Next Article in Journal
Genome-Wide Identification and Comparative Transcriptome Methods Reveal FaMDHAR50 Regulating Ascorbic Acid Regeneration and Quality Formation of Strawberry Fruits
Next Article in Special Issue
MsSPL9 Modulates Nodulation under Nitrate Sufficiency Condition in Medicago sativa
Previous Article in Journal
Fangs in the Ghats: Preclinical Insights into the Medical Importance of Pit Vipers from the Western Ghats
Previous Article in Special Issue
Genome-Wide Identification and Expression Analysis of Dendrocalamus farinosus CCoAOMT Gene Family and the Role of DfCCoAOMT14 Involved in Lignin Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Roles of S-Adenosylmethionine and Its Derivatives in Salt Tolerance of Cotton

1
College of Life Science, Yangtze University, Jingzhou 434025, China
2
State Key Laboratory of Cotton Biology, Institute of Cotton Research, Chinese Academy of Agricultural Sciences, Anyang 455000, China
3
Western Research Institute, Chinese Academy of Agricultural Sciences (CAAS), Changji 831100, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(11), 9517; https://doi.org/10.3390/ijms24119517
Submission received: 13 April 2023 / Revised: 19 May 2023 / Accepted: 25 May 2023 / Published: 30 May 2023
(This article belongs to the Special Issue Abiotic Stresses in Plants: From Molecules to Environment)

Abstract

:
Salinity is a major abiotic stress that restricts cotton growth and affects fiber yield and quality. Although studies on salt tolerance have achieved great progress in cotton since the completion of cotton genome sequencing, knowledge about how cotton copes with salt stress is still scant. S-adenosylmethionine (SAM) plays important roles in many organelles with the help of the SAM transporter, and it is also a synthetic precursor for substances such as ethylene (ET), polyamines (PAs), betaine, and lignin, which often accumulate in plants in response to stresses. This review focused on the biosynthesis and signal transduction pathways of ET and PAs. The current progress of ET and PAs in regulating plant growth and development under salt stress has been summarized. Moreover, we verified the function of a cotton SAM transporter and suggested that it can regulate salt stress response in cotton. At last, an improved regulatory pathway of ET and PAs under salt stress in cotton is proposed for the breeding of salt-tolerant varieties.

1. Introduction

Cotton is one of the most important cash crops in the world and the main source of natural fiber. With the continuous growth of the world’s population, cultivated land decreases year by year, and cotton cultivation has gradually shifted to saline land. It is estimated that more than 50% of arable land will be salinized by 2050 [1]. Although cotton is a moderately salt-tolerant crop with a threshold of 7.7 dSm−1, long-term and high salt stress affects its growth and yield, especially at the stages of germination and seedling [2,3]. Salt stress reduces the stomatal conductance in cotton, leading to a decrease in photosynthesis and reduced carbohydrate supply in developing boll, which ultimately affects the fiber quality and yield of cotton [4,5]. Currently, although a large number of studies have been conducted to analyze the genetic mechanism of salt tolerance in plants, the specific mechanism of salt tolerance in cotton is still unclear.
As genome sequences have increasingly become available for Gossypium, a large number of genes have been identified in response to abiotic stress in cotton [6,7]. For example, the NAC transcription factor gene GhNAC072 has been overexpressed to improve salt and drought tolerance in Arabidopsis [8], and overexpression of GhMPK3 alleviated the damage of salt stress in Arabidopsis [9]. S-adenosyl-L-methionine (SAM, also known as AdoMet) is a major methyl donor that not only participates in the methionine (Met) cycle, but also provides aminopropyl for the synthesis of PAs and ET [10,11,12]. A series of studies have revealed that ET is one of the important factors responding to salt stress in many species, as exemplified in Arabidopsis, rice, wheat, cucumber, tomato, and cotton [13,14,15,16,17,18]. PAs, a group of aliphatic amine compounds similar to phytohormones, exist mainly in three endogenous statuses (free, conjugated, bond) and function in the processes of plant development and stress response [19,20]. In soybean, PAs promote hypocotyl elongation by enhancing their own catabolism reactions and increasing the production of hydrogen peroxide (H2O2) under salt stress [21]. Put2 is a polyamine uptake protein, and overexpression of put2 in tomatoes can improve the antioxidant enzyme activity and salt tolerance of transgenic lines [22]. In addition, PAs respond to the saline–alkali stress in rapeseed through the dynamic adjustment of different PA statuses [23]. However, studies on the regulation of salt stress by PAs are still very few in cotton [24].
SAM, as the precursor of ET and PAs synthesis, needs to be transported to the corresponding organelles through SAM transporters to perform a biological function. SAM transporters were first identified in yeast and humans [25,26]. In Arabidopsis, only two SAM transporters (AtSAMC1 and AtSAMC2) have been identified. The impaired function of AtSAMC1 will reduce the level of prenyl lipids, which mainly affect the chlorophyll pathway, thereby affecting photosynthesis and reducing tolerance to stress [11]. Thus far, studies on SAM transporters have mainly focused on humans and micro-organisms, with the exception of a few studies on Arabidopsis [27,28,29].
This review focuses on the recent findings on salt tolerance in cotton and systematically introduces the important role of SAM in the response to salt stress through its metabolic derivatives ET and PAs. Furthermore, we identified a putative SAM transporter (GhSAMC) from cotton through the transcriptome data of different cotton species and found that it can positively regulate the salt tolerance of cotton. The knowledge presented in this review may be used for breeding elite cotton varieties with high salt tolerance.

2. Research Status of Cotton on Salt Stress

Soil salinization affects about 20% of the world’s arable land. This problem continues to exacerbate with global warming, the deterioration of the natural environment, and unreasonable irrigation methods [30,31]. Plants are sessile organisms and cannot escape extreme external environments as animals. Hence, plants have evolved a flexible system that can adjust their morphological, physiological, biochemical, and molecular mechanisms to respond to (a)biotic stresses. Salt stress causes plant cells to suffer from osmotic stress, ion stress, and other secondary stresses, especially oxidative stress [32]. Long-term salt stress will lead to soft and dark leaves in cotton, causing shortened functional periods, early shedding, weak stems, decreased fresh matter and plant height, increased salt concentration, and delayed flowering time [12,33]. In addition, increased relative shedding of flowers and bolls ultimately affects yield and fiber quality in cotton [34] (Figure 1).

2.1. Osmotic Regulation

In the early stage of salt stress, plants are subjected to osmotic stress and accumulate a large number of osmomodulating substances, such as inorganic ions (K+, Cl, and inorganic salts, etc.), and organic solutes (proline, betaine, soluble sugars, polyols, and polyamines, etc.). Plants try to keep osmotic homeostasis depending on the dynamic adjustment of these osmomodulating substances [35].

2.2. Ionic Regulation

With the increase in salt concentration, plant cells accumulate a large amount of Na+, causing a K+/Na+ homeostatic imbalance and Ca2+ dysfunction, resulting in ion stress. Plants mitigate ionic toxicity primarily by reducing Na+ influx, separating and excreting Na+ [36,37].
At present, dozens of studies have been carried out on ion transporters in cotton. SOS1, a plasma membrane Na+/H+ reverse transporter, is responsible for the secretion of Na+ in the cytoplasm [35]. Overexpression of the cotton SOS1 gene GhSOS1 in Arabidopsis revealed that the Na+/K+ ratio and malondialdehyde (MDA) content decreased in the leaves of transgenic lines treated with salt stress, and the salt-tolerant ability was enhanced [38,39]. NHX, a vacuole Na+/H+ anti-transporter, is present in roots and leaves and enables excess Na+ to be retained in the vacuole. The transport activity of NHX mainly depends on the H+ gradient maintained by H+-ATPase and H+-PPase in the vacuole [40]. In a previous investigation, the overexpression of two cotton homologous genes GhNHX1 and GhNHX3D enhanced the salt tolerance in cotton [41,42]. The K+ transporter HAK5 and K+ channel AKT1 are responsible for K+ uptake in soil, and the Na+/K+ transporter HKT is responsible for transferring Na+ from photosynthetic tissues to roots and extracting Na+ from xylem [43,44,45]. While the homologous gene SbHKT1 was overexpressed in cotton, it can improve the K+ absorption ability in transgenic lines and maintain the homeostasis of reactive oxygen species (ROS) [46]. The K+ efflux anti-transporter KEA enhances plant salt tolerance by regulating the homeostasis of K+ and H+ in cells. As an example, it was found that silencing of GhKEA4 and GhKEA12 reduced proline and soluble sugar content in cotton leaves and increased sensitivity to salt stress [47].
In addition to the above-mentioned genes that have been well studied in regulating Na+/K+ homeostasis, some new genes in the regulation of the Ca2+ and Na+ flux have been identified in cotton in recent years, such as the cationic amino acid transporter gene (GhCAT10D), cyclic nucleotide-gated channel gene (GhCNGC1/18/GhCNGC12/31/), calcium-binding protein gene (GhCLO6), annexin gene (GhANN1/GhANN8b), and sodium bile acid transporter gene (GhBASS2) [48,49,50,51,52]. These observations indicate that cotton has a complex regulatory network to keep ionic homeostasis under salt stress as other plants. Therefore, exploring the key genes in the ionic regulation process is crucial for comprehensively dissecting the regulatory mechanism of salt tolerance in cotton.

2.3. Oxidation Regulation

With the extension time of salt stress, osmotic and ionic stresses will induce oxidative stress, causing the accumulation of a large number of ROS in plants. Additionally, chloroplasts, mitochondria, and peroxisomes are the three main sites for ROS accumulation. ROS mainly includes singlet oxygen (1O2), superoxide (O2•−), H2O2, and hydroxyl radicals (OH•) [53]. Under normal growth conditions, low concentrations of ROS in plants can be used as second messengers, involved in seed germination, growth and development, root growth and gravitropism, programmed cell death, and other processes [54,55]. Under salt stress conditions, plants accumulate a large amount of ROS. High concentrations of ROS can lead to protein denaturation, lipid peroxidation, and nucleic acid damage [56,57]. Although cotton is a relatively salt-tolerant crop, when cotton is in a high-salt environment, a large amount of ROS is produced in the plant. Excessive ROS can affect the growth and development of cotton plants and eventually lead to a decrease in cotton yield and fiber quality [58].
In order to avoid the damage caused by ROS, plants synthesize a series of antioxidant enzymes, such as superoxide dismutase (SOD), peroxidase (POD), catalase (CAT), ascorbyl peroxidase (APX), glutathione peroxidase (GPX), and non-enzymatic antioxidants such as carotenoids, phenolic compounds, flavonoids, ascorbic acid (AsA), and glutathione (GSH) [59,60]. In addition, the expression of related genes in plants can regulate the level of ROS and improve the tolerance of plants to salt stress. Recent studies have shown that ghr-miR414c can further regulate ROS metabolism by regulating the expression level of GhFSD1 to affect the salt tolerance of cotton [61]. Cotton WRKY transcription factor GhWRKY17 can improve the salt sensitivity of transgenic tobacco through ABA signal transduction and regulation of ROS production in plant cells [62]. Under salt stress conditions, the aldehyde dehydrogenase 21 gene (ScALDH21) of S. canadensis can scavenge excessive ROS in transgenic cotton, thereby improving the salt tolerance of cotton [63]. Silencing the Raf-like MAPKKK gene GhRaf19 reduced the accumulation of ROS in cotton, thereby improving the tolerance of plants to salt stress [64].

2.4. Signal Transduction Regulation

Over the long evolution, plants have evolved various mechanisms to deal with adversity, such as the accumulation of permeable substances, the regulation of genes associated with Na+/K+ and Ca2+ homeostasis, and the production of antioxidant enzymes and non-enzymatic antioxidants. Each mechanism in above contains a complex signaling pathway, consisting of receptors, secondary messengers, phytohormones, and signal transductors [35]. Of these, the signal transductors include many enzymatic cascades such as mitogen-activated protein kinase (MAPK), calcium-dependent protein kinase (CDPK), G protein, phosphatase, and other components [65,66]. Various transcription factors (TFs), such as MYB [67,68], bZIP [69,70], WRKY [71,72,73], NAC [74,75,76], MYC [77], etc., act as a central regulator and a molecular switch in the signal transduction network under salt stress. These TFs regulate the expression pattern of downstream genes and affect the salt tolerance level by activating or inhibiting their interacting genes [78,79]. In addition, various phytohormones such as abscisic acid (ABA), jasmonic acid (JA), salicylic acid (SA), ethylene (ET), gibberellin (GA), and cytokinin (CK) also participate in the signal transduction of salt stress to regulate the response to salt stress [80,81].
At present, the salt tolerance of cotton can be improved to a certain extent through traditional breeding methods, molecular marker-assisted selection, transgenic breeding, and gene editing. However, salt tolerance in plants is a complex trait controlled by multiple quantitative trait loci (QTL), determined by the response of the whole plant [82]. It is very hard to significantly improve the salt-tolerant ability with changes in a single gene or protein. These require us to further explore the molecular mechanism of cotton salt tolerance to breed varieties with high salt-tolerant capability.

3. Role of SAM in Plant Salt Tolerance

As a precursor for the biosynthesis of ET and PAs, SAM is synthesized from the substrates ATP and Met by the catalysis of S-adenosylmethionine synthetase (SAMS) [83]. Related studies have shown that SAMS regulates plant responses to salt stress by promoting the synthesis of SAM. Heterologous expression of potato SbSAMS in Arabidopsis led to the accumulation of more SAM, S-adenosyl-L-homocysteine (SAHC), and ET in transgenic lines, showing higher salt and drought tolerance levels [83]. In another study, the overexpression of beet BvM14-SAMS2 in Arabidopsis improved SAM content and the tolerant ability to salt in transgenic plants, in which the antioxidant system and polyamines’ metabolism played an important role [84].. The advanced analysis found that the biosynthesis of PAs and 1-aminocyclopropane-1-carboxylic acid (ACC) was related to the CsSAMS1 expression level in transgenic plants [85,86].

4. Role of Ethylene (ET) in Plant Growth and Development of Cotton under Salt Stress

When plants are exposed to salt stress, it leads to elevated ROS content in the body, especially including H2O2 and O2•−; these increased ROS have a dual effect: when they are at low concentrations, they can act as signaling molecules that mediate salt tolerance, whereas they will produce oxidative damage to plant cells when at high concentrations, thereby disrupting the function of chloroplasts and even triggering plant cell death (PCD) [87,88,89]. As a volatile gas plant hormone, ET is usually considered as another stress hormone in addition to ABA, and its synthesis is induced by various (a)biotic environmental stresses [90,91]. ET also has a dual role in regulating salt stress: one is inhibiting the accumulation of ROS through ET signaling pathways, thereby maintaining the homeostasis of ROS in plants, and the other is promoting ROS production and signal transduction to activate Na+/K+ transport [92,93,94].

