ROS metabolism, and cellular redox balance [26, 164]. A plethora of reports supports the role of NO in the maintenance of the cellular redox balance via regulation of AsA-GSH cycle components (enzymatic and non-enzymatic) in salinity-impacted plants [123, 165, 166]. In different salinity-exposed plants, NO (or its donor, SNP)-supply resulted in significantly decreased cellular O2•− generation, H2O2, MDA content, and electrolyte leakage via maintaining a fine-tuning among antioxidant enzymes (including SOD, CAT, APX a H2O2-scavenging enzyme, MDHAR, DHAR, GR, GST, GPX, and CAT) and non-enzymatic antioxidants (including GSH and AsA) [15, 16, 151, 167].

3.9 Salicylic acid

Salicylic acid (SA), a phenolic plant hormone, is widely known for its involvement in plant growth and development and modulation of plant stress responses. Exogenous SA supply has strengthened salinity stress-tolerance mechanisms in extensive studies [168, 169]. Reports are also available on a high bio-stimulatory capacity of SA for salinity tolerance involving positive regulation of the AsA-GSH cycle, elevated accumulation of osmoprotectors, antioxidant enzyme activation, and increasing tolerance under ion toxicity and oxidative stress [34, 170, 171, 172]. In salinity-exposed Vigna radiata, the role of SA-induced accumulation of glycine betaine protected photosynthesis and growth against NaCl-accrued impacts in V. radiata as a result of the minimized accumulation of Na+ and Cl ions and oxidative stress and maintained high GSH level and eventually reduced cellular redox environment [25]. In many instances, SA-mediated plant-salinity tolerance has SA-dose dependency [25, 40, 173, 174]. In salinity (100 mM NaCl)-exposed pepper (Capsicum annuum) plants, an exogenous supply of SA (0.5 mM) reduced leaf Na+ content and oxidative stress-related traits [171]. SA-supply mediated regulation of ROS-metabolism and AsA-GSH cycle has also been reported in plants under salinity stress [175]. In earlier studies, supplied SA-assisted mitigation of salinity stress impacts in plants involved characteristic changes in the expression pattern of GST-gene family members such as SlGSTT2, SlGSTT3, and SlGSTF4 [176]; enhanced transcript level of antioxidant genes; GPX1, GPX2, DHAR, GR, GST1, GST2, MDHAR, and GS [177], and GORK channel-mediated control of K+ loss [178].

3.10 Strigolactones (SLs)

Synthesized in several plant species, strigolactones (SLs) are multifunctional β-carotene derivative molecules and are considered as an essential plant hormone in regulating plant functions. SLs have been considered an emerging growth regulator for developing resilience in plants [37, 41, 179]. In salinity-treated rapeseed (Brassica napus), the supplied SL increased plant growth, photosynthetic traits, and antioxidant enzyme activity [180]. Involvement of SL signal transduction and SL-biosynthetic mutants in more axillary growth3 (max3) and max4 was reported in SLs-mediated positive regulation of plant salt tolerance [181, 182]. In many instances, SL-mediated plant salt tolerance involved ABA accumulation [40, 183, 184]. In some studies on salinity-exposed plants, exogenous SLs improved ROS metabolism and decreased lipid peroxidation and cellular damage [185, 186]. The role of arbuscular mycorrhizal fungi and ABA in SL-assisted improvement in plant-salinity tolerance has also been reported [183, 184].

Advertisement

4. Conclusions and prospects

The salinity of soils is significantly limiting crop production in most arid and semi-arid regions of the world. Owing to their sensitivity to salinity stress, most agricultural crop plants exhibit soil salinization-accrued inhibition in growth, metabolism, development, and productivity (yield). Thus, salinity-caused serious impacts on crop health and productivity are challenging food security. Though very small molecules are produced in small quantities, phytohormones have versatile roles in plants as the major regulator of various signaling cascades interrelated with plant development, stress resilience, and coping mechanisms. Apart from presenting a brief overview, this chapter attempted to enlighten the major roles (and the basic mechanisms involved) of selected 10 phytohormones (ABA, auxins, BRs, CKs, ethylene, GAs, JA, NO, SA, and SLs) in plant-salinity tolerance. An optimum concentration and timely exogenous application of these phytohormones would be a promising approach for countering the salinity stress impacts in crop plants. There are still challenges to understanding how ABA, auxins, BRs, CKs, ethylene, GAs, JA, NO, SA, and SLs and associated close molecules can function at the molecular level and how the intimate mechanisms of interaction among these phytohormones and also with other emerging signaling molecules work. Additionally, a few contradictory results related to the involvement of CK in ABA-mediated stress signaling and the reported few ambiguous roles of ethylene in plant-salinity stress responses under osmotic/salinity stress warrant further molecular-genetic clarifications.

