You are currently viewing a beta version of our website. If you spot anything unusual, kindly let us know.

Preprint
Review

Recent Advancement on Physiological and Molecular Response to Cotton under Salt Stress: A Review

Altmetrics

Downloads

218

Views

144

Comments

0

Submitted:

19 July 2023

Posted:

21 July 2023

You are already at the latest version

Alerts
Abstract
The production of plants and crops is influenced by environmental stress, which is a serious scientific issue. Cotton is an essential crop for producing natural fibers that are used to make biofuel and edible oils. Salinity is the main element that influences cotton growth during the beginning of germination. The type of salt and the growth stage affects how sensitive a plant is to salt stress. Developing ways to enhance cotton performance in salty circumstances can be aided by an understanding of the response of cotton to salt, its mechanism of resistance, and its management methods. Osmotic and ionic imbalances originate due to the deposition of soluble salts under salinity stress in the plant's root zone. Soil salinity significantly reduces plant growth due to several factors, such as nutritional ion imbalance, which reduces K+, PO4-, and NO- absorption, excessive salt chloride concentrations, and osmotic stress, which hinders water availability. Research has revealed that compared to subsequent stages, the germination, emergence, and seedling phases are more vulnerable to salinity stress, which ultimately affects the seed cotton yield by delaying blooming, reducing fruiting positions, shedding fewer fruits, and reducing boll weight. The morphology, growth of cotton roots, shoots, yield, and fiber quality are all strongly impacted by salinity stress. It slows down plant feeding, cellular metabolism, and photosynthesis. The soil, water, and climate all affect how the plant responds to salinity stress. During salt stress, excessive exclusion of sodium or its compartmentalization is the key adaptation process in cotton. A major adaptive potential to create cotton types that can withstand salt is provided by the up-regulation of both physicochemical and non-enzymatic antioxidant genes. A successful strategy to increase cotton germination in saline soils is seed priming. Moreover, the transgenic strategy might be a viable choice for improving cotton yield in saline environments. Our review focuses on the impacts of salinity on cotton productivity as well as how plants react to salt stress. It also clarifies recent genetic advancements and molecular breeding for cotton's resistance to soil salt. To create salt-tolerant cultivars, a combination of traditional breeding and novel molecular approaches will be useful.
Keywords: 
Subject: Biology and Life Sciences  -   Agricultural Science and Agronomy

1. Introduction

Cotton (Gossypium hirsutum L.), is a cash crop for the textile industry, and contributes to 35% of the total fiber consumed globally (Salimath et al., 2021). China is considered as a leading country regarding cotton production followed by India, USA, Brazil and Pakistan (Ahmad et al., 2022). During growth and development, cotton crops are frequently subjected to abiotic stresses, which limits the lint yield and fiber quality (Zhang et al., 2021). Salt stress is the main hindrance to cotton production globally. Identification and functional characterization of salt-tolerant genes, molecular breeding, and understanding the functions of exogenous hormones under salt stress have become hot topics for researchers.
Salt stress affects over 30% of the world's arable land, and this problem worsens continously (Pitman and Läuchli, 2002;Shrivastava and Kumar, 2015). According to adaptive evolution, there are two major categories of plants: glycophytes and halophytes (which can tolerate salinity) (Abobatta, 2020). Salinity stress could severely influence on metabolic and physiological mechanisms and eventually reduce crop yields (Salwan and Vivek, 2023). The well-known process by which soil salinity affects plant development is initially through osmotic process and ion toxicity (Babayev et al.;Nadeem et al., 2022). One of the most alarming environmental problems of the twenty-first century is soil salinity (Phour and Sindhu, 2023). It is an important to explore the prospective solution to combat the soil salinity threats (Fróna et al., 2019). Currently ,only two main strategies for reducing soil salinization are utilizing chemical additives to repair soil degradation and applying recent advanced biotechnology tools to develop salt-tolerant plant varieties (Ashraf and Wu, 1994). When comparing these two techniques , the former method considered as a expensive and causing secondary salinization risk (Stigter et al., 2006). The ability of cotton plants to tolerate salinity varies genetically depending on the species. Cotton plant biologists can describe physiological mechanisms, pairs of genes, and transcription factors concerned with stress tolerance and combine them into selected crops to generate salt-tolerant cotton plants with various levels and variable responses to salinity stress.
A specific, as of yet unidentified Na+-sensing module recognizes Na+. Early perception is followed by the induction of upstream signalling g expressions. The complicated signal transduction network includes the following molecules: K+, Ca2+, H+, phospholipid, ROS, different protein kinases, and plant hormones. Transcriptional elements that control stress-responsive genes are subsequently activated. Functional genes, such as those for ion transporters and antioxidants, are consequently effectively expressed. Inference-based reactions, such as ROS scavenging, are explained by plants that can withstand and adapt to salt stress. In plants to develop resistance, each link in the chain is crucial. To address this, we can modify plants grown in saline soil. To use comparable techniques on plants that are not salt-tolerant, Scientists are investigating salinity tolerance mechanisms in salt-stress plants (Liang et al., 2018) (Figure 1).
The objective of current review is to present a brief overview on the situation of the art research on plants under salt stress. We described how cotton plants could become damaged by salt, the mechanism of salt stress signalling, and the mechanism used by cotton plants to withstand salinity in this review paper. The current study will help the readers better understand the background knowledge on cotton plants under salt stress. In addition, we offered our opinions on the future course of this field's study.

2. Salonization in Cotton

Salinity stress alters the physiological activity of the seed, which has a detrimental influence on seed germination (Ibrahim, 2016). Salinity reduces food storage , protein content, and water potential in various plant seeds During germination (Wu et al., 2019). Salt stress decreased phosphatase activity in Arabidopsis thaliana seeds, ultimately leading to inhibiting seed germination (Shi et al., 2003). By lowering the amount of gibberellic acid (GA3), and raising the composition of soluble sugar, and starch in cotton plants, the germination rate was reduced by salts (Bai et al., 2020). Cotton plant development and the formation of its essential components are significantly impacted by salinity stress (Brugnoli and Björkman, 1992). Biological mass, foliage area, yield, stem length, and root length all declined in high-saline soil (Arif et al., 2020). The growth parameters, including, the percentage of lint yield, the weight of the cotton ball, the number of cotton balls made per plant, and the crop quality are all significantly impacted by soil salinity (Sabagh et al., 2021).
The previous study indicated that cotton plants exhibit delayed growth, reduced branch growth, decreased plant height, lower germination rate, and wilted leaves when exposed to salt stress (Qin et al., 2022). Due to the salinity, the mechanisms of osmotic pressure become disturbed, which inhibits plant growth (Safdar et al., 2019). Ion toxicity is what happens when too many salty ions enter a plant's transpiration stream and interruption of photosynthesis, deterioration of ion homeostasis, and oxidation of membrane lipids all harm plant cells (Hasanuzzaman et al., 2013). In conclusion, understanding how plants physiologically to salt stress are essential for increasing plants' tolerance to salts (Figure 2).

3. Salonization and Photosynthetic Activity in Cotton

3.1. Effect of Salinity on Stomatal Conductance

Salinity hinders gas exchange and lowers stomatal conductance, which decrese the rate of photosynthetic activity. Due to the salinity halophyte grasses produce less photosynthetic pigments. High accumulation of salinity in broccoli results in decreased indicators of photosynthetic activity, total chlorophyll contents, and grana-reducible chloroplast structural degradation (Del Amor and Cuadra-Crespo, 2011;Hamza and AL-Taey, 2020). Increasing salt concentrations significantly impact Chickpea's photosynthetic electron transport chain (Çiçek et al., 2018). Due to high salinity levels, transpiration, photosynthesis, chlorophyll contents, and stomatal conductance were decreased in different pomegranate genotypes (Soori et al., 2019). Moreover, lowering carboxylation results in a decrease in mesophyll conductivity. Reduction of photosynthetic apparatus lead to reduced photosynthetic capability and photosynthetic characteristics because of soil salinization (Mbarki et al., 2018). High salt contents caused abnormalities in stomata structure, leading to shortened chlorophyll pigments, further reducing photosynthesis (Yildiztugay et al., 2020). salinity significantly influenced net photosynthesis, the Fv/Fm ratio, stomatal conductance, and rocket plants' transpiration rate (Hniličková et al., 2017).

3.2. Salinity Influence on Total Chlorophyll Contents

High salinity decreases the amount of total chlorophyll, chlorophyll fluorescence, and Rubisco activity in maize. Additionally, it alters the shape of the chloroplast and stomata (Xu et al., 2018). In mustard, prolonged soil salinity decreased photosynthetic properties, particularly by reducing PSII (Fv/Fm) activity (Wani et al., 2019). In lisianthus cultivars, salinity decreased gaseous exchange and transpiration (Ashrafi and Rezaei Nejad, 2018). Total closure and a decrease in photochemical and carbon metabolism are two ways that salinity inhibits photosynthesis (Ju et al., 2023). High salinity drastically reduces PSII photochemistry's maximum quantum yield (Liu et al., 2023b). As a result, salinity in cotton plants decreases photosynthesis, chlorophyll fluorescence, transpiration, and stomatal conductance by degrading ultrastructure, chlorophyll, and carotenoid production.

3.3. Osmotic Pressure Affected by Salinity

Salt stress has a major effect on the rate of photosynthetic activity, osmotic potential, water potential, transpiration rate, leaf temperature, and the overall water content of cotton plant leaves (Liu et al., 2023a;Naikwade, 2023). There were two negative impacts of salt stress on photosynthesis (Singh et al., 2022b). First off, salt stress causes plants' stomas to close, lowering the level of CO2 in their cells (Yildirim et al., 2022). Second, the system of photosynthetic membranes is harmed by salt stress, which also reduces the ability to absorb CO2, restrictions not connected to the stoma, like harmed photosynthetic pigments (Zahra et al., 2022). Specialized epidermal cells called a stoma, commonly called guard cells, create pores on the surface of cotton plant leaves (Wall et al., 2022). When the osmotic pressure within the cells falls, guard cell water outflow will result in stoma closure and the contraction of guard cells (Liu et al.). This often results in salt stress leaves having higher Na+ and Cl contents, lower K+ contents, and eventually closing their stomas (Jadidi et al., 2022). The reduced stoma opening significantly decreases the intercellular CO2 concentration and photosynthetic rate, as well as CO2 diffusion from the environment to the chloroplast (Zou et al., 2022).

3.4. Salt Stress Harms the Components of Photosynthesis

In addition to stomatal restrictions, salt stress harms several other components of photosynthesis but does not directly affect the stoma (Singh et al., 2022a). Chlorophyll, the main pigment involved in photosynthesis in salt-sensitive plants, usually decreases under salt stress. The same results were found after salt stress in rice (Somaddar et al., 2022), soybean (Begum et al., 2022), and pepper (Atzori et al., 2022), the ultrastructure of chloroplasts is similarly impacted by salt stress. The light reaction occurs on the thylakoid membrane, the most significant part of the chloroplast.

3.5. Salt stress Influence on Photosynthetic Enzyme Activities

The impact of salt stress extend to components of photosynthesis, especially the structural proteins and enzymes engaged in the transmission of photosynthetic electrons, the fixation of CO2, and the absorption of light energy (Zahra et al., 2022). (Wei et al., 2022), found indicate a connection between the photosynthetic process and glucose-6-phosphate dehydrogenase (G6PDH). Interfering with the oxygen-evolving complex's ability to function, salt stress reduces electron transfer efficiency by preventing electron transfer from plastoquinone A (QA) to plastoquinone B (QB), and by impairing the function of the pigment-protein complex on the thylakoid membrane (Sghaier et al., 2022;Yan et al., 2022). Salt stress also affects rubisco, a rate-limiting enzyme in photosynthesis(Li et al., 2022a). Salt stress reduces Rubisco's activity and restricts the restoration of the compound ribulose-1, 5-bisphosphate (RuBP), which has an impact on plants' ability to take up and utilize CO2 (Bhattacharya, 2022;Yang et al., 2022). According to previous research, salt stress prohibits salt-intolerant plants from photosynthesis due to stomatal and non-stomatal restrictions (Iqbal and Ashraf, 2005).
Some contend that stomatal closure results from physiological salt stress (Azevedo Neto et al., 2004). Moreover, some research has shown that salt stress does not cause reduce in photosynthetic activity to occur at the exact time as a decrease in plant water potential. It's a common misconception that stomatal closure doesn't appear to be brought on by a drop in leaf water potential (Williams et al., 2012). According to others, the feedback cycle resulting from plant development in salty soils restricts photosynthesis by raising the sugar level in plant tissues (Bernstein and Hayward, 1958;Ruiz-Lozano et al., 2012;Wang et al., 2023). This viewpoint, cannot explain the extremely low glucose concentration found in perennial woody plant stems that have been exposed to persistent salt stress. Consequently, a greater investigation is required into how salt stress is decreased by photosynthesis in plants (Hashemi et al., 2023).

4. Ion Homeostasis Disruption

Mineral elements, which normally reside as ions in cotton plant cells, make up the structural elements of cells (Naz, 2023). Minerals serve as coenzyme and enzyme building blocks that regulate enzyme activity (Panda and Zhou, 2023). They also engage in the electrochemistry of charge neutralization, colloidal stabilization, and osmotic regulation (Hou et al., 2023). The harmonization of ion metabolism is required for the consistency of the cell membrane and the promotion of plant growth (Adhikari et al., 2023;Phour and Sindhu, 2023). The major sources of salt damage are an excess of Na+ and Cl+ and a corresponding lack of important ions like Ca2+ and K+ (Isayenkov and Maathuis, 2019;Vivaldi et al., 2021) because there is so much Na+ in the soil, it accumulates in plants. Lowered membrane potential and promoted Cl absorption along a chemical gradient are both effects of high Na+ concentrations. High Na+ concentrations prevent cell division and growth because they lead to osmotic imbalance, membrane dysfunction, and enhanced ROS production (Ahmad et al., 2019). Too much Na+ can harm some enzymes and affect cellular metabolism (Nieves-Cordones et al., 2016). A high Cl concentration hinders plant growth, contributing to reducing the quantity of chlorophyll present and harming the organelle and cell membrane systems (El-Banna and Abdelaal, 2018). In addition to their toxic effects, Na+ and K+ compete since they have similar ionic radii and hydration energies (Kaya et al., 2007). To ensure their physiological function, the majority of cells maintain a cytoplasmic K+ concentration that is quite high and a cytoplasmic Na+ concentration that is relatively low (Leigh and Wyn Jones, 1984). As a result, a significant external Na+ influx will stop K+ from entering, preventing the K+ deficiency that damages plants (Kumari et al., 2021). According to the researcher, that Na+ competitive inhibition, the amount of Ca2+ in cells also drops (Rengel, 1992). Ca2+ is an important component in the signal transduction of some physiological mechanisms in crops. (Bernstein and Hayward, 1958), claimed that when there are high Na+ concentrations present, plants can adjust their osmotic pressure compensating reduction in Ca2+. Furthermore, (Dai et al., 2015) showed that salt stress decreases plants' capacity to absorb NH4+ and NO3. In a nutshell, the ion balance in plant cells is upset by salt stress, which influences cotton plant progress and growth (Ashraf et al., 2018b).

5. Salt stress and Membrane Permeability

The movement of chemicals, energy, and information transmission are all made possible by the cell membrane, which protects plant cells and serves as an essential protective barrier (Finkina et al., 2016). The cell membrane's capacity to control ion homeostasis in a specific way, and maintenance of plant physiological functions. However, increased ROS generation under salt stress causes significant membrane damage by raising relative permeability and lowering fluidity (Gill and Tuteja, 2010). However, it also produces many electrolyte exosmotic, leading to osmotic pressure problem.
Membrane-bound proteins and enzymes may become inactive due to MDA contents. MDA is a primary consequence of membrane lipid peroxidation, which could harm the biofilm's structure and functions (Pigeot-Rémy et al., 2012). The MDA contents in the plasma membrane is an important indicator of the extent of the damage. In previous study, it was observed that Carex rigescens' MDA content significantly increased following salt stress treatment (Li et al., 2018a).It was discovered that salt stress increased the MDA content of soybean. (Li et al., 2017) The establishment of the endoplasmic reticulum and the plasma membrane vacuole are just two examples of organelle structures within cells that increased ROS can harm. In other words, oxidation caused by excessive ROS formation due to salt stress, which prevents plant growth by damaging plant cells (Khan et al., 2011).