4.1. ET Biosynthesis

The biosynthesis of ET begins with SAMS catalyzing the conversion of Met to SAM. Next, the SAM is converted into ACC by 1-aminocyclopropane-1-carboxylic acid (ACC) synthetases (ACSs). Finally, using ACC as the substrate, ET is synthesized through ACC oxidases (ACOs) (Figure 2) [95,96,97].
ACSs and ACOs are commonly considered key biosynthetic enzymes in ET biosynthesis. Thus far, 12 ACS homologous genes have been identified in Arabidopsis [98]. Among these ACS genes, AtACS1 encodes non-functional homodimers or inactivating catalytic enzymes, AtACS3 is a pseudogene, AtACS10 and AtACS12 perform amino transfer functions, and the other 8 ACS genes (AtACS2, AtACS4-9, AtACS11) encode functional ACS [99]. Moreover, the presence of these eight functional active ACS enzymes and their ability to form active heterodimers will increase the diversity of ET biosynthetic reactions and improve the ability to regulate ET production at different developmental and environmental stages [100,101]. A total of 35 GhACS proteins with a conserved C-terminus were identified in upland cotton, whereas the N-terminus of ACS10 and ACS12 were divergent between cotton and Arabidopsis, indicating that they may have a different evolutionary trajectory. QRT-PCR analysis showed that the expression of GhACS1 was significantly upregulated under salt stress, whereas GhACS6.3, GhACS7.1, and GhACS10/12 could respond to cold stress [102]. ACO also plays an important role in ET biosynthesis. Arabidopsis encodes a total of five ACO genes, whose expression is regulated by growth and developmental and environmental stresses [103]. Through genome-wide identification, a total of 332 GhACOs were identified in upland cotton, and most of them co-expressed with other genes in response to salt and drought stresses. For example, the overexpression of GhACO106-At could improve salt tolerance in transgenic Arabidopsis [104]. At present, the ET biosynthetic pathway has been well elucidated in model species, but there is still little research on genes that promote ET biosynthesis and respond to salt stress in cotton, and further exploration is needed in the future.

4.2. ET Signaling Transduction

ET can be sensed and bound by ET receptors that are mainly located on the endoplasmic reticulum membrane to stimulate ET response [105,106]. In Arabidopsis, five ET receptors composed of Ethylene response 1 (ETR1), ETR2, Ethylene response sensor1 (ERS1), ERS2, and Ethylene insensitive 4 (EIN4) negatively regulate ET reactions [87,107,108]. ET activates downstream signaling pathways by inactivating these receptors. In the absence of ET, constitutive triple response factor 1 (CTR1) is activated by ethylene receptors and further phosphorylates the C-terminus of Ethylene-Insensitive 2 (EIN2-CEND), causing it to become inactive [109]. Ethylene-Insensitive 3 (EIN3) and EIL1 (EIN3-like1), downstream of EIN2, are degraded by two F-box proteins, EBF1/2 (EIN3 binds to F-box1/2) in the nucleus, thereby blocking the transcription of downstream target genes in the ethylene response [110,111,112]. Under stress conditions, ET increases in plants and is bound to ethylene receptors. As a result, CTR1 is inactivated. Subsequently, EIN2 is dephosphorylated and cleaved to release the EIN2-CEND, which can shuttle into the nucleus with the assistance of EIN2 nuclear-associated protein 1 (ENAP1) and enhance the activity of EIN3/EIL1 [113]. Due to the cascade reaction, downstream target genes that respond to ET, such as ethylene-response factors (ERFs), are regulated and then affect plant growth and development [114,115,116,117]. A summary of ethylene signal transduction pathways is shown in Figure 2.
CTR1 is a Raf-like Ser/Thr protein kinase that negatively regulates the ethylene signal transduction pathway, whereas GhCTR1 expression was not induced by ET or salt but by SA, GA, and ABA [118]. EIN3/EIL1 are positive regulators of the ethylene signaling pathways and have been characterized in a lot of plants. Under the stimulation of exogenous ACC, the homologous gene GhEIN3 of EIN3/EIL1 could be induced. A previous study showed that the transgenic Arabidopsis-containing GhEIN3 had strong salt tolerance, indicating that GhEIN3 may play a role in plant responses to salt stress by regulating the ROS and ABA pathways [119]. ERFs belong to the APETALA2(AP2)/ERFs family and play an important role in plant stress signaling pathways [120]. Thus far, a series of ERFs-related genes and their functions have been identified and analyzed in cotton, such as the overexpression of GhERF38 and GhERF13.12 in Arabidopsis, which improved salt tolerance and revealed dynamic changes in relevant biochemical parameters in transgenic lines [121,122]. In another study, virus-induced silencing of cotton GhERF12 enhanced the sensitivity of transgenic plants to salt stress and reduced the activity of SOD and POD [123]. Similarly, the silencing of GhERF4L and GhERF54L in cotton led to reduced resistance to salt stress [124]. In addition, studies had found that each ERF transcription factor contained an AP2 domain, which was the DNA binding domain of ERFs and the key to the regulation of ERFs transcription [125]. A total of 220 genes containing a single AP2 domain sequence has been identified in upland cotton. Transcriptome and qRT-PCR analysis found that Ghi-ERF2D.6, Ghi-ERF-12D.13, Ghi-ERF-6D.1, Ghi-ERF-7A.6, and Ghi-ERF11D.5 could respond to salt stress [126]. Transcriptome meta-analysis and network topology analysis showed that salt stress could increase the expression level of GhERF109, and further qRT-PCR analysis showed that GhERF109 could indeed respond to salt stress [127]. Previous studies have found that GhERF1, GhERF2, GhERF3, GhERF4, GhERF5, and GhERF6 mediate plant tolerance to salt and drought [128,129,130,131]. ET correlation factors in cotton are summarized in Table 1.

5. Roles of Polyamines (PAs) in Plant Growth and Development under Salt Stress

PAs are a class of low-molecular aliphatic polycations with strong biological activity, widely present in various prokaryotes and eukaryotes [132]. PAs mainly consist of putrescine (Put, a diamine), spermidine (Spd, a triamine), spermine (Spm, a tetramine), and thermospermine (a structural isomer of spermine) [133,134]. In plants, PAs participate in a variety of physiological processes such as organogenesis, embryogenesis, flower development, leaf aging, fruit ripening, etc. [20,135,136]. Under abiotic stress, PAs can not only bind to negatively charged membrane phospholipids to maintain the permeability of cell membranes and the function of membrane proteins but also bind to anionic macromolecules (nucleic acids, proteins, etc.) to affect their physiological functions [137]. Moreover, PAs can partially replace the function of antioxidant enzymes to remove ROS and reduce oxidative damage [138,139]. Up to date, several reports have revealed that PAs regulate plant morphological growth parameters, photosynthetic pigment contents, stress-related indicators, antioxidant enzyme contents, and non-enzymatic antioxidant contents in rapeseed [140], cucumber [141], wheat [142], rice [143] and Calendula officinalis L. [144].

5.1. PAs Biosynthesis

5.1.1. Put Biosynthesis

Put is a central product in PAs’ biosynthetic pathways [145]. Put mainly has three biosynthetic pathways. The first is the direct synthesis of Put from ornithine (ORN) catalyzed by ornithine decarboxylase (ODC) [146]. The second is to synthesize Put from arginine (Arg) under the ordered catalysis of arginine decarboxylase (ADC), argmatine iminohydrolase (AIH), and N-carbamoylputrescine amidohydrolase (CPA) [147]. In the last, Arg first degraded to citrulline (Cit) catalyzed by Arginine deaminase (ADI) and then catalyzed into Put by citrulline decarboxylase (CDC) decarboxylase (Figure 3) [148].
The Arabidopsis genome contains two ADCs genes (AtADC1 and AtADC2), which have different expression patterns. AtADC1 was almost not expressed in seeds, roots, and leaves, whereas AtADC2 was highly expressed in these tissues [149]. Under salt stress, the overexpression of AtADC2 can regulate the activity of SOD and CAT, thereby improving the salt tolerance of Arabidopsis [150]. The virus-induced silencing of cotton GhADC2 enhanced plant tolerance to salt stress and reduced H2O2 content [24].

5.1.2. Spd and Spm Biosynthesis

Put, as a synthetic precursor for Spd and Spm, is catalyzed into Spd by Spd synthetase (SPDS), and then Spm synthase (SPMS) catalyzes Spd into Spm [151,152]. During this process, decarboxyl-SAM (dcSAM), a product catalyzed from SAM by SAM decarboxylase (SAMDC), provides aminopropyl groups for SPDS and SPMS-catalyzing reactions [153] (Figure 3).
Among these genes encoding key enzymes, SAMDC plays a crucial role in regulating the biosynthesis of Spd and Spm [136]. For example, the overexpression of FvSAMDC enhanced salt tolerance in tobacco [154]. In Arabidopsis, the overexpression of CsSAMDC2 exhibited higher levels of salt and drought tolerance than its wild type [155]. Meanwhile, the ectopic expression of GhSAMDC3 can improve salt tolerance in Arabidopsis by increasing Spd content and activating salt-tolerance-related genes [156]. In addition, studies have shown that AhSAMDC and TrSAMDC1 can improve plant resistance to salt stress by increasing PAs content and antioxidant enzyme activities [157,158].

5.2. PAs Catabolism

The catabolism of PAs in plants relies on the oxidative deamination catalyzed by amino oxidases (AOs), mainly including cupramine oxidases (DAOs) and FAD-dependent polyamine oxidases (PAOs) [152,159]. DAOs convert Put into H2O2, amino, and Δ1-pyrroline. Δ1-pyrroline can be further converted to γ-aminobutyric acid (GABA) [160]. PAOs mainly decompose Spd and Spm to produce 1,3-diaminopropane and H2O2 (Figure 3) [161]. Five PAOs (AtPAO1 to AtPAO5) have been identified in Arabidopsis, of which AtPAO1 and AtPAO5 are located in the cytoplasm, and AtPAO2 to AtPAO4 are located in peroxisomes [162,163,164].
PAOs play an important role in plant growth, development, and stress response. For example, CsPAO2 can interact with CsPSA3 to improve salt tolerance of cucumber by affecting photosynthesis and promoting PAs conversion [165]. Furthermore, the overexpression of CsPAO3 in Arabidopsis can promote seed germination and root growth in NaCl-containing media and alleviate growth inhibition induced by salt stress [166]. Liu et al. [167] found that OsPAO3 in rice was upregulated under salt stress during the germination stage, which increased PAO activity and led to an increase in PAs content of in salt-tolerant strains. Meanwhile, the increased PAs could improve the activity of ROS-scavenging enzymes to eliminate the excessive accumulation of H2O2 and ultimately elevate the salt tolerance of rice during germination. When GhPAO3 was overexpressed in Arabidopsis, the salt tolerance was promoted [168]. In addition, PAs can also improve plant resistance to diseases and drought [169,170,171,172]. The above genes related to salt stress in P biosynthesis and catabolism in various crops are summarized in Table 2.

6. Cross-Talk between ET and PAs under Salt Stress

ET and PAs share the same precursor SAM, indicating that changes in ET content may affect the homeostasis of PAs [173]. An increasing number of studies have found that salt treatment led to an increase in ET, ACC, Spd, and Spm in peppers, lettuce, spinach, and beetroot, indicating that the SAM pool is high enough to support ET and PAs biosynthesis under conditions, and ET and PAs can coexist without competition [174]. It can be seen that the interaction between ET and PAs depends more on the content of SAM in plants. In addition, co-regulation of endogenous ET and PAs in plants can enhance their tolerances to salt stress. For example, the interaction between CsCDPK6 and CsSAMS1 negatively regulates the biosynthesis of plant ET, inhibiting SAMDC, resulting in a decrease in Put conversion to Spd and Spm [86]. Wu et al. [155] overexpressed the PAs related gene CsSAMDC2 in Arabidopsis to improve salt tolerance, and the expression levels of the ET and ABA responding genes also increased. In a recent report, Freitas et al. [175] found that ET regulated the H2O2 content derived from PA decomposition enzymes, inducing salt tolerance in maize. Moreover, ET participated in salt-induced oxidative stress in ripening tomato fruits and regulated the metabolic level of PAs [176]. In summary, a series of key genes in the biosynthesis pathways of ET and PAs have been studied to dissect the relationship between ET and PAs. However, the exact interaction mechanism between ET and PAs under salt stress still needs further research.

7. Roles of SAM Transporters in Plants

SAM is only synthesized in the cytoplasm and is necessary for the methylation of DNA, RNA, and protein in chloroplasts and mitochondria. Chloroplasts and mitochondria are semi-autonomous organelles and need the help of SAM transporters to transfer SAM into them [177,178]. Two genes, AtSAMC1 (At4g39460) and AtSAMC2 (At1g34065), homologous to yeast (Sam5p) and mammalian (SAMC) SAM transporters, are present in Arabidopsis. AtSAMC1 is located in the plastid, whereas AtSAMC2 is the dual localization of plastid and mitochondrial membranes [28]. The expression of AtSAMC2 was almost undetectable in roots, stems, leaves, flowers, and seedlings, whereas AtSAMC1 was expressed in the above organs and especially had a very high expression in seedlings [179]. Chloroplast SAM transporters are important bonds in many biological processes between cytoplasm and chloroplasts, such as the single-carbon metabolism for maintaining methylation reactions and other SAM-dependent functions in chloroplasts, as well as removing the byproduct SAHC derived from the methylation reaction [180]. Mitochondrial SAM transporters belong to the mitochondrial carrier (MC) family and are important for the methylation reaction in mitochondria [181]. SAM transporters are mainly responsible for the input of SAM in organelles and the output of SAHC to maintain the normal metabolism homeostasis of SAM in plant cells. Once the SAM transporter loses function, it will affect the synthesis of SAM, which is a precursor for ethylene, polyamines, glycine betaine, and lignin. Then, it may ultimately affect the survival of plants in adversity.
We screened out the salt-tolerant genes through comparative transcriptome analysis in G. hirsutum, G. hirsutum race marie-galante, G. tomentosum, and G. barbadense (PRJNA933089). Among them, Gh_A05G2087 (named GhSAMC) was found to be a homologous gene of AtSAMC1 and AtSAMC2 in cotton after sequence alignment (Supplementary Figure S1). To investigate whether GhSAMC was affected by salt stress in cotton leaves, the TM-1 seedlings of upland cotton were first treated with 250 mM NaCl solution and then sampled at 0 h, 3 h, 6 h, 12 h, and 24 h for qRT-PCR analysis(The primers in the Supplementary Table S1). The results showed that GhSAMC was affected by salt stress, and its expression revealed a trend that it was rising first, reaching its peak at 6 h, and then falling with the extension of salt stress time (Figure 4E). We further performed virus induced gene silencing (VIGS) to test whether GhSAMC was necessary for cotton to tolerate salt stress (Materials and Methods in Supplementary). When newly emerged leaves were infiltrated with TRV:CLA1, they exhibited an albino phenotype (Figure 4A). The expression levels of GhSAMC were measured by qRT-PCR to confirm the silencing efficiency, and the results showed that GhSAMC had been silenced completely (Figure 4B). Treated GhSAMC silencing (TRV:SAMC) and control (TRV:00) plants with 250 mM NaCl solution at the three-leaf stage, it was found that TRV:SAMC seedlings exhibited more serious symptoms of wilting and lodging than TRV:00 plants (Figure 4C,D). Subsequently, we detected the malondialdehyde (MDA) content and total antioxidant capacity (T-AOC) of VIGS cotton plant leaves (Figure 4F,G), and the results showed that TRV:SAMC accumulated more ROS than TRV:00 seedlings. Therefore, we speculated that GhSAMC could regulate the accumulation of ROS under salt stress conditions to improve the salt tolerance of cotton. However, further research is needed to determine whether GhSAMC affects the synthesis of ET and PAs by regulating the transport of SAM, thereby regulating salt stress.
Furthermore, we carried out the transcriptomic analysis for leaves of VIGS plants (TRV:SAMC, TRV:00) treated with 250 mM of saline solution (PRJNA937453). Compared with CK (TRV:00), a total of 1169 differentially expressed genes (DEGs) were identified (Supplementary Table S2), including 470 upregulated genes and 699 downregulated genes (Figure 5C). GO enrichment analysis revealed that the DEGs were mainly divided into three categories, exemplified as biological processes, cell compositions, and molecular functions (Figure 5A). Moreover, GO categories were concentrated in the functional groups of oxidoreductase activity, oxidative stress response, cell membrane components, and cell walls. These observations indicated that the cell walls and membranes first played a barrier role in response to salt stress, and then the oxidoreductase genes were mobilized into cells to cope with the accumulation of ROS. Thus, the transcriptome data implied that GhSAMC may alleviate salt damage by regulating ROS. Plants can respond to salt stress through three ways: osmotic regulation, ion regulation, and oxidative regulation. Therefore, we mapped the differential genes related to these three regulatory methods into a heat map (Supplementary Figure S2). Among them, there were many DEGs related to oxidation–reduction, mainly including the POD and cytochrome P450 genes. The KEGG results enriched 22 metabolic pathways, including lipid and amino acid metabolism, glycolysis/gluconeogenesis, starch and sucrose metabolism, and carbon metabolism, suggesting that GhSAMC may be involved in regulating these metabolic pathways to help cotton cope with salt stress (Figure 5B).
The tolerance of plants to salt stress is regulated by a series of TFs. Through TFs analysis, 58 TFs related to salt stress were screened, including HD-ZIP, DREB, MYC, Dof, HSF, C2H2, WRKY, MYB, NAC, ERF, and bHLH (Figure 5E). Among them, the TFs families of bHLH, ERF, NAC, and MYB contained significantly more DEGs than other families, including 13, 10, 8, and 8 DEGs, respectively. Previous studies have shown that HD-ZIP [182,183], DREB [184,185], MYC [186,187], Dof [188,189], HSF [190,191], C2H2 [192,193], WRKY [194,195], MYB [196,197], NAC [198,199], and bHLH [200,201] could work alone or synergistically to regulate plant tolerance to salt stress. Additionally, the above TFs could crosslink with ET [202,203,204,205,206,207,208]. For example, the apple NAC transcription factor (MdNAC047) enhances plant resistance to salt stress by increasing ET synthesis in plants [209]. MYB108A can activate the expression of ACS1 to promote the formation of ET and ultimately improve the salt tolerance of grapes [210]. The WRKY29 transcription factor in Arabidopsis positively regulates the expression of the ACS5, ACS6, ACS8, ACS11, and ACO5 genes, thereby promoting ET production [211].
In addition, phytohormones are also involved in the regulation of plant salt stress. Through the analysis of transcriptome data, a total of six phytohormones were screened, including ET, gibberellic acid (GA), auxin, salicylic acid (SA), ABA, and jasmonic acid (JA) (Figure 5D). Among them, ET and GA contained more differential genes, including 10 and 8 DEGs, respectively. The above six phytohormones play an important role in plant growth and development and can synergistically regulate plant tolerance to salt stress [212,213,214]. For example, under salt stress, the interaction between ET and JA inhibits the growth of rice seed roots [215]. Tomato WRKY23 gene can regulate ET and auxin pathways to improve the tolerance of transgenic Arabidopsis to salt stress [216]. In addition, under normal conditions, GA and auxin can regulate plant growth together with ET, whereas SA, JA, and ABA have antagonistic effects on ET in various stress responses [212].
Finally, combining experimental data and transcriptome data, we summarized the working model for GhSAMC in cotton salt stress regulation: GhSAMC can regulate the rapid accumulation of SAM transporters when cotton is under salt stress. These SAM transporters accelerate the transport of SAM from cytoplasm to organelle and use SAM as a precursor to further increase the synthesis of ET and PAs. The increase in ET and PAs can inhibit the accumulation of ROS and maintain its homeostasis to alleviate salt stress in cotton (Figure 5F). In this process, related TFs (such as MYB, ERF, WRKY and NAC, etc.) and phytohormones (such as ABA, GA and SA, etc.) can also regulate ET biosynthesis and thus affect plant response to salt stress. The mechanism by which GhSAMC regulates plant salt stress needs further exploration.