References

  1. 1. Tigchelaar M, Battisti DS, Naylor RL, Ray DK. Future warming increases probability of globally synchronized maize production shocks. Proceedings of the National Academy of Sciences, USA. 2018;115:644-6649
  2. 2. Vaughan MM, Block A, Christensen SA, Allen LH, Schmelz EA. The effects of climate change associated abiotic stresses on maize phytochemical defenses. Phytochemistry Reviews. 2018;17:37-49
  3. 3. Zörb C, Geilfus CM, Dietz KJ. Salinity and crop yield. Plant Biology. 2019;21:31-38
  4. 4. Intergovernmental Panel on Climate Change (IPCC). Climate Change 2022. Impacts, adaptation and vulnerability. In: Pörtner HO, Roberts DC, Tignor M, et al., editors. Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. UK and New York: Cambridge University Press; 2022. DOI: 10.1017/9781009325844
  5. 5. Zeifman L, Hertog S, Kantorova V, Wilmoth J. A World of 8 Billion. Policy Brief No. 140, Population Division, UN Department of Economic and Social Affairs. 2022. pp. 1-4. Available from: https://www.un.org/development/desa/dpad/wp-content/uploads/sites/45/publication/PB_140.pdf
  6. 6. Cassidy ES, West PC, Gerber JS, Foley JA. Redefining agricultural yields: From tonnes to people nourished per hectare. Environmental Research Letters. 2013;8(3):034015
  7. 7. Hemathilake DM, Gunathilake DM. Agricultural productivity and food supply to meet increased demands. In: Future Foods. Cambridge, Massachusetts, United States: Academic Press; 2022. pp. 539-553
  8. 8. Satir O, Berberoglu S. Crop yield prediction under soil salinity using satellite derived vegetation indices. Field Crops Research. 2016;192:134-143
  9. 9. Wicke B, Smeets E, Dornburg V, Vashev B, Gaiser T, Turkenburg W, et al. The global technical and economic potential of bioenergy from salt-affected soils. Energy and Environmental Science. 2011;4:26692681
  10. 10. Ayub MA, Ahmad HR, Ali M, Rizwan M, Ali S, Rehman MZ, et al. Salinity and its tolerance strategies in plants. In: Plant Life under Changing Environment. Cambridge, Massachusetts, United States: Academic Press; 2020. pp. 47-76
  11. 11. Rasheed F, Anjum NA, Masood A, Sofo A, Khan NA. The key roles of salicylic acid and sulfur in plant salinity stress tolerance. Journal of Plant Growth Regulation. 2022;41:1891-1904
  12. 12. Rasheed F, Sehar Z, Fatma M, Iqbal N, Masood A, Anjum NA, et al. Involvement of ethylene in reversal of salt stress by salicylic acid in the presence of sulfur in mustard (Brassica juncea L.). Journal of Plant Growth Regulation. 2022;41:3449-3466
  13. 13. Rengasamy P. Soil processes affecting crop production in salt-affected soils. Functional Plant Biology. 2010;37:613-620
  14. 14. Ludwiczak A, Osiak M, Cárdenas-Pérez S, Lubińska-Mielińska S, Piernik A. Osmotic stress or ionic composition: Which affects the early growth of crop species more? Agronomy. 2021;11(3):435
  15. 15. Hasanuzzaman M, Oku H, Nahar K, Bhuyan MB, Mahmud JA, Baluska F, et al. Nitric oxide-induced salt stress tolerance in plants: ROS metabolism, signaling, and molecular interactions. Plant Biotechnology Reports. 2018;12:77-92
  16. 16. Hasanuzzaman M, Hossain MA, Fujita M. Nitric oxide modulates antioxidant defense and methylglyoxal detoxification system and reduces salinity induced damage in wheat seedling. Plant Biotechnology Reports. 2011;5:353-365
  17. 17. Food and Agricultural Organization (FAO). Global Soil Resources. 2015. Available from: https://hal.archives-ouvertes.fr/hal-01241064/
  18. 18. Hossain MS. Present scenario of global salt affected soils, its management and importance of salinity research. International Research Journal of Biological Sciences. 2019;1:1-3
  19. 19. Qadir M, Quillerou E, Nangia V, Murtaza G, Singh M, Thomas RJ, et al. Economics of salt-induced land degradation and restoration. Natural Resources Forum. 2014;38:282-295
  20. 20. Shabala S, Wu H, Bose J. Salt stress sensing and early signalling events in plant roots: Current knowledge and hypothesis. Plant Science. 2015;241:109-119
  21. 21. Singh AK, Gupta KJ, Singla-Pareek SL, Foyer CH, Pareek A. Raising crops for dry and saline lands: Challenges and the way forward. Physiologia Plantarum. 2022;174:1-4
  22. 22. Wang CF, Han GL, Yang ZR, Li YX, Wang BS. Plant salinity sensors: Current understanding and future directions. Frontiers in Plant Science. 2022;13:859224
  23. 23. Awlia M, Alshareef N, Saber N, Korte A, Oakey H, Panzarová K, et al. Genetic mapping of the early responses to salt stress in Arabidopsis thaliana. The Plant Journal. 2021;107(2):544-563
  24. 24. Yang Y, Guo Y. Unraveling salt stress signaling in plants. Journal of Integrative Plant Biology. 2018;60:796-804
  25. 25. Syeed S, Sehar Z, Masood A, Anjum NA, Khan NA. Control of elevated ion accumulation, oxidative stress, and lipid peroxidation with salicylic acid-induced accumulation of glycine betaine in salinity-exposed Vigna radiata L. Applied Biochemistry and Biotechnology. 2021;193(10):3301-3320
  26. 26. Anjum NA, Umar S, Ahmad A. Oxidative Stress in Plants: Causes, Consequences and Tolerance. New Delhi: IK International Publishing House; 2012
  27. 27. Anjum NA, Sofo A, Scopa A, Roychoudhury A, Gill SS, Iqbal M, et al. Lipids and proteins - major targets of oxidative modifications in abiotic stressed plants. Environmental Science and Pollution Research. 2015;22:4099-4121