6. Characteristics of Salt Stress in Cotton

The first physiological sign of salinity stress in plants is the measurement of Na+ or Cl in cotton leaves and roots; it's common to use both the K+ content and the K+/Na+ ratio (Abbas et al., 2010). K+ level in Arabidopsis decreased after 4 hours of application of 100 mM NaCl (Demidchik et al., 2010). Additionally, because they affect nutrient intake, other studies are conducted on the ratios of other ions, such as Na+/Ca2+, Ca2+/Mg2+, and Cl/NO3 (Malik, 2022). The photosynthetic process is also slowed down by salt stress (Liu et al., 2019). While the cotton plant is under salt stress, stomatal conductance, chlorophyll fluorescence, and chlorophyll concentrations are also measured (Hatam et al., 2020). Chlorophyll contents in wheat reduced after three days of salt application (100 mM) (Tian et al., 2015). Osmotic pressure frequently coexists with salt stress, usually established by changes in water potential, relative water content (RWC), turgor pressure, and osmotic pressure (Yin et al., 2013). For instance, tomatoes' water potential decreases after two hours of exposure to 100 mM NaCl (Mäkelä et al., 1998). The ROS is significantly elevated when under salt stress (Shi et al., 2007). Typically, a rise in ROS production under salt stress damages cell membranes (Tanveer and Ahmed, 2020). Thus, electrolyte leakage and water loss are indicators of cell membrane breakdown, another salt stress symptom (Khan et al., 2013). As an illustration, when 100 mM NaCl was applied for two days, electrolyte leakage from the sunflower increased (Aziz et al., 2019).

7. Cotton Tolerance to Salt Stress: Physiological and Biochemical Mechanisms

Salinity stress causes physiological and morphological problems in cotton leading to decrasing its yield (Etesami and Noori, 2019). Plant nutritional and ionic imbalances occur in response physiological and morphological alteration under salts stress (Shahid et al., 2020). The combined effects of salinity stress on cotton plant development, physiological processes, and biochemical features have not been the subject of any meta-analyses (Malik et al., 2022).
Salinity stress often causes cotton plants to produce less biomass (Galić et al., 2020). In numerous investigations, salinity stress was demonstrated to have an additive influence on the buildup of dry matter (Prodjinoto et al., 2021). Some studies claim that shoot length is decreased because salinity raises the concentration of NaCl (Chauhan et al., 2019;Kumar et al., 2021). Varying water stress levels did not significantly impact shoot length, SDW, and RDW (Masuda et al., 2021;Angon et al., 2022). The total amount of assimilates decreased due to the negative effects of salt stress on plant photosynthetic rate and biomass accumulation across the growth duration (Yadav et al., 2019;Raziq et al., 2022). According to research, salt stress drastically lowers the amount of chlorophyll (Bibi et al., 2019;Gerona et al., 2019). However, consistent irrigation is linked to the highest levels of chlorophyll (Khalvandi et al., 2021).

7.1. Osmotic and Salinity

The cotton plant maintains its capacity to absorb water through osmotic adjustment, raising its cells' solute content (Hou et al., 2021). Under osmotic adjustment, there are two osmotic correction pathways. One potential is that the solute itself needs to serve as an osmolyte. However, solutes serve as a defence force by preserving the structural integrity of biological macromolecules. This poor attachment of water molecules to the hydrophobic regions of the protein surface, especially in hypertonic circumstances, may be the mechanism underlying the phenomena, in a hypertonic environment, cells initially lose this percentage of water. Proteins' surface polarity can be increased by dissolving osmotic regulators in the water that binds to them, improving the proteins' ability to adhere to water (Schobert and Tschesche, 1978). The two main categories of osmotic regulators are organic compounds and inorganic ions. In reaction to osmotic, these substances will be present in cells to variable degrees.

7.2. Organic Substances

Proline is quite capable of hydrating. Its hydrophilic end can cling to water molecules, while its hydrophobic end can bind to proteins. Proline-linked proteins are better able to bind to water, which protects them against dehydration and osmotic causes proteins to denatured. Plants can increase proline synthesis and decrease proline breakdown to accumulate proline in response to stress (Verbruggen and Hermans, 2008). Proline may also start to function as an antioxidant to stop the oversupply of reactive oxygen species that occurs in reaction to stress. It can assist plants in overcoming stress by playing the role of a quick source of nitrogen and carbon for regeneration (Hayat et al., 2012). It results in some stress-related reactions as a sign of stress. Several studies have found that proline can play significantly role in recovering plants from salinity (Banu et al., 2009;Ahmad et al., 2012). Proline treatments to different cultivars of Brassica juncea reduced the adverse effect of salinity on photosynthetic activities leading to improve their growth (Mittal et al., 2012).
Additionally, glycine betaine acts as an osmotic regulator. The enzymes choline monooxygenase (CMO) and betaine aldehyde dehydrogenase oxidize choline to produce betaine (BADH) (Bao et al., 2011). Plants have 12 different types of betaine, with glycine betaine being the most well-known and fundamental (Hafez et al., 2021). To decease the damage, which caused by salt stress, it may promote the integrity of plant cell membranes and enzyme activity. It has been found that many higher plants, especially those in the Gramineae and Chenopodiaceae families, acquire betaine exposed to salt stress (Flowers and Colmer, 2008). (Gupta et al., 2018), observed an increase in glycine betaine synthesis in halophytic seashore paspalum exposed to salt stress caused by choline. Numerous studies have also shown the benefits of topically administering betaine to plants experiencing salt stress (Hirt and Shinozaki, 2003;Ilyas et al., 2022).
Osmotic regulators can also be made from polyols and soluble carbohydrates. The metabolism of plants depends heavily on non-structural carbohydrates like glucose, sucrose, fructan, starch, and others. In particular, sucrose synthesis is increased when plants are under osmotic stress by raising the activity of sucrose phosphate synthase, which is closely associated with plant development and environmental sensitivity (SPS). According to (Peng et al., 2016), the levels of sucrose and starch increased along with the activities of sucrose phosphate synthase (SPS) and sucrose synthase in the main-stem leaves and subtending leaf bolls of cotton (Gossypium hirsutum L.). Even with a significant decrease in starch synthesis, the sucrose concentration in the plant cells of Phaseolus vulgaris may remain stable. Sucrose phosphate synthase activity was reduced by osmotic stress (SPS), but it also greatly reduced the synthesis of starch, allowing sucrose concentration to stay relatively high.

7.3. Inorganic Ions

Most inorganic ions that make up the osmotic pressure in cotyledons' cells are K+, Na+, and Cl. K+ is a crucial element for plant development and it’s essential in preventing the effect on plant cells caused by salt stress. Greater plant hydration status led to increased biomass and improved salt tolerance in stressed peanuts after exogenous K+ treatment, as shown by (Chakraborty et al., 2016).
In non-halophytes, rapid Cl uptake by plants during the earlier than-usual stages of salt stress elevated osmotic regulation of the root scheme. Moreover, for the final stages of osmotic pressure regulation, Cl still requires Na+ or K+. The presence of Ca2+ in the cytoplasm is required for signal transduction. Ca2+ channels are encouraged to open while plants endure being exposed to salt stress. Cell metabolism and gene translation are controlled by Ca2+ binding with calmodulin or other calcic binding proteins, and this is referred to as a calcic binding protein, plants are better able to adapt to adversity. To put it another way, Plants survive their capacity to accumulate water from their surroundings when they are affected by salt stress by rising the composition of inorganic ions, such as K+, and organic compounds, Proline and betaine are two examples.

7.4. Antioxidants Enzymatic System against Oxidation

SOD is the initial line of defence in plants' entire antioxidant enzymatic system against oxidative stress (Ighodaro and Akinloye, 2018). O2 is dismutase into H2O. SOD can indeed be divided into CuZn-SODs, Mn-SODs, and Fe-SODs based on the variations in the metal ions that bind to the SOD. Fe-SODs, Mn-SODs, and CuZn-SODs are primarily discovered in chloroplasts, mitochondria, and extracellular space, respectively (Corpas et al., 2006). CuZnSOD was introduced into a salt-sensitive sweet potato by Yan et al. and found using the stress- reflect the actual SWPA2 promoter. Under salt stress, transgenic plants have shown a significantly higher expression level than wild-type (Baghour et al., 2019).
To form phenoxyl radicals (PhO), phenolic compounds (PhOH) must be oxidized by POD. In this process, H2O2 acts as an electron acceptor and is then transformed into 2H2O (Gross et al., 2006). POD is present in many different plant tissues. As a high-activity adaptive enzyme, it can show the development and growth of plants, in vivo metabolism status, and environmental adaptation. Research conducted by (Yang et al., 2023), that POD activity enhanced with the aid of AMF (arbuscular mycorrhizal fungus) in response to the ROS effect brought on by NaCl treatment.
CAT may break down H2O2 quickly, releasing H2O and O2. The ability of CAT to reduce oxidative damage is demonstrated by the fact that the better the CAT activity in plants, the lower the H2O2 rise ratio. The impact of H2O2 pretreatment on maize leaves was investigated by (Bakala et al., 2023). The highest initial activity and the strongest response to H2O2 were displayed by CAT, showing the role it plays in plants' ROS scavenging (Jiang et al., 2023).
H2O2 can be used as an oxidant in APX to catalyze the conversion of AsA to MDHA. Higher plants' APX enzymes are divided into chloroplast APX and cytosolic APX (cAPX) (chlAPX). Stroma APX (sAPX) and thylakoid-bound APX are chloroplast APX (tAPX) types. Microsomes (peroxisomes, glyoxysomes) and some enzymes whose cytoplasmic locations are still unknown are among the cytoplasmic kinds. The oxidation of AsA to form MDHA and, consequently, DHA, as well as the detoxification of H2O2, is processed in which all enzymes can take part. In the AsA-GSH cycle, several enzymes play a part in the process regeneration of AsA and GSH but do not directly interact with ROS (Figure 3). AsA is generated from MDHA by the NADH or NADPH-dependent enzyme MDHAR. To achieve AsA regeneration, DHAR catalyzes the oxidation of DHA (Ding et al., 2023). The AsA-GSH cycle regulates redox homeostasis by converting GSSG to GSH and depends on the enzyme GR. Enzyme GPX may decrease H2O2 and shield cells from oxidative effects with the aid of GST by using GSH and thioredoxin. They discovered strong evidence within the research of Gaafar and Seyam to support the AsA-GSH cycle's participation in salt tolerance and the mechanism of Egyptian lentil varieties (Lens culinaris Medikus) (Gaafar and Seyam, 2018).

7.5. Non-Enzymatic Antioxidants

Active growth regions like meristems, photosynthetic cells, root tips, flowers, and immature fruits are rich sources of AsA, a potent water-soluble antioxidant. Normally, AsA plays a substantial role in scavenging ROS under stress by acting as a co-enzyme and giving electrons. Additionally, AsA plays a role in regenerating the crucial antioxidant tocopherol. Additionally, AsA is regarded as a key signal substance in controlling the redox state of cells and is crucial for the mitochondrial electron transport chain and photosynthetic system (Mittler, 2002). In numerous investigations, Exogenous AsA has been indicated to stimulate plant growth and enhance plant resilience to salt stress. Exogenous AsA, for instance, dramatically increased the development of wheat seedlings when there was a water constraint (Elkeilsh et al., 2019).
Plants can combine lowered glutathione (GSH) and oxidized glutathione (GSSG). Scavenging ROS allows GSH to play a crucial part in controlling intracellular defence. Additionally, as a part of the AsA-GSH cycle, GSH supports redox equilibrium. In the first phase of the AsA-GSH cycle, which uses AsA as an electron donor, APX reduces H2O2 to water. MDHAR regenerates the oxidized AsA (MDHA). MDHA, on the other hand, is a free radical that, if not rapidly simplified, is converted to DHA. DHAR converts DHA to AsA at the expense of GSH, a way that results in GSSG. In Jiang DX. et al. research, Exogenous AsA, and GSH increased the levels of endogenous AsA and GSH, decreased the levels of H2O2 and MDA, and Exogenous AsA and GSH improved the contracts of SOD, APX, and GR in the chloroplasts of two rice cultivars under salt stress, implying that exogenous AsA and GSH may be beneficial for salt stress resistance and ROS scavenging (Jiang et al., 2020).
CAR can be utilized in chloroplasts as supplemental pigments for light collection. Additionally, it can eliminate ROS generated by photosynthesis equipment. CAR security measures the photosystem in a total of four ways: they work with lipid peroxidation materials to stop the chain reaction; they eliminate singlet oxygen and release it as heat energy; they interact with triplet or activated chlorophyll molecules to prevent singlet oxygen formation; and they release extra energy via the lutein cycle. Tocopherol is also a significant ROS scavenger in plants, particularly tocopherol (Ve). To control the stabilization of membrane lipids, it snares ROS and free radicals generated by lipid peroxidation. Active oxygen scavengers known as flavonoids have received significant attention in recent years; however, it has been discovered that they only function at the close site of ROS production, like a vacuole or a cell wall. Proline, for example, is an osmotic regulator (Kahraman et al., 2019), and mannitol, also (Ejaz et al., 2020), discovered to have the capacity to scavenge ROS To reduce the harm caused by salinity; plants scavenge ROS using both an enzymatic and non-enzymatic system.

8. Additional Physiological Control in Salt Stress

Ion compartmentation appears to be a universal ability of plants because it occurs in both halophytes and non-halophytes. Halophytes have a different ion separation function than non-halophytes. In most cases, halophytes divide the cytoplasm from the assimilated saline ions in vacuoles, protecting the cytoplasm from the saline ions' toxicity. On the other hand, non-halophytes typically reduce the absorption of potentially damaging saline ions while also transporting The consumed saline ions are transferred to older tissues for capacity, defending young tissues at the expense of older tissues. The transmembrane proteins H+-ATPase, PPase, Ca2+-ATPase, secondary transporters, and numerous ion channel proteins are required for saline ion compartmentation. To pump H+ out of the cell and create a transmembrane potential gradient, H+-ATPase and H+-PPase on the membrane or vacuolar membrane hydrolyze ATP or pyrophosphate (PPI) to produce energy. The proton pumps-related Na+/H+ antiporters and vacuolar membrane then transfer Na+ into the cell the potential gradient. To lower the osmotic possibilities in the cytoplasm, ions can also be transferred into the vacuole. Na+/H+ function is decreased by salt stress, but this effect enhances the amount of unsaturated fatty acids in the cell membrane. The Na+/H+ antiporters that ensure Na+ in the vacuole significantly lower the vacuole's osmotic possibilities and lessen the harm that Na+ causes to the cytoplasm.
One of the traits of greater plants' salt tolerance is the presence of salt vesicles. It is a type of trichome-derived big, extremely vacuolized cell. Different species have different salt vesicles with different shapes and sizes. Water storage to secure plants from short-term osmotic stress, retrieval of organic osmolytes, collection of ROS scavengers, storage of photosynthetic protein, and so on. A significant portion of the salt vesicle's mechanism has been elucidated, although only at the cellular level. The entire salt secretion process's mechanism must be augmented because it is incomplete. There isn't much research on the formation of salt vesicles at the molecular level. Identifying the essential genes that govern salt vesicle generation has not been possible. It is also important to consider whether the genes connected to those governing salt vesicle development are linked to salt tolerance.

8.1. Ca2+-Dependent Signalling Pathways Pathway

Plant cells experience a rapid rise in Ca2+ content while under salt stress. Ca2+ has been recognized as a signal molecule associated with the transduction of salt stress signals being an osmotic regulator. Changes primarily influence ca+2 involvement as a second messenger in cell Ca2+ concentration. Ca2+ input and outflow in plant cells are regulated by various Ca2+ channels, Ca2+ pumps, and H+/Ca2+ antiporters found on cell and organelle membranes. As a result, cells have the flexibility to control adjustments in Ca2+ intensity and concentration and initiate various signalling pathways.

8.2. SOS Pathway

The SOS3 gene produced a protein that binds Ca2+. To stimulate the action of SOS2 kinase, the SOS3 protein can bind to the SOS2 protein's regulatory area at the C-terminus. SOS1 is phosphorylated by the activated SOS2 protein, which consequently increases SOS1 transporter activity.
As a result, a high external Na+ increases internal Ca2+, which is the SOS signal transduction pathway. Before binding to SOS2, SOS3 initially binds to Ca2+. SOS3 disables SOS2 self-inhibition to activate SOS2 kinase activity. The SOS1 transporter on the plasma membrane is subsequently phosphorylated by the complicated SOS3 and SOS2, improving its capacity to transport Na+ out of the cell. Another Na+/H+ antiporter found on the vacuole membrane is NHX1. It can transport K+ and Na+ into the vacuole using a proton gradient. The protein NHX1 controls protein localization, vacuolar pH, K+ concentration, vesicle trafficking, and the salt response mechanism. SOS kinases control NHX1, which is also controlled by ABA, to control the entry of Na+ into vacuoles. NHX requires a proton gradient to generate energy during Na+ transport, the vacuole membrane contains several H+-ATPase and H+-PPases to promote Na+ compartmentation. (Qi and Spalding, 2004) the research found that the SOS3 is unnecessary for the procedure that moves Na+ into vacuoles. The SOS3 homologous protein ScaBPs can also fulfil the same function.