8. Conclusions and Future Perspectives

Salt stress is considered to be one of the main factors limiting crop growth and yield. Therefore, plants have developed various strategies for survival under salt stress, such as osmotic regulation, ion, and ROS homeostasis [217]. Previous studies have shown that oxidative stress caused by salt stress is the most harmful to plants, and our analysis of transcriptome data also shows that a large number of oxidation–reduction-related factors are mobilized when plants are subjected to salt stress. In addition, plant tolerance to salt is regulated by multiple genes, not only a variety of transcription factors (such as EFR, WRKY, MYB, etc.) but also a variety of phytohormones (such as ET, ABA, SA, etc.).
As a general methyl donor for various methylation metabolites, SAM participates in the salt stress response through its derived metabolites (PAs and ET) [218,219]. This review focused on the biosynthesis and signal transduction of ET and PAs in salt stress, as well as the factors involved in these processes. In addition, the cotton SAM transporter gene GhSAMC, which is homologous to Arabidopsis AtSAMC1 and AtSAMC2, was also identified. The virus-induced gene silencing experiment suggested that it may have positively regulated the salt tolerance of cotton.
Cotton cultivation has gradually shifted towards saline–alkali land because of the salinization and the need for more arable land to support the ever-increasing population. To improve the tolerance of cotton to saline–alkali stress and increase yield and fiber quality, the following aspects can be considered in the future. First, a certain number of ion transporters have been discovered in cotton, which can alleviate salt stress damage by regulating ion homeostasis. Therefore, it is necessary to further dissect the exact regulatory mechanisms of these ion transporters, and then the knowledge learned from them can be used to improve the salt tolerance of cotton. Second, the impact of ET on cotton growth and development is multifaceted. To date, the regulatory mechanism of ET on cotton fiber development has been well studied, but the specific molecular mechanism of ET in salt stress regulation remains unclear [117,220,221]. Third, it has been proven that PAs can regulate the response to salt stress in many species; however, only a few reports have been found in cotton. Thus, it remains to be studied how PAs function in response to salt stress in cotton. Moreover, as a synthetic precursor for the synthesis of ET and PAs, SAM regulates plant salt stress by controlling the dynamic homeostasis of ET and PAs, thereby enabling SAM biosynthesis to affect responses to salt stress. Our study suggests that silencing GhSAMC enhances salt stress damage in cotton. Therefore, in the future, we can use transgenic technology to transfer GhSAMC into the cotton or use gene editing technology to knock out the GhSAMC gene in cotton, so as to further study the specific molecular mechanism of GhSAMC regulating cotton salt stress in cotton. Taken together, SAM and its derivates had very important roles in the response to salt stress, and the investigations summarized in this review revealed the complex biosynthesis and regulatory network of salt tolerance in plants and cotton.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24119517/s1. References [222,223,224,225,226,227] are cited in the Supplementary Materials.

Author Contributions

X.W., X.M., F.Z. (Fuyong Zhao) and D.Y. conceived and designed the experiments. L.Y., X.Z., J.H., X.R., X.P., Y.L., K.H. and F.Z. (Fei Zhang) performed the experiments and analyzed the data. L.Y. drafted the manuscript. X.W., W.L., F.Z. (Fuyong Zhao), X.M. and D.Y. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key R&D Program of China (2021YFF1000100), Science and Technology Program of Changji Hui Autonomous Prefecture (Grant No. 2021Z01-01), Xinjiang Tianshan Talents Program (2021), China Agriculture Research System (CARS-15-07), Innovation Project of the Chinese Academy of Agricultural Sciences (Grant No. CAAS-ASTIP-ICR-KP-2021-01), and Natural Science Foundation of Henan (Grant No. 202300410550).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The sequencing data for this study have been deposited into the Sequence Read Archive under accession PRJNA933089 and PRJNA937453.