  28. 28. Mittler R. Oxidative stress, antioxidants and stress tolerance. Trends in Plant Science. 2002;7:405-410
  29. 29. Gill SS, Tuteja N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant Physiology and Biochemistry. 2010;48:909-930
  30. 30. Anjum NA, Gill SS, Gill R, Hasanuzzaman M, Duarte AC, Pereira E, et al. Metal/metalloid stress tolerance in plants: Role of ascorbate, its redox couple, and associated enzymes. Protoplasma. 2014;251:1265-1283
  31. 31. Anjum NA, Sharma P, Gill SS, Hasanuzzaman M, Khan EA, Kachhap K, et al. Catalase and ascorbate peroxidase-representative H2O2-detoxifying heme enzymes in plants. Environmental Science and Pollution Research. 2016;23:19002-19029
  32. 32. Anjum NA, Gill SS, Corpas FJ, Ortega-Villasante C, Hernandez LE, Tuteja N, et al. Editorial: Recent insights into the double role of hydrogen peroxide in plants. Frontiers in Plant Science. 2022;13:843274
  33. 33. Gill SS, Anjum NA, Gill R, Yadav S, Hasanuzzaman M, Fujita M, et al. Superoxide dismutase – Mentor of abiotic stress tolerance in crop plants. Environmental Science and Pollution Research. 2015;22:10375-10394
  34. 34. Hasanuzzaman M, Bhuyan MB, Anee TI, Parvin K, Nahar K, Mahmud JA, et al. Regulation of ascorbate-glutathione pathway in mitigating oxidative damage in plants under abiotic stress. Antioxidants. 2019;8(9):384
  35. 35. Khan NA, Nazar R, Iqbal N, Anjum NA. Phytohormones and Abiotic Stress Tolerance in Plants. Berlin, Heidelberg, Germany: Springer-Verlag; 2012
  36. 36. Ryu H, Cho YG. Plant hormones in salt stress tolerance. Journal of Plant Biology. 2015;58:147-155
  37. 37. El-Esawi MA. Introductory chapter: Hormonal regulation in plant development and stress tolerance. In: Phytohormones-Signaling Mechanisms and Crosstalk in Plant Development and Stress Responses. London, UK: IntechOpen; 2017. pp. 1-9
  38. 38. Sharma A, Shahzad B, Kumar V, Kohli SK, Sidhu GP, Bali AS, et al. Phytohormones regulate accumulation of osmolytes under abiotic stress. Biomolecules. 2019;9(7):285
  39. 39. Verma V, Ravindran P, Kumar PP. Plant hormone-mediated regulation of stress responses. BMC Plant Biology. 2016;16:86
  40. 40. Yu Z, Duan X, Luo L, Dai S, Ding Z, Xia G. How plant hormones mediate salt stress responses. Trends in Plant Science. 2020;25(11):1117-1130
  41. 41. Wani SH, Kumar V, Shriram V, Sah SK. Phytohormones and their metabolic engineering for abiotic stress tolerance in crop plants. The Crop Journal. 2016;4:162-176
  42. 42. Sreenivasulu N, Harshavardhan VT, Govind G, Seiler C, Kohli A. Contrapuntal role of ABA: Does it mediate stress tolerance or plant growth retardation under long-term drought stress? Gene. 2012;506:265-273
  43. 43. Sah SK, Reddy KR, Li J. Abscisic acid and abiotic stress tolerance in crop plants. Frontiers in Plant Science. 2016;7:571
  44. 44. Cutler SR, Rodriguez PL, Finkelstein RR, Abrams SR. Abscisic acid: Emergence of a core signaling network. Annual Review of Plant Biology. 2010;61:651-679
  45. 45. Davies PJ. Plant Hormones: Biosynthesis, Signal Transduction, Action! 3rd ed. Dordrecht: Springer; 2010
  46. 46. Ruggiero A, Landi S, Punzo P, Possenti M, Van Oosten MJ, Costa A, et al. Salinity and ABA seed responses in pepper: Expression and interaction of ABA core signaling components. Frontiers in Plant Science. 2019;10:304
  47. 47. Majid A, Rather BA, Masood A, Sehar Z, Anjum NA, Khan NA. Abscisic acid in coordination with nitrogen alleviates salinity-inhibited photosynthetic potential in mustard by improving proline accumulation and antioxidant activity. Stress. 2021;1(3):162-180
  48. 48. Chen G, Zheng D, Feng N, Zhou H, Mu D, Zhao L, et al. Physiological mechanisms of ABA-induced salinity tolerance in leaves and roots of rice. Scientific Reports. 2022;12(1):8228
  49. 49. Parveen A, Ahmar S, Kamran M, Malik Z, Ali A, Riaz M, et al. Abscisic acid signaling reduced transpiration flow, regulated Na+ ion homeostasis and antioxidant enzyme activities to induce salinity tolerance in wheat (Triticum aestivum L.) seedlings. Environmental Technology and Innovation. 2021;24:101808
  50. 50. Sripinyowanich S, Klomsakul P, Boonburapong B, Bangyeekhun T, Asami T, 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
  51. 51. Carrillo-Carrasco VP, Hernandez-Garcia J, Mutte SK, Weijers D. The birth of a giant: Evolutionary insights into the origin of auxin responses in plants. The EMBO Journal. 2023;42(6):e113018
  52. 52. Jing H, Wilkinson EG, Sageman-Furnas K, Strader LC. Auxin and abiotic stress responses. Journal of Experimental Botany. 2023;17:erad325
  53. 53. Zhao Y. Auxin biosynthesis and its role in plant development. Annual Review of Plant Biology. 2010;61:49-64
  54. 54. Weijers D, Nemhauser J, Yang Z. Auxin, small molecule, big impact. Journal of Experimental Botany. 2018;5:133-136
  55. 55. Iglesias MJ, Terrile MC, Windels D, Lombardo MC, Bartoli CG, Vazquez F, et al. MiR393 regulation of auxin signaling and redox-related components during acclimation to salinity in Arabidopsis. PLoS One. 2014;9:e107678
  56. 56. Liu W, Li RJ, Han TT, Cai W, Fu ZW, Lu YT. Salt stress reduces root meristem size by nitric oxide-mediated modulation of auxin accumulation and signaling in Arabidopsis. Plant Physiology. 2015;168:343-356
  57. 57. Petersson SV, Johansson AI, Kowalczyk M, Makoveychuk A, Wang JY, et al. An auxin gradient and maximum in the Arabidopsis root apex shown by high-resolution cell-specific analysis of IAA distribution and synthesis. The Plant Cell. 2009;21:1659-1668