8.3. CDPK

Cotton plants have several Ca2+-dependent protein kinases (CDPKs). Both membrane-bound and membrane-free cells have CDPKs, depending on their subcellular location. They are engaged in both the process of resistance to abiotic stress and control throughout the entire plant cycle of growth and development. In general, accepted that CDPKs primarily control the stomatal and the expression of genes associated with ion channels to control plant resistance to salt stress. The protein kinase and other downstream signalling components are activated by the signal transduction transcription factors that the CDPK genes decode in various plants. CDPKs can control the equilibrium of ROS by controlling the expression of NADPH oxidase and encouraging the synthesis of antioxidant genes, contributing to the responding to oxidative stress and enhancing plants' tolerance to salt.

8.4. Ca2+- (MAPKs Cascade)

Protein phosphorylation mediated by protein kinases, MAPK cascades can gradually increase plants' environmental signal transduction pathway, which is one of the most significant. It can transmit a signal to the cell's DNA in the nucleus from a receptor on its surface. To control plants' physiological responses to salt stress, there is an intricate network made up of multiple stress-inducible signal pathways. The only four pathways that we are more familiar with and have more significant roles are the SOS, ABA, CDPK, and MAPK pathways. In recent years, more elements of salt stress signal transduction have been found and characterized, several questions still need to answer. There is still much to learn about the selectivity and interactions of salt-stressed plants' many signal transduction pathways. Future research should also focus on how different abiotic pressures alter signal transduction networks, and how different signals interact with one another when several abiotic stresses are present.

9. Genes Involved in Salt Tolerance

A molecular process supports plants' physiological ability to tolerate salt. In higher plants, genes involved in osmolytes, autoxidation, ion transport, signal transmission, and regulatory functions are the main subjects of salt-responsive gene experiments, according to Figure 5.
Figure 4. The MAPK cascade's mode.
Figure 4. The MAPK cascade's mode.
Preprints 79971 g004

9.1. Genes Involved in Osmotic Adjustment

Plants that are stressed by the salt can collect molecule solutes or osmotic agents. Osmotic regulators, including proline, betaine, treehouse, mannitol, fructan, and glycogen, can be produced by plants when they are under water stress since osmotic regulator synthase genes have been inserted into their genomes. Plants currently have defined metabolic routes for proline. The Glu and Orn pathways are the two processes plants use to synthesize proline. Both Glu and Orn are the first substrates. Some essential rate-limiting enzymes control each process. The Glu pathway's key enzymes are 1 pyrroline-5-carboxylic acid synthetase (P5CS) and glutamate dehydrogenase (GDH); the Orn pathway's key Ornithine aminotransferase (OAT) and arginase are examples of enzymes. The main enzyme involved in the breakdown of proline is proline dehydrogenase (ProDH). For example, Chen et al. proposed that proline was produced excessively and salt tolerance was increased in transgenic Arabidopsis thaliana when P5CS1 and P5CS2 cDNA were introduced. In plants, two stages of enzyme processes are primarily responsible for synthesising glycine betaine. Choline is initially converted into betaine aldehyde by an oxidation reaction that CMO catalyzes. Wang et al. observed that the Suaeda liaodonggensis was used to clone the BADH gene, which was then introduced into tomato (Solanum Lycopersicum). Additionally, the soybean was revealed to contain the unique tonoplast intrinsic protein GmTIP2, which may be crucial for osmotic control.
The biosynthesis of sucrose is a challenging process. 1, 6-diphosphate fructose (FBP) and sucrose phosphate synthase are the key rate-limiting enzymes in the synthesis of sucrose (SPS) (Li et al., 2012). Under salt stress, three differentially FBP proteins were discovered using two-dimensional electrophoresis and internal peptide sequence analysis. According to the findings of this study mechanism through which mangrove plants withstand salt involves FBP.
Plant salt tolerance is closely linked to HKT, a Na+ or Na+-K+ transporter. It can reduce the amount of Na+ in the shoot, discharge extra Na+ into the surrounding parenchyma cells, and maintain K+ homeostasis in vivo. HKT proteins can be split into two categories based on the distinct ways that Na+ and K+ are transported in the heterologous expression system: When the external K+ level is low, HKT1 primarily affects the transport of Na+ and mediates Na+ absorption; HKT2 serves as a K+-Na+ cotransporter (Noble and Rogers, 1992). Han et al.'s study on wild barley in Tibet revealed that HvHKT1; 1 knockdown caused Na+ accumulation in roots and leaves. When HvHKT1; 1 was overexpressed in salt-sensitive Arabidopsis hkt1-4 and sos1-12 mutant strains, Na+ content in root and shoot was significantly reduced, demonstrating that HVHKT1; 1 is important for Na+ transfer in roots. However, recent research found that HvHKT1; 5 knockdown barley lines had much higher K+/Na+ ratios than the wild type and significantly less Na+ translocation from root to shoot. In contrast to other HKT1; 5 members, HvHKT1; 5 exhibits salt tolerance is negatively regulated, this clearly shows that barley has a different Na+ transport mechanism. Zhang, Y.; et al. Yao, X.; et al., and Kader, M.A.; et al., found that under salt stress, Rice transporter genes OsHKT1, OsHKT2, and OsVHA were engineered to express and were found to be capable of lowering Na+ content by regulating the Na+/K+ ratio.

9.2. Genes Involved in Antioxidants

A significant amount of ROS is produced because of salt stress. ROS can be rapidly eliminated by enhanced expression of antioxidant enzyme genes like SOD, CAT, POD, and APX. The variable expression of numerous plant antioxidant enzyme gene families under abiotic stress in the recent period, there has been much research. Ma, Li, et al., carried out the Brassica juncea and Brassica oleracea genome-wide classification and identification of the abiotic-stress-responsive SOD gene family in Rapa. Wu et al. observed through classifying and evaluating Cassava's class III POD gene family's expression. Wang et al., found in Cotton's CAT gene family: genome-wide classification and bioinformatics analysis. Tao et al. implemented the Gossypium hirsutum APX gene family expression profiling and genome-wide analysis.
Other genes are associated with antioxidants in introducing the families of genes encoding antioxidant enzymes. Additionally, significant antioxidants in plants are AsA and GSH. Through the AsA-GSH cycle, they are capable of synergism and renewal. We can increase the production of AsA and GSH by overexpressing crucial enzymes in this cycle, which would increase plants' resistance to oxidative stress. Glutamate cysteine ligase (GCL) first catalyzes the formation of γ-glutamylcysteine from Glu and Cys, and then glutathione synthase (GS) catalyzes the structure of GSH from Glycine and γ-glutamylcysteine [164]. APX helps to detoxify H2O2 and oxidizes AsA to deliver MDHA. MDHA can also spontaneously transform into DHA or AsA with the support of MDHAR. DHAR further converts DHA to AsA. Other significant enzymes comprise GR, GSTs, etc. For example, Horváth et al. found that AtGSTF8 and AtGSTU19 play a role in fine-tuning redox homeostasis, contributing to Arabidopsis salt stress resistance. CAR's primary function is to eliminate ROS generated during photosynthesis. CARs come in a variety of forms. As a result, each step of their varied synthesis pathways uses a different enzyme. Sun et al., experiments in transgenic Arabidopsis, drought and salt stress tolerance are improved by overexpression of the Cerasus humilis ChVDE gene, which encodes a violaxanthin de-epoxidase.

9.3. Plant Reaction to Salinity Stress in Genes Upregulating and Downregulating

Plants use various techniques to either upregulate or downregulate the transcription of genes in reaction to salinity (mRNA) (Table 1). A new strategy in salt tolerance involves several regulators in genes through core dogma, transcription preparation to the participation of TF engaged in the creation of mRNA and protein produced. Salinity alone generates two types of stress in plants: first, it lowers the soil's water potential, which lowers water intake, and increases osmotic and drought stress. Second, it results in an ion buildup that results in ionic stress Nongpiur et al. In responding to salt, several genes in plants are stimulated to produce proteins, amino acids, carbohydrates (CCOMT, SAM, GAPDH, AND LAX AP2/ERF), sugars (SWEET15), and proteins (catalase, CAT; glutaredoxin, GRXC1; glutathione-S-transferase, GST), Zhao, W.; et al., a significant number of transcription factor genes that is either up or downregulated are reliable with the development of a sophisticated signal transduction system that underlies the reaction to salt stress (Wang et al., 2019).Different plants identified with salt stress have distinct up and downregulated genes (Table 2).

9.4. Genes Related to Signal Transduction

Previous research has related six genes in the Arabidopsis SOS gene family to salt tolerance, either directly or indirectly. Controlling SOS1 on the plasma membrane, SOS3 and SOS2 found in the cytoplasm can control K+ and Na+ ion homeostasis, improving salt tolerance. SOS1 is a SOS family gene directly linked to plant salt tolerance. Under salt stress, it regulates the activity of SOS1 and ion transporters, as well as the growth of root hair, SOS4 improves salt tolerance. The SOS5 protein, which is situated back side of the plasma membrane, contributes significantly by fostering the formation of cell junctions and cell walls. SOS6 is crucial for controlling oxidative and osmotic stress in cells. Simultaneously, significant advancement has been made in locating SOS gene families in different plant species. Dong, X. and his collaborators found the identification of SOS family genes across the genome and study of their gene expression in tuber mustard (Brassica juncea var. tumida). Zhao et al. observed that SOS family genes from spinach had been identified and characterized (Spinacia oleracea).
Protein kinases, particularly CDPKs and MAPKs, are crucial signal transducers that regulate how plants react to abiotic stress. Many plant species have had their CDPK gene families found and investigated (Li et al., 2022b). Interacting with CAT3 and phosphorylating its 261st serine residue, At CPK8 modulates the activity of CAT3 and, in turn, the capacity of plants to scavenge ROS. MAPK members involved in response to salt stress have been identified in some plants over the last few decades (Sadau et al., 2022). As a result of severe salt stress, AtMEKK1 in Arabidopsis phosphorylates AtMKK2 and AtMEK1 to activate downstream kinases AtMPK4 and AtMPK6. Sun, H.; et al. claimed that Populus trichocarpa PtMAPKK4 overexpression improves tobacco's ability to salt. Based on research conducted by Wang et al., when the grape VvMKK2 and VvMKK4 genes were overexpressed in transgenic Arabidopsis thaliana, seedling development under stress conditions outperformed the wild type. Although overexpression of VvMKK4 only increased tolerance to salt stress, overexpression of VvMKK2 raised response to both drought and salt stress in Arabidopsis. Overexpression of genes such as maize ZmMKK4, Khan, M.N.; et al. Chen, Y.; et al., and Talaat, N.B.; and others can improve transgenic plants' ability to withstand high salt and alkali concentrations.

9.5. Regulatory of Several Genes

Salt tolerance is a quantitative genetic trait that involves multiple genes. In reaction to salt stress, transcriptional factors change gene expression to control gene levels. Core transcription factor families, including bZIP, WRKY, AP2/ERF, MYB, DREB, bHLH, CBF, and NAC, have been identified. In hairy soybean roots, GmbZIP2 may be responsible for the overexpression of stress-responsive genes such as GmMYB48, GmWD40, GmDHN15, GmGST1, and GmLEA. Transcription factors bHLH and WRKY contribute to Arabidopsis thaliana's enhanced salt tolerance by managing the expression of the salt-responsive gene. According to Nguyen, N.H.; et al. (2019), controlling ABA signalling pathways, AtMYB20 in Arabidopsis thaliana influences plants resistant to salt stress for sale. ABI1, ABI2, and AtPP2CA gene expression levels decrease under salt stress when the AtMYB20 gene is overexpressed, which enhances transgenic Arabidopsis' resistance to salt stress. ThDREB transcription factor can increase Tamarix hispid's ability to tolerate salt by continuing to increase the antioxidant enzyme's activity. According to research by Chen, J. et al., In Arabidopsis thaliana, overexpression of VvNAC17 increases plant tolerance to salinity and drought through enhancing the expression of ABA and stress-related genes such as ABI5, AREB1, COR15A, COR47, and P5CS.
In addition to transcription factors, some regulatory genes are linked to plant salt tolerance. Numerous short non-coding RNAs, such as microRNAs (miRNAs), have emerged in recent years as crucial regulators of chromatin remodelling, translation inhibition, and mRNA degradation MiRNA targets include the metabolic pathway, the ubiquitination pathway, transcription factors, signal transduction elements, and genes concerned in RNA manufacturing and protein synthesis. Transgenic Osa-miR393a increased K+ uptake and salt stress tolerance compared to wild-type controls.
Although recent advances in research on the plant genes attributed to salt tolerance, it is still primarily concentrated on locating and modifying the homologous genes responsible for salt tolerance in Arabidopsis. Finding new salt-responsive genes can be done with an appealing approach. Comparative genomics or transcriptomic analysis can be used to investigate halophytes and find new salt-responsive genes. For instance, using transcriptome analysis, discovered a few more salt-tolerance genes in the halophyte A triplex canescens; although a complete genome is required for a comparative genomic and transcriptomic study, it is expected that as sequencing technologies advance, more and more genome sequences will become available, This will be useful for molecular breeding in the coming years.Different salt stress genes' roles in different plants (Table 2) .

10. Techniques for Non-Genetic Improvement

Traditional breeding is a successful strategy for increasing a plant's tolerance to salt. Recurrent crossbreeding, selective breeding through artificial means, cutting, grafting, tissue culture, and other forms of asexual proliferation are among the techniques used (Suman, 2017). New salt-tolerant plant genotypes can be created. Applying exogenous chemicals can also help plants be better able to withstand salt stress ; exogenous substances are classified as follows: Those that act as osmotic regulators, like betaine, sugar, organic acids, and gibberellin, and those that do not; those that promote plant growth; and those that lessen cell membrane permeability, like salicylic acid, humic acid, and Ca2+. Other examples include those that stimulate the interaction of antioxidant enzymes in plants, like NO, silicon, selenium, and others; those that can act as osmotic regulators. For instance, under conditions of salt stress, plants treated with exogenous melatonin showed higher antioxidant enzyme activity, higher rates of photosynthetic activity, lower rates of electrolyte leakage, and lower MDA contents than the control group; Melatonin suggests improved plant salt resistance by improving photosynthetic capacity, oxidation resistance, and ion homeostasis (Rajput et al.). Another recent approach has been to increase plants' tolerance to salt through the relationship of fungi and roots to create symbionts with specific cells and characteristics. In their study, (Azad and Kaminskyj, 2016;Hao et al., 2021), the influence of systemic fungal endophytes isolated from naturally developing plants in determine soil on plant salt tolerance. Compared to non-colonized plants, endophyte-colonized plants had higher root and shoot biomass, better water usage efficiency, higher photosynthetic efficiency, and lower ROS levels (Dastogeer and Wylie, 2017). The harm caused to plants by salt stress can also be effectively reduced through rational fertilization. For instance, phosphorus fertilizer increased Phaseolus Vulgari's ability to tolerate salt.

Genetic Improvement Techniques

Genetic engineering is the most straightforward and efficient way to increase the tolerance of plants to salt, which introduces genes relevant to salt stress into the plant (Shahbaz and Ashraf, 2013). In addition to the genes mentioned earlier, the tobacco osmotin gene has recently received much attention for its potential to help plants under salt stress. Plants react to biotic and abiotic challenges by applying stress-responsive protein osmotic widely. It is a member of the PR-5 pathogenesis-related protein family. It protects plants by preventing lipid peroxidation, initiating programmed cell death (PCD), rising proline content, and enhancing antioxidant enzyme activity. as stated by the researcher, overexpressing the tobacco osmotin protein in transgenic plants protects from a variety of stresses, fungal infections, and salt stress (Das et al., 2011). Xu, Fang, et al. found that the putative transgenic chili pepper plants' morphology resembled wild-type plants but were more salt tolerant (Xu et al., 2012). They overexpressed the osmotic gene. Soybeans exhibited the same results in plants with overexpressed tobacco osmotin gene. (Bashir et al., 2021) have researched whether increasing the tobacco osmotin protein in Olives to increase salt tolerance has anything to do with the transgenic plants' sulfur metabolism. Although there is some evidence that the tobacco osmotin protein contributes to increased plant stress resistance, the full mechanism underlying osmotin activity is still unknown and needs to be clarified.
There is a complicated network of levels and plants have genes that help them tolerate salt. Using a single transgenic technique, like upping a specific osmolyte's concentration or a particular oxidase's action, plants' ability to tolerate salt can be increased. The degree of progress is constrained, though. As a result, we must determine how to prioritize each tactic. When significantly raising a plant's resistance to salt, it's best to prioritize our strategies to obtain better results.