Acknowledgments

We sincerely thank Xiongfeng Ma (Cotton research institute) for his valuable advice and financial support of this research. To the entire research group, friends, and any other person who contributed, we are grateful for your help so much.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ashraf, M. Breeding for Salinity Tolerance in Plants. Crit. Rev. Plant Sci. 1994, 13, 17–42. [Google Scholar] [CrossRef]
  2. Maas, E.V.; Hoffman, G.J. Crop salt tolerance–current assessment. J. Irrig. Drain. Div. 1977, 103, 115–134. [Google Scholar] [CrossRef]
  3. Ahmad, S.; Khan, N.I.; Iqbal, M.Z.; Hussain, A.; Hassan, M. Salt Tolerance of Cotton (Gossypium hirsutum L.). Asian J. Plant Sci. 2002, 1, 715–719. [Google Scholar] [CrossRef]
  4. Abdelraheem, A.; Esmaeili, N.; O’Connell, M.; Zhang, J.F. Progress and perspective on drought and salt stress tolerance in cotton. Ind. Crops Prod. 2019, 130, 118–129. [Google Scholar] [CrossRef]
  5. Brugnoli, E.; Lauteri, M. Effects of Salinity on Stomatal Conductance, Photosynthetic Capacity, and Carbon Isotope Discrimination of Salt-Tolerant (Gossypium hirsutum L.) and Salt-Sensitive (Phaseolus vulgaris L.) C(3) Non-Halophytes. Plant Physiol. 1991, 95, 628–635. [Google Scholar] [CrossRef] [PubMed]
  6. Huang, G.; Wu, Z.; Percy, R.G.; Bai, M.; Li, Y.; Frelichowski, J.E.; Hu, J.; Wang, K.; Yu, J.Z.; Zhu, Y. Genome sequence of Gossypium herbaceum and genome updates of Gossypium herbaceum and Gossypium hirsutum provide insights into cotton A-genome evolution. Nat. Genet. 2020, 52, 516–524. [Google Scholar] [CrossRef]
  7. Wang, K.; Wang, Z.; Li, F.; Ye, W.; Wang, J.; Song, G.; Yue, Z.; Cong, L.; Shang, H.; Zhu, S.; et al. The draft genome of a diploid cotton Gossypium raimondii. Nat. Genet. 2012, 44, 1098–1103. [Google Scholar] [CrossRef]
  8. Mehari, T.G.; Hou, Y.; Xu, Y.; Umer, M.J.; Shiraku, M.L.; Wang, Y.; Wang, H.; Peng, R.; Wei, Y.; Cai, X.; et al. Overexpression of cotton GhNAC072 gene enhances drought and salt stress tolerance in transgenic Arabidopsis. BMC Genom. 2022, 23, 648. [Google Scholar] [CrossRef]
  9. Sadau, S.B.; Ahmad, A.; Tajo, S.M.; Ibrahim, S.; Kazeem, B.B.; Wei, H.; Yu, S. Overexpression of GhMPK3 from Cotton Enhances Cold, Drought, and Salt Stress in Arabidopsis. Agronomy 2021, 11, 1049. [Google Scholar] [CrossRef]
  10. Fontecave, M.; Atta, M.; Mulliez, E. S-adenosylmethionine: Nothing goes to waste. Trends Biochem. Sci. 2004, 29, 243–249. [Google Scholar] [CrossRef]
  11. Bouvier, F.; Linka, N.; Isner, J.C.; Mutterer, J.; Weber, A.P.; Camara, B. Arabidopsis SAMT1 defines a plastid transporter regulating plastid biogenesis and plant development. Plant Cell 2006, 18, 3088–3105. [Google Scholar] [CrossRef]
  12. Munawar, W.; Hameed, A.; Khan, M. Differential Morphophysiological and Biochemical Responses of Cotton Genotypes under Various Salinity Stress Levels during Early Growth Stage. Front. Plant Sci. 2021, 12, 622309. [Google Scholar] [CrossRef]
  13. Vaseva, I.I.; Simova-Stoilova, L.; Kirova, E.; Mishev, K.; Depaepe, T.; Van Der Straeten, D.; Vassileva, V. Ethylene signaling in salt-stressed Arabidopsis thaliana ein2-1 and ctr1-1 mutants A dissection of molecular mechanisms involved in acclimation. Plant Physiol. Biochem. 2021, 167, 999–1010. [Google Scholar] [CrossRef] [PubMed]
  14. Choudhury, A.R.; Roy, S.K.; Trivedi, P.; Choi, J.; Cho, K.; Yun, S.H.; Walitang, D.I.; Park, J.; Kim, K.; Sa, T. Label-free proteomics approach reveals candidate proteins in rice (Oryza sativa L.) important for ACC deaminase producing bacteria-mediated tolerance against salt stress. Environ. Microbiol. 2022, 8, 3612–3624. [Google Scholar] [CrossRef] [PubMed]
  15. Khan, S.; Sehar, Z.; Fatma, M.; Mir, I.R.; Iqbal, N.; Tarighat, M.A.; Abdi, G.; Khan, N.A. Involvement of ethylene in melatonin-modified photosynthetic-N use efficiency and antioxidant activity to improve photosynthesis of salt grown wheat. Plant Physiol. 2022, 174, e13832. [Google Scholar] [CrossRef] [PubMed]
  16. Shakar, M.; Yaseena, M.; Mahmoodb, R.; Ahmad, I. Calcium carbide induced ethylene modulate biochemical profile of Cucumis sativus at seed germination stage to alleviate salt stress. Sci. Hortic. 2016, 213, 179–185. [Google Scholar] [CrossRef]
  17. Gharbi, E.; Martínez, J.P.; Benahmed, H.; Lepoint, G.; Vanpee, B.; Quinet, M.; Lutts, S. Inhibition of ethylene synthesis reduces salt-tolerance in tomato wild relative species Solanum chilense. J. Plant Physiol. 2017, 210, 24–37. [Google Scholar] [CrossRef] [PubMed]
  18. Ahmad, Z.; Tahir, S.; Rehman, A.; Niazi, N.K.; Abid, M.; Amanullah, M. Effect of Substrate Dependent Ethylene on Cotton (Gossypium hirsutum L.) at Physiological and Molecular Levels under Salinity Stress. J. Plant Nutr. 2015, 38, 1913–1928. [Google Scholar] [CrossRef]
  19. Gill, S.S.; Tuteja, N. Polyamines and abiotic stress tolerance in plants. Plant Signal. Behav. 2010, 5, 26–33. [Google Scholar] [CrossRef]
  20. Gholami, M.; Fakhari, A.R.; Ghanati, F. Selective Regulation of Nicotine and Polyamines Biosynthesis in Tobacco Cells by Enantiomers of Ornithine. Chirality 2013, 25, 22–27. [Google Scholar] [CrossRef] [PubMed]
  21. Campestre, M.P.; Bordenave, C.D.; Origone, A.C.; Menéndez, A.B.; Ruiz, O.A.; Rodríguez, A.A.; Maiale, S.J. Polyamine catabolism is involved in response to salt stress in soybean hypocotyls. J. Plant Physiol. 2011, 168, 1234–1240. [Google Scholar] [CrossRef] [PubMed]
  22. Zhong, M.; Yue, L.; Liu, W.; Qin, H.; Lei, B.; Huang, R.; Yang, X.; Kang, Y. Genome-Wide Identification and Characterization of the Polyamine Uptake Transporter (Put) Gene Family in Tomatoes and the Role of Put2 in Response to Salt Stress. Antioxidants 2023, 12, 228. [Google Scholar] [CrossRef] [PubMed]
  23. Lechowska, K.; Wojtyla, A.; Quinet, M.; Kubala, S.; Lutts, S.; Garnczarska, M. Endogenous Polyamines and Ethylene Biosynthesis in Relation to Germination of Osmoprimed Brassica napus Seeds under Salt Stress. Int. J. Mol. Sci. 2022, 23, 349. [Google Scholar] [CrossRef] [PubMed]
  24. Gu, Q.; Ke, H.; Liu, C.; Lv, X.; Sun, Z.; Liu, Z.; Rong, W.; Yang, J.; Zhang, Y.; Wu, L.; et al. A stable QTL qSalt-A04-1 contributes to salt tolerance in the cotton seed germination stage. Theor. Appl. Genet. 2021, 134, 2399–2410. [Google Scholar] [CrossRef] [PubMed]
  25. Agrimi, G.; Di Noia, M.A.; Marobbio, C.M.; Fiermonte, G.; Lasorsa, F.M.; Palmieri, F. Identification of the human mitochondrial S-adenosylmethionine transporter: Bacterial expression, reconstitution, functional characterization and tissue distribution. Biochem. J. 2004, 379, 183–190. [Google Scholar] [CrossRef] [PubMed]
  26. Petrotta-Simpson, T.F.; Talmadge, J.E.; Spence, K.D. Specificity and genetics of S-adenosylmethionine transport in Saccharomyces cerevisiae. J. Bacteriol. 1975, 123, 516–522. [Google Scholar] [CrossRef] [PubMed]
  27. Dridi, L.; Ahmed Ouameur, A.; Ouellette, M. High Affinity S-Adenosylmethionine Plasma Membrane Transporter of Leishmania Is a Member of the Folate Biopterin Transporter (FBT) Family. J. Biol. Chem. 2010, 285, 19767–19775. [Google Scholar] [CrossRef] [PubMed]
  28. Palmieri, L.; Arrigoni, R.; Blanco, E.; Carrari, F.; Zanor, M.I.; Studart-Guimaraes, C.; Fernie, A.R.; Palmieri, F. Molecular Identification of an Arabidopsis S-Adenosylmethionine Transporter. Analysis of Organ Distribution, Bacterial Expression, Reconstitution into Liposomes, and Functional Characterization. Plant Physiol. 2006, 142, 855–865. [Google Scholar] [CrossRef]
  29. Kraidlova, L.; Schrevens, S.; Tournu, H.; Van Zeebroeck, G.; Sychrova, H.; Van Dijck, P. Characterization of the Candida albicans Amino Acid Permease Family: Gap2 Is the Only General Amino Acid Permease and Gap4 Is an S-Adenosylmethionine (SAM) Transporter Required for SAM-Induced Morphogenesis. Msphere 2016, 1, e00284-16. [Google Scholar] [CrossRef]
  30. Abiala, M.A.; Abdelrahman, M.; Burritt, D.J.; Tran, L. Salt stress tolerance mechanisms and potential applications of legumes for sustainable reclamation of salt-degraded soils. Land Degrad. Dev. 2018, 29, 3812–3822. [Google Scholar] [CrossRef]
  31. Park, H.J.; Kim, W.Y.; Yun, D.J. A New Insight of Salt Stress Signaling in Plant. Mol. Cells. 2016, 39, 447–459. [Google Scholar] [CrossRef]
  32. Muchate, N.S.; Nikalje, G.C.; Rajurkar, N.S.; Suprasanna, P.; Nikam, T.D. Plant Salt Stress: Adaptive Responses, Tolerance Mechanism and Bioengineering for Salt Tolerance. Bot. Rev. 2016, 82, 371–406. [Google Scholar] [CrossRef]
  33. Sharif, I.; Aleem, S.; Farooq, J.; Rizwan, M.; Younas, A.; Sarwar, G.; Chohan, S.M. Salinity stress in cotton: Effects, mechanism of tolerance and its management strategies. Physiol. Mol. Biol. Plants 2019, 25, 807–820. [Google Scholar] [CrossRef] [PubMed]
  34. Maryum, Z.; Luqman, T.; Nadeem, S.; Khan, S.; Wang, B.; Ditta, A.; Khan, M. An overview of salinity stress, mechanism of salinity tolerance and strategies for its management in cotton. Front. Plant Sci. 2022, 13, 907937. [Google Scholar] [CrossRef]
  35. Sanchez, D.H.; Siahpoosh, M.R.; Roessner, U.; Udvardi, M.; Kopka, J. Plant metabolomics reveals conserved and divergent metabolic responses to salinity. Physiol. Plant. 2008, 132, 209–219. [Google Scholar] [CrossRef] [PubMed]
  36. Flowers, T.J.; Colmer, T.D. Plant salt tolerance: Adaptations in halophytes. Ann. Bot. 2015, 115, 327–331. [Google Scholar] [CrossRef]
  37. Mishra, A.; Tanna, B. Halophytes: Potential Resources for Salt Stress Tolerance Genes and Promoters. Front. Plant Sci. 2017, 8, 829. [Google Scholar] [CrossRef]
  38. Feki, K.; Quintero, F.J.; Khoudi, H.; Leidi, E.O.; Masmoudi, K.; Pardo, J.M.; Brini, F. A constitutively active form of a durum wheat Na+/H+ antiporter SOS1 confers high salt tolerance to transgenic Arabidopsis. Plant Cell Rep. 2014, 33, 277–288. [Google Scholar] [CrossRef] [PubMed]
  39. Che, B.N.; Cheng, C.; Fang, J.J.; Liu, Y.M.; Jiang, L.; Yu, B.J. The Recretohalophyte Tamarix TrSOS1 Gene Confers Enhanced Salt Tolerance to Transgenic Hairy Root Composite Cotton Seedlings Exhibiting Virus-Induced Gene Silencing of GhSOS1. Int. J. Mol. Sci. 2019, 20, 2930. [Google Scholar] [CrossRef] [PubMed]
  40. Fahmideh, L.; Fooladvand, Z. Isolation and Semi Quantitative PCR of Na+/H+ Antiporter (SOS1 and NHX) Genes under Salinity Stress in Kochia scoparia. Biol. Proced. Online 2018, 20, 11. [Google Scholar] [CrossRef] [PubMed]
  41. Cushman, K.R.; Pabuayon, I.C.M.; Hinze, L.L.; Sweeney, M.E.; de Los Reyes, B.G. Networks of Physiological Adjustments and Defenses, and Their Synergy with Sodium (Na+) Homeostasis Explain the Hidden Variation for Salinity Tolerance across the Cultivated Gossypium hirsutum Germplasm. Front. Plant Sci. 2020, 11, 588854. [Google Scholar] [CrossRef] [PubMed]
  42. Feng, J.; Ma, W.; Ma, Z.; Ren, Z.; Zhou, Y.; Zhao, J.; Li, W.; Liu, W. GhNHX3D, a Vacuolar-Localized Na(+)/H(+) Antiporter, Positively Regulates Salt Response in Upland Cotton. Int. J. Mol. Sci. 2021, 22, 4047. [Google Scholar] [CrossRef] [PubMed]
  43. Nieves-Cordones, M.; Miller, A.J.; Aleman, F.; Martinez, V.; Rubio, F. A putative role for the plasma membrane potential in the control of the expression of the gene encoding the tomato high-affinity potassium transporter HAK5. Plant Mol. Biol. 2008, 68, 521–532. [Google Scholar] [CrossRef] [PubMed]
  44. Pilot, G.; Gaymard, F.; Mouline, K.; Chérel, I.; Sentenac, H. Regulated expression of Arabidopsis Shaker K+ channel genes involved in K+ uptake and distribution in the plant. Plant Mol. Biol. 2003, 51, 773–787. [Google Scholar] [CrossRef] [PubMed]
  45. Ali, A.; Raddatz, N.; Pardo, J.M.; Yun, D.J. HKT sodium and potassium transporters in Arabidopsis thaliana and related halophyte species. Physiol. Plant. 2021, 171, 546–558. [Google Scholar] [CrossRef] [PubMed]
  46. Guo, Q.; Meng, S.; Tao, S.; Feng, J.; Fan, X.; Xu, P.; Xu, Z.; Shen, X. Overexpression of a samphire high-affinity potassium transporter gene SbHKT1 enhances salt tolerance in transgenic cotton. Acta Physiol. Plant. 2020, 42, 36. [Google Scholar] [CrossRef]
  47. Li, Y.; Feng, Z.; Wei, H.; Cheng, S.; Hao, P.; Yu, S.; Wang, H. Silencing of GhKEA4 and GhKEA12 Revealed Their Potential Functions under Salt and Potassium Stresses in Upland Cotton. Front. Plant Sci. 2021, 12, 789775. [Google Scholar] [CrossRef]
  48. Chen, X.G.; Wu, Z.; Yin, Z.J.; Zhang, Y.X.; Rui, C.; Wang, J.; Malik, W.A.; Lu, X.K.; Wang, D.L.; Wang, J.J.; et al. Comprehensive genomic characterization of cotton cationic amino acid transporter genes reveals that GhCAT10D regulates salt tolerance. BMC Plant Biol. 2022, 22, 441. [Google Scholar] [CrossRef]
  49. Zhao, J.; Peng, S.; Cui, H.; Li, P.; Li, T.; Liu, L.; Zhang, H.; Tian, Z.; Shang, H.; Xu, R. Dynamic Expression, Differential Regulation and Functional Diversity of the CNGC Family Genes in Cotton. Int. J. Mol. Sci. 2022, 23, 2041. [Google Scholar] [CrossRef]
  50. Fu, X.; Yang, Y.; Kang, M.; Wei, H.; Lian, B.; Wang, B.; Ma, L.; Hao, P.; Lu, J.; Yu, S.; et al. Evolution and Stress Responses of CLO Genes and Potential Function of the GhCLO06 Gene in Salt Resistance of Cotton. Front. Plant Sci. 2021, 12, 801239. [Google Scholar] [CrossRef]