  58. 58. Wang Y, Li K, Li X. Auxin redistribution modulates plastic development of root system architecture under salt stress in Arabidopsis thaliana. Journal of Plant Physiology. 2009;166:1637-1645
  59. 59. Akbari G, Sanavy SA, Yousefzadeh S. Effect of auxin and salt stress (NaCl) on seed germination of wheat cultivars (Triticum aestivum L.). Pakistan Journal of Biological Sciences. 2007;10:2557-2561
  60. 60. Rao SSR, Vardhini BV, Sujatha E, Anuradha S. Brassinosteroids – New class of phytohormones. Current Science. 2002;82:1239-1245
  61. 61. Vriet C, Russinova E, Reuzeau C. Boosting crop yields with plant steroids. The Plant Cell. 2012;24:842-857
  62. 62. Vardhini BV, Anjum NA. Brassinosteroids make plant life easier under abiotic stresses mainly by modulating major components of antioxidant defense system. Frontiers in Environmental Science. 2015;2:67
  63. 63. Zhu T, Deng X, Zhou X, Zhu L, Zou L, Li P, et al. Ethylene and hydrogen peroxide are involved in brassinosteroid-induced salt tolerance in tomato. Scientific Reports. 2016;6:35392
  64. 64. Zhu T, Deng XG, Tan WR, Zhou X, Luo SS, Han XY, et al. Nitric oxide is involved in brassinosteroid-induced alternative respiratory pathway in Nicotiana benthamiana seedlings’ response to salt stress. Physiologia Plantarum. 2016;156:150-163
  65. 65. Wei LJ, Deng XG, Zhu T, Zheng T, Li PX, Wu JQ , et al. Ethylene is involved in brassinosteroids induced alternative respiratory pathway in cucumber (Cucumis sativus L.) seedlings response to abiotic stress. Frontiers in Plant Science. 2015;6:982
  66. 66. Su Q , Zheng X, Tian Y, Wang C. Exogenous brassinolide alleviates salt stress in Malus hupehensis Rehd. by regulating the transcription of NHX-type Na+(K+)/H+ antiporters. Frontiers in Plant Science. 2020;11:38
  67. 67. Yan J, Guan L, Sun Y, Zhu Y, Liu L, Lu R, et al. Calcium and ZmCCaMK are involved in brassinosteroid-induced antioxidant defense in maize leaves. Plant and Cell Physiology. 2015;56:883-896
  68. 68. Arora N, Bhardwaj R, Sharma P, Arora HK. Effects of 28- homobrassinolide on growth, lipid peroxidation and antioxidative enzyme activities in seedlings of Zea mays L. under salinity stress. Acta. Physiologia Plantarum. 2008;30:833-839
  69. 69. Hayat S, Hasan SA, Yusuf M, Hayat Q , Ahmad A. Effect of 28- homobrassinolide on photosynthesis, fluorescence and antioxidant system in the presence or absence of salinity and temperature in Vigna radiata. Environmental and Experimental Botany. 2010;69:105-112
  70. 70. Zhang S, Hu J, Zhang Y, Xie XJ, Knapp A. Seed priming with brassinolide improves lucerne (Medicago sativa L.) seed germination and seedling growth in relation to physiological changes under salinity stress. Australian Journal of Agricultural Research. 2007;58:811-815
  71. 71. Nunez M, Mazzafera P, Mazorra LM, Siqueira WJ, Zullo MAT. Influence of a brassinosteroid analog on antioxidant enzymes in rice grown in culture medium with NaCl. Biologia Plantarum. 2003;47:67-70
  72. 72. Özdemir F, Bor M, Demiral T, Türkan I. Effects of 24-epibrassinolide on seed germination, seedling growth, lipid peroxidation, proline content and antioxidative system of rice (Oryza sativa L.) under salinity stress. Plant Growth Regulation. 2004;42:203-211
  73. 73. Shahbaz M, Ashraf M. Influence of exogenous application of brassinosteroids on growth and mineral nutrients of wheat (Triticum aestivum L.) under saline conditions. Pakistan Journal of Botany. 2007;39:513-522
  74. 74. Ali B, Hasan SA, Hayat S, Hayat Q , Yadav S, Fariduddin Q , et al. A role for brassinosteroids in the amelioration of aluminium stress through antioxidant system in mung bean (Vigna radiata L. Wilczek). Environmental and Experimental Botany. 2008;62:153-159
  75. 75. Shahid MA, Pervez MA, Balal RM, Mattson NS, Rashid A, Ahmad R, et al. Brassinosteroid (24-epibrassinolide) enhances growth and alleviates the deleterious effects induced by salt stress in pea (Pisum sativum L.). Austrian Journal of Crop Science. 2011;5:500-510
  76. 76. Sharma I, Ching E, 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
  77. 77. Abbas S, Latif HH, Elsherbiny EA. Effect of 24-epibrassinolide on the physiological and genetic changes on two varieties of pepper under salt stress conditions. Pakistan Journal of Botany. 2013;45:1273-1284
  78. 78. Ding HD, Zhu XH, Zhu ZW, Yang SJ, Zha DS, et al. Amelioration of salt-induced oxidative stress in eggplant by application of 24-epibrassinolide. Biologia Plantarum. 2012;56:767-770
  79. 79. Fariduddin Q , Mir BA, Yusuf M, Ahmad A. Comparative roles of brassinosteroids and polyamines in salt stress tolerance. Acta Physiologia Plantarum. 2013;35:2037-2053
  80. 80. Mouchel CF, Osmont KS, Hardtke CS. BRX mediates feedback between brassinosteroid levels and auxin signalling in root growth. Nature. 2006;443(7110):458-461
  81. 81. Djemal R, Hanin M, Ebel C. Control of plant responses to salt stress: Significance of auxin and brassinosteroids. In: Anjum NA, Masood A, Thangavel P, Khan NA, editors. Making Plant Life Easier and Productive Under Salinity – Updates and Prospects. London, UK: IntechOpen; 2023. DOI: 10.5772/intechopen.111449