11. Conclusions Future Research Perspectives

In our review, we discussed what occurs to cotton plants under salt stress and the physiological reaction that emerges, which involves four primary signalling pathways, the related genes that respond to salt stress, and a few methods for boosting plants' resistance to salt. The main components of the physiological response are osmotic adjustment and ROS scavenging.
Because phenotyping is difficult because of the complexity of stress conditions, which are impacted by genes with small and variable impacts and environmental elements, it is unclear what genetic factors contribute to these tolerances. But salt stresses harm physiological, biochemical, and molecular processes, ultimately preventing cotton from growing and developing normally. Some examples include reduced photosynthetic rate, plant height, leaf and root size, biomass, economic yield, yield components, and inferior fiber quality. Because of this, regardless of whether cotton is grown in a greenhouse or a field (based on yield, components yield, and fiber quality), it is possible to screen it for salt tolerance (according to plant height and biomass).
Many researchers have conducted experiments over the last few decades to understand better the mechanism that regulates plant salt tolerance. Nevertheless, future research and discoveries still have a lot of room. It is not yet clear how salt stress affects plant photosynthesis. The development of salt-tolerant cotton cultivars offers a practical management approach to cotton under salt stress. When selecting cotton genotypes that are salt tolerant, increase Ca, Na, and K/Na ratios utilized with classification factors. The existing genetic diversity could be effectively used to create salt tolerant genotypes. Because QTLs control salt tolerance genetically, marker-assisted classification is an exciting method for indirectly selecting salt-tolerant genotypes. So far, no mutual understanding has been reached. Second, further research is necessary to integrate to identify plant stress, and morphological, physiological, and molecular methods are used. Thirdly, the specific salt-tolerance mechanisms used by halophytes, like salt vesicles, are still largely unknown. Fourth, there is still much to learn about how plants react to salt stress, and it is not yet clear how various signal transduction pathways interact with one another. Fifthly, In the future, comparative genomic and transcriptomic approaches may be more important in identifying genes that respond to salt stress. Finally, we must determine how to prioritize each improvement strategy to achieve better improvements. Overall, Understanding the mechanism underlying plant salt tolerance will take a lot of additional work. Future research should concentrate on the molecular interactions between and within cells that contribute to the response to salinity stress.

Author Contributions

All of the authors of this review contributed to its design and writing, and they all approved of it for publishing after reading it.

Funding

The Authors Gratefully Acknowledge Agricultural Science and Technology Innovation Program of the Chinese Academy of Agricultural Sciences and This Research for Financial Support by the Natural Science Foundation of Xinjiang Uygur Autonomous Region (Grant number 2020D01B61).

Acknowledgments

This work was funded by the Agricultural Science and Technology Innovation Program of the Chinese Academy of Agricultural Sciences and Natural Science Foundation of Xinjiang Uygur Autonomous Region.