  51. Myo, T.; Wei, F.; Zhang, H.; Hao, J.; Zhang, B.; Liu, Z.; Cao, G.; Tian, B.; Shi, G. Genome-wide identification of the BASS gene family in four Gossypium species and functional characterization of GhBASSs against salt stress. Sci. Rep. 2021, 11, 11342. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, D.; Li, J.; Niu, X.; Deng, C.; Song, X.; Li, W.; Cheng, Z.; Xu, Q.A.; Zhang, B.; Guo, W. GhANN1 modulates the salinity tolerance by regulating ABA biosynthesis, ion homeostasis and phenylpropanoid pathway in cotton. Environ. Exp. Bot. 2021, 185, 104427. [Google Scholar] [CrossRef]
  53. Hasanuzzaman, M.; Raihan, M.R.H.; Masud, A.A.C.; Rahman, K.; Nowroz, F.; Rahman, M.; Nahar, K.; Fujita, M. Regulation of Reactive Oxygen Species and Antioxidant Defense in Plants under Salinity. Int. J. Mol. Sci. 2021, 22, 9326. [Google Scholar] [CrossRef] [PubMed]
  54. Singh, R.; Singh, S.; Parihar, P.; Mishra, R.K.; Tripathi, D.K.; Singh, V.P.; Chauhan, D.K.; Prasad, S.M. Reactive Oxygen Species (ROS): Beneficial Companions of Plants’ Developmental Processes. Front. Plant Sci. 2016, 7, 1299. [Google Scholar] [CrossRef] [PubMed]
  55. Choudhury, S.; Panda, P.; Sahoo, L.; Panda, S.K. Reactive oxygen species signaling in plants under abiotic stress. Plant Signal. Behav. 2013, 8, e23681. [Google Scholar] [CrossRef] [PubMed]
  56. Mohsin, S.; Hasanuzzaman, M.; Bhuyan, M.; Parvin, K.; Fujita, M. Exogenous Tebuconazole and Trifloxystrobin Regulates Reactive Oxygen Species Metabolism toward Mitigating Salt-Induced Damages in Cucumber Seedling. Plants 2019, 8, 428. [Google Scholar] [CrossRef]
  57. Mangal, V.; Lal, M.K.; Tiwari, R.K.; Altaf, M.A.; Sood, S.; Kumar, D.; Bharadwaj, V.; Singh, B.; Singh, R.K.; Aftab, T. Molecular Insights into the Role of Reactive Oxygen, Nitrogen and Sulphur Species in Conferring Salinity Stress Tolerance in Plants. J. Plant Growth Regul. 2023, 42, 554–574. [Google Scholar] [CrossRef]
  58. Xie, F.; Wang, Q.; Sun, R.; Zhang, B. Deep sequencing reveals important roles of microRNAs in response to drought and salinity stress in cotton. J. Exp. Bot. 2015, 66, 789–804. [Google Scholar] [CrossRef]
  59. Rahman, M.M.; Mostofa, M.G.; Keya, S.S.; Siddiqui, M.N.; Ansary, M.M.U.; Das, A.K.; Rahman, M.A.; Tran, L.S. Adaptive Mechanisms of Halophytes and Their Potential in Improving Salinity Tolerance in Plants. Int. J. Mol. Sci. 2021, 22, 10733. [Google Scholar] [CrossRef]
  60. Khalid, M.F.; Huda, S.; Yong, M.; Li, L.; Li, L.; Chen, Z.; Ahmed, T. Alleviation of drought and salt stress in vegetables: Crop responses and mitigation strategies. Plant Growth Regul. 2023, 99, 177–194. [Google Scholar] [CrossRef]
  61. Wang, W.; Liu, D.; Chen, D.; Cheng, Y.; Zhang, X.; Song, L.; Hu, M.; Dong, J.; Shen, F. MicroRNA414c affects salt tolerance of cotton by regulating reactive oxygen species metabolism under salinity stress. RNA Biol. 2019, 16, 362–375. [Google Scholar] [CrossRef]
  62. Yan, H.; Jia, H.; Chen, X.; Hao, L.; An, H.; Guo, X. The Cotton WRKY Transcription Factor GhWRKY17 Functions in Drought and Salt Stress in Transgenic Nicotiana benthamiana through ABA Signaling and the Modulation of Reactive Oxygen Species Production. Plant Cell Physiol. 2014, 55, 2060–2076. [Google Scholar] [CrossRef] [PubMed]
  63. Yang, H.; Yang, Q.; Zhang, D.; Wang, J.; Cao, T.; Bozorov, T.A.; Cheng, L.; Zhang, D. Transcriptome Reveals the Molecular Mechanism of the ScALDH21 Gene from the Desert Moss Syntrichia caninervis Conferring Resistance to Salt Stress in Cotton. Int. J. Mol. Sci. 2023, 24, 5822. [Google Scholar] [CrossRef] [PubMed]
  64. Jia, H.; Hao, L.; Guo, X.; Liu, S.; Yan, Y.; Guo, X. A Raf-like MAPKKK gene, GhRaf19, negatively regulates tolerance to drought and salt and positively regulates resistance to cold stress by modulating reactive oxygen species in cotton. Plant Sci. 2016, 252, 267–281. [Google Scholar] [CrossRef] [PubMed]
  65. Atkinson, N.J.; Urwin, P.E. The interaction of plant biotic and abiotic stresses: From genes to the field. J. Exp. Bot. 2012, 63, 695–709. [Google Scholar] [CrossRef] [PubMed]
  66. Mishra, S.; Kumar, S.; Saha, B.; Awasthi, J.; Dey, M.; Panda, S.K.; Sahoo, L. Crosstalk between Salt, Drought, and Cold Stress in Plants: Toward Genetic Engineering for Stress Tolerance; Tuteja, N., Gill, S.S., Eds.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2016; pp. 57–88. [Google Scholar] [CrossRef]
  67. Kim, D.; Jeon, S.J.; Yanders, S.; Park, S.C.; Kim, H.S.; Kim, S. MYB3 plays an important role in lignin and anthocyanin biosynthesis under salt stress condition in Arabidopsis. Plant Cell Rep. 2022, 41, 1549–1560. [Google Scholar] [CrossRef]
  68. Ullah, A.; Ul, Q.M.; Nisar, M.; Hazrat, A.; Rahim, G.; Khan, A.H.; Hayat, K.; Ahmed, S.; Ali, W.; Khan, A.; et al. Characterization of a novel cotton MYB gene, GhMYB108-like responsive to abiotic stresses. Mol. Biol. Rep. 2020, 47, 1573–1581. [Google Scholar] [CrossRef] [PubMed]
  69. Khanale, V.; Bhattacharya, A.; Satpute, R.; Char, B. Brief bioinformatics identification of cotton bZIP transcription factors family from Gossypium hirsutum, Gossypium arboreum and Gossypium raimondii. Plant Biotechnol. Rep. 2021, 15, 493–511. [Google Scholar] [CrossRef]
  70. Azeem, F.; Tahir, H.; Usman, I.; Tayyaba, S. A genome-wide comparative analysis of bZIP transcription factors in G. arboreum and G. raimondii (Diploid ancestors of present-day cotton). Physiol. Mol. Biol. Plants 2020, 3, 433–444. [Google Scholar] [CrossRef] [PubMed]
  71. Ullah, A.; Sun, H.; Hakim; Yang, X.; Zhang, X. A novel cotton WRKY gene, GhWRKY6-like, improves salt tolerance by activating the ABA signaling pathway and scavenging of reactive oxygen species. Physiol. Plant. 2018, 162, 439–454. [Google Scholar] [CrossRef]
  72. Çelik, Ö.; Meriç, S.; Ayan, A.; Atak, Ç. Epigenetic analysis of WRKY transcription factor genes in salt stressed rice (Oryza sativa L.) plants. Environ. Exp. Bot. 2019, 159, 121–131. [Google Scholar] [CrossRef]
  73. Guo, X.; Ullah, A.; Siuta, D.; Kukfisz, B.; Iqbal, S. Role of WRKY Transcription Factors in Regulation of Abiotic Stress Responses in Cotton. Life 2022, 12, 1410. [Google Scholar] [CrossRef] [PubMed]
  74. Shah, S.T.; Pang, C.; Fan, S.; Song, M.; Arain, S.; Yu, S. Isolation and expression profiling of GhNAC transcription factor genes in cotton (Gossypium hirsutum L.) during leaf senescence and in response to stresses. Gene 2013, 531, 220–234. [Google Scholar] [CrossRef] [PubMed]
  75. Latif, A.; Azam, S.; Shahid, N.; Javed, M.R.; Haider, Z.; Yasmeen, A.; Sadaqat, S.; Shad, M.; Husnain, T.; Rao, A.Q. Overexpression of the AGL42 gene in cotton delayed leaf senescence through downregulation of NAC transcription factors. Sci. Rep. 2022, 12, 21093. [Google Scholar] [CrossRef] [PubMed]
  76. Trishla, V.S.; Kirti, P.B. Structure-function relationship of Gossypium hirsutum NAC transcription factor, GhNAC4 with regard to ABA and abiotic stress responses. Plant Sci. 2021, 302, 110718. [Google Scholar] [CrossRef] [PubMed]
  77. Sun, H.; Chen, L.; Li, J.; Hu, M.; Ullah, A.; He, X.; Yang, X.; Zhang, X. The JASMONATE ZIM-Domain Gene Family Mediates JA Signaling and Stress Response in Cotton. Plant Cell Physiol. 2017, 58, 2139–2154. [Google Scholar] [CrossRef]
  78. Abid, M.A.; Liang, C.; Malik, W.; Meng, Z.; Tao, Z.; Meng, Z.; Ashraf, J.; Guo, S.; Zhang, R. Cascades of Ionic and Molecular Networks Involved in Expression of Genes Underpin Salinity Tolerance in Cotton. J. Plant Growth Regul. 2018, 37, 668–679. [Google Scholar] [CrossRef]
  79. Shah, W.H.; Rasool, A.; Saleem, S.; Mushtaq, N.U.; Tahir, I.; Hakeem, K.R.; Rehman, R.U. Understanding the Integrated Pathways and Mechanisms of Transporters, Protein Kinases, and Transcription Factors in Plants under Salt Stress. Int. J. Genom. 2021, 2021, 5578727. [Google Scholar] [CrossRef]
  80. Choudhary, P.; Pramitha, L.; Rana, S.; Verma, S.; Aggarwal, P.R.; Muthamilarasan, M. Hormonal crosstalk in regulating salinity stress tolerance in graminaceous crops. Physiol. Plant. 2021, 173, 1587–1596. [Google Scholar] [CrossRef]
  81. Singh, P.; Choudhary, K.K.; Chaudhary, N.; Gupta, S.; Sahu, M.; Tejaswini, B.; Sarkar, S. Salt stress resilience in plants mediated through osmolyte accumulation and its crosstalk mechanism with phytohormones. Front. Plant Sci. 2022, 13, 1006617. [Google Scholar] [CrossRef]
  82. Hanin, M.; Ebel, C.; Ngom, M.; Laplaze, L.; Masmoudi, K. New Insights on Plant Salt Tolerance Mechanisms and Their Potential Use for Breeding. Front. Plant Sci. 2016, 7, 1787. [Google Scholar] [CrossRef] [PubMed]
  83. Kim, S.H.; Kim, S.H.; Palaniyandi, S.A.; Yang, S.H.; Suh, J.W. Expression of potato S-adenosyl-L-methionine synthase (SbSAMS) gene altered developmental characteristics and stress responses in transgenic Arabidopsis plants. Plant Physiol. Biochem. 2015, 87, 84–91. [Google Scholar] [CrossRef]
  84. Ma, C.; Wang, Y.; Gu, D.; Nan, J.; Chen, S.; Li, H. Overexpression of S-Adenosyl-l-Methionine Synthetase 2 from Sugar Beet M14 Increased Arabidopsis Tolerance to Salt and Oxidative Stress. Int. J. Mol. Sci. 2017, 18, 847. [Google Scholar] [CrossRef] [PubMed]
  85. He, M.W.; Wang, Y.; Wu, J.Q.; Shu, S.; Sun, J.; Guo, S.R. Isolation and characterization of S-Adenosylmethionine synthase gene from cucumber and responsive to abiotic stress. Plant Physiol. Biochem. 2019, 141, 431–445. [Google Scholar] [CrossRef]
  86. Zhu, H.; He, M.; Jahan, M.S.; Wu, J.; Gu, Q.; Shu, S.; Sun, J.; Guo, S. CsCDPK6, a CsSAMS1-Interacting Protein, Affects Polyamine/Ethylene Biosynthesis in Cucumber and Enhances Salt Tolerance by Overexpression in Tobacco. Int. J. Mol. Sci. 2021, 22, 11133. [Google Scholar] [CrossRef]
  87. Riyazuddin, R.; Verma, R.; Singh, K.; Nisha, N.; Keisham, M.; Bhati, K.K.; Kim, S.T.; Gupta, R. Ethylene: A Master Regulator of Salinity Stress Tolerance in Plants. Biomolecules 2020, 10, 959. [Google Scholar] [CrossRef] [PubMed]
  88. Rossi, F.R.; Krapp, A.R.; Bisaro, F.; Maiale, S.J.; Pieckenstain, F.L.; Carrillo, N. Reactive oxygen species generated in chloroplasts contribute to tobacco leaf infection by the necrotrophic fungus Botrytis cinerea. Plant J. 2017, 92, 761–773. [Google Scholar] [CrossRef]
  89. Petrov, V.; Hille, J.; Mueller-Roeber, B.; Gechev, T.S. ROS-mediated abiotic stress-induced programmed cell death in plants. Front. Plant Sci. 2015, 6, 69. [Google Scholar] [CrossRef] [PubMed]
  90. Ecker, J.R.U.O. The ethylene signal transduction pathway in plants. Science 1995, 268, 667–675. [Google Scholar] [CrossRef] [PubMed]
  91. Mehrotra, R.; Bhalothia, P.; Bansal, P.; Basantani, M.K.; Bharti, V.; Mehrotra, S. Abscisic acid and abiotic stress tolerance different tiers of regulation. J. Plant Physiol. 2014, 171, 486–496. [Google Scholar] [CrossRef] [PubMed]
  92. Naing, A.H.; Campol, J.R.; Kang, H.; Xu, J.; Chung, M.Y.; Kim, C.K. Role of Ethylene Biosynthesis Genes in the Regulation of Salt Stress and Drought Stress Tolerance in Petunia. Front. Plant Sci. 2022, 13, 844449. [Google Scholar] [CrossRef] [PubMed]
  93. Binder, B.M. Ethylene signaling in plants. J. Biol. Chem. 2020, 295, 7710–7725. [Google Scholar] [CrossRef]
  94. Lockhart, J. Salt of the Earth: Ethylene Promotes Salt Tolerance by Enhancing Na+/K+ Homeostasis. Plant Cell 2013, 25, 3150. [Google Scholar] [CrossRef] [PubMed]
  95. Pattyn, J.; Vaughan-Hirsch, J.; Van de Poel, B. The regulation of ethylene biosynthesis: A complex multilevel control circuitry. New Phytol. 2021, 229, 770–782. [Google Scholar] [CrossRef] [PubMed]
  96. Boller, T.; Herner, R.C.; Kende, H. Assay for and enzymatic formation of an ethylene precursor, 1-aminocyclopropane-1-carboxylic acid. Planta 1979, 145, 293–303. [Google Scholar] [CrossRef]
  97. Druege, U. Ethylene and Plant Responses to Abiotic Stress. In Ethylene Action in Plants; Springer: Berlin/Heidelberg, Germany, 2006; pp. 81–118. [Google Scholar] [CrossRef]
  98. Jakubowicz, M.; Sadowski, J. 1-Aminocyclopropane-l-carboxylate synthase-genes and expression. Acta Physiol. Plant. 2002, 24, 459–478. [Google Scholar] [CrossRef]
  99. Yamagami, T.; Tsuchisaka, A.; Yamada, K.; Haddon, W.F.; Harden, L.A.; Theologis, A. Biochemical Diversity among the 1-Amino-cyclopropane-1-Carboxylate Synthase Isozymes Encoded by the Arabidopsis Gene Family. J. Biol. Chem. 2003, 278, 49102–49112. [Google Scholar] [CrossRef] [PubMed]
  100. Lee, H.Y.; Chen, Y.C.; Kieber, J.J.; Yoon, G.M. Regulation of the turnover of ACC synthases by phytohormones and heterodimerization in Arabidopsis. Plant J. 2017, 91, 491–504. [Google Scholar] [CrossRef]
  101. Tsuchisaka, A.; Theologis, A. Heterodimeric interactions among the 1-amino-cyclopropane-1-carboxylate synthase polypeptides encoded by the Arabidopsis gene family. Proc. Natl. Acad. Sci. USA 2004, 101, 2275–2280. [Google Scholar] [CrossRef]
  102. Li, J.; Zou, X.; Chen, G.; Meng, Y.; Ma, Q.; Chen, Q.; Wang, Z.; Li, F. Potential Roles of 1-Aminocyclopropane-1-carboxylic Acid Synthase Genes in the Response of Gossypium Species to Abiotic Stress by Genome-Wide Identification and Expression Analysis. Plants 2022, 11, 1524. [Google Scholar] [CrossRef]
  103. Gómez-Lim, M.A.; Valdés-López, V.; Cruz-Hernandez, A.; Saucedo-Arias, L.J. Isolation and characterization of a gene involved in ethylene biosynthesis from Arabidopsis thaliana. Gene 1993, 134, 217–221. [Google Scholar] [CrossRef] [PubMed]
  104. Wei, H.; Xue, Y.; Chen, P.; Hao, P.; Wei, F.; Sun, L.; Yang, Y. Genome-Wide Identification and Functional Investigation of 1-Aminocyclopropane-1-carboxylic Acid Oxidase (ACO) Genes in Cotton. Plants 2021, 10, 1699. [Google Scholar] [CrossRef] [PubMed]
  105. Schott-Verdugo, S.; Müller, L.; Classen, E.; Gohlke, H.; Groth, G. Structural Model of the ETR1 Ethylene Receptor Transmembrane Sensor Domain. Sci. Rep. 2019, 9, 8869. [Google Scholar] [CrossRef] [PubMed]