  82. 82. Zhang S, Cai Z, Wang X. The primary signaling outputs of brassinosteroids are regulated by abscisic acid signaling. Proceedings of the National Academy of Sciences, USA. 2009;106:4543-4548
  83. 83. Zeng H, Tang Q , Hua X. Arabidopsis brassinosteroid mutants det2-1 and bin2-1 display altered salt tolerance. Journal of Plant Growth Regulation. 2010;29:44-52
  84. 84. Li ZY, Xu ZS, He GY, Yang GX, Chen M, Li LC, et al. A mutation in Arabidopsis BSK5 encoding a brassinosteroid-signaling kinase protein affects responses to salinity and abscisic acid. Biochemical and Biophysical Research Communications. 2012;426:522-527
  85. 85. Sun F, Yu H, Qu J, Cao Y, Ding L, Feng W, et al. Maize ZmBES1/BZR1-5 decreases ABA sensitivity and confers tolerance to osmotic stress in transgenic Arabidopsis. International Journal of Molecular Sciences. 2020;21:996
  86. 86. Wang X, Chen X, Wang Q , Chen M, Liu X, Gao D, et al. MdBZR1 and MdBZR1-2like transcription factors improves salt tolerance by regulating gibberellin biosynthesis in apple. Frontiers in Plant Science. 2019;10:1473
  87. 87. Divi UK, Rahman T, Krishna P. Brassinosteroid-mediated stress tolerance in Arabidopsis shows interactions with abscisic acid, ethylene and salicylic acid pathways. BMC Plant Biology. 2010;10:1-4
  88. 88. Geng Y, Wu R, Wee CW, Xie F, Wei X, Chan PM, et al. A spatio-temporal understanding of growth regulation during the salt stress response in Arabidopsis. The Plant Cell. 2013;25:2132-2154
  89. 89. Duan F, Ding J, Lee D, Lu X, Feng Y, Song W. Overexpression of SoCYP85A1, a spinach cytochrome p450 gene in transgenic tobacco enhances root development and drought stress tolerance. Frontiers in Plant Science. 2017;8:1909
  90. 90. Srivastava M, Srivastava AK, Orosa-Puente B, Campanaro A, Zhang C, Sadanandom A. SUMO conjugation to BZR1 enables brassinosteroid signaling to integrate environmental cues to shape plant growth. Current Biology. 2020;30:1410-1423
  91. 91. George E, Hall M, Klerk GJ. Plant growth regulators II: Cytokinins, their analogues and antagonists. In: George E, Hall M, Klerk GJ, editors. Plant Propagation by Tissue Culture. Dordrecht, The Netherlands: Springer; 2008. pp. 205-226
  92. 92. Miller CO. A kinetin-like compound in maize. Proceeding of National Academy of Sciences, USA. 1961;47:170-174
  93. 93. Mok DW, Mok MC. Cytokinin metabolism and action. Annual Review of Plant Physiology and Plant Molecular Biology. 2001;52:89-118
  94. 94. Kieber JJ, Schaller GE. Cytokinin signaling in plant development. Development. 2018;145(4):dev149344
  95. 95. O’Brien JA, Benková E. Cytokinin cross-talking during biotic and abiotic stress responses. Frontiers in Plant Science. 2013;4:451
  96. 96. Asgher M, Khan MIR, Anjum NA, et al. Minimising toxicity of cadmium in plants - role of plant growth regulators. Protoplasma. 2015;252:399-413
  97. 97. Cortleven A, Leuendorf JE, Frank M, Pezzetta D, Bolt S, Schmülling T. Cytokinin action in response to abiotic and biotic stresses in plants. Plant, Cell & Environment. 2019;42(3):998-1018
  98. 98. Zalabák D, Pospíšilová H, Šmehilová M, Mrízová K, Frébort I, Galuszka P. Genetic engineering of cytokinin metabolism: Prospective way to improve agricultural traits of crop plants. Biotechnology Advances. 2013;31:97-117
  99. 99. Yin P, Liang X, Zhao H, Xu Z, Chen L, Yang X, et al. Cytokinin signaling promotes salt tolerance by modulating shoot chloride exclusion in maize. Molecular Plant. 2023;16(6):1031-1047
  100. 100. Azzam CR, Zaki SN, Bamagoos AA, Rady MM, Alharby HF. Soaking maize seeds in zeatin-type cytokinin biostimulators improves salt tolerance by enhancing the antioxidant system and photosynthetic efficiency. Plants. 2022;11(8):1004
  101. 101. Merchan F, de Lorenzo L, Rizzo SG, Niebel A, Manyani H, Frugier F, et al. Identification of regulatory pathways involved in the reacquisition of root growth after salt stress in Medicago truncatula. Plant Journal. 2007;51:1-17
  102. 102. Tran LS, Shinozaki K, Yamaguchi-Shinozaki K. Role of cytokinin responsive two-component system in ABA and osmotic stress signalings. Plant Signaling and Behavior. 2010;5:148-150
  103. 103. Nishiyama R, Watanabe Y, Fujita Y, Le DT, 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;23:2169-2183
  104. 104. Wang KLC, Li H, Ecker JR. Ethylene biosynthesis and signaling networks. Plant Cell (Suppl). 2002;14:S131-S151
  105. 105. Yang C, Lu X, Ma B, Chen SY, Zhang JS. Ethylene signaling in rice and Arabidopsis: Conserved and diverged aspects. Molecular Plant. 2015;8:495-505
  106. 106. Zhang M, Smith JA, 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
  107. 107. Cao YR, Chen SY, Zhang JS. Ethylene signaling regulates salt stress response: An overview. Plant Signaling and Behavior. 2008;3(10):761-763
  108. 108. An JP, Yao JF, Xu RR, You CX, Wang XF, Hao YJ. An apple NAC transcription factor enhances salt stress tolerance by modulating the ethylene response. Physiologia Plantarum. 2018;164(3):279-289
  109. 109. Jiang C, Belfield EJ, Cao Y, Smith JA, Harberd NP. An Arabidopsis soil salinity-tolerance mutation confers ethylene-mediated enhancement of sodium/potassium homeostasis. The Plant Cell. 2013;25:3535-3552
  110. 110. Peng J, Li Z, Wen X, Li W, Shi H, Yang L, et al. Salt-induced stabilization of EIN3/EIL1 confers salinity tolerance by deterring ROS accumulation in Arabidopsis. PLoS Genetics. 2014;10(10):e1004664
  111. 111. Yang C, Ma B, He SJ, Xiong Q , Duan KX, Yin CC, et al. MAOHUZI6/ETHYLENE INSENSITIVE3-LIKE1 and ETHYLENE INSENSITIVE3-LIKE2 regulate ethylene response of roots and coleoptiles and negatively affect salt tolerance in rice. Plant Physiology. 2015;169:148-165
  112. 112. Yang L, Zu YG, Tang ZH. Ethylene improves Arabidopsis salt tolerance mainly via retaining K+ in shoots and roots rather than decreasing tissue Na+ content. Environmental and Experimental Botany. 2013;86:60-69
  113. 113. Khan MIR, Asgher M, Khan NA. Alleviation of salt-induced photosynthesis and growth inhibition by salicylic acid involves glycinebetaine and ethylene in mungbean (Vigna radiata L.). Plant Physiology and Biochemistry. 2014;80:67-74
  114. 114. Silva PO, Medina EF, Barros RS, Ribeiro DM. Germination of salt-stressed seeds as related to the ethylene biosynthesis ability in three Stylosanthes species. Journal of Plant Physiology. 2014;171:14-22
  115. 115. 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. 2017;143:707-719
  116. 116. Freitas VS, de Souza MR, Costa JH, de Oliveira DF, de Oliveira PS, et al. Ethylene triggers salt tolerance in maize genotypes by modulating polyamine catabolism enzymes associated with H2O2 production. Environmental and Experimental Botany. 2018;145:75-86
  117. 117. Liu M, Wei JW, Liu W, Gong B. S-nitrosylation of ACO homolog 4 improves ethylene synthesis and salt tolerance in tomato. New Phytologist. 2023;239:159-173
  118. 118. Yang L, Wang X, Zhao F, Zhang X, Li W, Huang J, et al. Roles of S-adenosylmethionine and its derivatives in salt tolerance of cotton. International Journal of Molecular Sciences. 2023;24(11):9517
  119. 119. An JP, Zhang XW, Xu RR, You CX, Wang XF, Hao YJ. Apple MdERF4 negatively regulates salt tolerance by inhibiting MdERF3 transcription. Plant Science. 2018;276:181-188
  120. 120. Chen K, Song M, Guo Y, Liu L, Xue H, Dai H, et al. MdMYB46 could enhance salt and osmotic stress tolerance in apple by directly activating stress-responsive signals. Plant Biotechnology Journal. 2019;17:2341-2355
  121. 121. Liu Y, Li D, Yan J, Wang K, Luo H, Zhang W. MiR319 mediated salt tolerance by ethylene. Plant Biotechnology Journal. 2019;17(12):2370-2383
  122. 122. Ma B, Chen SY, Zhang JS. Ethylene signaling in rice. Chinese Science Bulletin. 2010;55:2204-2210
  123. 123. Chen D, Ma X, Li C, Zhang W, Xia G, Wang M. A wheat aminocyclopropane-1-carboxylate oxidase gene, TaACO1, negatively regulates salinity stress in Arabidopsis thaliana. Plant Cell Reports. 2014;33:1815-1827
  124. 124. Tao JJ, Chen HW, Ma B, Zhang WK, Chen SY, Zhang JS. The role of ethylene in plants under salinity stress. Frontiers in Plant Science. 2015;6:1059
  125. 125. Matsuoka M. Gibberellins signaling: How do plant cells respond to GA signals? Journal of Plant Growth Regulation. 2003;22:123-125
  126. 126. Gupta R, Chakrabarty SK. Gibberellic acid in plant – Still a mystery unresolved. Plant Signaling and Behavior. 2013;8:e25504
  127. 127. Zawaski C, Busov VB. Roles of gibberellin catabolism and signaling in growth and physiological response to drought and short-day photoperiods in Populus trees. PLoS One. 2014;9:e86217
  128. 128. Uçarli C. Physiological and molecular effects of exogenous gibberellin (GA3) treatment on germination of barley seeds under salt stress. Adıyaman University Journal of Science. 2021;11(2):227-243
  129. 129. Shahzad K, Hussain S, Arfan M, Hussain S, Waraich EA, Zamir S, et al. Exogenously applied gibberellic acid enhances growth and salinity stress tolerance of maize through modulating the morpho-physiological, biochemical and molecular attributes. Biomolecules. 2021;11(7):1005
  130. 130. 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
  131. 131. Ghodrat V, Rousta MJ. Effect of priming with gibberellic acid (GA3) on germination and growth of corn (Zea mays L.) under saline conditions. International Journal of Agricultural and Crop Science. 2012;4:882-885
  132. 132. Iqbal M, Ashraf M. Gibberellic acid mediated induction of salt tolerance in wheat plants: Growth, ionic partitioning, photosynthesis, yield and hormonal homeostasis. Environmental and Experimental Botany. 2013;86:76-85
  133. 133. Iftikhar A, Ali S, Yasmeen T, Arif MS, Zubair M, Rizwan M, et al. Effect of gibberellic acid on growth, photosynthesis and antioxidant defense system of wheat under zinc oxide nanoparticle stress. Environmental Pollution. 2019;254:113109
  134. 134. Achard P, Renou J, Berthome R, Harberd N, Genschik P. Plant DELLAs restrain growth and promote survival of adversity by reducing the levels of reactive oxygen species. Current Biology. 2008;18:656-660
  135. 135. Colebrook EH, Thomas SG, Phillips AL, Hedden P. The role of gibberellin signalling in plant responses to abiotic stress. Journal of Experimental Biology. 2014;217(1):67-75
  136. 136. Ahmad F, Kamal A, Singh A, Ashfaque F, Alamri S, Siddiqui MH, et al. Seed priming with gibberellic acid induces high salinity tolerance in Pisum sativum through antioxidants, secondary metabolites and up-regulation of antiporter genes. Plant Biology. 2021;23:113-121
  137. 137. Li Z, Xu J, Gao Y, Wang C, Guo G, Luo Y, et al. The synergistic priming effect of exogenous salicylic acid and H2O2 on chilling tolerance enhancement during maize (Zea mays L.) seed germination. Frontiers in Plant Science. 2017;8:1153
  138. 138. Iqbal H, Yaning C, Waqas M, Rehman HM, Sharee FM, Iqbal S. Hydrogen peroxide application improves quinoa performance by affecting physiological and biochemical mechanisms under water-deficit conditions. Journal of Agronomy and Crop Sciences. 2018;204:541-553
  139. 139. Al-Harthi MM, Bafeel SO, El-Zohri M. Gibberellic acid and jasmonic acid improve salt tolerance in summer squash by modulating some physiological parameters symptomatic for oxidative stress and mineral nutrition. Plants. 2021;10:2768
  140. 140. Rady MM, Talaat NB, Abdelhamid MT, Shawky BT, Desoky EM. Maize (Zea mays L.) grains extract mitigates the deleterious effects of salt stress on common bean (Phaseolus vulgaris L.) growth and physiology. Journal of Horticultural Science and Biotechnology. 2021;94:777-789
  141. 141. Kazan K, Manners JM. JAZ repressors and the orchestration of phytohormone crosstalk. Trends in Plant Science. 2012;17(1):22-31
  142. 142. Zainordin ZF, San CT, Ahmad A. Foliar sprays of methyl jasmonic acid shift the endogenous fatty acid levels to support rice plant growth in saline soil. Plant Stress. 2023;13:100221
  143. 143. Sheteiwy MS, Ulhassan Z, Qi W, Lu H, AbdElgawad H, Minkina T, et al. Association of jasmonic acid priming with multiple defense mechanisms in wheat plants under high salt stress. Frontiers in Plant Science. 2022;13:886862
  144. 144. Taheri Z, Vatankhah E, Jafarian V. Methyl jasmonate improves physiological and biochemical responses of Anchusa italica under salinity stress. South African Journal of Botany. 2020;130:375-382
  145. 145. Kramell R, Miersch O, Atzorn R, Parthier B, Wasternack C. Octadecanoid-derived alteration of gene expression and the ‘oxylipin signature’ in stressed barley leaves. Implications for different signaling pathways. Plant Physiology. 2000;123:177-187
  146. 146. Pedranzani H, Racagni G, Alemano S, Miersch O, Ramirez I, Pena-Cortes H, et al. Salt tolerant tomato plants show increased levels of jasmonic acid. Plant Growth Regulation. 2003;41:149-158
  147. 147. Shinozaki K, Yamaguchi-Shinozaki K. Gene networks involved in drought stress response and tolerance. Journal of Experimental Botany. 2007;58(2):221-227
  148. 148. Khan MS, Hemalatha S. Autophagy and programmed cell death are critical pathways in Jasmonic acid mediated saline stress tolerance in Oryza sativa. Applied Biochemistry and Biotechnology. 2022;194(11):5353-5366
  149. 149. Asfaw KG, Liu Q , Eghbalian R, Purper S, Akaberi S, Dhakarey R, et al. The jasmonate biosynthesis Gene OsOPR7 can mitigate salinity induced mitochondrial oxidative stress. Plant Science. 2022;316:111156
  150. 150. Valenzuela CE, Acevedo-Acevedo O, Miranda GS, Vergara-Barros P, Holuigue L, Figueroa CR, et al. Salt stress response triggers activation of the jasmonate signaling pathway leading to inhibition of cell elongation in Arabidopsis primary root. Journal of Experimental Botany. 2016;67:4209-4220
  151. 151. Rather BA, Masood A, Sehar Z, Majid A, Anjum NA, Khan NA. Mechanisms and role of nitric oxide in phytotoxicity-mitigation of copper. Frontiers in Plant Science. 2020;11:675
  152. 152. Ahmad P, Latef AAA, Hashem A, Allah EF, Gucel S, Tran L. Nitric oxide mitigates salt stress by regulating levels of osmolytes and antioxidant enzymes in chickpea. Frontiers in Plant Science. 2016;7:347
  153. 153. Liu A, Fan J, Gitau MM, Chen L, Fu J. Nitric oxide involvement in Bermuda grass response to salt stress. Journal of the American Society for Horticultural Science. 2016;141:425-433
  154. 154. Sehar Z, Masood A, Khan NA. Nitric oxide reverses glucose-mediated photosynthetic repression in wheat (Triticum aestivum L.) under salt stress. Environmental and Experimental Botany. 2019;161:277-289
  155. 155. Zheng C, Jiang D, Liu F, Dai T, Liu W, Jing Q , et al. Exogenous nitric oxide improves seed germination in wheat against mitochondrial oxidative damage induced by high salinity. Environmental and Experimental Botany. 2009;67:222-227
  156. 156. Marvasi M. Potential use and perspectives of nitric oxide donors in agriculture. Journal of Science of Food and Agriculture. 2017;97:1065-1072
  157. 157. Kausar F, Shahbaz M, Ashraf M. Protective role of foliar-applied nitric oxide in Triticum aestivum under saline stress. Turkish Journal of Botany. 2013;37:1155-1165
  158. 158. Fatma M, Khan NA. Nitric oxide protects photosynthetic capacity inhibition by salinity in Indian mustard. Journal of Functional and Environmental Botany. 2014;4:106-116
  159. 159. Chen WW, Yang JL, Qin C, Jin CW, Mo JH, Ye T, et al. Nitric oxide acts downstream of auxin to trigger root ferric-chelate reductase activity in response to iron deficiency in Arabidopsis. Plant Physiology. 2010;154:810-819
  160. 160. Zhang Y, Wang L, Liu Y, Zhang Q , Wei Q , Zhang W. Nitric oxide enhances salt tolerance in maize seedlings through increasing activities of proton-pump and Na+/H+ antiport in the tonoplast. Planta. 2006;224:545-555
  161. 161. Wodala B, Deák Z, Vass I, Erdei L, Altorjay I, Horváth F. In vivo target sites of nitric oxide in photosynthetic electron transport as studied by chlorophyll fluorescence in pea leaves. Plant Physiology. 2008;146:1920-1927
  162. 162. Wu XX, Ding HD, Chen JL, Zhang HJ, Zhu WM. Attenuation of salt-induced changes in photosynthesis by exogenous nitric oxide in tomato (Lycopersicon esculentum Mill. L.) seedlings. African Journal of Biotechnology. 2010;9:7837-7846
  163. 163. Fatma M, Masood A, Per TS, Khan NA. Nitric oxide alleviates salt stress inhibited photosynthetic performance by interacting with sulfur assimilation in mustard. Frontiers in Plant Science. 2016;7:521
  164. 164. Anjum NA, Umar S, Chan MT. Ascorbate-Glutathione Pathway and Stress Tolerance in Plants. Dordrecht: Springer; 2010
  165. 165. Hasanuzzaman M, Gill SS, Fujita M. Physiological role of nitric oxide in plants grown under adverse environmental conditions. In: Tuteja N, Singh Gill S, editors. Plant Acclimation to Environmental Stress. New York: Springer; 2013b. pp. 269-322
  166. 166. da Silva CJ, Fontes EPB, Modolo LV. Salinity-induced accumulation of endogenous H2S and NO is associated with modulation of the antioxidant and redox defense systems in Nicotiana tabacum L. cv. Havana. Plant Science. 2017;256:148-159
  167. 167. Ali Q , Daud MK, Haider MZ, Ali S, Aslam N, Noman A, et al. Seed priming by sodium nitroprusside improves salt tolerance in wheat (Triticum aestivum L.) by enhancing physiological and biochemical parameters. Plant Physiology and Biochemistry. 2017;119:50-58
  168. 168. Miura K, Tada Y. Regulation of water, salinity, and cold stress responses by salicylic acid. Frontiers in Plant Science. 2014;5:4
  169. 169. Khan MIR, Fatma M, Per TS, Anjum NA, Khan NA. Salicylic acid-induced abiotic stress tolerance and underlying mechanisms in plants. Frontiers in Plant Science. 2015;6:462
  170. 170. Asensi-Fabado MA, Munné-Bosch S. The aba3-1 mutant of Arabidopsis thaliana withstands moderate doses of salt stress by modulating leaf growth and salicylic acid levels. Journal of Plant Growth Regulation. 2011;30:456-466
  171. 171. Kaya C, Ashraf M, Alyemeni MN, Ahmad P. The role of endogenous nitric oxide in salicylic acid-induced up-regulation of ascorbate-glutathione cycle involved in salinity tolerance of pepper (Capsicum annuum L.) plants. Plant Physiology and Biochemistry. 2020;147:10-20
  172. 172. Sousa VF, Santos AS, Sales WS, Silva AJ, Gomes FA, Dias TJ, et al. Exogenous application of salicylic acid induces salinity tolerance in eggplant seedlings. Brazilian Journal of Biology. 2022;84:e257739
  173. 173. Quilis J, Peñas G, Messeguer J, Brugidou C, Segundo BS. The Arabidopsis AtNPR1 inversely modulates defense responses against fungal, bacterial, or viral pathogens while conferring hypersensitivity to abiotic stresses in transgenic rice. Molecular Plant-Microbe Interaction. 2008;21:1215-1231
  174. 174. Silva KJ, Mahna N, Mou Z, Folta KM. NPR1 as a transgenic crop protection strategy in horticultural species. Horticulture Research. 2018;5:15
  175. 175. Yan Y, Pan C, Du Y, Li D, Liu W. Exogenous salicylic acid regulates reactive oxygen species metabolism and ascorbate-glutathione cycle in Nitraria tangutorum Bobr. under salinity stress. Physiology and Molecular Biology of Plants. 2018;24:577-589
  176. 176. Csiszár J, Horváth E, Váry Z, Gallé Á, Bela K, Brunner S, et al. Glutathione transferase supergene family in tomato: Salt stress-regulated expression of representative genes from distinct GST classes in plants primed with salicylic acid. Plant Physiology and Biochemistry. 2014;78:15-26
  177. 177. Li G, Peng X, Wei L, Kang G. Salicylic acid increases the contents of glutathione and ascorbate and temporally regulates the related gene expression in salt-stressed wheat seedlings. Gene. 2013;529:321-325
  178. 178. 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
  179. 179. Alvi AF, Sehar Z, Fatma M, Masood A, Khan NA. Strigolactone: An emerging growth regulator for developing resilience in plants. Plants. 2022;11(19):2604
  180. 180. Ma N, Hu C, Wan L, Hu Q , Xiong J, Zhang C. Strigolactones improve plant growth, photosynthesis, and alleviate oxidative stress under salinity in rapeseed (Brassica napus L.) by regulating gene expression. Frontiers in Plant Science. 2017;8:1671
  181. 181. Ha CV, Leyva-González MA, Osakabe Y, Tran UT, Nishiyama R, Watanabe Y, et al. Positive regulatory role of strigolactone in plant responses to drought and salt stress. Proceedings of the National Academy of Sciences. 2014;111(2):851-856
  182. 182. Burger M, Chory J. The many models of strigolactone signaling. Trends in Plant Science. 2020;25:395-405
  183. 183. Aroca R, Ruiz-Lozano JM, Zamarreño ÁM, Paz JA, García-Mina JM, Pozo MJ, et al. Arbuscular mycorrhizal symbiosis influences strigolactone production under salinity and alleviates salt stress in lettuce plants. Journal of Plant Physiology. 2013;170:47-55
  184. 184. Ren CG, Kong CC, Xie ZH. Role of abscisic acid in strigolactone-induced salt stress tolerance in arbuscular mycorrhizal Sesbania cannabina seedlings. BMC Plant Biology. 2018;18:74
  185. 185. Sharifi P, Bidabadi SS. Strigolactone could enhances gas exchange through augmented antioxidant defense system in Salvia nemorosa L. plants subjected to saline conditions stress. Industrial Crop Production. 2020;151:112460
  186. 186. Zhang X, Zhang L, Ma C, Su M, Wang J, et al. Exogenous strigolactones alleviate the photosynthetic inhibition and oxidative damage of cucumber seedlings under salt stress. Scientia Horticulturae. 2022;297:110962

Written By

Naser A. Anjum, Asim Masood, Faisal Rasheed, Palaniswamy Thangavel and Nafees A. Khan

Submitted: 09 October 2023 Reviewed: 13 October 2023 Published: 05 November 2023