References

  1. Abbas, W.; Ashraf, M.; Akram, N.A. Alleviation of salt-induced adverse effects in eggplant (Solanum melongena L.) by glycinebetaine and sugarbeet extracts. Scientia horticulturae 2010, 125, 188–195. [Google Scholar] [CrossRef]
  2. Abobatta, W.F. Plant responses and tolerance to extreme salinity. In Salt and Drought Stress Tolerance in Plants; Springer: Berlin/Heidelberg, Germany, 2020; pp. 177–210. [Google Scholar]
  3. Adhikari, S.; Anuragi, H.; Chandra, K.; Tarte, S.H.; Dhaka, S.R.; Jatav, H.S.; Hingonia, K. Molecular basis of plant nutrient use efficiency-concepts and challenges for its improvement. In Sustainable Plant Nutrition; Elsevier: Amsterdam, The Netherlands, 2023; pp. 107–151. [Google Scholar]
  4. Ahmad, I.; Zhu, G.; Zhou, G.; Song, X.; Hussein Ibrahim, M.E.; Ibrahim Salih, E.G.; Hussain, S.; Younas, M.U. Pivotal role of phytohormones and their responsive genes in plant growth and their signaling and transduction pathway under salt stress in cotton. International Journal of Molecular Sciences 2022, 23, 7339. [Google Scholar] [CrossRef] [PubMed]
  5. Ahmad, P.; John, R.; Sarwat, M.; Umar, S. Responses of proline, lipid peroxidation and antioxidative enzymes in two varieties of Pisum sativum L. under salt stress. International Journal of Plant Production 2012, 2, 353–366. [Google Scholar]
  6. Ahmad, R.; Hussain, S.; Anjum, M.A.; Khalid, M.F.; Saqib, M.; Zakir, I.; Hassan, A.; Fahad, S.; Ahmad, S. Oxidative stress and antioxidant defense mechanisms in plants under salt stress. In Plant abiotic stress tolerance; Springer: Berlin/Heidelberg, Germany, 2019; pp. 191–205. [Google Scholar]
  7. Albaladejo, I.; Egea, I.; Morales, B.; Flores, F.B.; Capel, C.; Lozano, R.; Bolarin, M.C. Identification of key genes involved in the phenotypic alterations of res (restored cell structure by salinity) tomato mutant and its recovery induced by salt stress through transcriptomic analysis. BMC plant biology 2018, 18, 1–19. [Google Scholar] [CrossRef]
  8. Angon, P.B.; Tahjib-Ul-Arif, M.; Samin, S.I.; Habiba, U.; Hossain, M.A.; Brestic, M. How Do Plants Respond to Combined Drought and Salinity Stress?—A Systematic Review. Plants 2022, 11, 2884. [Google Scholar] [CrossRef]
  9. Ansari, H.H.; Siddiqui, A.; Wajid, D.; Tabassum, S.; Umar, M.; Siddiqui, Z.S. Profiling of energy compartmentalization in photosystem II (PSII), light harvesting complexes and specific energy fluxes of primed maize cultivar (P1429) under salt stress environment. Plant Physiology and Biochemistry 2022, 170, 296–306. [Google Scholar] [CrossRef]
  10. Arif, Y.; Singh, P.; Siddiqui, H.; Bajguz, A.; Hayat, S. Salinity induced physiological and biochemical changes in plants: An omic approach towards salt stress tolerance. Plant Physiology and Biochemistry 2020, 156, 64–77. [Google Scholar] [CrossRef]
  11. Ashraf, J.; Zuo, D.; Wang, Q.; Malik, W.; Zhang, Y.; Abid, M.A.; Cheng, H.; Yang, Q.; Song, G. Recent insights into cotton functional genomics: progress and future perspectives. Plant biotechnology journal 2018, 16, 699–713. [Google Scholar] [CrossRef]
  12. Ashraf, M.; Wu, L. Breeding for salinity tolerance in plants. Critical Reviews in Plant Sciences 1994, 13, 17–42. [Google Scholar] [CrossRef]
  13. Ashraf, M.A.; Akbar, A.; Askari, S.H.; Iqbal, M.; Rasheed, R.; Hussain, I. Recent advances in abiotic stress tolerance of plants through chemical priming: an overview. Advances in seed priming 2018, 51–79. [Google Scholar]
  14. Ashrafi, N.; Rezaei Nejad, A. Lisianthus response to salinity stress. Photosynthetica 2018, 56, 487–494. [Google Scholar] [CrossRef]
  15. Atzori, G.; Guidi Nissim, W.; Mancuso, S.; Palm, E. Intercropping Salt-Sensitive Lactuca sativa L. and Salt-Tolerant Salsola soda L. in a Saline Hydroponic Medium: An Agronomic and Physiological Assessment. Plants 2022, 11, 2924. [Google Scholar] [CrossRef]
  16. Azad, K.; Kaminskyj, S. A fungal endophyte strategy for mitigating the effect of salt and drought stress on plant growth. Symbiosis 2016, 68, 73–78. [Google Scholar] [CrossRef]
  17. Azevedo Neto, A.D.D.; Prisco, J.T.; Enéas-Filho, J.; Lacerda, C.F.D.; Silva, J.V.; Costa, P.H.a.D. , and Gomes-Filho, E. Effects of salt stress on plant growth, stomatal response and solute accumulation of different maize genotypes. Brazilian Journal of Plant Physiology 2004, 16, 31–38. [Google Scholar] [CrossRef]
  18. Aziz, A.; Ashraf, M.; Sikandar, S.; Asif, M.; Akhtar, N.; Shahzad, S.M.; Wasaya, A.; Raza, A.; Babar, B.H. Optimizing sulfur for improving salt tolerance of sunflower (Helianthus annuus L.). Soil Environ 2019, 38, 222–233. [Google Scholar] [CrossRef]
  19. Azooz, M.; Ismail, A.; Elhamd, M.A. Growth, lipid peroxidation and antioxidant enzyme activities as a selection criterion for the salt tolerance of maize cultivars grown under salinity stress. Int. J. Agric. Biol 2009, 11, 21–26. [Google Scholar]
  20. Babayev, D.; Rahimova, N.; Arzyamova, O.; Yollibayev, A. EFFECT OF FERTILIZERS INTO PHYSIOLOGICAL AND BIOCHEMICAL INDICATORS OF COTTON IN SALINE SOILS.
  21. Baghour, M.; Gálvez, F.J.; Sánchez, M.E.; Aranda, M.N.; Venema, K.; Rodríguez-Rosales, M.P. Overexpression of LeNHX2 and SlSOS2 increases salt tolerance and fruit production in double transgenic tomato plants. Plant Physiology and Biochemistry 2019, 135, 77–86. [Google Scholar] [CrossRef] [PubMed]
  22. Bai, H.; Chen, J.; Gao, T.; Tang, Z.; Li, H.; Gong, S.; Du, Y.; Yu, Y.; Wang, W. A Na+/H+ antiporter localized on the Golgi-to-vacuole transport system from Camellia sinensis, CsNHX6, plays a positive role in salt tolerance. Scientia Horticulturae 2023, 309, 111704. [Google Scholar] [CrossRef]
  23. Bai, Y.; Xiao, S.; Zhang, Z.; Zhang, Y.; Sun, H.; Zhang, K.; Wang, X.; Bai, Z.; Li, C.; Liu, L. Melatonin improves the germination rate of cotton seeds under drought stress by opening pores in the seed coat. PeerJ 2020, 8, e9450. [Google Scholar] [CrossRef] [PubMed]
  24. Bakala, H.S.; Devi, J.; Sarao, L.K.; Kaur, S. Utilization of Wheat and Maize Waste as Biofuel Source. In Agroindustrial Waste for Green Fuel Application; Springer: Berlin/Heidelberg, Germany, 2023; pp. 27–66. [Google Scholar]
  25. Banu, M.N.A.; Hoque, M.A.; Watanabe-Sugimoto, M.; Matsuoka, K.; Nakamura, Y.; Shimoishi, Y.; Murata, Y. Proline and glycinebetaine induce antioxidant defense gene expression and suppress cell death in cultured tobacco cells under salt stress. Journal of Plant Physiology 2009, 166, 146–156. [Google Scholar] [CrossRef]
  26. Bao, Y.; Zhao, R.; Li, F.; Tang, W.; Han, L. Simultaneous expression of Spinacia oleracea chloroplast choline monooxygenase (CMO) and betaine aldehyde dehydrogenase (BADH) genes contribute to dwarfism in transgenic Lolium perenne. Plant Molecular Biology Reporter 2011, 29, 379–388. [Google Scholar] [CrossRef]
  27. Bartels, D.; Sunkar, R. Drought and salt tolerance in plants. Critical reviews in plant sciences 2005, 24, 23–58. [Google Scholar] [CrossRef]
  28. Bashir, M.A.; Silvestri, C.; Coppa, E.; Brunori, E.; Cristofori, V.; Rugini, E.; Ahmad, T.; Hafiz, I.A.; Abbasi, N.A.; Nawaz Shah, M.K. Response of olive shoots to salinity stress suggests the involvement of sulfur metabolism. Plants 2021, 10, 350. [Google Scholar] [CrossRef] [PubMed]
  29. Begum, N.; Hasanuzzaman, M.; Li, Y.; Akhtar, K.; Zhang, C.; Zhao, T. Seed germination behavior, growth, physiology and antioxidant metabolism of four contrasting cultivars under combined drought and salinity in soybean. Antioxidants 2022, 11, 498. [Google Scholar] [CrossRef]
  30. Bernstein, L.; Hayward, H. Physiology of salt tolerance. Annual review of plant physiology 1958, 9, 25–46. [Google Scholar] [CrossRef]
  31. Bhattacharya, A. Effect of Low Temperature Stress on Photosynthesis and Allied Traits: A Review. Physiological Processes in Plants Under Low Temperature Stress 2022, 199–297. [Google Scholar]
  32. Bibi, N.; Jan, G.; Jan, F.G.; Hamayun, M.; Iqbal, A.; Hussain, A.; Rehman, H.; Tawab, A.; Khushdil, F. Cochliobolus sp. acts as a biochemical modulator to alleviate salinity stress in okra plants. Plant physiology and biochemistry 2019, 139, 459–469. [Google Scholar] [CrossRef]
  33. Billah, M.; Aktar, S.; Brestic, M.; Zivcak, M.; Khaldun, A.B.M.; Uddin, M.S.; Bagum, S.A.; Yang, X.; Skalicky, M.; Mehari, T.G. Progressive genomic approaches to explore drought-and salt-induced oxidative stress responses in plants under changing climate. Plants 2021, 10, 1910. [Google Scholar] [CrossRef]
  34. Bor, M.; Özdemir, F. Manipulating metabolic pathways for development of salt-tolerant crops. In Salinity Responses and Tolerance in Plants, Volume 1; Springer: Berlin/Heidelberg, Germany, 2018; pp. 235–256. [Google Scholar]
  35. Brenya, E. (2020). Elucidating thigmomorphogenesis: An epigenetic phenomenon of mechanical stress acclimation in plants. University of Western Sydney.
  36. Brugnoli, E.; Björkman, O. Growth of cotton under continuous salinity stress: influence on allocation pattern, stomatal and non-stomatal components of photosynthesis and dissipation of excess light energy. Planta 1992, 187, 335–347. [Google Scholar] [CrossRef]
  37. Çakır, B.; Kılıçkaya, O. Mitogen-activated protein kinase cascades in Vitis vinifera. Frontiers in plant science 2015, 6, 556. [Google Scholar] [CrossRef]
  38. Chai, C.; Wang, Y.; Valliyodan, B.; Nguyen, H.T. Comprehensive analysis of the soybean (Glycine max) GmLAX auxin transporter gene family. Frontiers in plant science 2016, 7, 282. [Google Scholar] [CrossRef] [PubMed]
  39. Chakraborty, K.; Bhaduri, D.; Meena, H.N.; Kalariya, K. External potassium (K+) application improves salinity tolerance by promoting Na+-exclusion, K+-accumulation and osmotic adjustment in contrasting peanut cultivars. Plant Physiology and Biochemistry 2016, 103, 143–153. [Google Scholar] [CrossRef]
  40. Chauhan, A.; Abuamarah, B.A.; Kumar, A.; Verma, J.; Ghramh, H.A.; Khan, K.A.; Ansari, M.J. Influence of gibberellic acid and different salt concentrations on germination percentage and physiological parameters of oat cultivars. Saudi journal of biological sciences 2019, 26, 1298–1304. [Google Scholar] [CrossRef] [PubMed]
  41. Chen, J.; Zhang, X.; Jing, R.; Blair, M.W.; Mao, X.; Wang, S. Cloning and genetic diversity analysis of a new P5CS gene from common bean (Phaseolus vulgaris L.). Theoretical and applied genetics 2010, 120, 1393–1404. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, Y.; Wang, J.; Yao, L.; Li, B.; Ma, X.; Si, E.; Yang, K.; Li, C.; Shang, X.; Meng, Y. Combined proteomic and metabolomic analysis of the molecular mechanism underlying the response to salt stress during seed germination in barley. International Journal of Molecular Sciences 2022, 23, 10515. [Google Scholar] [CrossRef] [PubMed]
  43. Çiçek, N.; Oukarroum, A.; Strasser, R.J.; Schansker, G. Salt stress effects on the photosynthetic electron transport chain in two chickpea lines differing in their salt stress tolerance. Photosynthesis research 2018, 136, 291–301. [Google Scholar] [CrossRef]
  44. Corpas, F.J.; Fernandez-Ocana, A.; Carreras, A.; Valderrama, R.; Luque, F.; Esteban, F.J.; Rodríguez-Serrano, M.; Chaki, M.; Pedrajas, J.R.; Sandalio, L.M. The expression of different superoxide dismutase forms is cell-type dependent in olive (Olea europaea L.) leaves. Plant and cell physiology 2006, 47, 984–994. [Google Scholar] [CrossRef]
  45. Dai, J.; Duan, L.; Dong, H. Comparative effect of nitrogen forms on nitrogen uptake and cotton growth under salinity stress. Journal of Plant Nutrition 2015, 38, 1530–1543. [Google Scholar] [CrossRef]
  46. Das, M.; Chauhan, H.; Chhibbar, A.; Haq, R.; Mohd, Q.; Khurana, P. High-efficiency transformation and selective tolerance against biotic and abiotic stress in mulberry, Morus indica cv. K2, by constitutive and inducible expression of tobacco osmotin. Transgenic research 2011, 20, 231–246. [Google Scholar] [CrossRef]
  47. Dastogeer, K.M.; Wylie, S.J. Plant–fungi association: role of fungal endophytes in improving plant tolerance to water stress. In Plant-microbe interactions in agro-ecological perspectives; Springer: Berlin/Heidelberg, Germany, 2017; pp. 143–159. [Google Scholar]
  48. Del Amor, F.M.; Cuadra-Crespo, P. Alleviation of salinity stress in broccoli using foliar urea or methyl-jasmonate: analysis of growth, gas exchange, and isotope composition. Plant Growth Regulation 2011, 63, 55–62. [Google Scholar] [CrossRef]
  49. Delgado, C.; Mora-Poblete, F.; Ahmar, S.; Chen, J.-T. , and Figueroa, C.R. Jasmonates and plant salt stress: Molecular players, physiological effects, and improving tolerance by using genome-associated tools. International Journal of Molecular Sciences 2021, 22, 3082. [Google Scholar] [CrossRef] [PubMed]
  50. Demidchik, V.; Cuin, T.A.; Svistunenko, D.; Smith, S.J.; Miller, A.J.; Shabala, S.; Sokolik, A.; Yurin, V. Arabidopsis root K+-efflux conductance activated by hydroxyl radicals: single-channel properties, genetic basis and involvement in stress-induced cell death. Journal of cell science 2010, 123, 1468–1479. [Google Scholar] [CrossRef] [PubMed]
  51. Deng, H.; Li, Q.; Cao, R.; Ren, Y.; Wang, G.; Guo, H.; Bu, S.; Liu, J.; Ma, P. Overexpression of SmMYC2 enhances salt resistance in Arabidopsis thaliana and Salvia miltiorrhiza hairy roots. Journal of Plant Physiology 2023, 280, 153862. [Google Scholar] [CrossRef] [PubMed]
  52. Ding, X.; Yao, Q.; Zhu, D.; Dong, B. Hydrogen-rich water maintains the quality of Rosa roxburghii by regulating AsA biosynthesis and regeneration. Postharvest Biology and Technology 2023, 195, 112136. [Google Scholar] [CrossRef]
  53. Dong, X.; Wang, F. Edited by: Loredana F. Ciarmiello, University of Campania Luigi Vanvitelli, Italy Reviewed by. Salt Tolerance: Molecular and Physiological Mechanisms and Breeding Applications.
  54. Ejaz, S.; Fahad, S.; Anjum, M.A.; Nawaz, A.; Naz, S.; Hussain, S.; Ahmad, S. Role of osmolytes in the mechanisms of antioxidant defense of plants. In Sustainable Agriculture Reviews 39; Springer: Berlin/Heidelberg, Germany, 2020; pp. 95–117. [Google Scholar]
  55. El-Banna, M.; Abdelaal, K.A. Response of strawberry plants grown in the hydroponic system to pretreatment with H2O2 before exposure to salinity stress. Journal of Plant Production 2018, 9, 989–1001. [Google Scholar] [CrossRef]
  56. El-Katony, T.M.; El-Bastawisy, Z.M.; El-Ghareeb, S.S. Timing of salicylic acid application affects the response of maize (Zea mays L.) hybrids to salinity stress. Heliyon 2019, 5, e01547. [Google Scholar] [CrossRef]
  57. Elkeilsh, A.; Awad, Y.M.; Soliman, M.H.; Abu-Elsaoud, A.; Abdelhamid, M.T.; El-Metwally, I.M. Exogenous application of β-sitosterol mediated growth and yield improvement in water-stressed wheat (Triticum aestivum) involves up-regulated antioxidant system. Journal of plant research 2019, 132, 881–901. [Google Scholar] [CrossRef]
  58. Etesami, H.; Noori, F. Soil salinity as a challenge for sustainable agriculture and bacterial-mediated alleviation of salinity stress in crop plants. In Saline soil-based agriculture by halotolerant microorganisms; Springer: Berlin/Heidelberg, Germany, 2019; pp. 1–22. [Google Scholar]
  59. Finkina, E.; Melnikova, D.; Bogdanov, I. Lipid transfer proteins as components of the plant innate immune system: structure, functions, and applications. Acta Naturae (англoязычная версия) 2016, 8, 47–61. [Google Scholar] [CrossRef]
  60. Flowers, T.J.; Colmer, T.D. Salinity tolerance in halophytes. New Phytologist.
  61. Fróna, D.; Szenderák, J.; Harangi-Rákos, M. The challenge of feeding the world. Sustainability 2019, 11, 5816. [Google Scholar] [CrossRef]
  62. Gaafar, R.M.; Seyam, M.M. Ascorbate–glutathione cycle confers salt tolerance in Egyptian lentil cultivars. Physiology and Molecular Biology of Plants 2018, 24, 1083–1092. [Google Scholar] [CrossRef]
  63. Galić, V.; Mazur, M.; Šimić, D.; Zdunić, Z.; Franić, M. Plant biomass in salt-stressed young maize plants can be modelled with photosynthetic performance. Photosynthetica 2020, 58, 194–204. [Google Scholar] [CrossRef]
  64. Gao, H.; Yu, C.; Liu, R.; Li, X.; Huang, H.; Wang, X.; Zhang, C.; Jiang, N.; Li, X.; Cheng, S. The Glutathione S-Transferase PtGSTF1 Improves Biomass Production and Salt Tolerance through Regulating Xylem Cell Proliferation, Ion Homeostasis and Reactive Oxygen Species Scavenging in Poplar. International Journal of Molecular Sciences 2022, 23, 11288. [Google Scholar] [CrossRef] [PubMed]
  65. Gerona, M.E.B.; Deocampo, M.P.; Egdane, J.A.; Ismail, A.M.; Dionisio-Sese, M.L. Physiological responses of contrasting rice genotypes to salt stress at reproductive stage. Rice Science 2019, 26, 207–219. [Google Scholar] [CrossRef]
  66. Gill, S.S.; Tuteja, N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant physiology and biochemistry 2010, 48, 909–930. [Google Scholar] [CrossRef] [PubMed]