  106. Bleecker, A.B. The ethylene-receptor family from Arabidopsis: Structure and function. Philos. Trans. R. Soc. B Biol. Sci. 1998, 353, 1405–1412. [Google Scholar] [CrossRef] [PubMed]
  107. Sakai, H.; Hua, J.; Chen, Q.G.; Chang, C.; Medrano, L.J.; Bleecker, A.B.; Meyerowitz, E.M. ETR2 is an ETR1-like gene involved in ethylene signaling in Arabidopsis. Proc. Natl. Acad. Sci. USA 1998, 95, 5812–5817. [Google Scholar] [CrossRef] [PubMed]
  108. Gallie, D.R. Appearance and elaboration of the ethylene receptor family during land plant evolution. Plant Mol. Biol. 2015, 87, 521–539. [Google Scholar] [CrossRef]
  109. Park, H.L.; Seo, D.H.; Lee, H.Y.; Bakshi, A.; Park, C.; Chien, Y.; Kieber, J.J.; Binder, B.M.; Yoon, G.M. Ethylene-triggered subcellular trafficking of CTR1 enhances the response to ethylene gas. Nat. Commun. 2023, 14, 365. [Google Scholar] [CrossRef]
  110. Bisson, M.M.A.; Groth, G. New Insight in Ethylene Signaling: Autokinase Activity of ETR1 Modulates the Interaction of Receptors and EIN2. Mol. Plant 2010, 3, 882–889. [Google Scholar] [CrossRef]
  111. Merchante, C.; Brumos, J.; Yun, J.; Hu, Q.; Spencer, K.R.; Enríquez, P.; Binder, B.M.; Heber, S.; Stepanova, A.N.; Alonso, J.M. Gene-Specific Translation Regulation Mediated by the Hormone-Signaling Molecule EIN2. Cell 2015, 163, 684–697. [Google Scholar] [CrossRef]
  112. Iqbal, N.; Trivellini, A.; Masood, A.; Ferrante, A.; Khan, N.A. Current understanding on ethylene signaling in plants: The influence of nutrient availability. Plant Physiol. Biochem. 2013, 73, 128–138. [Google Scholar] [CrossRef]
  113. Dolgikh, V.A.; Pukhovaya, E.M.; Zemlyanskaya, E.V. Shaping Ethylene Response: The Role of EIN3/EIL1 Transcription Factors. Front. Plant Sci. 2019, 10, 1030. [Google Scholar] [CrossRef] [PubMed]
  114. Debbarma, J.; Sarki, Y.N.; Saikia, B.; Boruah, H.; Singha, D.L.; Chikkaputtaiah, C. Ethylene Response Factor (ERF) Family Proteins in Abiotic Stresses and CRISPR-Cas9 Genome Editing of ERFs for Multiple Abiotic Stress Tolerance in Crop Plants: A Review. Mol. Biotechnol. 2019, 61, 153–172. [Google Scholar] [CrossRef] [PubMed]
  115. Klay, I.; Pirrello, J.; Riahi, L.; Bernadac, A.; Cherif, A.; Bouzayen, M.; Bouzid, S. Ethylene Response Factor Sl-ERF.B.3 Is Responsive to Abiotic Stresses and Mediates Salt and Cold Stress Response Regulation in Tomato. Sci. World J. 2014, 2014, 167681. [Google Scholar] [CrossRef] [PubMed]
  116. Fatma, M.; Asgher, M.; Iqbal, N.; Rasheed, F.; Sehar, Z.; Sofo, A.; Khan, N.A. Ethylene Signaling under Stressful Environments: Analyzing Collaborative Knowledge. Plants 2022, 11, 2211. [Google Scholar] [CrossRef] [PubMed]
  117. Yu, D.; Li, X.; Li, Y.; Ali, F.; Li, F.; Wang, Z. Dynamic roles and intricate mechanisms of ethylene in epidermal hair development in Arabidopsis and cotton. New Phytol. 2022, 234, 375–391. [Google Scholar] [CrossRef] [PubMed]
  118. Yu, F.; Guo, R.; Wu, C.; Li, H.; Guo, X. Molecular cloning and expression characteristics of a novel MAPKKK gene, GhCTR1, from cotton (Gossypium hirsutum L.). S. Afr. J. Bot. 2012, 78, 211–219. [Google Scholar] [CrossRef]
  119. Wang, X.Q.; Han, L.H.; Zhou, W.; Tao, M.; Hu, Q.Q.; Zhou, Y.N.; Li, X.B.; Li, D.D.; Huang, G.Q. GhEIN3, a cotton (Gossypium hirsutum) homologue of AtEIN3, is involved in regulation of plant salinity tolerance. Plant Physiol. Biochem. 2019, 143, 83–93. [Google Scholar] [CrossRef]
  120. Phukan, U.J.; Jeena, G.S.; Tripathi, V.; Shukla, R.K. Regulation of Apetala2/Ethylene Response Factors in Plants. Front. Plant Sci. 2017, 8, 150. [Google Scholar] [CrossRef]
  121. Ma, L.; Hu, L.; Fan, J.; Amombo, E.; Khaldun, A.; Zheng, Y.; Chen, L. Cotton GhERF38 gene is involved in plant response to salt/drought and ABA. Ecotoxicology 2017, 26, 841–854. [Google Scholar] [CrossRef]
  122. Lu, L.; Qanmber, G.; Li, J.; Pu, M.; Chen, G.; Li, S.; Liu, L.; Qin, W.; Ma, S.; Wang, Y.; et al. Identification and Characterization of the ERF Subfamily B3 Group Revealed GhERF13.12 Improves Salt Tolerance in Upland Cotton. Front. Plant Sci. 2021, 12, 705883. [Google Scholar] [CrossRef]
  123. Zhang, J.; Zhang, P.; Huo, X.; Gao, Y.; Chen, Y.; Song, Z.; Wang, F.; Zhang, J. Comparative Phenotypic and Transcriptomic Analysis Reveals Key Responses of Upland Cotton to Salinity Stress during Postgermination. Front. Plant Sci. 2021, 12, 639104. [Google Scholar] [CrossRef] [PubMed]
  124. Long, L.; Yang, W.; Liao, P.; Guo, Y.; Kumar, A.; Gao, W. Transcriptome analysis reveals differentially expressed ERF transcription factors associated with salt response in cotton. Plant Sci. 2019, 281, 72–81. [Google Scholar] [CrossRef] [PubMed]
  125. Wessler, S.R. Homing into the origin of the AP2 DNA binding domain. Trends Plant Sci. 2005, 10, 54–56. [Google Scholar] [CrossRef] [PubMed]
  126. Zafar, M.M.; Rehman, A.; Razzaq, A.; Parvaiz, A.; Mustafa, G.; Sharif, F.; Mo, H.; Youlu, Y.; Shakeel, A.; Ren, M. Genome-wide characterization and expression analysis of Erf gene family in cotton. BMC Plant Biol. 2022, 22, 134. [Google Scholar] [CrossRef]
  127. Bano, N.; Fakhrah, S.; Mohanty, C.S.; Bag, S.K. Transcriptome Meta-Analysis Associated Targeting Hub Genes and Pathways of Drought and Salt Stress Responses in Cotton (Gossypium hirsutum): A Network Biology Approach. Front. Plant Sci. 2022, 13, 818472. [Google Scholar] [CrossRef]
  128. Qiao, Z.X.; Huang, B.; Liu, J.Y. Molecular cloning and functional analysis of an ERF gene from cotton (Gossypium hirsutum). Biochim Biophys Acta 2008, 1779, 122–127. [Google Scholar] [CrossRef] [PubMed]
  129. Jin, L.; Li, H.; Liu, J. Molecular Characterization of Three Ethylene Responsive Element Binding Factor Genes from Cotton. J. Integr. Plant Biol. 2010, 52, 485–495. [Google Scholar] [CrossRef]
  130. Jin, L.G.; Liu, J.Y. Molecular cloning, expression profile and promoter analysis of a novel ethylene responsive transcription factor gene GhERF4 from cotton (Gossypium hirstum). Plant Physiol. Biochem. 2008, 46, 46–53. [Google Scholar] [CrossRef] [PubMed]
  131. Jin, L.; Huang, B.; Li, H.; Liu, J. Expression profiles and transactivation analysis of a novel ethylene-responsive transcription factor gene GhERF5 from cotton. Prog. Nat. Sci. 2009, 19, 563–572. [Google Scholar] [CrossRef]
  132. Kusano, T.; Yamaguchi, K.; Berberich, T.; Takahashi, Y. Advances in polyamine research in 2007. J. Plant Res. 2007, 120, 345–350. [Google Scholar] [CrossRef]
  133. Alcazar, R.; Altabella, T.; Marco, F.; Bortolotti, C.; Reymond, M.; Koncz, C.; Carrasco, P.; Tiburcio, A.F. Polyamines: Molecules with regulatory functions in plant abiotic stress tolerance. Planta 2010, 231, 1237–1249. [Google Scholar] [CrossRef]
  134. Takano, A.; Kakehi, J.; Takahashi, T. Thermospermine is not a minor polyamine in the plant kingdom. Plant Cell Physiol. 2012, 53, 606–616. [Google Scholar] [CrossRef] [PubMed]
  135. Vuosku, J.; Suorsa, M.; Ruottinen, M.; Sutela, S.; Muilu-Makela, R.; Julkunen-Tiitto, R.; Sarjala, T.; Neubauer, P.; Haggman, H. Polyamine metabolism during exponential growth transition in Scots pine embryogenic cell culture. Tree Physiol. 2012, 32, 1274–1287. [Google Scholar] [CrossRef] [PubMed]
  136. Tiburcio, A.F.; Altabella, T.; Bitrian, M.; Alcazar, R. The roles of polyamines during the lifespan of plants: From development to stress. Planta 2014, 240, 1–18. [Google Scholar] [CrossRef] [PubMed]
  137. Mansour, M. Plasma membrane permeability as an indicator of salt tolerance in plants. Biol. Plant 2013, 57, 1–10. [Google Scholar] [CrossRef]
  138. Lightfoot, H.L.; Hall, J. Endogenous polyamine function-the RNA perspective. Nucleic Acids Res. 2014, 42, 11275–11290. [Google Scholar] [CrossRef] [PubMed]
  139. Bueno, M.; Cordovilla, M.P. Polyamines in Halophytes. Front. Plant Sci. 2019, 10, 439. [Google Scholar] [CrossRef] [PubMed]
  140. ElSayed, A.I.; Mohamed, A.H.; Rafudeen, M.S.; Omar, A.A.; Awad, M.F.; Mansour, E. Polyamines mitigate the destructive impacts of salinity stress by enhancing photosynthetic capacity, antioxidant defense system and upregulation of calvin cycle-related genes in rapeseed (Brassica napus L.). Saudi J. Biol. Sci. 2022, 29, 3675–3686. [Google Scholar] [CrossRef] [PubMed]
  141. Korbas, A.; Kubiś, J.; Rybus-Zając, M.; Chadzinikolau, T. Spermidine Modify Antioxidant Activity in Cucumber Exposed to Salinity Stress. Agronomy 2022, 12, 1554. [Google Scholar] [CrossRef]
  142. Pál, M.; Ivanovska, B.; Oláh, T.; Tajti, J.; Hamow, K.Á.; Szalai, G.; Khalil, R.; Vanková, R.; Dobrev, P.; Misheva, S.P.; et al. Janda Role of polyamines in plant growth regulation of Rht wheat mutants. Plant Physiol. Biochem. 2019, 137, 189–202. [Google Scholar] [CrossRef]
  143. Sagor, G.H.M.; Inoue, M.; Kusano, T.; Berberich, T. Expression profile of seven polyamine oxidase genes in rice (Oryza sativa) in response to abiotic stresses, phytohormones and polyamines. Physiol. Mol. Biol. Plants 2021, 27, 1353–1359. [Google Scholar] [CrossRef] [PubMed]
  144. Baniasadi, F.; Saffari, V.R.; Moud, A. Physiological and growth responses of Calendula officinalis L. plants to the interaction effects of polyamines and salt stress. Sci. Hortic. 2018, 234, 312–317. [Google Scholar] [CrossRef]
  145. Michael, A.J. Biosynthesis of polyamines and polyamine-containing molecules. Biochem. J. 2016, 473, 2315–2329. [Google Scholar] [CrossRef] [PubMed]
  146. Fuell, C.; Elliott, K.A.; Hanfrey, C.C.; Franceschetti, M.; Michael, A.J. Polyamine biosynthetic diversity in plants and algae. Plant Physiol. Biochem. 2010, 48, 513–520. [Google Scholar] [CrossRef] [PubMed]
  147. Hanfrey, C.; Sommer, S.; Mayer, M.J.; Burtin, D.; Michael, A.J. Arabidopsis polyamine biosynthesis: Absence of ornithine decarboxylase and the mechanism of arginine decarboxylase activity. Plant J. 2001, 27, 551–560. [Google Scholar] [CrossRef] [PubMed]
  148. Pegg, A.E. Functions of Polyamines in Mammals. J. Biol. Chem. 2016, 291, 14904–14912. [Google Scholar] [CrossRef]
  149. Hummel, I.; Bourdais, G.; Gouesbet, G.; Couee, I.; Malmberg, R.L.; El, A.A. Differential gene expression of ARGININE DECARBOXYLASE ADC1 and ADC2 in Arabidopsis thaliana: Characterization of transcriptional regulation during seed germination and seedling development. New Phytol. 2004, 163, 519–531. [Google Scholar] [CrossRef]
  150. Fu, Y.; Guo, C.; Wu, H.; Chen, C. Arginine decarboxylase ADC2 enhances salt tolerance through increasing ROS scavenging enzyme activity in Arabidopsis thaliana. Plant Growth Regul. 2017, 83, 253–263. [Google Scholar] [CrossRef]
  151. Marco, F.; Alcazar, R.; Tiburcio, A.F.; Carrasco, P. Interactions between polyamines and abiotic stress pathway responses unraveled by transcriptome analysis of polyamine overproducers. OMICS 2011, 15, 775–781. [Google Scholar] [CrossRef]
  152. Saha, J.; Giri, K. Molecular phylogenomic study and the role of exogenous spermidine in the metabolic adjustment of endogenous polyamine in two rice cultivars under salt stress. Genes 2017, 609, 88–103. [Google Scholar] [CrossRef]
  153. Napieraj, N.; Reda, M.G.; Janicka, M.G. The role of NO in plant response to salt stress: Interactions with polyamines. Funct. Plant Biol. 2020, 47, 865–879. [Google Scholar] [CrossRef] [PubMed]
  154. Kovacs, L.; Mendel, A.; Szentgyorgyi, A.; Fekete, S.; Sore, F.; Posta, K.; Kiss, E. Comparative analysis of overexpressed Fragaria vesca S-adenosyl-l-methionine synthase (FvSAMS) and decarboxylase (FvSAMDC) during salt stress in transgenic Nicotiana benthamiana. Plant Growth Regul. 2020, 91, 53–73. [Google Scholar] [CrossRef]
  155. Wu, F.; Muvunyi, B.P.; Yan, Q.; Kanzana, G.; Ma, T.; Zhang, Z.; Wang, Y.; Zhang, J. Comprehensive genome-wide analysis of polyamine and ethylene pathway genes in Cleistogenes songorica and CsSAMDC2 function in response to abiotic stress. Environ. Exp. Bot. 2022, 202, 105029. [Google Scholar] [CrossRef]
  156. Tang, X.; Wu, L.; Wang, F.; Tian, W.; Hu, X.; Jin, S.; Zhu, H. Ectopic Expression of GhSAMDC3 Enhanced Salt Tolerance Due to Accumulated Spd Content and Activation of Salt Tolerance-Related Genes in Arabidopsis thaliana. DNA Cell Biol. 2021, 40, 1144–1157. [Google Scholar] [CrossRef] [PubMed]
  157. Meng, D.Y.; Yang, S.; Xing, J.Y.; Ma, N.N.; Wang, B.Z.; Qiu, F.T.; Guo, F.; Meng, J.; Zhang, J.L.; Wan, S.B.; et al. Peanut (Arachis hypogaea L.) S-adenosylmethionine decarboxylase confers transgenic tobacco with elevated tolerance to salt stress. Plant Biol. 2021, 23, 341–350. [Google Scholar] [CrossRef] [PubMed]
  158. Jia, T.; Hou, J.; Iqbal, M.Z.; Zhang, Y.; Cheng, B.; Feng, H.; Li, Z.; Liu, L.; Zhou, J.; Feng, G.; et al. Overexpression of the white clover TrSAMDC1 gene enhanced salt and drought resistance in Arabidopsis thaliana. Plant Physiol. Biochem. 2021, 165, 147–160. [Google Scholar] [CrossRef] [PubMed]
  159. Cona, A.; Rea, G.; Angelini, R.; Federico, R.; Tavladoraki, P. Functions of amine oxidases in plant development and defence. Trends Plant Sci. 2006, 11, 80–88. [Google Scholar] [CrossRef]
  160. Moschou, P.N.; Wu, J.; Cona, A.; Tavladoraki, P.; Angelini, R.; Roubelakis-Angelakis, K.A. The polyamines and their catabolic products are significant players in the turnover of nitrogenous molecules in plants. J. Exp. Bot. 2012, 63, 5003–5015. [Google Scholar] [CrossRef]
  161. Khajuria, A.; Sharma, N.; Bhardwaj, R.; Ohri, P. Emerging Role of Polyamines in Plant Stress Tolerance. Curr. Protein Pept. Sci. 2018, 19, 1114–1123. [Google Scholar] [CrossRef]
  162. Kim, D.W.; Watanabe, K.; Murayama, C.; Izawa, S.; Niitsu, M.; Michael, A.J.; Berberich, T.; Kusano, T. Polyamine Oxidase5 Regulates Arabidopsis Growth through Thermospermine Oxidase Activity. Plant Physiol. 2014, 165, 1575–1590. [Google Scholar] [CrossRef]
  163. Fincato, P.; Moschou, P.N.; Ahou, A.; Angelini, R.; Roubelakis-Angelakis, K.A.; Federico, R.; Tavladoraki, P. The members of Arabidopsis thaliana PAO gene family exhibit distinct tissue and organ-specific expression pattern during seedling growth and flower development. Amino Acids 2012, 42, 831–841. [Google Scholar] [CrossRef] [PubMed]
  164. Sagor, G.H.; Zhang, S.; Kojima, S.; Simm, S.; Berberich, T.; Kusano, T. Reducing Cytoplasmic Polyamine Oxidase Activity in Arabidopsis Increases Salt and Drought Tolerance by Reducing Reactive Oxygen Species Production and Increasing Defense Gene Expression. Front. Plant Sci. 2016, 7, 214. [Google Scholar] [CrossRef] [PubMed]
  165. Wu, J.; Zhu, M.; Liu, W.; Jahan, M.S.; Gu, Q.; Shu, S.; Sun, J.; Guo, S. CsPAO2 Improves Salt Tolerance of Cucumber through the Interaction with CsPSA3 by Affecting Photosynthesis and Polyamine Conversion. Int. J. Mol. Sci. 2022, 23, 12413. [Google Scholar] [CrossRef] [PubMed]
  166. Wu, J.; Liu, W.; Jahan, M.S.; Shu, S.; Sun, J.; Guo, S. Characterization of polyamine oxidase genes in cucumber and roles of CsPAO3 in response to salt stress. Environ. Exp. Bot. 2022, 194, 104696. [Google Scholar] [CrossRef]
  167. Liu, G.; Jiang, W.; Tian, L.; Fu, Y.; Tan, L.; Zhu, Z.; Sun, C.; Liu, F. Polyamine oxidase 3 is involved in salt tolerance at the germination stage in rice. J. Genet. Genom. 2022, 49, 458–468. [Google Scholar] [CrossRef]
  168. Cheng, X.Q.; Zhu, X.F.; Tian, W.G.; Cheng, W.H.; Hakim; Sun, J.; Jin, S.X.; Zhu, H.G. Genome-wide identification and expression analysis of polyamine oxidase genes in upland cotton (Gossypium hirsutum L.). Plant Cell Tissue Organ Cult. 2017, 129, 237–249. [Google Scholar] [CrossRef]
  169. Gerlin, L.; Baroukh, C.; Genin, S. Polyamines: Double agents in disease and plant immunity. Trends Plant Sci. 2021, 26, 1061–1071. [Google Scholar] [CrossRef]
  170. Majumdar, R.; Minocha, R.; Lebar, M.D.; Rajasekaran, K.; Long, S.; Carter-Wientjes, C.; Minocha, S.; Cary, J.W. Contribution of Maize Polyamine and Amino Acid Metabolism toward Resistance against Aspergillus flavus Infection and Aflatoxin Production. Front. Plant Sci. 2019, 10, 692. [Google Scholar] [CrossRef] [PubMed]
  171. Nandy, S.; Mandal, S.; Gupta, S.K.; Anand, U.; Ghorai, M.; Mundhra, A.; Rahman, M.H.; Ray, P.; Mitra, S.; Ray, D.; et al. Role of Polyamines in Molecular Regulation and Cross-Talks against Drought Tolerance in Plants. J. Plant Growth Regul. 2022. [Google Scholar] [CrossRef]
  172. Momtaz, O.A.; Hussein, E.M.; Fahmy, E.M.; Ahmed, S.E. Expression of S-adenosyl methionine decarboxylase gene for polyamine accumulation in Egyptian cotton Giza 88 and Giza 90. GM Crops 2010, 1, 257–266. [Google Scholar] [CrossRef] [PubMed]
  173. Grzesiak, M.; Filek, M.; Barbasz, A.; Kreczmer, B.; Hartikainen, H. Relationships between polyamines, ethylene, osmoprotectants and antioxidant enzymes activities in wheat seedlings after short-term PEG- and NaCl-induced stresses. Plant Growth Regul. 2013, 69, 177–189. [Google Scholar] [CrossRef]
  174. Zapata, P.J.; Serrano, M.; Garcia-Legaz, M.F.; Pretel, M.T.; Botella, M.A. Short Term Effect of Salt Shock on Ethylene and Polyamines Depends on Plant Salt Sensitivity. Front. Plant Sci. 2017, 8, 855. [Google Scholar] [CrossRef] [PubMed]
  175. Freitas, V.S.; Miranda, R.D.S.; Costa, J.H.; Oliveira, D.F.D.; Paula, S.D.O.; Miguel, E.D.C.; Freire, R.S.; Prisco, J.T.; Gomes-Filho, E. Ethylene triggers salt tolerance in maize genotypes by modulating polyamine catabolism enzymes associated with H2O2 production. Environ. Exp. Bot. 2018, 145, 75–86. [Google Scholar] [CrossRef]
  176. Takacs, Z.; Czekus, Z.; Tari, I.; Poor, P. The role of ethylene signalling in the regulation of salt stress response in mature tomato fruits: Metabolism of antioxidants and polyamines. J. Plant Physiol. 2022, 277, 153793. [Google Scholar] [CrossRef] [PubMed]
  177. Cheng, Z.; Sattler, S.; Maeda, H.; Sakuragi, Y.; Bryant, D.A.; DellaPenna, D. Highly divergent methyltransferases catalyze a conserved reaction in tocopherol and plastoquinone synthesis in cyanobacteria and photosynthetic eukaryotes. Plant Cell 2003, 15, 2343–2356. [Google Scholar] [CrossRef] [PubMed]
  178. Yasuno, R.; Wada, H. The biosynthetic pathway for lipoic acid is present in plastids and mitochondria in Arabidopsis thaliana. FEBS Lett. 2002, 517, 110–114. [Google Scholar] [CrossRef]
  179. Monné, M.; Marobbio, C.M.T.; Agrimi, G.; Palmieri, L.; Palmieri, F. Mitochondrial transport and metabolism of the major methyl donor and versatile cofactor S-adenosylmethionine, and related diseases: A review. IUBMB Life 2022, 74, 573–591. [Google Scholar] [CrossRef]
  180. Ravanel, S.; Block, M.A.; Rippert, P.; Jabrin, S.; Curien, G.; Rébeillé, F.; Douce, R. Methionine metabolism in plants: Chloroplasts are autonomous for de novo methionine synthesis and can import S-adenosylmethionine from the cytosol. J. Biol. Chem. 2004, 279, 22548–22557. [Google Scholar] [CrossRef]
  181. Nunes-Nesi, A.; Cavalcanti, J.; Fernie, A.R. Characterization of In Vivo Function(s) of Members of the Plant Mitochondrial Carrier Family. Biomolecules 2020, 10, 1226. [Google Scholar] [CrossRef]
  182. Li, Y.; Yang, Z.; Zhang, Y.; Guo, J.; Liu, L.; Wang, C.; Wang, B.; Han, G. The roles of HD-ZIP proteins in plant abiotic stress tolerance. Front. Plant Sci. 2022, 13, 1027071. [Google Scholar] [CrossRef]
  183. Sharif, R.; Raza, A.; Chen, P.; Li, Y.; El-Ballat, E.M.; Rauf, A.; Hano, C.; El-Esawi, M.A. HD-ZIP Gene Family: Potential Roles in Improving Plant Growth and Regulating Stress-Responsive Mechanisms in Plants. Genes 2021, 12, 1256. [Google Scholar] [CrossRef] [PubMed]
  184. Rai, K.K.; Rai, N.; Rai, S.P. Prediction and validation of DREB transcription factors for salt tolerance in Solanum lycopersicum L.: An integrated experimental and computational approach. Environ. Exp. Bot. 2019, 165, 1–18. [Google Scholar] [CrossRef]
  185. Hassan, S.; Berk, K.; Aronsson, H. Evolution and identification of DREB transcription factors in the wheat genome: Modeling, docking and simulation of DREB proteins associated with salt stress. J. Biomol. Struct. Dyn. 2022, 40, 7191–7204. [Google Scholar] [CrossRef] [PubMed]
  186. Wang, S.; Wang, Y.; Yang, R.; Cai, W.; Liu, Y.; Zhou, D.; Meng, L.; Wang, P.; Huang, B. Genome-Wide Identification and Analysis Uncovers the Potential Role of JAZ and MYC Families in Potato under Abiotic Stress. Int. J. Mol. Sci. 2023, 24, 6706. [Google Scholar] [CrossRef] [PubMed]
  187. Valenzuela, C.E.; Acevedo-Acevedo, O.; Miranda, G.S.; Vergara-Barros, P.; Holuigue, L.; Figueroa, C.R.; Figueroa, P.M. Salt stress response triggers activation of the jasmonate signaling pathway leading to inhibition of cell elongation in Arabidopsis primary root. J. Exp. Bot. 2016, 67, 4209–4220. [Google Scholar] [CrossRef] [PubMed]
  188. Su, Y.; Liang, W.; Liu, Z.; Wang, Y.; Zhao, Y.; Ijaz, B.; Hua, J. Overexpression of GhDof1 improved salt and cold tolerance and seed oil content in Gossypium hirsutum. J. Plant Physiol. 2017, 218, 222–234. [Google Scholar] [CrossRef]
  189. Zou, X.; Sun, H. DOF transcription factors: Specific regulators of plant biological processes. Front. Plant Sci. 2023, 14, 1044918. [Google Scholar] [CrossRef]
  190. Iqbal, M.Z.; Jia, T.; Tang, T.; Anwar, M.; Ali, A.; Hassan, M.J.; Zhang, Y.; Tang, Q.; Peng, Y. A Heat Shock Transcription Factor TrHSFB2a of White Clover Negatively Regulates Drought, Heat and Salt Stress Tolerance in Transgenic Arabidopsis. Int. J. Mol. Sci. 2022, 23, 12769. [Google Scholar] [CrossRef]
  191. Guo, M.; Liu, J.; Ma, X.; Luo, D.; Gong, Z.; Lu, M. The Plant Heat Stress Transcription Factors (HSFs): Structure, Regulation, and Function in Response to Abiotic Stresses. Front. Plant Sci. 2016, 7, 114. [Google Scholar] [CrossRef]
  192. Han, G.; Li, Y.; Qiao, Z.; Wang, C.; Zhao, Y.; Guo, J.; Chen, M.; Wang, B. Advances in the Regulation of Epidermal Cell Development by C2H2 Zinc Finger Proteins in Plants. Front. Plant Sci. 2021, 12, 754512. [Google Scholar] [CrossRef]
  193. Liu, Y.; Khan, A.R.; Gan, Y. C2H2 Zinc Finger Proteins Response to Abiotic Stress in Plants. Int. J. Mol. Sci. 2022, 23, 2730. [Google Scholar] [CrossRef] [PubMed]
  194. Bankaji, I.; Sleimi, N.; Vives-Peris, V.; Gómez-Cadenas, A.; Pérez-Clemente, R.M. Identification and expression of the Cucurbita WRKY transcription factors in response to water deficit and salt stress. Sci. Hortic. 2019, 256, 108562. [Google Scholar] [CrossRef]
  195. Hichri, I.; Muhovski, Y.; Žižková, E.; Dobrev, P.I.; Gharbi, E.; Franco-Zorrilla, J.M.; Lopez-Vidriero, I.; Solano, R.; Clippe, A.; Errachid, A.; et al. The Solanum lycopersicum WRKY3 Transcription Factor SlWRKY3 Is Involved in Salt Stress Tolerance in Tomato. Front. Plant Sci. 2017, 8, 1343. [Google Scholar] [CrossRef] [PubMed]
  196. Beathard, C.; Mooney, S.; Al-Saharin, R.; Goyer, A.; Hellmann, H. Characterization of Arabidopsis thaliana R2R3 S23 MYB Transcription Factors as Novel Targets of the Ubiquitin Proteasome-Pathway and Regulators of Salt Stress and Abscisic Acid Response. Front. Plant Sci. 2021, 12, 629208. [Google Scholar] [CrossRef] [PubMed]
  197. Dossa, K.; Mmadi, M.A.; Zhou, R.; Liu, A.; Yang, Y.; Diouf, D.; You, J.; Zhang, X. Ectopic expression of the sesame MYB transcription factor SiMYB305 promotes root growth and modulates ABA-mediated tolerance to drought and salt stresses in Arabidopsis. Aob Plants 2020, 12, z81. [Google Scholar] [CrossRef]
  198. Punia, H.; Tokas, J.; Malik, A.; Sangwan, S.; Rani, A.; Yashveer, S.; Alansi, S.; Hashim, M.J.; El-Sheikh, M.A. Genome-Wide Transcriptome Profiling, Characterization, and Functional Identification of NAC Transcription Factors in Sorghum under Salt Stress. Antioxidants 2021, 10, 1605. [Google Scholar] [CrossRef]
  199. Alshareef, N.O.; Wang, J.Y.; Ali, S.; Al-Babili, S.; Tester, M.; Schmöckel, S.M. Overexpression of the NAC transcription factor JUNGBRUNNEN1 (JUB1) increases salinity tolerance in tomato. Plant Physiol. Biochem. 2019, 140, 113–121. [Google Scholar] [CrossRef]
  200. Ahmad, A.; Niwa, Y.; Goto, S.; Ogawa, T.; Shimizu, M.; Suzuki, A.; Kobayashi, K.; Kobayashi, H. bHLH106 Integrates Functions of Multiple Genes through Their G-Box to Confer Salt Tolerance on Arabidopsis. PLoS ONE 2015, 10, e126872. [Google Scholar] [CrossRef]
  201. Krishnamurthy, P.; Vishal, B.; Khoo, K.; Rajappa, S.; Loh, C.; Kumar, P.P. Expression of AoNHX1 increases salt tolerance of rice and Arabidopsis, and bHLH transcription factors regulate AtNHX1 and AtNHX6 in Arabidopsis. Plant Cell Rep. 2019, 38, 1299–1315. [Google Scholar] [CrossRef]
  202. Manavella, P.A.; Dezar, C.A.; Bonaventure, G.; Baldwin, I.T.; Chan, R.L. HAHB4, a sunflower HD-Zip protein, integrates signals from the jasmonic acid and ethylene pathways during wounding and biotic stress responses. Plant J. 2008, 56, 376–388. [Google Scholar] [CrossRef]
  203. Rehman, S.; Mahmood, T. Functional role of DREB and ERF transcription factors: Regulating stress-responsive network in plants. Acta Physiol. Plant. 2015, 37, 178. [Google Scholar] [CrossRef]
  204. Hu, Y.; Sun, H.; Han, Z.; Wang, S.; Wang, T.; Li, Q.; Tian, J.; Wang, Y.; Zhang, X.; Xu, X.; et al. ERF4 affects fruit ripening by acting as a JAZ interactor between ethylene and jasmonic acid hormone signaling pathways. Hortic. Plant J. 2022, 8, 689–699. [Google Scholar] [CrossRef]
  205. Feng, B.; Han, Y.; Xiao, Y.; Kuang, J.; Fan, Z.; Chen, J.; Lu, W. The banana fruit Dof transcription factor MaDof23 acts as a repressor and interacts with MaERF9 in regulating ripening-related genes. J. Exp. Bot. 2016, 67, 2263–2275. [Google Scholar] [CrossRef] [PubMed]
  206. Wang, Y.; Zhou, Y.; Wang, R.; Xu, F.; Tong, S.; Song, C.; Shao, Y.; Yi, M.; He, J. Ethylene Response Factor LlERF110 Mediates Heat Stress Response via Regulation of LlHsfA3A Expression and Interaction with LlHsfA2 in Lilies (Lilium longiflorum). Int. J. Mol. Sci. 2022, 23, 16135. [Google Scholar] [CrossRef] [PubMed]
  207. Han, Y.; Fu, C.; Kuang, J.; Chen, J.; Lu, W. Two banana fruit ripening-related C2H2 zinc finger proteins are transcriptional repressors of ethylene biosynthetic genes. Postharvest Biol. Technol. 2016, 116, 8–15. [Google Scholar] [CrossRef]
  208. Alessio, V.M.; Cavaçana, N.; Dantas, L.L.D.B.; Lee, N.; Hotta, C.T.; Imaizumi, T.; Menossi, M. The FBH family of bHLH transcription factors controls ACC synthase expression in sugarcane. J. Exp. Bot. 2018, 69, 2511–2525. [Google Scholar] [CrossRef] [PubMed]
  209. An, J.P.; Yao, J.F.; Xu, R.R.; You, C.X.; Wang, X.F.; Hao, Y.J. An apple NAC transcription factor enhances salt stress tolerance by modulating the ethylene response. Physiol. Plant. 2018, 164, 279–289. [Google Scholar] [CrossRef]
  210. Xu, L.; Xiang, G.; Sun, Q.; Ni, Y.; Jin, Z.; Gao, S.; Yao, Y. Melatonin enhances salt tolerance by promoting MYB108A-mediated ethylene biosynthesis in grapevines. Hortic. Res.-Engl. 2019, 6, 114. [Google Scholar] [CrossRef]
  211. Wang, Z.; Wei, X.; Wang, Y.; Sun, M.; Zhao, P.; Wang, Q.; Yang, B.; Li, J.; Jiang, Y. WRKY29 transcription factor regulates ethylene biosynthesis and response in Arabidopsis. Plant Physiol. Biochem. 2023, 194, 134–145. [Google Scholar] [CrossRef]