  67. Gong, B.; Wen, D.; Vandenlangenberg, K.; Wei, M.; Yang, F.; Shi, Q.; Wang, X. Comparative effects of NaCl and NaHCO3 stress on photosynthetic parameters, nutrient metabolism, and the antioxidant system in tomato leaves. Scientia Horticulturae 2013, 157, 1–12. [Google Scholar] [CrossRef]
  68. Goodman, A.; Ennos, A. A comparative study of the response of the roots and shoots of sunflower and maize to mechanical stimulation. Journal of Experimental Botany 1996, 47, 1499–1507. [Google Scholar] [CrossRef]
  69. Graus, D.; Li, K.; Rathje, J.M.; Ding, M.; Krischke, M.; Müller, M.J.; Cuin, T.A.; Al-Rasheid, K.A.; Scherzer, S.; Marten, I. Tobacco leaf tissue rapidly detoxifies direct salt loads without activation of calcium and SOS signaling. New Phytologist 2022. [Google Scholar] [CrossRef]
  70. Gross, E.; Sevier, C.S.; Heldman, N.; Vitu, E.; Bentzur, M.; Kaiser, C.A.; Thorpe, C.; Fass, D. Generating disulfides enzymatically: reaction products and electron acceptors of the endoplasmic reticulum thiol oxidase Ero1p. Proceedings of the National Academy of Sciences 2006, 103, 299–304. [Google Scholar] [CrossRef]
  71. Gupta, S.; Goyal, M.R.; Singh, A. Physiological and biochemical changes in plants under soil salinity stress: a review. Engineering practices for management of soil salinity 2018, 159–200. [Google Scholar]
  72. Hafez, E.M.; Gowayed, S.M.; Nehela, Y.; Sakran, R.M.; Rady, A.M.; Awadalla, A.; Omara, A.E.-D. , and Alowaiesh, B.F. Incorporated biochar-based soil amendment and exogenous glycine betaine foliar application ameliorate rice (Oryza sativa L.) tolerance and resilience to osmotic stress. Plants 2021, 10, 1930. [Google Scholar] [CrossRef]
  73. Hamza, O.M.; Al-Taey, D.K. A study on the effect of Glutamic acid and benzyl adenine application upon growth and yield parameters and active components of two broccoli hybrids. Int. J. Agricult. Stat. Sci 2020, 16, 1163–1167. [Google Scholar]
  74. Hao, S.; Wang, Y.; Yan, Y.; Liu, Y.; Wang, J.; Chen, S. A review on plant responses to salt stress and their mechanisms of salt resistance. Horticulturae 2021, 7, 132. [Google Scholar] [CrossRef]
  75. Haro, R.; Bañuelos, M.A.; Senn, M.E.; Barrero-Gil, J.; Rodríguez-Navarro, A. HKT1 mediates sodium uniport in roots. Pitfalls in the expression of HKT1 in yeast. Plant Physiology 2005, 139, 1495–1506. [Google Scholar] [CrossRef] [PubMed]
  76. Hasanuzzaman, M.; Nahar, K.; Fujita, M. Plant response to salt stress and role of exogenous protectants to mitigate salt-induced damages. In Ecophysiology and responses of plants under salt stress; Springer: Berlin/Heidelberg, Germany, 2013; pp. 25–87. [Google Scholar]
  77. Hashemi, S.E.; Madahhosseini, S.; Pirasteh-Anosheh, H.; Sedaghati, E.; Race, M. The Role of Nitrogen in Inducing Salt Stress Tolerance in Crocus sativus L.: Assessment Based on Plant Growth and Ions Distribution in Leaves. Sustainability 2023, 15, 567. [Google Scholar] [CrossRef]
  78. Hatam, Z.; Sabet, M.S.; Malakouti, M.J.; Mokhtassi-Bidgoli, A.; Homaee, M. Zinc and potassium fertilizer recommendation for cotton seedlings under salinity stress based on gas exchange and chlorophyll fluorescence responses. South African Journal of Botany 2020, 130, 155–164. [Google Scholar] [CrossRef]
  79. Hayat, S.; Hayat, Q.; Alyemeni, M.N.; Wani, A.S.; Pichtel, J.; Ahmad, A. Role of proline under changing environments: a review. Plant signaling & behavior 2012, 7, 1456–1466. [Google Scholar]
  80. Hirt, H.; Shinozaki, K. Plant responses to abiotic stress; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2003. [Google Scholar]
  81. Hniličková, H.; Hnilička, F.; Martinkova, J.; Kraus, K. Effects of salt stress on water status, photosynthesis and chlorophyll fluorescence of rocket. Plant, Soil and Environment 2017, 63, 362–367. [Google Scholar] [CrossRef]
  82. Horváth, E.; Bela, K.; Holinka, B.; Riyazuddin, R.; Gallé, Á.; Hajnal, Á.; Hurton, Á.; Fehér, A.; Csiszár, J. The Arabidopsis glutathione transferases, AtGSTF8 and AtGSTU19 are involved in the maintenance of root redox homeostasis affecting meristem size and salt stress sensitivity. Plant Science 2019, 283, 366–374. [Google Scholar] [CrossRef]
  83. Hosseinifard, M.; Stefaniak, S.; Ghorbani Javid, M.; Soltani, E.; Wojtyla, Ł.; Garnczarska, M. Contribution of Exogenous Proline to Abiotic Stresses Tolerance in Plants: A Review. International Journal of Molecular Sciences 2022, 23, 5186. [Google Scholar] [CrossRef]
  84. Hou, C.; Cheng, X.; Zhang, X.; Zhu, X.; Xu, J.; Luo, X.; Wu, D.; Liang, H. Effect of ferrous-activated calcium peroxide oxidation on forward osmosis treatment of algae-laden water: Membrane fouling mitigation and mechanism. Science of The Total Environment 2023, 858, 160100. [Google Scholar] [CrossRef]
  85. Hou, P.; Wang, F.; Luo, B.; Li, A.; Wang, C.; Shabala, L.; Ahmed, H.a.I. , Deng, S.; Zhang, H.; Song, P. Antioxidant enzymatic activity and osmotic adjustment as components of the drought tolerance mechanism in Carex duriuscula. Plants 2021, 10, 436. [Google Scholar] [CrossRef] [PubMed]
  86. Hrabak, E.M.; Dickmann, L.J.; Satterlee, J.S.; Sussman, M.R. Characterization of eight new members of the calmodulin-like domain protein kinase gene family from Arabidopsis thaliana. Plant molecular biology 1996, 31, 405–412. [Google Scholar] [CrossRef] [PubMed]
  87. Hussain, Q.; Asim, M.; Zhang, R.; Khan, R.; Farooq, S.; Wu, J. Transcription factors interact with ABA through gene expression and signaling pathways to mitigate drought and salinity stress. Biomolecules 2021, 11, 1159. [Google Scholar] [CrossRef] [PubMed]
  88. Ibrahim, E.A. Seed priming to alleviate salinity stress in germinating seeds. Journal of plant physiology 2016, 192, 38–46. [Google Scholar] [CrossRef]
  89. Ighodaro, O.; Akinloye, O. First line defence antioxidants-superoxide dismutase (SOD), catalase (CAT) and glutathione peroxidase (GPX): Their fundamental role in the entire antioxidant defence grid. Alexandria journal of medicine 2018, 54, 287–293. [Google Scholar] [CrossRef]
  90. Ilyas, M.; Ahad, A.; Batool, T.; Jabbar, W.; Ejaz, M.; Gul, A.; Ozturk, M. Salinity induced stress in wheat. Special Issue on “Cereals and Stress Tolerance” 2022, 126891, 67. [Google Scholar] [CrossRef]
  91. Iqbal, M.; Ashraf, M. Changes in growth, photosynthetic capacity and ionic relations in spring wheat (Triticum aestivum L.) due to pre-sowing seed treatment with polyamines. Plant Growth Regulation 2005, 46, 19–30. [Google Scholar] [CrossRef]
  92. Isayenkov, S.V.; Maathuis, F.J. Plant salinity stress: many unanswered questions remain. Frontiers in plant science 2019, 10, 80. [Google Scholar] [CrossRef]
  93. Jadidi, O.; Etminan, A.; Azizi-Nezhad, R.; Ebrahimi, A.; Pour-Aboughadareh, A. Physiological and molecular responses of barley genotypes to salinity stress. Genes 2022, 13, 2040. [Google Scholar] [CrossRef]
  94. Jia, C.; Zhao, S.; Bao, T.; Zhao, P.; Peng, K.; Guo, Q.; Gao, X.; Qin, J. Tomato BZR/BES transcription factor SlBZR1 positively regulates BR signaling and salt stress tolerance in tomato and Arabidopsis. Plant Science 2021, 302, 110719. [Google Scholar] [CrossRef]
  95. Jiang, D.; Chu, X.; Li, M.; Hou, J.; Tong, X.; Gao, Z.; Chen, G. Exogenous spermidine enhances salt-stressed rice photosynthetic performance by stabilizing structure and function of chloroplast and thylakoid membranes. Photosynthetica 2020, 58, 61–71. [Google Scholar] [CrossRef]
  96. Jiang, W.; Ye, Q.; Wu, Z.; Zhang, Q.; Wang, L.; Liu, J.; Hu, X.; Guo, D.; Wang, X.; Zhang, Z. Analysis of CAT Gene Family and Functional Identification of OsCAT3 in Rice. Genes 2023, 14, 138. [Google Scholar] [CrossRef] [PubMed]
  97. Johnson, S. (2016). The Mechanisms of drought stress tolerance in the crop Sorghum Bicolor. Durham University.
  98. Ju, F.; Pang, J.; Sun, L.; Gu, J.; Wang, Z.; Wu, X.; Ali, S.; Wang, Y.; Zhao, W.; Wang, S. Integrative transcriptomic, metabolomic and physiological analyses revealed the physiological and molecular mechanisms by which potassium regulates the salt tolerance of cotton (Gossypium hirsutum L.) roots. Industrial Crops and Products 2023, 193, 116177. [Google Scholar] [CrossRef]
  99. Ju, Y.-L. , Yue, X.-F., Min, Z.; Wang, X.-H., Fang, Y.-L., and Zhang, J.-X. VvNAC17, a novel stress-responsive grapevine (Vitis vinifera L.) NAC transcription factor, increases sensitivity to abscisic acid and enhances salinity, freezing, and drought tolerance in transgenic Arabidopsis. Plant Physiology and Biochemistry 2020, 146, 98–111. [Google Scholar] [PubMed]
  100. Kader, M.A.; Seidel, T.; Golldack, D.; Lindberg, S. Expressions of OsHKT1, OsHKT2, and OsVHA are differentially regulated under NaCl stress in salt-sensitive and salt-tolerant rice (Oryza sativa L.) cultivars. Journal of Experimental Botany 2006, 57, 4257–4268. [Google Scholar] [CrossRef] [PubMed]
  101. Kahraman, M.; Sevim, G.; Bor, M. The role of proline, glycinebetaine, and trehalose in stress-responsive gene expression. In osmoprotectant-mediated abiotic stress tolerance in plants; Springer: Berlin/Heidelberg, Germany, 2019; pp. 241–256. [Google Scholar]
  102. Kaushal, M. Insights into microbially induced salt tolerance and endurance mechanisms (STEM) in plants. Frontiers in Microbiology 2020, 11, 1518. [Google Scholar] [CrossRef] [PubMed]
  103. Kaya, C.; Sonmez, O.; Aydemir, S.; Ashraf, M.; Dikilitas, M. Exogenous application of mannitol and thiourea regulates plant growth and oxidative stress responses in salt-stressed maize (Zea mays L.). Journal of plant interactions 2013, 8, 234–241. [Google Scholar] [CrossRef]
  104. Kaya, C.; Tuna, A.L.; Ashraf, M.; Altunlu, H. Improved salt tolerance of melon (Cucumis melo L.) by the addition of proline and potassium nitrate. Environmental and Experimental Botany 2007, 60, 397–403. [Google Scholar] [CrossRef]
  105. Ke, D.; He, Y.; Fan, L.; Niu, R.; Cheng, L.; Wang, L.; Zhang, Z. The soybean TGA transcription factor GmTGA13 plays important roles in the response to salinity stress. Plant Biology 2022, 24, 313–322. [Google Scholar] [CrossRef]
  106. Khalvandi, M.; Amerian, M.; Pirdashti, H.; Keramati, S. Does co-inoculation of mycorrhiza and Piriformospora indica fungi enhance the efficiency of chlorophyll fluorescence and essential oil composition in peppermint under irrigation with saline water from the Caspian Sea? PloS one 2021, 16, e0254076. [Google Scholar] [CrossRef]
  107. Khan, I.; Awan, S.A.; Raza, M.A.; Rizwan, M.; Tariq, R.; Ali, S.; Huang, L. Silver nanoparticles improved the plant growth and reduced the sodium and chlorine accumulation in pearl millet: a life cycle study. Environmental Science and Pollution Research 2021, 28, 13712–13724. [Google Scholar] [CrossRef] [PubMed]
  108. Khan, M.M.; Al-Mas’oudi, R.S.; Al-Said, F.; Khan, I. Salinity effects on growth, electrolyte leakage, chlorophyll content and lipid peroxidation in cucumber (Cucumis sativus L.). In International Conference on Food and Agricultural Sciences Malaysia; IACSIT Press: Singapore, 2013; pp. 28–32. [Google Scholar]
  109. Khan, M.N.; Fu, C.; Li, J.; Tao, Y.; Li, Y.; Hu, J.; Chen, L.; Khan, Z.; Wu, H.; Li, Z. Seed nanopriming: How do nanomaterials improve seed tolerance to salinity and drought? Chemosphere 2022, 136911. [Google Scholar] [CrossRef] [PubMed]
  110. Khan, T.; Mazid, M.; Mohammad, F. A review of ascorbic acid potentialities against oxidative stress induced in plants. Journal of agrobiology 2011, 28, 97. [Google Scholar] [CrossRef]
  111. Kong, X.; Pan, J.; Zhang, M.; Xing, X.; Zhou, Y.; Liu, Y.; Li, D.; Li, D. ZmMKK4, a novel group C mitogen-activated protein kinase kinase in maize (Zea mays), confers salt and cold tolerance in transgenic Arabidopsis. Plant, cell & environment 2011, 34, 1291–1303. [Google Scholar]
  112. Krasensky, J.; Broyart, C.; Rabanal, F.A.; Jonak, C. The redox-sensitive chloroplast trehalose-6-phosphate phosphatase AtTPPD regulates salt stress tolerance. Antioxidants & redox signaling 2014, 21, 1289–1304. [Google Scholar]
  113. Kumar, S.; Singh, T.B.; Agnihotri, R.K.; Chaturvedi, P. Comparative effect of nacl and peg on physiological and biochemical attributes during vegetative stage of tomato. Res J Agric Sci 2021, 12, 955–961. [Google Scholar]
  114. Kumari, S.; Chhillar, H.; Chopra, P.; Khanna, R.R.; Khan, M.I.R. Potassium: A track to develop salinity tolerant plants. Plant Physiology and Biochemistry 2021, 167, 1011–1023. [Google Scholar] [CrossRef]
  115. Leigh, R.A.; Wyn Jones, R. A hypothesis relating critical potassium concentrations for growth to the distribution and functions of this ion in the plant cell. New Phytologist 1984, 97, 1–13. [Google Scholar] [CrossRef]
  116. Li, M.; Zhang, K.; Sun, Y.; Cui, H.; Cao, S.; Yan, L.; Xu, M. Growth, physiology, and transcriptional analysis of two contrasting Carex rigescens genotypes under salt stress reveals salt-tolerance mechanisms. Journal of plant physiology 2018, 229, 77–88. [Google Scholar] [CrossRef]
  117. Li, P.; Yang, H.; Liu, G.; Ma, W.; Li, C.; Huo, H.; He, J.; Liu, L. PpSARK regulates moss senescence and salt tolerance through ABA related pathway. International journal of molecular sciences 2018, 19, 2609. [Google Scholar] [CrossRef]
  118. Li, Q.-L. , Gao, X.-R., Yu, X.-H., Wang, X.-Z., and An, L.-J. Molecular cloning and characterization of betaine aldehyde dehydrogenase gene from Suaeda liaotungensis and its use in improved tolerance to salinity in transgenic tobacco. Biotechnology letters 2003, 25, 1431–1436. [Google Scholar] [CrossRef] [PubMed]
  119. Li, Q.; Liu, R.; Li, Z.; Fan, H.; Song, J. Positive effects of NaCl on the photoreaction and carbon assimilation efficiency in Suaeda salsa. Plant Physiology and Biochemistry 2022, 177, 32–37. [Google Scholar] [CrossRef] [PubMed]
  120. Li, Y.; Chen, Q.; Nan, H.; Li, X.; Lu, S.; Zhao, X.; Liu, B.; Guo, C.; Kong, F.; Cao, D. Overexpression of GmFDL19 enhances tolerance to drought and salt stresses in soybean. PLoS One 2017, 12, e0179554. [Google Scholar] [CrossRef] [PubMed]
  121. Li, Y.; Zhang, H.; Liang, S.; Chen, X.; Liu, J.; Zhang, Y.; Wang, A. Identification of CDPK Gene Family in Solanum habrochaites and Its Function Analysis under Stress. International journal of molecular sciences 2022, 23. [Google Scholar] [CrossRef]
  122. Li, Z.; Palmer, W.M.; Martin, A.P.; Wang, R.; Rainsford, F.; Jin, Y.; Patrick, J.W.; Yang, Y.; Ruan, Y.-L. High invertase activity in tomato reproductive organs correlates with enhanced sucrose import into, and heat tolerance of, young fruit. Journal of experimental botany 2012, 63, 1155–1166. [Google Scholar] [CrossRef]
  123. Liang, W.; Ma, X.; Wan, P.; Liu, L. Plant salt-tolerance mechanism: A review. Biochemical and biophysical research communications 2018, 495, 286–291. [Google Scholar] [CrossRef] [PubMed]
  124. Lin, F.; Li, S.; Wang, K.; Tian, H.; Gao, J.; Zhao, Q.; Du, C. A leucine-rich repeat receptor-like kinase, OsSTLK, modulates salt tolerance in rice. Plant Science 2020, 296, 110465. [Google Scholar] [CrossRef]
  125. Liu, A.; Xiao, Z.; Li, M.W.; Wong, F.L.; Yung, W.S.; Ku, Y.S.; Wang, Q.; Wang, X.; Xie, M.; Yim, A.K.Y. Transcriptomic reprogramming in soybean seedlings under salt stress. Plant, Cell & Environment 2019, 42, 98–114. [Google Scholar]
  126. Liu, H.; Xiao, C.; Qiu, T.; Deng, J.; Cheng, H.; Cong, X.; Cheng, S.; Rao, S.; Zhang, Y. Selenium Regulates Antioxidant, Photosynthesis, and Cell Permeability in Plants under Various Abiotic Stresses: A Review. Plants 2023, 12. [Google Scholar] [CrossRef]
  127. Liu, Q.; Lin, Y.; Liu, J.; Barrett, S.C.; Wang, Y. A New Type of Cell for Organ Movement in Plants. Available at SSRN 416 4238. [CrossRef]
  128. Liu, X.; Xu, X.; Ji, M.; Amombo, E.; Fu, J. Growth, Photosynthesis, and Gene Expression of Bermudagrass in Response to Salinity and Shade Stress. Journal of the American Society for Horticultural Science 2023, 148, 1–8. [Google Scholar] [CrossRef]
  129. Liu, Y.; Song, Y.; Zeng, S.; Patra, B.; Yuan, L.; Wang, Y. Isolation and characterization of a salt stress-responsive betaine aldehyde dehydrogenase in Lycium ruthenicum Murr. Physiologia plantarum 2018, 163, 73–87. [Google Scholar] [CrossRef] [PubMed]
  130. Mahajan, S.; Pandey, G.K.; Tuteja, N. Calcium-and salt-stress signaling in plants: shedding light on SOS pathway. Archives of biochemistry and biophysics 2008, 471, 146–158. [Google Scholar] [CrossRef] [PubMed]
  131. Mäkelä, P.; Munns, R.; Colmer, T.; Condon, A.; Peltonen-Sainio, P. Effect of foliar applications of glycinebetaine on stomatal conductance, abscisic acid and solute concentrations in leaves of salt-or drought-stressed tomato. Functional Plant Biology 1998, 25, 655–663. [Google Scholar] [CrossRef]
  132. Malik, A. Groundwater hydro-geochemistry, irrigation and drinking quality, and source apportionment in the intensively cultivated area of Sutlej sub-basin of main Indus basin. Environmental Earth Sciences 2022, 81, 1–26. [Google Scholar] [CrossRef]
  133. Malik, L.; Sanaullah, M.; Mahmood, F.; Hussain, S.; Siddique, M.H.; Anwar, F.; Shahzad, T. Unlocking the potential of co-applied biochar and plant growth-promoting rhizobacteria (PGPR) for sustainable agriculture under stress conditions. Chemical and biological technologies in agriculture 2022, 9, 1–29. [Google Scholar] [CrossRef]
  134. Manandhar, A. Making space for the pore: Structural characterization of stomatal movements when guard cells and epidermal cells interact. 2022.
  135. Mansour, M.M.F. Role of Vacuolar Membrane Transport Systems in Plant Salinity Tolerance. Journal of Plant Growth Regulation 2022, 1–38. [Google Scholar] [CrossRef]
  136. Masuda, M.S.; Azad, M.a.K. , Hasanuzzaman, M.; Arifuzzaman, M. Evaluation of salt tolerance in maize (Zea mays L.) at seedling stage through morphological characters and salt tolerance index. Plant Physiology Reports 2021, 26, 419–427. [Google Scholar] [CrossRef]
  137. Mbarki, S.; Sytar, O.; Cerda, A.; Zivcak, M.; Rastogi, A.; He, X.; Zoghlami, A.; Abdelly, C.; Brestic, M. Strategies to mitigate the salt stress effects on photosynthetic apparatus and productivity of crop plants. In Salinity responses and tolerance in plants, volume 1; Springer: Berlin/Heidelberg, Germany, 2018; pp. 85–136. [Google Scholar]
  138. Meloni, D.A.; Oliva, M.A.; Ruiz, H.A.; Martinez, C.A. Contribution of proline and inorganic solutes to osmotic adjustment in cotton under salt stress. Journal of Plant Nutrition 2001, 24, 599–612. [Google Scholar] [CrossRef]
  139. Mitra, G. Molecular approaches to nutrient uptake and cellular homeostasis in plants under abiotic stress. In Plant nutrients and abiotic stress tolerance; Springer: Berlin/Heidelberg, Germany, 2018; pp. 525–590. [Google Scholar]
  140. Mittal, S.; Kumari, N.; Sharma, V. Differential response of salt stress on Brassica juncea: photosynthetic performance, pigment, proline, D1 and antioxidant enzymes. Plant Physiology and Biochemistry 2012, 54, 17–26. [Google Scholar] [CrossRef]
  141. Mittler, R. Oxidative stress, antioxidants and stress tolerance. Trends in plant science 2002, 7, 405–410. [Google Scholar] [CrossRef] [PubMed]
  142. Mohamed, A.K.S.; Qayyum, M.F.; Abdel-Hadi, A.M.; Rehman, R.A.; Ali, S.; Rizwan, M. Interactive effect of salinity and silver nanoparticles on photosynthetic and biochemical parameters of wheat. Archives of Agronomy and Soil Science 2017, 63, 1736–1747. [Google Scholar] [CrossRef]
  143. Nadeem, M.; Anwar-Ul-Haq, M.; Saqib, M.; Maqsood, M.; He, Z. Ameliorative effect of silicic acid and silicates on oxidative, osmotic stress, and specific ion toxicity in spring wheat (Triticum aestivum L.) genotypes. Journal of Soil Science and Plant Nutrition.
  144. Naikwade, P.V. Plant Responses to Drought Stress: Morphological, Physiological, Molecular Approaches, and Drought Resistance. In Plant Metabolites under Environmental Stress; Apple Academic Press: Cambridge, MA, USA, 2023; pp. 149–183. [Google Scholar]
  145. Naz, F. Plant nutrition, transport, mechanism and sensing in plants. In Sustainable Plant Nutrition; Elsevier: Amsterdam, The Netherlands, 2023; pp. 209–228. [Google Scholar]
  146. Nguyen, N.H.; Jung, C.; Cheong, J.-J. Chromatin remodeling for the transcription of type 2C protein phosphatase genes in response to salt stress. Plant Physiology and Biochemistry 2019, 141, 325–331. [Google Scholar] [CrossRef] [PubMed]
  147. Nieves-Cordones, M.; Shiblawi, A.; Razzaq, F.; Sentenac, H. Roles and transport of sodium and potassium in plants. In The alkali metal ions: Their role for life; Springer: Berlin/Heidelberg, Germany, 2016; pp. 291–324. [Google Scholar]
  148. Noble, C.; Rogers, M. Arguments for the use of physiological criteria for improving the salt tolerance in crops. Plant and soil 1992, 146, 99–107. [Google Scholar] [CrossRef]
  149. Panda, S.; Zhou, K. Engineering microbes to overproduce natural products as agrochemicals. Synthetic and Systems Biotechnology 2023, 8, 79–85. [Google Scholar] [CrossRef]
  150. Pavlović, I.; Mlinarić, S.; Tarkowská, D.; Oklestkova, J.; Novák, O.; Lepeduš, H.; Bok, V.V.; Brkanac, S.R.; Strnad, M.; Salopek-Sondi, B. Early Brassica crops responses to salinity stress: a comparative analysis between Chinese cabbage, white cabbage, and kale. Frontiers in plant science 2019, 10, 450. [Google Scholar] [CrossRef] [PubMed]
  151. Pei, L.; Li, H.; Zhou, Y.; Li, W.; Jiang, Y.; Li, H. Exogenous glycinebetaine application contributes to abiotic stress tolerance in maize. Journal of Plant Biology 2020, 1–13. [Google Scholar] [CrossRef]
  152. Peng, J.; Liu, J.; Zhang, L.; Luo, J.; Dong, H.; Ma, Y.; Zhao, X.; Chen, B.; Sui, N.; Zhou, Z. Effects of soil salinity on sucrose metabolism in cotton leaves. PLoS One 2016, 11, e0156241. [Google Scholar] [CrossRef]
  153. Phour, M.; Sindhu, S.S. Soil Salinity and Climate Change: Microbiome-Based Strategies for Mitigation of Salt Stress to Sustainable Agriculture. In Climate Change and Microbiome Dynamics; Springer: Berlin/Heidelberg, Germany, 2023; pp. 191–243. [Google Scholar]
  154. Pigeot-Rémy, S.; Simonet, F.; Atlan, D.; Lazzaroni, J.; Guillard, C. Bactericidal efficiency and mode of action: a comparative study of photochemistry and photocatalysis. Water research 2012, 46, 3208–3218. [Google Scholar] [CrossRef]
  155. Pitman, M.G.; Läuchli, A. Global impact of salinity and agricultural ecosystems. In Salinity: environment-plants-molecules; Springer: Berlin/Heidelberg, Germany, 2002; pp. 3–20. [Google Scholar]
  156. Prodjinoto, H.; Irakoze, W.; Gandonou, C.; Lepoint, G.; Lutts, S. Discriminating the impact of Na+ and Cl− in the deleterious effects of salt stress on the African rice species (Oryza glaberrima Steud.). Plant Growth Regulation 2021, 94, 201–219. [Google Scholar] [CrossRef]
  157. Qi, Z.; Spalding, E.P. Protection of plasma membrane K+ transport by the salt overly sensitive1 Na+-H+ antiporter during salinity stress. Plant Physiology 2004, 136, 2548–2555. [Google Scholar] [CrossRef] [PubMed]
  158. Qin, L.; Zhang, H.; Li, J.; Zhu, Y.; Jiao, G.; Wang, C.; Wu, S. Down-regulation of GhADF1 in cotton (Gossypium hirsutum) improves plant drought tolerance and increases fiber yield. The Crop Journal 2022, 10, 1037–1048. [Google Scholar] [CrossRef]
  159. Rahman, A.; Nahar, K.; Al Mahmud, J.; Hasanuzzaman, M.; Hossain, M.S.; Fujita, M. Salt stress tolerance in rice: Emerging role of exogenous phytoprotectants. Advances in international rice research 2017, 9, 139–174. [Google Scholar]
  160. Rajappa, S.; Krishnamurthy, P.; Kumar, P.P. Regulation of AtKUP2 expression by bHLH and WRKY transcription factors helps to confer increased salt tolerance to Arabidopsis thaliana plants. Frontiers in plant science 2020, 11, 1311. [Google Scholar] [CrossRef] [PubMed]
  161. Rajput, V.; Minkina, T.; Kumari, A. Harish. Singh, RK.
  162. Rauf, M.; Shahzad, K.; Ali, R.; Ahmad, M.; Habib, I.; Mansoor, S.; Berkowitz, G.A.; Saeed, N.A. Cloning and characterization of Na+/H+ antiporter (LfNHX1) gene from a halophyte grass Leptochloa fusca for drought and salt tolerance. Molecular Biology Reports 2014, 41, 1669–1682. [Google Scholar] [CrossRef]
  163. Raziq, A.; Din, A.M.U.; Anwar, S.; Wang, Y.; Jahan, M.S.; He, M.; Ling, C.G.; Sun, J.; Shu, S.; Guo, S. Exogenous spermidine modulates polyamine metabolism and improves stress responsive mechanisms to protect tomato seedlings against salt stress. Plant Physiology and Biochemistry 2022, 187, 1–10. [Google Scholar] [CrossRef]
  164. Ren, W.; Chen, L.; Peng, X. Combined transcriptome and metabolome analysis revealed pathways involved in improved salt tolerance of Gossypium hirsutum L. seedlings in response to exogenous melatonin application. BMC Plant Biology 2022, 22, 1–19. [Google Scholar] [CrossRef]
  165. Rengel, Z. The role of calcium in salt toxicity. Plant, Cell & Environment 1992, 15, 625–632. [Google Scholar]
  166. Riley, J.J. Review of halophyte feeding trials with ruminants. In Halophytic and Salt-Tolerant Feedstuffs; CRC Press: Boca Raton, FL, USA, 2015; pp. 177–217. [Google Scholar]
  167. Rios-Gonzalez, K.; Erdei, L.; Lips, S.H. The activity of antioxidant enzymes in maize and sunflower seedlings as affected by salinity and different nitrogen sources. Plant Science 2002, 162, 923–930. [Google Scholar] [CrossRef]
  168. Ruiz-Lozano, J.M.; Porcel, R.; Azcón, C.; Aroca, R. Regulation by arbuscular mycorrhizae of the integrated physiological response to salinity in plants: new challenges in physiological and molecular studies. Journal of Experimental Botany 2012, 63, 4033–4044. [Google Scholar] [CrossRef]
  169. Sabagh, A.E.; Islam, M.S.; Iqbal, M.A.; Hossain, A.; Mubeen, M.; Jabeen, T.; Waleed, M.; Wasaya, A.; Rahman, M.H.U.; Ratnasekera, D. Salinity Stress in Cotton: Adverse Effects, Survival Mechanisms and Management Strategies. In Engineering Tolerance in Crop Plants Against Abiotic Stress; CRC Press: Boca Raton, FL, USA, 2021; pp. 59–80. [Google Scholar]
  170. Sadau, S.B.; Mehari, T.G.; Ahmad, A.; Tajo, S.M.; Ibrahim, S.; Iqbal, M.S.; Elasad, M.; Zhang, J.; Wei, H.; Yu, S. Genome wide identification and characterization of MAPK genes reveals their potential in enhancing drought and salt stress tolerance in Gossypium hirsutum. Journal of Cotton Research 2022, 5, 1–13. [Google Scholar] [CrossRef]
  171. Safdar, H.; Amin, A.; Shafiq, Y.; Ali, A.; Yasin, R.; Shoukat, A.; Hussan, M.U.; Sarwar, M.I. A review: Impact of salinity on plant growth. Nat. Sci 2019, 17, 34–40. [Google Scholar]
  172. Salimath, S.S.; Romsdahl, T.B.; Konda, A.R.; Zhang, W.; Cahoon, E.B.; Dowd, M.K.; Wedegaertner, T.C.; Hake, K.D.; Chapman, K.D. Production of tocotrienols in seeds of cotton (Gossypium hirsutum L.) enhances oxidative stability and offers nutraceutical potential. Plant Biotechnology Journal 2021, 19, 1268–1282. [Google Scholar] [CrossRef] [PubMed]
  173. Salwan, R.; Vivek, S. Salinity Stress in Plants and Role of Microbes in Its Alleviation. In Agricultural Biocatalysis; Jenny Stanford Publishing: Dubai, United Arab Emirates, 2023; pp. 175–214. [Google Scholar]
  174. Schobert, B.; Tschesche, H. Unusual solution properties of proline and its interaction with proteins. Biochimica et Biophysica Acta (BBA)-General Subjects 1978, 541, 270–277. [Google Scholar] [CrossRef]
  175. Sghaier, D.B.; Pedro, S.; Duarte, B.; Caçador, I.; Sleimi, N. Photosynthetic Responses of Two Salt-Tolerant Plants, Tamarix gallica and Arthrocnemum indicum Against Arsenic Stress: A Case Study. Arsenic in Plants: Uptake, Consequences and Remediation Techniques.
  176. 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. International Journal of Genomics 2021. [Google Scholar] [CrossRef]
  177. Shahbaz, M.; Ashraf, M. Improving salinity tolerance in cereals. Critical reviews in plant sciences 2013, 32, 237–249. [Google Scholar] [CrossRef]
  178. Shahid, M.A.; Sarkhosh, A.; Khan, N.; Balal, R.M.; Ali, S.; Rossi, L.; Gómez, C.; Mattson, N.; Nasim, W.; Garcia-Sanchez, F. Insights into the physiological and biochemical impacts of salt stress on plant growth and development. Agronomy 2020, 10, 938. [Google Scholar] [CrossRef]
  179. Shi, H.; Lee, B.-H. , Wu, S.-J., and Zhu, J.-K. Overexpression of a plasma membrane Na+/H+ antiporter gene improves salt tolerance in Arabidopsis thaliana. Nature biotechnology 2003, 21, 81–85. [Google Scholar] [CrossRef]
  180. Shi, Q.; Ding, F.; Wang, X.; Wei, M. Exogenous nitric oxide protect cucumber roots against oxidative stress induced by salt stress. Plant physiology and biochemistry 2007, 45, 542–550. [Google Scholar] [CrossRef]
  181. Shrivastava, P.; Kumar, R. Soil salinity: A serious environmental issue and plant growth promoting bacteria as one of the tools for its alleviation. Saudi journal of biological sciences 2015, 22, 123–131. [Google Scholar] [CrossRef]
  182. Sikder, R.K.; Wang, X.; Jin, D.; Zhang, H.; Gui, H.; Dong, Q.; Pang, N.; Zhang, X.; Song, M. Screening and evaluation of reliable traits of upland cotton (Gossypium hirsutum L.) genotypes for salt tolerance at the seedling growth stage. Journal of Cotton Research 2020, 3, 1–13. [Google Scholar] [CrossRef]
  183. Singh, A. Potassium (K+) transporters in plants: regulation and functional role in K+ uptake and homeostasis. In Cation Transporters in Plants; Elsevier: Amsterdam, The Netherlands, 2022; pp. 29–47. [Google Scholar]
  184. Singh, A.; Sengar, R.S.; Rajput, V.D.; Minkina, T.; Singh, R.K. Zinc Oxide Nanoparticles Improve Salt Tolerance in Rice Seedlings by Improving Physiological and Biochemical Indices. Agriculture 2022, 12, 1014. [Google Scholar] [CrossRef]
  185. Singh, P.; Kumar, V.; Sharma, J.; Saini, S.; Sharma, P.; Kumar, S.; Sinhmar, Y.; Kumar, D.; Sharma, A. Silicon supplementation alleviates the salinity stress in wheat plants by enhancing the plant water status, photosynthetic pigments, proline content and antioxidant enzyme activities. Plants 2022, 11, 2525. [Google Scholar] [CrossRef]
  186. Solebo, O.R. (2021). The Biochemical Functionality and Cellular Importance of the Vacuolar H+ Translocating Pyrophosphatase 1, PfVP1, in Plasmodium Falciparum. Drexel University.
  187. Soltabayeva, A.; Ongaltay, A.; Omondi, J.O.; Srivastava, S. Morphological, physiological and molecular markers for salt-stressed plants. Plants 2021, 10, 243. [Google Scholar] [CrossRef]
  188. Somaddar, U.; Dey, H.C.; Mim, S.K.; Sarker, U.K.; Uddin, M.R.; Ahmed, N.U.; Mostofa, M.G.; Saha, G. Assessing Silicon-Mediated Growth Performances in Contrasting Rice Cultivars under Salt Stress. Plants 2022, 11, 1831. [Google Scholar] [CrossRef]
  189. Song, Y.; Ci, D.; Tian, M.; Zhang, D. Comparison of the physiological effects and transcriptome responses of Populus simonii under different abiotic stresses. Plant molecular biology 2014, 86, 139–156. [Google Scholar] [CrossRef]
  190. Soori, N.; Bakhshi, D.; Rezaei, N.A.; Faizian, M. Effect of salinity stress on some physiological characteristics and photosynthetic parameters of several Iranian commercial pomegranate genotypes. 2019.
  191. Sottosanto, J.B.; Gelli, A.; Blumwald, E. DNA array analyses of Arabidopsis thaliana lacking a vacuolar Na+/H+ antiporter: impact of AtNHX1 on gene expression. The Plant Journal 2004, 40, 752–771. [Google Scholar] [CrossRef]
  192. Stigter, T.; Ribeiro, L.; Dill, A. Evaluation of an intrinsic and a specific vulnerability assessment method in comparison with groundwater salinisation and nitrate contamination levels in two agricultural regions in the south of Portugal. Hydrogeology journal 2006, 14, 79–99. [Google Scholar] [CrossRef]
  193. Suman, S. Plant tissue culture: A promising tool of quality material production with special reference to micropropagation of banana. Biochem. Cell. Arch 2017, 17, 1–26. [Google Scholar]
  194. Sun, H.; Sun, X.; Wang, H.; Ma, X. Advances in salt tolerance molecular mechanism in tobacco plants. Hereditas 2020, 157, 1–6. [Google Scholar] [CrossRef]
  195. Sun, L.J.; Zhao, X.Y.; Ren, J.; Yan, S.P.; Zhao, X.Y.; Song, X.S. Overexpression of Cerasus humilis ChAOX2 improves the tolerance of Arabidopsis to salt stress. 3 Biotech 2021, 11, 1–11. [Google Scholar]
  196. Talaat, N.B.; Mostafa, A.A.; El-Rahman, S.N.A. A novel plant growth–promoting agent mitigates salt toxicity in barley (Hordeum vulgare L.) by activating photosynthetic, antioxidant defense, and methylglyoxal detoxification machineries. Journal of Soil Science and Plant Nutrition.
  197. Tanveer, M.; Ahmed, H.a.I. ROS signalling in modulating salinity stress tolerance in plants. In Salt and drought stress tolerance in plants; Springer: Berlin/Heidelberg, Germany, 2020; pp. 299–314. [Google Scholar]
  198. Tian, C.; Feng, G.; Li, X.; Zhang, F. Different effects of arbuscular mycorrhizal fungal isolates from saline or non-saline soil on salinity tolerance of plants. Applied Soil Ecology 2004, 26, 143–148. [Google Scholar] [CrossRef]
  199. Tian, X.; He, M.; Wang, Z.; Zhang, J.; Song, Y.; He, Z.; Dong, Y. Application of nitric oxide and calcium nitrate enhances tolerance of wheat seedlings to salt stress. Plant Growth Regulation 2015, 77, 343–356. [Google Scholar] [CrossRef]
  200. Verbruggen, N.; Hermans, C. Proline accumulation in plants: a review. Amino acids 2008, 35, 753–759. [Google Scholar] [CrossRef] [PubMed]
  201. Vivaldi, G.A.; Camposeo, S.; Romero-Trigueros, C.; Pedrero, F.; Caponio, G.; Lopriore, G.; Álvarez, S. Physiological responses of almond trees under regulated deficit irrigation using saline and desalinated reclaimed water. Agricultural Water Management 2021, 258, 107172. [Google Scholar] [CrossRef]
  202. Wall, S.; Vialet-Chabrand, S.; Davey, P.; Van Rie, J.; Galle, A.; Cockram, J.; Lawson, T. Stomata on the abaxial and adaxial leaf surface contribute differently to leaf gas exchange and photosynthesis in wheat. New Phytologist 2022. [Google Scholar] [CrossRef] [PubMed]
  203. Wang, G.; Fu, X.; Zhao, W.; Zhang, M.; Chen, F. Ectopic Expression of Maize Plastidic Methionine Sulfoxide Reductase ZmMSRB1 Enhances Salinity Stress Tolerance in Arabidopsis thaliana. Plant Molecular Biology Reporter 2022, 40, 284–295. [Google Scholar] [CrossRef]
  204. Wang, G.; Liang, Y.-H. , Zhang, J.-Y., and Cheng, Z.-M.M. Cloning, molecular and functional characterization by overexpression in Arabidopsis of MAPKK genes from grapevine (Vitis vinifera). BMC plant biology 2020, 20, 1–15. [Google Scholar]
  205. Wang, G.; Zhang, L.; Zhang, S.; Li, B.; Li, J.; Wang, X.; Zhang, J.; Guan, C.; Ji, J. The combined use of a plant growth promoting Bacillus sp. strain and GABA promotes the growth of rice under salt stress by regulating antioxidant enzyme system, enhancing photosynthesis and improving soil enzyme activities. Microbiological Research 2023, 266, 127225. [Google Scholar] [CrossRef]
  206. Wang, X.; Gao, B.; Liu, X.; Dong, X.; Zhang, Z.; Fan, H.; Zhang, L.; Wang, J.; Shi, S.; Tu, P. Salinity stress induces the production of 2-(2-phenylethyl) chromones and regulates novel classes of responsive genes involved in signal transduction in Aquilaria sinensis calli. BMC plant biology 2016, 16, 1–20. [Google Scholar] [CrossRef]
  207. Wani, A.S.; Ahmad, A.; Hayat, S.; Tahir, I. Epibrassinolide and proline alleviate the photosynthetic and yield inhibition under salt stress by acting on antioxidant system in mustard. Plant physiology and biochemistry 2019, 135, 385–394. [Google Scholar]
  208. Wani, S.H.; Kumar, V.; Shriram, V.; Sah, S.K. Phytohormones and their metabolic engineering for abiotic stress tolerance in crop plants. The Crop Journal 2016, 4, 162–176. [Google Scholar]
  209. Wei, X.; Huang, X.; Yang, W.; Wang, X.; Guan, T.; Kang, Z.; Liu, J. A Chloroplast-Localized Glucose-6-Phosphate Dehydrogenase Positively Regulates Stripe Rust Resistance in Wheat. International Journal of Molecular Sciences 2022, 24, 459. [Google Scholar] [PubMed]
  210. Wei, Y.; Xu, Y.; Lu, P.; Wang, X.; Li, Z.; Cai, X.; Zhou, Z.; Wang, Y.; Zhang, Z.; Lin, Z. Salt stress responsiveness of a wild cotton species (Gossypium klotzschianum) based on transcriptomic analysis. PLoS One 2017, 12, e0178313. [Google Scholar]
  211. Williams, L.; Baeza, P.; Vaughn, P. Midday measurements of leaf water potential and stomatal conductance are highly correlated with daily water use of Thompson Seedless grapevines. Irrigation Science 2012, 30, 201–212. [Google Scholar]
  212. Wu, L.; Huo, W.; Yao, D.; Li, M. Effects of solid matrix priming (SMP) and salt stress on broccoli and cauliflower seed germination and early seedling growth. Scientia horticulturae 2019, 255, 161–168. [Google Scholar]
  213. Xu, F.; Xu, S.; Wiermer, M.; Zhang, Y.; Li, X. The cyclin L homolog MOS12 and the MOS4-associated complex are required for the proper splicing of plant resistance genes. The Plant Journal 2012, 70, 916–928. [Google Scholar]
  214. Xu, H.; Lu, Y.; Tong, S. Effects of arbuscular mycorrhizal fungi on photosynthesis and chlorophyll fluorescence of maize seedlings under salt stress. Emirates Journal of Food and Agriculture.
  215. Yadav, S.P.; Bharadwaj, R.; Nayak, H.; Mahto, R.; Singh, R.K.; Prasad, S.K. Impact of salt stress on growth, productivity and physicochemical properties of plants: A Review. International Journal of Chemical Studies 2019, 7, 1793–1798. [Google Scholar]
  216. Yan, K.; Mei, H.; Dong, X.; Zhou, S.; Cui, J.; Sun, Y. Dissecting photosynthetic electron transport and photosystems performance in Jerusalem artichoke (Helianthus tuberosus L.) under salt stress. Frontiers in Plant Science 2022, 13, 905100. [Google Scholar]
  217. Yang, G.; Yu, L.; Zhang, K.; Zhao, Y.; Guo, Y.; Gao, C. A ThDREB gene from Tamarix hispida improved the salt and drought tolerance of transgenic tobacco and T. hispida. Plant Physiology and Biochemistry 2017, 113, 187–197. [Google Scholar]
  218. Yang, X.; Lu, M.; Wang, Y.; Wang, Y.; Liu, Z.; Chen, S. "Response Mechanism of Plants to Drought Stress. Horticulturae 2021, 2021, 7–50. [Google Scholar]
  219. Yang, Y.; Xie, J.; Li, J.; Zhang, J.; Zhang, X.; Yao, Y.; Wang, C.; Niu, T.; Bakpa, E.P. Trehalose alleviates salt tolerance by improving photosynthetic performance and maintaining mineral ion homeostasis in tomato plants. Frontiers in Plant Science 2022, 13, 974507. [Google Scholar] [PubMed]
  220. Yang, Y.M.; Zhu, Y.; Naseer, M.; Wang, Q.; Li, G.; Tao, H.Y.; Zhu, S.G.; Wang, B.Z.; Wang, W.; Xiong, Y.C. Rhizosphere effect of nanoscale zero-valent iron on mycorrhiza-dependent maize assimilation. Plant, Cell & Environment 2023, 46, 251–267. [Google Scholar]
  221. Yao, X.; Horie, T.; Xue, S.; Leung, H.-Y. , Katsuhara, M.; Brodsky, D.E.; Wu, Y.; Schroeder, J.I. Differential sodium and potassium transport selectivities of the rice OsHKT2; 1 and OsHKT2; 2 transporters in plant cells. Plant physiology 2010, 152, 341–355. [Google Scholar] [CrossRef] [PubMed]
  222. Yildirim, E.; Ekinci, M.; Turan, M.; Ağar, G.; Ors, S.; Dursun, A.; Kul, R.; Akgül, G. Physiological and Biochemical Changes of Pepper Cultivars Under Combined Salt and Drought Stress. Gesunde Pflanzen 2022, 1–9. [Google Scholar] [CrossRef]
  223. Yildiztugay, E.; Ozfidan-Konakci, C.; Kucukoduk, M.; Turkan, I. Flavonoid naringenin alleviates short-term osmotic and salinity stresses through regulating photosynthetic machinery and chloroplastic antioxidant metabolism in Phaseolus vulgaris. Frontiers in plant science 2020, 11, 682. [Google Scholar]
  224. Yin, L.; Wang, S.; Li, J.; Tanaka, K.; Oka, M. Application of silicon improves salt tolerance through ameliorating osmotic and ionic stresses in the seedling of Sorghum bicolor. Acta physiologiae plantarum 2013, 35, 3099–3107. [Google Scholar] [CrossRef]
  225. Yu, Y.; Guo, D.-D. , Min, D.-H., Cao, T.; Ning, L.; Jiang, Q.-Y., Sun, X.-J., Zhang, H.; Tang, W.-S., and Gao, S.-Q. (2022). Foxtail millet MYB-like transcription factor SiMYB16 confers salt tolerance in transgenic rice by regulating phenylpropane pathway. Plant Physiology and Biochemistry.
  226. Yue, J.-Y. , Jiao, J.-L., Wang, W.-W., and Wang, H.-Z. The Calcium-Dependent Protein Kinase TaCDPK27 Positively Regulates Salt Tolerance in Wheat. International Journal of Molecular Sciences 2022, 23, 7341. [Google Scholar] [CrossRef]
  227. Zahra, N.; Al Hinai, M.S.; Hafeez, M.B.; Rehman, A.; Wahid, A.; Siddique, K.H.; Farooq, M. Regulation of photosynthesis under salt stress and associated tolerance mechanisms. Plant Physiology and Biochemistry 2022, 178, 55–69. [Google Scholar] [CrossRef]
  228. Zhang, D.; Jiang, S.; Pan, J.; Kong, X.; Zhou, Y.; Liu, Y.; Li, D. The overexpression of a maize mitogen-activated protein kinase gene (Z m MPK 5) confers salt stress tolerance and induces defence responses in tobacco. Plant biology 2014, 16, 558–570. [Google Scholar] [CrossRef]
  229. Zhang, X.; Long, Y.; Huang, J.; Xia, J. OsNAC45 is involved in ABA response and salt tolerance in rice. Rice 2020, 13, 1–13. [Google Scholar] [CrossRef] [PubMed]
  230. Zhang, X.; Wang, L.; Xu, X.; Cai, C.; Guo, W. Genome-wide identification of mitogen-activated protein kinase gene family in Gossypium raimondii and the function of their corresponding orthologs in tetraploid cultivated cotton. BMC plant biology 2014, 14, 1–17. [Google Scholar] [CrossRef] [PubMed]
  231. Zhang, Y.; Fang, J.; Wu, X.; Dong, L. Na+/K+ balance and transport regulatory mechanisms in weedy and cultivated rice (Oryza sativa L.) under salt stress. BMC plant biology 2018, 18, 1–14. [Google Scholar] [CrossRef] [PubMed]
  232. Zhang, Y.; Liu, G.; Dong, H.; Li, C. Waterlogging stress in cotton: Damage, adaptability, alleviation strategies, and mechanisms. The Crop Journal 2021, 9, 257–270. [Google Scholar] [CrossRef]
  233. Zhao, C.; Zhang, H.; Song, C.; Zhu, J.-K. , and Shabala, S. Mechanisms of plant responses and adaptation to soil salinity. The innovation 2020, 1, 100017. [Google Scholar] [CrossRef]
  234. Zhao, W.; Jung, S.; Schubert, S. Transcription profile analysis identifies marker genes to distinguish salt shock and salt stress after stepwise acclimation in Arabidopsis thaliana and Zea mays. Plant Physiology and Biochemistry 2019, 143, 232–245. [Google Scholar] [CrossRef]
  235. Zhao, X.; Wu, T.; Guo, S.; Hu, J.; Zhan, Y. Ectopic Expression of AeNAC83, a NAC Transcription Factor from Abelmoschus esculentus, Inhibits Growth and Confers Tolerance to Salt Stress in Arabidopsis. International Journal of Molecular Sciences 2022, 23, 10182. [Google Scholar] [CrossRef]
  236. Zhao, Z.; Zhang, G.; Zhou, S.; Ren, Y.; Wang, W. The improvement of salt tolerance in transgenic tobacco by overexpression of wheat F-box gene TaFBA1. Plant Science 2017, 259, 71–85. [Google Scholar] [CrossRef]
  237. Zhu, J.; Guo, Y.; Zhou, H.; Hong, Y.; Zhao, C.; Shabala, S.; Lv, C.; Guo, B.; Zhou, M.; Xu, R. Genome wide association study and haplotype analysis reveals the role of HvHKT1; 5 in potassium retention but not Na+ exclusion in barley (Hordeum vulgare L.). Environmental and Experimental Botany 2022, 201, 104973. [Google Scholar] [CrossRef]
  238. Zou, J.; Hu, W.; Li, Y.; Zhu, H.; He, J.; Wang, Y.; Meng, Y.; Chen, B.; Zhao, W.; Wang, S. Leaf anatomical alterations reduce cotton’s mesophyll conductance under dynamic drought stress conditions. The Plant Journal 2022. [Google Scholar] [CrossRef]
Figure 1. The general diagram describes the developing process of salt tolerance in plants.
Figure 1. The general diagram describes the developing process of salt tolerance in plants.
Preprints 79971 g001
Figure 2. The mechanism of salt-related plant injuries.
Figure 2. The mechanism of salt-related plant injuries.
Preprints 79971 g002
Figure 3. Processes of antioxidant defence Additional information can be found in the text.
Figure 3. Processes of antioxidant defence Additional information can be found in the text.
Preprints 79971 g003
Figure 5. .
Figure 5. .
Preprints 79971 g005
Table 1. Different plants identified with salt stress have distinct genes that are upregulated and downregulated.
Table 1. Different plants identified with salt stress have distinct genes that are upregulated and downregulated.
Name of Genes Function Upregulated/ Downregulated References
ETR1/ ETR2/ EIN4 Increase salt tolerance Downregulated (Arif et al., 2020)
B-ARR, CRE1 Improve plant development with ion homeostasis Downregulated (Song et al., 2014)
LHC Adapt the light-harvesting system Downregulated (Ansari et al., 2022)
BRI1/2 Ion homeostasis improves plant development Downregulated (Brenya, 2020)
MPK3, MEKK2 Ion homeostasis, osmoprotectant Downregulated (Billah et al., 2021)
NPR1 Reduce the input, exclusion, and sequestration of salt Upregulated (Bor and Özdemir, 2018)
GmST1 Exclusion of salt and improvement of the immune and antioxidant systems Upregulated (Kaushal, 2020)
SAUR, ARF Promotes plant growth and balance Upregulated (Zhao et al., 2022)
CaM, CDPK, CML, CBL Ion homeostasis enhances plant development and growth Upregulated (Shah et al., 2021)
MYC2, TGA Increase the production of jasmonates and activate plant defenses Upregulated (Delgado et al., 2021)
CCOMT Promote the biosynthesis of lignin and suberin. Upregulated (Yu et al., 2022)
NRT1/PTR, CAX1 Enhance the antioxidant system and control calcium homeostasis Upregulated (Mitra, 2018)
SOS3/,SOS2/ CIPK24 Salt exclusion and sequestration Upregulated (Mahajan et al., 2008)
SDIR1 Salt exclusion and sequestration Upregulated (Zhao et al., 2017)
SDIR1 Salt homeostasis, decrease senescence Upregulated (Johnson, 2016)
BAK1 Exclusion and sequestration of salt Upregulated (Lin et al., 2020)
NCED Osmoprotectant, closure of the stomata, development of lateral roots, and salt acclimation Upregulated (Wani et al., 2016)
LAX Vascular development Upregulated (Chai et al., 2016)
SWEET15 Modify vacuolar transit and sugar storage Upregulated (Zhao et al., 2019)
RCA1, AOX1A Modulate photosynthesis Upregulated (Albaladejo et al., 2018)
LRR-RLK, CRR-RLK, Plant development and growth Upregulated (Wang et al., 2016)
NHX1 Transport of salt from the cytosol to the vacuole Upregulated (Graus et al., 2022)
SOS, VDAC Salt exclusion, ion homeostasis, and homeostasis Upregulated (Dong and Wang, 2022)
AtNHX1 Enhanced germination and increased biomass Upregulated (Ren et al., 2022)
OsNHX1 Enhanced biomass Upregulated (Mansour, 2022)
GutD Enhanced biomass Upregulated (Khan et al., 2022)
BADH Better resiliency to salt stress and development Upregulated (Chen et al., 2022)
AtNHX1 Improved germination and increased biomass Upregulated (Deng et al., 2023)
Bt Improved chlorophyll growth and stability Upregulated (Talaat et al., 2022)
CAX1 Enhance the antioxidant system and control calcium homeostasis Upregulated (Wei et al., 2017)
ZEP, ABA 8′- OH, ABI5 Osmoprotectant, closure of the stomata, development of lateral roots, and salt acclimation Upregulated (Bartels and Sunkar, 2005)
Table 2. Different salt stress genes' roles in different plants.
Table 2. Different salt stress genes' roles in different plants.
Gene name Role in Gene Origin Gene References
VvNAC17 An osmotic regulation-related tonoplast intrinsic protein Glycine max (Yang et al., 2021)
VvNAC17 Increases ABA and stress-related gene expression Vitis vinifera (Ju et al., 2020)
PpSARK ABA-related senescence-associated receptor-like kinase Physcomitrella patens (Li et al., 2018b)
AtGSTU19 Effectively fine-tune redox homeostasis at the root Arabidopsis thaliana (Horváth et al., 2019)
AtGSTF8 A root-level process that fine-tunes the redox homeostasis Arabidopsis thaliana (Wang et al., 2022)
SlMYB102 A transcriptional regulator of stress reactions Solanum Lycopersicum (Jia et al., 2021)
ThDREB A transcription factor that plays a role in stress responses Tamarix his pida (Yang et al., 2017)
AtMYB20 Negatively controls serine/threonine protein phosphatases of type 2C Arabidopsis thaliana (Nguyen et al., 2019)
AtWRKY33 Regulated salt responsive gene improved salt tolerance AtKUP2 Arabidopsis thaliana (Rajappa et al., 2020)
AtbHLH122 Enhanced salt sensitivity via controlling the salt-sensitive gene AtKUP2 Arabidopsis thaliana (Singh, 2022)
GmbZIP2 A gene transcription factor that plays a role in the response to salt stress. Glycine max (Ke et al., 2022)
GhMPK2 MAPK signaling via protein kinase Gossypium herbaceum (Zhang et al., 2014b)
GhMAP3K40 In MAPK signaling, protein kinase Gossypium herbaceum (Ashraf et al., 2018a)
ZmMPK5 MAPK signaling via protein kinase Zea mays (Zhang et al., 2014a)
ZmMKK4 Protein kinase in MAPKs signaling Zea mays (Kong et al., 2011)
VvMKK4 MAPK signaling protein kinase Vitis vinifera (Wang et al., 2020)
VvMKK2 protein kinase in MAPKs signaling Vitis vinifera (Çakır and Kılıçkaya, 2015)
PtMAPKK4 protein kinase in MAPKs signaling Populus trichocarpa (Sun et al., 2020)
AtCPK12 protein kinase in CDPKs signaling Arabidopsis thaliana (Hrabak et al., 1996)
AtCDPK27 CDPK signaling involves a membrane-localized protein kinase Arabidopsis thaliana (Yue et al., 2022)
OsMADS25 A transcription factor that participates in ABA-mediated regulatory pathways Oryza sativa (Hussain et al., 2021)
OsNAC45 A transcription factor is involved in several stress and ABA signaling responses Oryza sativa (Zhang et al., 2020)
ChVDE Violaxanthin de-epoxidase catalyzes the transformation Cerasus humilis (Sun et al., 2021)
AVP1 Vacuolar H+ phosphorylase Arabidopsis thaliana (Solebo, 2021)
TaTVP1 Vacuolar H+ phosphorylase Triticum aestivuml (Hao et al., 2021)
PtVP1.1 vacuolar H+ phosphorylase Populus trichocarpa (Gao et al., 2022)
VrNHX1 Na+/H+ vacuolar antiporter Vigna radiata (Bai et al., 2023)
P5CS1 A protein that regulates the rate of proline biosynthesis Phaseolus vulgaris (Hosseinifard et al., 2022)
P5CS2 The enzyme responsible for limiting the rate of proline biosynthesis Phaseolus vulgaris (Chen et al., 2010)
LrAMADH1 Betaine aldehyde to betaine converting catalyzed Lycium ruthenicum Murr (Liu et al., 2018)
SlBADH Promote the transformation of betaine aldehyde to betaine Suaeda liaodonggensis (Li et al., 2003)
AtTPPD The dephosphorylation of trehalose 6-phosphate to form trehalose is instigated Arabidopsis thaliana (Krasensky et al., 2014)
HvHKT1;1 When the external K+ level is low, transports Na+ and mediates Na+ absorption. Hordeum vulgare (Haro et al., 2005)
HvHKT1;5 Negatively transport Na+ in Barley Hordeum vulgare (Zhu et al., 2022)
OsHKT1 Transport Na+ Oryza sativa (Zhang et al., 2018)
OsHKT2 Transport Na+ Oryza sativa (Yao et al., 2010)
OsVHA Transport Na+ Oryza sativa (Kader et al., 2006)
AtNHX1 Vacuolar Na+/H+ antiporter Arabidopsis thaliana (Sottosanto et al., 2004)
LfNHX1 Na+/H+ vacuolar Leptochloa fusca (Rauf et al., 2014)
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.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

© 2024 MDPI (Basel, Switzerland) unless otherwise stated