  212. Zhou, Y.; Xiong, Q.; Yin, C.C.; Ma, B.; Chen, S.Y.; Zhang, J.S. Ethylene Biosynthesis, Signaling, and Crosstalk with Other Hormones in Rice. Small Methods 2020, 4, 1900278. [Google Scholar] [CrossRef]
  213. Yu, Z.; Duan, X.; Luo, L.; Dai, S.; Ding, Z.; Xia, G. How Plant Hormones Mediate Salt Stress Responses. Trends Plant Sci. 2020, 25, 1117–1130. [Google Scholar] [CrossRef] [PubMed]
  214. Ryu, H.; Cho, Y. Plant hormones in salt stress tolerance. J. Plant Biol. 2015, 58, 147–155. [Google Scholar] [CrossRef]
  215. Zou, X.; Liu, L.; Hu, Z.; Wang, X.; Zhu, Y.; Zhang, J.; Li, X.; Kang, Z.; Lin, Y.; Yin, C. Salt-induced inhibition of rice seminal root growth is mediated by ethylene-jasmonate interaction. J. Exp. Bot. 2021, 72, 5656–5672. [Google Scholar] [CrossRef] [PubMed]
  216. Singh, D.; Debnath, P.; Sane, A.P.; Sane, V.A. Tomato (Solanumlycopersicum) WRKY23 enhances salt and osmoticstress tolerance by modulating the ethylene and auxin pathways in transgenic Arabidopsis. Plant Physiol. Biochem. 2023, 195, 11. [Google Scholar] [CrossRef] [PubMed]
  217. Afzal, M.; Hindawi, S.E.S.; Alghamdi, S.S.; Migdadi, H.H.; Khan, M.A.; Hasnain, M.U.; Arslan, M.; Habib Ur Rahman, M.; Sohaib, M. Potential Breeding Strategies for Improving Salt Tolerance in Crop Plants. J. Plant Growth Regul. 2022, 42, 3365–3387. [Google Scholar] [CrossRef]
  218. Ogawa, S.; Mitsuya, S. S-methylmethionine is involved in the salinity tolerance of Arabidopsis thaliana plants at germination and early growth stages. Physiol. Plant. 2012, 144, 13–19. [Google Scholar] [CrossRef]
  219. Watanabe, M.; Chiba, Y.; Hirai, M.Y. Metabolism and Regulatory Functions of O-Acetylserine, S-Adenosylmethionine, Homocysteine, and Serine in Plant Development and Environmental Responses. Front. Plant Sci. 2021, 12, 643403. [Google Scholar] [CrossRef]
  220. Verma, P.; Venugopalan, M.V.; Blaise, D.; Waghmare, V.N. Ethylene mediated regulation of fiber development in Asiatic cotton (Gossypium arboreum L.). S. Afr. J. Bot. 2020, 135, 349–354. [Google Scholar] [CrossRef]
  221. Yousaf, S.; Rehman, T.; Tabassum, B.; Aftab, F.; Qaisar, U. Genome scale analysis of 1-aminocyclopropane-1-carboxylate oxidase gene family in G. barbadense and its functions in cotton fiber development. Sci. Rep. 2023, 13, 4004. [Google Scholar] [CrossRef]
  222. Liu, Z.; Ge, X.; Yang, Z.; Zhang, C.; Zhao, G.; Chen, E.; Liu, J.; Zhang, X.; Li, F. Genome-wide identification and characteriza tion of SnRK2 gene family in cotton (Gossypium hirsutum L.). BMC Genet. 2017, 18, 54. [Google Scholar] [CrossRef]
  223. Livak, K.J.; Schmittgen, T.D. Analysis of Relative Gene Expression Data Using Real-Time Quantitative PCR and the 2−ΔΔCT Method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef] [PubMed]
  224. Tuttle, J.R.; Haigler, C.H.; Robertson, D.N. Virus-induced gene silencing of fiber-related genes in cotton. Methods Mol. Biol. 2015, 1287, 219–234. [Google Scholar] [CrossRef] [PubMed]
  225. Pertea, M.; Kim, D.; Pertea, G.M.; Leek, J.T.; Salzberg, S.L. Transcript-level expression analysis of RNA-seq experiments with HISAT, StringTie and Ballgown. Nat. Protoc. 2016, 11, 1650–1667. [Google Scholar] [CrossRef] [PubMed]
  226. Anders, S.; Pyl, P.T.; Huber, W. HTSeq-a Python framework to work with high-throughput sequencing data. Bioinformatics 2015, 31, 166–169. [Google Scholar] [CrossRef] [PubMed]
  227. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools: An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef]
Figure 1. The response to salt stress in cotton. The left part describes the effects of salt stress on the growth and development of cotton; the right part illustrates the related mechanisms and factors of cotton salt tolerance.
Figure 1. The response to salt stress in cotton. The left part describes the effects of salt stress on the growth and development of cotton; the right part illustrates the related mechanisms and factors of cotton salt tolerance.
Ijms 24 09517 g001
Figure 2. Ethylene biosynthesis and signal transduction. Met: methionine; SAM: S-adenosylmethionine; ACC: 1-Aminocyclopropanecarboxylic Acid; SAMS: SAM synthetase; ACSs: ACC synthases; ACOs: ACC oxidases; CTR1: Constitutive triple response1; EIN2: Ethylene-Insensitive2; EIN3/EIL1: Ethylene-Insensitive3/EIN3-like1; ENAP1: EIN2 nuclear-associated protein1; EBF1/2: EIN3 binding F-box protein1/2; ERFs: Ethylene-response factors.
Figure 2. Ethylene biosynthesis and signal transduction. Met: methionine; SAM: S-adenosylmethionine; ACC: 1-Aminocyclopropanecarboxylic Acid; SAMS: SAM synthetase; ACSs: ACC synthases; ACOs: ACC oxidases; CTR1: Constitutive triple response1; EIN2: Ethylene-Insensitive2; EIN3/EIL1: Ethylene-Insensitive3/EIN3-like1; ENAP1: EIN2 nuclear-associated protein1; EBF1/2: EIN3 binding F-box protein1/2; ERFs: Ethylene-response factors.
Ijms 24 09517 g002
Figure 3. Polyamines biosynthesis and catabolism in plants. Met: methionine; SAM: S-adenosylmethionine; dcSAM: decarboxylated S-adenosylmethionine; GABA: γ-aminobutyric acid; ODC: ornithine decarboxylase; ADC: arginine decarboxylase; CDC: citrulline decarboxylase; AIH: argmatine iminohydrolase; CPA: N-carbamoylputrescine amidohydrolase; ADI: Agmatine deiminase; SAMDC: S-adenosylmethio-Nine decarboxylase; SPDS: spermidine synthase; SPMS: spermine synthase; DAO: Diamine oxidase; PAO: Polymine oxidase.
Figure 3. Polyamines biosynthesis and catabolism in plants. Met: methionine; SAM: S-adenosylmethionine; dcSAM: decarboxylated S-adenosylmethionine; GABA: γ-aminobutyric acid; ODC: ornithine decarboxylase; ADC: arginine decarboxylase; CDC: citrulline decarboxylase; AIH: argmatine iminohydrolase; CPA: N-carbamoylputrescine amidohydrolase; ADI: Agmatine deiminase; SAMDC: S-adenosylmethio-Nine decarboxylase; SPDS: spermidine synthase; SPMS: spermine synthase; DAO: Diamine oxidase; PAO: Polymine oxidase.
Ijms 24 09517 g003
Figure 4. GhSAMC silencing in cotton increases the sensitivity to salt stress. (A) Albino phenotype of TRV:CLA1 about two weeks after infiltration. (B) The silencing efficiency of GhSAMC. (C) Phenotype of silencing cotton seedlings before salt treatment. (D) Phenotype of silencing cotton seedlings after salt treatment. (E) Expression of GhSAMC over time under treatment with 250 mM saline solution. (F) MDA content of silencing cotton seedlings after salt treatment. (G) T-AOC content of silencing cotton seedlings after salt treatment (Values are means ± s.e.m (n = 3 biological replicates). Error bars represent the SD of three biological replicates. * p < 0.05 (Statistically significant), ** p < 0.01 (Statistically highly significant)).
Figure 4. GhSAMC silencing in cotton increases the sensitivity to salt stress. (A) Albino phenotype of TRV:CLA1 about two weeks after infiltration. (B) The silencing efficiency of GhSAMC. (C) Phenotype of silencing cotton seedlings before salt treatment. (D) Phenotype of silencing cotton seedlings after salt treatment. (E) Expression of GhSAMC over time under treatment with 250 mM saline solution. (F) MDA content of silencing cotton seedlings after salt treatment. (G) T-AOC content of silencing cotton seedlings after salt treatment (Values are means ± s.e.m (n = 3 biological replicates). Error bars represent the SD of three biological replicates. * p < 0.05 (Statistically significant), ** p < 0.01 (Statistically highly significant)).
Ijms 24 09517 g004
Figure 5. Transcriptome analysis of cotton leaves under salt stress. (A) GO enrichment analysis of differentially expressed genes. In the red boxs are some oxidation-reduction related pathways (B) KEGG enrichment analysis of differentially expressed genes. The three pathways in the red box, tyrosine metabolism, cytochrome P450, and phenylpropanoid metabolism, play an important role in salt stress. (C) The heat map of differentially expressed genes in cotton leaves under salt stress. (D) Heatmap of differentially expressed genes related to phytohormone. (E) Statistics of the number of differentially expressed transcription factors. (F) Model diagram of GhSAMC regulation in cotton under salt stress. The red upward arrow indicates an increase in the content of the substance. The red downward arrow indicates an decrease in the content of the substance. ? means that it is uncertain whether the content of the substance is increase or decrease.
Figure 5. Transcriptome analysis of cotton leaves under salt stress. (A) GO enrichment analysis of differentially expressed genes. In the red boxs are some oxidation-reduction related pathways (B) KEGG enrichment analysis of differentially expressed genes. The three pathways in the red box, tyrosine metabolism, cytochrome P450, and phenylpropanoid metabolism, play an important role in salt stress. (C) The heat map of differentially expressed genes in cotton leaves under salt stress. (D) Heatmap of differentially expressed genes related to phytohormone. (E) Statistics of the number of differentially expressed transcription factors. (F) Model diagram of GhSAMC regulation in cotton under salt stress. The red upward arrow indicates an increase in the content of the substance. The red downward arrow indicates an decrease in the content of the substance. ? means that it is uncertain whether the content of the substance is increase or decrease.
Ijms 24 09517 g005aIjms 24 09517 g005b
Table 1. The functions of ethylene-related factors in cotton.
Table 1. The functions of ethylene-related factors in cotton.
Gene NameExperimental MethodsBiological FunctionRef.
GhACS1RNA-Seq data analysis and qRT-PCR analysisResponsed
to salt stress
(Li J et al., 2022 [102])
GhACO106-AtOverexpression
in Arabidopsis
Promoted flowering and increased salt tolerance(Wei H et al.,2021 [104])
GhEIN3Overexpression
in Arabidopsis
VIGS in Cotton
Regulated ROS pathway and ABA signaling to response to salt(Wang X et al., 2019 [119])
GhERF38Overexpression
in Arabidopsis
Responsed to salt/drought stress and ABA signaling(Ma L et al., 2017 [121])
GhERF13.12Overexpression
in Arabidopsis
VIGS in Cotton
Regulated ROS pathway and ABA signaling to response to salt stress(Lu L et al., 2021 [122])
GhERF12VIGS in CottonRegulated ROS pathway to response to salt stress(Zhang J et al., 2021 [123])
GhERF4L/54LVIGS in CottonResponsed to salt stress(Long L et al., 2019 [124])
Ghi-ERF-2D.6/12D.13/6D.1/7A.6/11D.5RNA-Seq data analysis and qRT-PCR analysisResponsed to salt stress(Zafar M et al., 2022 [126])
GhERF109RNA-Seq data analysis and qRT-PCR analysisResponsed to salt stress(Bano N et al., 2022 [127])
GhERF1Semi-qRT-PCR analysisResponsed to ET, ABA, salt, cold, and drought stress(Qiao Z et al., 2008 [128])
GhERF2/3/6Semi-qRT-PCR analysisResponsed to ET, ABA, salt, cold, and drought stress(Jin L et al., 2010 [129])
GhERF4Semi-qRT-PCR analysisResponsed to ET, ABA, salt, cold, and drought stress(Jin L and Liu J., 2008 [130])
GhERF5Semi-qRT-PCR analysisResponsed to ET, ABA, salt, cold, and drought stress(Jin L et al., 2009 [131])
Table 2. Polyamines (PAs) biosynthesis and catabolism genes in mitigating salt stress in various crops.
Table 2. Polyamines (PAs) biosynthesis and catabolism genes in mitigating salt stress in various crops.
PAsCropsGenesGenes Response to
Salt Stress
Ref.
Biosynthesis
genes
Arabidopsis
thaliana
AtADC2Improved SOD and CAT activities(Fu Y et al., 2017 [150])
CottonGhADC2Increased H2O2 content and oxidative stress(Gu Q et al., 2021. [24])
Fragaria vescaFvSAMDCReduced H2O2 and O2•− content(Kov’acs L et al., 2020. [154])
Cleistogenes songoricaCsSAMDC2Improved chlorophyll content and Photosynthetic capability(Wu F et al., 2022. [155])
CottonGhSAMDC3Increased Spd content(Tang X et al., 2021. [156])
PeanutAhSAMDCImproved activities of
antioxidant enzymes
Increased Spd and Spm content
(Meng D et al., 2021. [157])
White cloverTrSAMDC1Improved SOD, POD, and CAT activities
Reduced MDA and H2O2 content
(Jia T et al., 2021. [158])
Catabolic
genes
Arabidopsis
thaliana
AtPAO1Increased ROS and H2O2 content(Sagor G et al., 2016. [164])
CucumberCsPAO2Improved activities of
antioxidant enzymes
Reduced MDA content
(Wu J et al., 2022. [165])
CucumberCsPAO3Improved POD and CAT activities
Reduced MDA and H2O2 content
(Wu J et al., 2022. [166])
RiceOsPAO3Increased PAs content Improved Polyamine oxidase activities(Liu G et al., 2022. [167])
CottonGhPAO3Increased PAs content(Cheng X et al., 2017. [168])
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, L.; Wang, X.; Zhao, F.; Zhang, X.; Li, W.; Huang, J.; Pei, X.; Ren, X.; Liu, Y.; He, K.; et al. Roles of S-Adenosylmethionine and Its Derivatives in Salt Tolerance of Cotton. Int. J. Mol. Sci. 2023, 24, 9517. https://doi.org/10.3390/ijms24119517

AMA Style

Yang L, Wang X, Zhao F, Zhang X, Li W, Huang J, Pei X, Ren X, Liu Y, He K, et al. Roles of S-Adenosylmethionine and Its Derivatives in Salt Tolerance of Cotton. International Journal of Molecular Sciences. 2023; 24(11):9517. https://doi.org/10.3390/ijms24119517

Chicago/Turabian Style

Yang, Li, Xingxing Wang, Fuyong Zhao, Xianliang Zhang, Wei Li, Junsen Huang, Xiaoyu Pei, Xiang Ren, Yangai Liu, Kunlun He, and et al. 2023. "Roles of S-Adenosylmethionine and Its Derivatives in Salt Tolerance of Cotton" International Journal of Molecular Sciences 24, no. 11: 9517. https://doi.org/10.3390/ijms24119517

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop