Open access peer-reviewed chapter - ONLINE FIRST

Underlying Mechanisms of Action to Improve Plant Growth and Fruit Quality in Crops under Alkaline Stress

Written By

Fabián Pérez-Labrada, José Luis Espinoza-Acosta, Daniel Bárcenas-Santana, Elizabeth García-León and Mari Carmen López-Pérez

Submitted: 22 December 2023 Reviewed: 21 February 2024 Published: 24 April 2024

DOI: 10.5772/intechopen.114335

Abiotic Stress in Crop Plants IntechOpen
Abiotic Stress in Crop Plants Edited by Mirza Hasanuzzaman

From the Edited Volume

Abiotic Stress in Crop Plants [Working Title]

Prof. Mirza Hasanuzzaman and MSc. Kamrun Nahar

Chapter metrics overview

14 Chapter Downloads

View Full Metrics

Abstract

The high content of carbonates (CO32−), bicarbonate (HCO3−), and high pH (>7.5) causes environmental pressure and alkaline stress, impairs plant growth and development, and limits fruit quality by causing osmotic alterations and hindering nutrient absorption. Because of alkaline stress, plants are in an oxidative environment that alters their metabolic processes, impairing their growth, development, and fruit quality. In response to this situation, plants use several mechanisms to cope, including the alteration of osmolytes, induction of transcription factors, signal transduction, hormone synthesis, alteration of the antioxidant system, and differential gene expression. Current knowledge and understanding of the underlying mechanisms that promote alkalinity tolerance in crops may lead to new production strategies to improve crop quality under these conditions, while ensuring food security.

Keywords

  • antioxidants
  • calcium
  • pH
  • oxidative stress
  • potassium
  • transcriptomic

1. Introduction

Natural alkalinity (associated with arid and semi-arid regions) or alkalinity induced by agricultural activity is serious because it affects food productivity [1, 2]. Of the total global cultivated land, this stress affects 30–37% (438 million ha) [3]. High carbonate (CO32−) and bicarbonate (HCO3) contents [4] promote the soil buffering capacity, inducing a high pH [5]. Natural highly alkaline conditions (pH ≥ 9) are associated with parts of Western Asia, Eastern Europe, southern South America, and southern Australia, with increasing incidence in India and Egypt as a consequence of irrigation by the Ganges River and the Nile (Figure 1) [6]. This alkaline environment impairs plant and fruit growth and quality by negatively altering the root system, modifying the nutrient uptake system, and altering the relative water, chlorophyll (Chl), and soluble sugar contents [7]. In addition, there is usually a reduction in the microbiological activity and bioavailability of mineral nutrients [5, 8]. Likewise, plants show damage to electron transport, and thus, photosynthetic activity [9]. Similarly, high pH induces the synthesis of reactive oxygen species (ROS) and malondialdehyde, which cause significant damage to the cell membrane and intracellular components [10].

Figure 1.

Probable distribution of alkaline soils.

Alkalinity has diversified impacts on plants, so the response and tolerance mechanisms are more complex [10]. Initially, the plant senses the stressful condition (elevated pH derived from carbonate content) via root apoplast, thanks to a sulfotyrosine of the endogenous peptide root growth factor (RGF) being deprotonated, destabilizing the RGF-RGFR receptor-SERK coreceptor complex. Considering that the target of RGF is a master regulator of root development (PLETHORA) [11], inhibition of this complex reduces root apical meristem growth [12, 13]. The same authors pointed out that this pH sensing system procures a growth or defense system, which significantly affects plant quality.

To combat this stress, plants develop mechanisms that allow them to cope with high pH, high carbonate content, ionic toxicity, and oxidizing environments. Initially, they promote an osmoregulatory system to reduce cellular Ψ [9] and maintain a high K+/Na+ ratio [14]. Likewise, it promotes the activity of enzymatic and non-enzymatic antioxidant systems to combat oxidative stress [15, 16] as well as the expression of hormones and a diversity of metabolites [17] derived from transcriptional modifications [18, 19, 20, 21]. Alternatives in the management of crops developed in these environments include grafting [22], application of Ca2+-based fertilizers (to replace Na+ ions) [23], micronutrients [24], nanotechnology [25], biochar [26], microorganisms [27], growth regulators [28], and genetic engineering [29, 30].

In that sense, this study aimed to establish the main impacts of alkalinity on plants and to review the underlying mechanisms that could promote tolerance to this type of stress. It also highlights management alternatives that can potentiate crop quality in the face of environmental stress. This knowledge seeks to promote the development of innovative strategies to mitigate the negative impacts of alkalinity and safeguard the production of high-quality food.

Advertisement

2. Alkalinity environments

Alkaline stress in soils is characterized by a high pH (generally >7.5) caused by the dominance of [HCO3] + 2[CO32−], salts, and carbonate alkalinity [8, 31]. These anions naturally originate from the dissolution of carbonate minerals from the parent material and by the combination of water and (H2O + CO2disolved + CO32− ↔ 2HCO3) which in turn can react with calcium (Ca2+) in the soil: Ca2+ + CO32− ↔ CaCO3; [Ca2+][CO32−] [8, 32]. Another pathway is high evaporation and low precipitation [33], which are common in arid and semi-arid areas. In soil solutions, when organic species are omitted, the alkalinity can be defined as [HCO3] + 2[CO32−] + [H2PO4] + 2[HPO4] + 3[PO43−] + [B(OH)4] + [OH] − [H+] [8, 34]. Generally, soils can have between 0.4 and 3 mM HCO3, a condition that increases and buffers the edaphic pH and alkalinizes the xylem medium [13]. However, high alkalinity (pH ≥ 9) can occur in soil solutions with HCO3 + CO32− > Ca2+ + Mg2+. Likewise, when sodium chloride (NaCl) salts react in the soil to release H+, they tend to react with CaCO3 to form sodium bicarbonate (NaHCO3) or sodium carbonate (Na2CO3) salts. In addition to increasing exchangeable Na+, it is important to mention that alkalinity is regulated by the concentrations of HCO3 and CO32− and not by Na+ [8, 35]. In this context, the increase in soil pH is associated with an increase in Na+ concentration, an environment that reduces Ca2+ and magnesium (Mg2+) cations, causing the dispersion of clays and loss of soil structure [8].

In water bodies, alkalinity is defined as [HCO3] + 2[CO32−] + [OH] − [H+] and is regulated by the dynamics of CO2dissolved, carbonic acid (H2CO3) [CO2 + H2O = H2CO3], HCO3, and CO32−, as H2CO3 ↔ H+ + HCO3 and HCO3 ↔ H+ + CO32−, and CO2 can react with CaCO3 + H2O dissociating into Ca2+ + 2HCO3 [36]. Therefore, water alkalinity is a determinant of agricultural production, as it can modify the pH of the substrate (H2O + CO2dissolved + CO32− ↔ 2HCO3), altering the growth, development, and quality of plants and fruits [37]. Although the negative effects of carbonates vary according to pH, the greatest damage by HCO3 occurs at pH 8–9 while CO32− occurs at pH >9 [8]. When alkaline soils (with a sodic subsoil) present transient waterlogging or anoxic environments, the CO2 concentration increases with increasing pH because of the increase in HCO3 + CO32− > Ca2+ + Mg2+ [8]. Waterlogging can induce toxicity in Mn, Fe, Al, and B.

Plants in alkaline environments are simultaneously exposed to salt action and elevated pH [9], causing osmotic stress, ionic toxicity, and elevated pH, which severely affect plant growth [1, 38]. High pH reduces cellular water content by altering the root system by reducing turgor pressure, which minimizes the extensibility processes of microfibrils and thus cell extension, water and nutrient uptake, and solute accumulation [39]. Similarly, the buffering capacity of HCO3 affects the solubility of essential nutrients, inducing deficits in iron (Fe2+), manganese (Mn2+), boron (B), zinc (Zn2+), and phosphate (PO43−) [8, 9, 40]. However, attenuation of alkaline effects in soil can be mitigated when pH < 9; soils with pH 8.5 or > 9 demand less H+ to induce a reduction due to lower CaCO3 content, which buffers the pH. In contrast, soils with pH ≈ 8.5 demand more H+ because the concentration of CaCO3 dissolves and buffers pH alteration [8, 41]. This is because CaCO3 ↔ Ca2+ + CO32−, and CO32− + H2O ↔ HCO3 + OH. Thus, one of the most feasible criteria for mitigating the negative effects of alkalinity is HCO3 + CO32− > Ca2+ + Mg2+, where one would ideally seek to reduce HCO3 + CO32− and maintain soil pH ≤ 8.5 [8].

Advertisement

3. Impact of alkalinity on agricultural crops and fruit quality

Alkaline stress has a significant impact on yield, plants, and fruit quality because it hinders the smooth uptake of nutrients [42]. The damage caused varies according to the type and phenological stage of the plant, time of exposure, and level of alkalinity of the medium; however, the main disorders are oxidative damage caused by alterations in ionic homeostasis and impairment of the photosynthetic process (Figure 2a) [43, 44].

Figure 2.

Endogenous alterations (a) and response mechanisms (b) in plants under alkaline stress.

3.1 Process photosynthetic alterations

The photosynthetic capacity of plants under alkaline stress is negatively altered by reduced leaf area and Chl content [25, 45]. In Medicago sativa plants, Chl synthesis and photosystem activity (PSII/PSI) are altered, promoting non-photochemical quenching (NPQ) and cyclic electron transfer, and negative regulation of PetA, PetB, petE, and petF genes associated with the photosynthetic mechanism was observed [46]. A decrease in the maximum quantum yield of PSII photochemical parameters (Fv/Fm), PI, Fv, Fm, Fv/Fo, RC/ABS, and (1-Vj)/Vj, Chl fluorescence, and relative leaf water content in relation to NaHCO3 levels during the vegetative and reproductive growth stages has been documented in strawberry plants. These conditions reduce fruit quality parameters, such as electrical conductivity, pH, total soluble solids, and color [9]. Likewise, a decrease in gas exchange rates (reduced transpiration, stomatal conductance, and CO2 assimilation) can occur, reducing the maximum quantum yield of PSII primary photochemistry (φPo) and causing an increase in PSII energy dissipation and inhibiting PSII electron transport. This environment leads to overproduction and overaccumulation of ROS, mainly hydrogen peroxide (H2O2), hydroxyl ions (OH), singlet oxygen (1O2) [47], and superoxide radicals (O2). Increased concentrations of H2O2, OH, 1O2, and O2•− cause lipid peroxidation of the thylakoid and chloroplast membranes, degrading Chl and inhibiting the electron transport chain, leading to the collapse of the photosynthetic apparatus.

The growth, development, and quality of plants and fruits strongly depend on the synthesis of sugars (carbohydrate metabolism), such as hexose, glucose, and fructose, as they are the basis for the generation of metabolites involved in plant growth [48, 49]. The collapse of the photosynthetic process causes a reduction or no synthesis of sugars; therefore, alkalinity decreases the chemical characteristics of the fruit, which is cited in strawberry fruits under alkaline conditions, where the increase in NaHCO3 concentration was associated with a decrease in the content of soluble sugars [9] coupled with intercellular ion imbalance [50]. Alkalinity stress decreases production, as well as total soluble solids and acidity, in Blanc Du Bois vines with their own roots, derived from the low nutritional status [51]. Similarly, reduction of the photosynthetic process by carbonate salts minimizes the synthesis of some pigments, such as anthocyanins, which may affect the quality of leaves, flowers, and fruits [9, 52]. In addition, the aforementioned soluble sugars participate in carbon stores and are critical for osmoregulation, osmotic adjustment, and ROS scavenging [14, 53].

3.2 Ionic disbalance

The affection in the ionic balance of the plant by alkalinity (NaHCO3) is due to a differential modification in the absorption and storage of calcium (Ca2+), potassium (K+) and sodium (Na+) or by the increase in pH that reduces the availability and absorption of phosphorus (H2PO4), magnesium (Mg2+), K+, iron (Fe), copper (Cu), zinc (Zn), and manganese (Mn) [25, 54, 55, 56]. For example, in oilseed rape (Brassica napus L.), inhibition of Ca2+ uptake in stems and leaves has been reported, which leads to ionic imbalance between K+ and Na+ [57].

Under 40 mM Na2CO3, a reduction in the growth of B. napus shoots was documented because of the high pH and pH-salt ion interaction, showing a high K+ concentration and K+/Na+ ratio in leaves as well as an increase in Ca2+ and Mg2+ in roots [58]. In blueberry “Jewel” (Vaccinium corymbosum) plants, a reduction in H+ flux from the root was documented, which may be associated with inhibition of the VcHA gene encoding H+-ATPases derived from the alkaline condition, pH 8.0 [59]. In strawberry plants, increased Na+ absorption was observed, accompanied by a decrease in K+, Ca2+, Mg2+, and Fe, which negatively affected fruit color [47]. In Ricinus communis L. seeds exposed to 50, 100, and 150 mM NaHCO3, the germination rate was reduced by an increase in Na+ content and reduction in K+ [60].

3.3 Oxidative stress

The burst in ROS production and concentration is a generalized response in plants subjected to alkaline stress, which causes lipid peroxidation, increases malondialdehyde (MDA) concentration, and electrolyte leakage [16, 28, 61], creating an oxidative environment [56] that can damage plant genetic material by arresting the cell cycle in the root [14]. This oxidative environment can reduce the germination process and root development in seedlings [62]. In this regard, inhibition of the germination power of Hordeum vulgare L. under alkaline stress has been cited, decreasing seedling vigor, which may be due to starch degradation, suppression of sugar metabolism, and reduction of enzymes involved in the Embden-Meyerhof-Parnas process, tricarboxylic acid cycle, pentose phosphate pathway, and the enzymes hexokinase and malic dehydrogenase [63].

In maize, increased lipid peroxidation, higher Na+ content, and elevated H2O2 levels were documented, which is in agreement with the reduction of K+ and Chl, decreased shoot and root fresh weight, root dry weight, plant length, number of leaves, and maximum leaf width, when grown under NaHCO3:Na2CO3 (9:1) [64]. The oxidative environment prevailing in plants under alkaline stress can significantly reduce the level of phosphatidylcholine (the main phospholipid of the cell membrane) in addition to inducing alterations in the expression of non-specific phospholipase A and phospholipase D/phospholipase C encoding genes [65]. This phenomenon can lead to marked deterioration in plant and fruit quality.

Under alkaline conditions (Na2CO3 75 mM), Trifolium alexandrinum L. showed oxidative damage by H2O2, O2•−, methylglyoxal, MDA, and oxidized glutathione, reducing shoot and root length, as well as Chl content, water, and nutrient uptake [54], reducing plant quality. In Onobrychis viciaefolia, there was an increase in MDA derived from the generation of ROS (mainly O2•− and H2O2), accompanied by an increase in proline content. This environment reduces fresh and dry biomass, Chl content, K+ concentration, K+/Na+ ratio, and fructose and glucose contents [66]. This stress also affects fruit quality; it decreases nucleic acids, lipids, hydroxycinnamoyl derivatives, organic acids and derivatives, and vitamins in wolfberry (Lycium barbarum) under saline-alkaline conditions [67].

3.4 Enzyme activity and transcriptomic responses

Under alkaline stress, radical cells are severely damaged by the accumulation of ROS (mainly O2•− and H2O2), which inhibit plant growth. Although this condition increases the activation of antioxidant enzymes such as superoxide dismutase (SOD, EC 1.15.1.1), catalase (CAT, EC 1.11.1.6), peroxidase (POD, EC 1.11.1.7), ascorbate peroxidase (APX, EC 1.11.1.11), ascorbic acid (AsA), and glutathione (GSH), their synthesis is overcome by excess ROS. This phenomenon was associated with the positive expression of cell death-associated genes OsKOD1, OsHsr203j, OsCP1, and OsNAC4 and negative regulation of the cell death suppressor gene (OsBI1) [68].

Suppression of enzyme activity implies a potential reduction in plant and fruit growth, development, and quality, although this response is cultivar- and alkalinity level-dependent. In mulberry seedlings, oxidative stress (O2•− and H2O2) induced by NaHCO3 (100 mM) reduced CAT activity and CAT gene expression (W9RJ43), but improved POD and the AsA-GSH acid cycle, allowing the elimination of H2O2. The genes involved in the expression of the electron donor ferredoxin-thioredoxin reductase (FTR) and proteins such as thioredoxin M (TrxM), thioredoxin M4 (TrxM4), thioredoxin X (TrxX), TrxF, and Trx CSDP decreased their expression, that is, there was an inhibition of the thioredoxin-peroxiredoxin (Trx-Prx) elimination pathway, which did not decrease oxidative damage [69]. Higher POD, SOD, and CAT activities have been reported in maize plants [70].

Transcriptomic regulation related to the oxidation-reduction process and lipid metabolism, as well as the expression of genes encoding D-3-phosphoglycerate dehydrogenase 1, glutamyl-tRNA reductase 1, fatty acid hydroperoxide lyase, ethylene-insensitive protein 2, metal tolerance protein 11, and magnesium-chelatase subunit ChlI, are altered in sugar beets under this stress [10]. These transcriptional modifications can significantly reduce crop quality; however, some red cherry tomato varieties susceptible to alkaline soils (pH 7.9) exhibit acceptable fruit quality [71].

Advertisement

4. Plant mechanism against alkalinity stress

Plants sense alkaline environments through the root, where the sulfotyrosine of the endogenous peptide RGF is deprotonated and destabilizes the RGF-RGFR receptor-SERK receptor complex, causing alterations in the root apical meristem [12, 13]. Thus, RGF is a pH sensor that reacts to H+ concentration in the medium, which could be the first signaling mechanism for plants under alkaline stress. Considering that this stress impacts plants in a multifaceted manner, plant response mechanisms are complex [10]. However, as shown in Figure 2b, the tolerance systems developed by plants can be generalized to osmoregulation, induction of the antioxidant defense system, and metabolites through transcriptomic modifications.

4.1 Osmoregulation

As previously mentioned, alkaline environments cause alterations in ionic homeostasis, and plants can extrude hydrogen ions (H+) into the rhizosphere by means of proton pumps associated with the root plasma membrane (PM H+-ATPases - HAs) acidifying the rhizosphere zone, allowing nutrient uptake [59] to achieve an adequate ionic balance. Although Na+ can enter through non-selective K+ channels, the latter (K+) tends to be translocated to the cytoplasm to generate osmotic regulation [54], thereby reducing toxicity and generating an adequate flow of water and nutrients [72].

In canola roots under alkaline conditions (Na2CO3, 40 mM), the concentrations of K+, Ca2+, and Mg2+ are elevated, similar to the K+/Na+ ratio in the leaves [58]. In alkaline soils, Mn deficiency triggers the kinase pathway (CPK21 and CPK23) and increases Ca2+ uptake mediated by the NRAMP1 transporter [73], safeguarding the cation balance. A good ionic supply and K+/Na+ regulation can improve transpiration, Chl content, PSII quantum efficiency, relative water content, and water saturation [74], in addition to photosynthetic indicators [18].

Another osmoregulation alternative can be the increase in sugar concentration, in this regard it is mentioned that in fruits of Jujube “Junzao” irrigated with a moderate alkaline solution (90 mM of NaCl:NaHCO3 (3:1)), an increase in sucrose, glucose, and fructose content was observed, in addition to an increase in sucrose phosphate synthase (SPS) and sucrose synthase (SS-I) activities [53] allowing a partial attenuation of stress and maintaining plant and fruit quality. In addition to the accumulation of soluble sugars [14], proline is associated with osmoregulatory mechanisms in plants under alkaline conditions [46].

4.2 Stimulation of antioxidant biocompounds

One of the mechanisms developed by plants to census and scavenge ROS under alkaline stress conditions is the promotion of the antioxidant system by SOD, POD, CAT, APX, glutathione reductase (GR, EC 1.6.4.2), dehydroascorbate reductase (DHAR, EC 1.8.5.1), and monodehydroascorbate reductase (MDHAR, EC 1.6.5.4) [16], as well as phenolic acids and flavonoids. Increased SOD, CAT, POD, and APX enzyme activities are in line with higher levels of SOD, CAT, POD, and APX gene expression [64]; however, this mechanism only operates when the ROS burst is below the enzyme response threshold and in line with osmotic regulation. For example, some cultivars of broom millet show tolerance to alkaline stress (80 and 40 mM) because of high antioxidant enzyme activity that reduces MDA concentration, as well as a high content of soluble osmolytes, which reduces electrolyte leakage [15]. Similarly, in soybeans grown in hydroponics (NaHCO3 and Na2CO3; Na+ from 25 to 50 mM), activation of the antioxidant enzymes SOD, POD, and APX was found [14]. In contrast, rapeseed line 2205 (B. napus) showed reduced plasma membrane damage due to osmoregulation and higher enzyme activity [75]. In the case of autotetraploid rice 93-11 T grown on saline-alkali soil, tolerance was found to be derived from soluble sugar accumulation and increased SOD and POD activity in leaves [17]. It appears that the level of polyploidization in rice may trigger alkaline tolerance by stimulating peroxide-associated genes, transcription factors [76], and possibly the antioxidant system.

Another type of compound that promotes ROS attenuation is phenolic acids, polyphenols, flavonoids, and alkaloids; in grapes, their content increases under alkaline conditions [16]. In addition to these compounds, some essential oils [25], amino acids, and organic acids (in Lycium barbarum L. berry) promote regulatory mechanisms [67]. In alkalinity-tolerant Arabidopsis thaliana, an increase in methionine-derived glucosinolates via jasmonate and salicylate has been reported [77]. Likewise, salicylic acid (SA) can increase alkalinity by promoting signaling mechanisms [17].

4.3 Transcriptomic regulation

Plants under alkaline stress can exhibit alterations in signaling and response processes through a complex transcriptomic network. In sugar beet (Beta vulgaris L.), an alkaline salt (Na2CO3) attenuating mechanism, differential expression of genes related to cutin, suberin, wax biosynthesis, sesquiterpenoid and triterpenoid biosynthesis, flavonoid biosynthesis and monoterpenoid biosynthesis, amino acid metabolism, and starch and sucrose metabolism have been documented [18]. Histologically, B. napus showed more integrated mesophyll cells as well as higher expression of genes associated with carbohydrate metabolism, photosynthetic processes, and ROS regulation [75].

Inhibition of ion uptake could be a mechanism used by plants to signal alkaline environments through the differential expression of genes related to hormone signal transduction, mitogen-activated protein kinase (MAPK) signaling, phenylpropanoid biosynthesis, and ATP-binding cassette (ABC) transport. Although phenylpropanoid and flavonoid biosynthesis pathways are exacerbated, the accumulation of metabolites varies according to the level of alkalinity in the environment [78]. Likewise, there may be positive regulation of specific root transporters (e.g., NTR1, NRT2.1, and NRT2.5) [79].

Surprisingly, alkalinity of alfalfa (Medicago truncatula) generates a redistribution of nutrients, a mechanism controlled by the differential expression of 985 genes associated with nutrient cycling, cellular amino acid metabolic, and catabolic processes of organic substances, and 1180 genes associated with the oxidoreductase complex, aerobic respiration, and ion transport. In addition to several transcription factors, MYB, WRKY, bHLH, and NAC are involved in stress resistance and senescence regulation [80]. In contrast, recombinant inbred lines (RILs) of rice showed tolerance to alkalinity due to a low Na+/K+ ratio derived from a reduction in Na+ accumulation and increased K+ accumulation in roots and shoots. Tolerance may also be due to the expression of expressed protein genes, glucan endo-1,3-beta-glucosidase precursor, F-box domain-containing proteins, double-stranded RNA-binding motif-containing proteins, aquaporins, receptor kinase-like proteins, semialdehyde, hydrogenase, and NAD-binding domain-containing protein genes [81].

Sodium-hydrogen exchangers (NHX) are fundamental regulators of cellular Na+/K+ and pH homeostasis. In soybean, the expression of the GmNHX6 gene encodes a sodium/hydrogen exchanger (NHX) in the Golgi apparatus via a single nucleotide polymorphism in the NHX6 promoter region. In A. thaliana, the expression of GmNHX6 enhances tolerance to alkaline salts derived from a high K+ content and low Na+/K+ ratio [31, 82]. In contrast, the Gγ protein seems to regulate sensitivity to alkalinity, with the alkaline tolerance 1 (AT1) locus, which encodes the γ subunit of the atypical G protein. In this context, a model for the regulation of crop alkaline tolerance mediated by the AT1 subunit of the G protein is proposed because under alkaline stress conditions, artificially or naturally removing AT1, as well as its homologs, modulates the phosphorylation of PIP2, which allows the outflow of H2O2, reducing its cellular concentration, improving tolerance, and promoting crop yield [61]. Several aquaporins can interact with AT1, including SbPIP2;1 with SbAT1 (sorghum) and OsPIP2 with 1-OsGS3, a homolog of AT1 (rice). These findings indicate a vital role for aquaporins in ROS regulation. Thus, SbAT1, which encodes the γ subunit of the G protein, inhibits aquaporin phosphorylation, thus regulating H2O2 concentration and levels under alkaline stress [83]. This information applies to monocotyledonous crops; therefore, future research should be conducted on dicotyledonous plants. Similarly, when TaAT1 was modified, higher wheat yields were observed in alkaline soils [84]. In sorghum plants tolerant to alkaline stress, ATP1 (alkaline tolerance 1) encodes the G protein by negatively modulating the phosphorylation level of PIP2 (aquaporin with H2O2 transport activity), that is, when ATP1 is eliminated, PIP2 is phosphorylated from the cell membrane, allowing greater transport of H2O2 from the cytosol to the apoplast, reducing oxidative stress [85].

Oilseed rape grown in alkalinized soil (Na2CO3; 8 g  kg−1) shows strong genetic regulation associated with ionic K+ transport as well as alterations in the amino acid metabolic pathway, leading to reduced Na+, MDA, and relative conductance (REC), increased K+ and Mg2+, and the enzymes SOD, POD, and CAT. Interestingly, the positive regulation of phenylalanine ammonia lyase (PAL) transcription factors shows greater importance in mitigating alkaline stress [86]. Similarly, higher nitrogen content, soluble sugar, proline, and higher SOD and POD activities have been reported in plants under alkaline conditions [79].

In soybean (Glycine max, cultivar Talar), exposure to alkalinity (50 mmol L−1 NaHCO3) affected the transcriptomic level associated with abiotic stress response processes, metabolic regulation, response to cellular stimuli, and signaling, highlighting 321 overexpressed genes [87]. A. thaliana is sensitive to alkalinity and induces genes associated with Fe deficiency in leaves [77]. Molecular regulation enhances the defense system and osmotic adjustment, by modifying genes associated with ubiquitin-mediated proteolysis, glycolysis/gluconeogenesis pathways, transcription factors (C2H2, bZIP, and bHLH), and protein kinases CaMK (calcium/calmodulin-dependent protein kinase), CDPK (calcium-dependent protein kinase), MAPK (mitogen-activated protein kinase), and RLK (receptor-like kinase) [88]. In wheat plants, approximately 20 genes are associated with salt stress control, abiotic stress tolerance, response to oxidative stress, and response to osmotic stress, including TraesCS2B02G276500, TraesCS2B02G504900, and TraesCS2B02G276300 [55].

Hormone content may also be regulated in response to alkalinity; cotton plants differentially express 18,230 and 11,177 genes in the root and leaf tissue, respectively. These genes are associated with plant hormone signal transduction, amino acid biosynthesis, and secondary metabolites, the latter of which is related to ROS scavenging [89]. In broomcorn millet (Panicum miliaceum L.), there is negative regulation of gibberellic acid (GA) synthesis genes and positive regulation of GA inactivation and abscisic acid (ABA) synthesis genes; therefore, resistant genotypes have higher GA/ABA ratios, in addition to having a more robust antioxidant system [90]. Similarly, polyploid rice plants under high salt–alkali stress conditions (50 mM Na2CO3, pH = 11.39) showed positive hormone regulation, particularly of genes associated with auxin (AUX), SA signal transduction (roots), lignin biosynthesis (leaves or roots), and wax biosynthesis (leaves) [17].

Some plant 3-ketoacyl-CoA synthase gene families (KCS; GBKCS3, GBKCS8, GBKCS20, GBKCS34) may be involved in hormonal signaling, defense, and stress responses that allow Gossypium barbadense to improve its fiber quality [91]. In lentil plants grown in alkaline soil, an increased number of genes related to ABA signaling (dehydrin 1, 9-cis-epoxycarotenoid dioxygenase, ABA-responsive protein 18, and BEL1-like homeodomain protein 1) as well as increased synthesis of secondary metabolites were found [20], which could promote a higher mitotic rate, allowing the organization of the internal structures (root and shoot) of the plant, avoiding the disintegration of the internal structure and the distortion of the root and shoot architecture. This environment allows for an increase in K+ accumulation, osmolytes, antioxidant enzymes, relative water content, membrane stability index, and abscisic acid. In canola, tolerance to alkaline salts is mediated by modifications in Ca2+ signaling, ABA, and ROS pathways, which together promote gene regulation, organic acid synthesis (osmoregulators), and the transcription factors bHLH, WRKY, ERF, MYB, and NAC [92]. In the cereal Chenopodium quinoa Willd, the differential expression of genes associated with ROS homeostasis, hormone signaling (auxin and ethylene), cell wall synthesis (hemicellulose), and transcription factors has been documented [93].

Metabolite accumulation is a plant response to alkaline events; however, these transcriptomic alterations can reduce plant production and fruit quality [28]. In this regard, canola plants stressed with Na2CO3 showed accumulation of L-proline, L-glutamate, L-histidine, L-phenylalanine, L-citrulline, L-tyrosine, L-saccharopine, L-tryptophan, linoleic acid, dihomo gamma linolenic acid, alpha linolenic acid, Eric acid, oleic acid, neuronic acid, sucrase, alpha, alpha trehalose, and polyol (ribitol), associated with a decrease in UDP-D-galactose, D-mannose, D-fructose, and D-glucose 6-phosphate [21]. In broom millet, differential accumulation of volatile and non-volatile metabolites has been reported, namely phenylpropanoids, flavonoids, flavones, flavanols, valine, leucine, isoleucine, arginine, proline, tryptophan, and ascorbate [78]. Under alkaline conditions (50, 100, and 150 mM NaHCO3), castor seeds promoted between 31 and 113 differential metabolites and 138–694 genes associated with amino acid, carbohydrate, lipid, and nucleotide metabolism. In this sense, castor seeds in alkaline conditions improve amino acid metabolism and the tricarboxylic acid cycle, allowing the regulation of ROS levels and promotion of osmotic balance [60].

Alkalinity induced higher expression of GmABF4, GmAAO3, GmABI5, GmNCED5, and GmNCED9 genes in soybean [62]; however, ribonucleic acid editing in rice chloroplast and mitochondria mRNA increased under alkaline stress; four types of RNA editing were notable—A-G (A-I), C-T (C-U), G-A, and T-C. Pentatricopeptide repeats (PPR) and organelle zinc-fingers (OZI) are involved in various biological functions and molecular processes. This phenomenon may contribute to amino acid changes in genes, which could lead to tolerance to this type of stress [94].

In Glycine max (L.) Merri, the gene coding for DNAJ chaperone proteins (coding for GmDNAJC7 gene) showed overexpression allowing stress tolerance; higher Chl content and relative water content [82]. In soybean, the gene associated with chaperone proteins GmDnaJA6 seems to coordinate resistance to alkaline stress; likewise, DnaJ interacting with Hsp70 and GmHsp70 generates greater tolerance to alkaline stress (100 mmol L−1 NaHCO3) [95]. For Gossypium hirsutum under alkaline conditions, differential transcriptional regulation is associated with carbohydrate metabolism and the oxidation-reduction process (stage SS48), while in roots, there is an alteration of the oxidation-reduction process (stages SS3, SS12, and SS48). Specifically, the GhSOS3/GhCBL10-SOS2 network is notoriously related to the response to this type of stress [96]. Finally, in apples, overexpression of the MdATG18a gene under alkaline conditions provides tolerance to alkaline stress by increasing biomass, photosynthetic rate, and antioxidant capacity. Similarly, an increase in glutamate promotes the derivation of γ-aminobutyric acid. Similarly, upregulation of MdATG18a increases autophagosomes, allowing the recovery of small molecules from substances damaged by alkaline stress [2].

Recently, it has been reported that long non-coding RNAs (lncRNAs) can act as competitive endogenous RNAs (ceRNAs) that participate in the response to alkaline stress. Rice plants showed alterations in differentially expressed lncRNAs, miRNAs, and mRNA (75, 34, and 1725, respectively). Likewise, the triplet NONOSAT000455-osa_miR5809b-LOC_Os11g01210 was found to be a central node in the response to saline-alkali stress. These lncRNAs act as competitive endogenous RNAs (ceRNAs) and constitute a novel set of genetic controls under saline-alkaline stress conditions [97]. It is important to note that the response to this type of stress may exhibit differences according to the level of polyploidization and expression of transcription factors (phytohormone pathways, salt resistance, signal transduction, and physiological and biochemical metabolism) [98].

Advertisement

5. Alkalinity management strategies

Several alternatives have been studied to reduce alkali stress damage in crops. The use of grafting ameliorates this stress by enhancing the desired characteristics of the rootstock and improving the Ca, Mg, P, S, Cu, Fe, and Mn contents [22]. The use of rootstocks can increase the total soluble solids and nutritional status of grapevine fruits grown in alkaline environments [51]. In contrast, the use of light-emitting diodes in monochromatic blue (460 nm), monochromatic red (660 nm), and dichromatic blue/red (1:3) light at 200 μmol m−2 s−1 enhances electron transport per reaction center and the probability of transferring electrons from the exciton into the electron transport system, promoting gas exchange in alkalinity-stressed strawberry plants [47, 99].

The efficient use of fertilizers (macro- and micronutrients) can ameliorate the negative effects of alkalinity by stimulating plant and soil enzymatic activity, improving nutrient content, microbiological activity, total Chl, protein [26, 42, 100], antioxidant activity [70], ion homeostasis (K+/Na+) [57], photosynthetic activity [24, 101, 102], and gene expression [103]. Potassium-ferrite-nanocoated diammonium phosphate [104], organic polymer composite materials [105], carbon nanodots [106], zinc nanocomplexes [25], and nano-Ca2+ [107] can ameliorate alkalinity damage. Similarly, biochar combined with daily fertigation in alkaline soil can improve cucumber yield [101]. Similarly, microorganisms are another window of action to mitigate alkalinity damage, as they increase antioxidant enzymes and modify plant gene expression [108], regulate enzyme activity and soil microbial load [27, 109], decrease pH [110], and regulate soil ionic balance [111, 112], improving nutrient availability [113]. All of these management techniques improve plant productivity and promote better fruit quality by increasing the total soluble solids, acid ratio, sugar, and phenol content [42, 114].

Some compounds, such as spermine [115], taurine [54], auxin (IAA, indole-3-acetic acid)/TIBA (2,3,5-triiodobenzoic acid, which inhibits the transport of auxins) [116], hormone-like growth regulator (AUT) [117], yeast extract and gibberellic acid (GA) [118], ABA [119], strigolactone [62], and brassinosteroid [28] improve plant development under alkaline conditions by inducing alterations at the transcriptomic level, promoting Chl content, photosynthetic rate, enzymatic activity, and reduction of prooxidant compounds. Finally, transgenic plants can show stress tolerance [29, 30, 56, 61, 83, 84, 120, 121], which is a further strategy to promote quality foods under alkaline conditions. The use of vinas and molases favors yield, total soluble solids, and total acidity in strawberries grown in calcareous soils [122]. In addition, several genotypes (VRO-112, VRO-110, Kashi Kranti, VROB-178, AE-70, and VRO-108) can tolerate alkaline conditions (pH 8–9.5), thereby increasing yield [123].

Advertisement

6. Conclusions and outlook

Alkalinity stress as an environmental stress, which is associated with high pH and high CO32− and HCO3 content, causes considerable decreases in crop quality. To attenuate stress, plants censor the edaphic environment and promote osmotic regulation (soluble sugars and proline), ionic homeostasis (K+), activation of secondary messengers, differential gene expression (associated with the antioxidant system), and accumulation of secondary metabolites. Such mechanisms may promote tolerance to alkalinity by reducing the burst and accumulation of ROS, consistent with the high efficiency of the photosynthetic process. Plant type, phenological stage, and the level and time of exposure to an alkaline environment are fundamental to the response to this stress.

The climatic alterations of the last decades, the excessive use of fertilizers, and poor management of water use can increase alkalinity stress, so it is important to understand the underlying mechanisms that plants develop to cope with this stress. Future research should be conducted to mitigate the negative effects of alkaline salts and produce high-quality crops in a sustained manner.

Advertisement

Acknowledgments

The authors are highly thankful to the researchers for their findings in this area, which served directly or indirectly in the preparation of this manuscript.

Advertisement

Conflict of interest

The authors declare no conflicts of interest.

Advertisement

Funding

The authors received no direct funding for this study.

Advertisement

Data availability

All the data generated are included in this article reference′s part.

References

  1. 1. Liu X, Liang W, Xing Li Y, Jun Li M, Quan Ma B, Hai Liu C, et al. Transcriptome analysis reveals the effects of alkali stress on root system architecture and endogenous hormones in apple rootstocks. Journal of Integrative Agriculture. 2019;18(10):2264-2271. DOI: 10.1016/S2095-3119(19)62706-1
  2. 2. Li Y, Liu C, Sun X, Liu B, Zhang X, Liang W, et al. Overexpression of MdATG18a enhances alkaline tolerance and GABA shunt in apple through increased autophagy under alkaline conditions. Tree Physiology. 2020;40(11):1509-1519. DOI: 10.1093/treephys/tpaa075
  3. 3. Wei WF, Wang C, Sun Y, Wang N, Wei LX, Yuan DY, et al. Overexpression of vacuolar proton pump ATPase (V-H+-ATPase) subunits B, C and H confers tolerance to salt and saline-alkali stresses in transgenic alfalfa (Medicago sativa L.). Journal of Integrative Agriculture. 2016;15(10):2279-2289. Available from: https://linkinghub.elsevier.com/retrieve/pii/S209531191661399-0
  4. 4. Singh K, Trivedi P, Singh G, Singh B, Patra DD. Effect of different leaf litters on carbon, nitrogen and microbial activities of sodic soils. Land Degradation and Development. 2016;27(4):1215-1226. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/ldr.2313
  5. 5. Liu J, Tang L, Gao H, Zhang M, Guo C. Enhancement of alfalfa yield and quality by plant growth-promoting rhizobacteria under saline-alkali conditions. Journal of the Science of Food and Agriculture. 2019;99(1):281-289. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/jsfa.9185
  6. 6. Jobbágy EG, Tóth T, Nosetto MD, Earman S. On the fundamental causes of high environmental alkalinity (pH ≥ 9): An assessment of its drivers and global distribution. Land Degradation and Development. 2017;28(7):1973-1981. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/ldr.2718
  7. 7. Damodaran T, Mishra VK, Jha SK, Pankaj U, Gupta G, Gopal R. Identification of rhizosphere bacterial diversity with promising salt tolerance, PGP traits and their exploitation for seed germination enhancement in sodic soil. Agricultural Research. 2019;8(1):36-43. Available from: https://link.springer.com/article/10.1007/s40003-018-0343-5
  8. 8. Rengasamy P, De Lacerda CF, Gheyi HR. Salinity, sodicity and alkalinity. In: Subsoil Constraints for Crop Production. Berlin, Heidelberg, Germany: Springer International Publishing; 2022. pp. 83-108. Available from: https://link.springer.com/chapter/10.1007/978-3-031-00317-2_4
  9. 9. Malekzadeh Shamsabad MR, Roosta HR, Esmaeilizadeh M. Responses of seven strawberry cultivars to alkalinity stress under soilless culture system. Journal of Plant Nutrition. 2021;44(2):166-180. Available from: https://www.tandfonline.com/doi/abs/10.1080/01904167.2020.1822401
  10. 10. Zou C, Liu D, Wu P, Wang Y, Gai Z, Liu L, et al. Transcriptome analysis of sugar beet (Beta vulgaris L.) in response to alkaline stress. Plant Molecular Biology. 2020;102(6):645-657. Available from: https://link.springer.com/article/10.1007/s11103-020-00971-7
  11. 11. Galinha C, Hofhuis H, Luijten M, Willemsen V, Blilou I, Heidstra R, et al. PLETHORA proteins as dose-dependent master regulators of Arabidopsis root development. Nature. 2007;449(7165):1053-1057. Available from: https://www.nature.com/articles/nature06206
  12. 12. Liu L, Song W, Huang S, Jiang K, Moriwaki Y, Wang Y, et al. Extracellular pH sensing by plant cell-surface peptide-receptor complexes. Cell. 2022;185(18):3341-3355.e13. Available from: http://www.cell.com/article/S0092867422009151/fulltext
  13. 13. Vélez-Bermúdez IC, Schmidt W. Plant strategies to mine iron from alkaline substrates. Plant and Soil. 2023;483(1–2):1-25. Available from: https://link.springer.com/article/10.1007/s11104-022-05746-1
  14. 14. Wang G, Shen W, Zhang Z, Guo S, Hu J, Feng R, et al. The effect of neutral salt and alkaline stress with the same Na+ concentration on root growth of soybean (Glycine max (L.) Merr.) seedlings. Agronomy. 2022;12(11):2708. Available from: https://www.mdpi.com/2073-4395/12/11/2708/htm
  15. 15. Ma Q, Wu C, Liang S, Yuan Y, Liu C, Liu J, et al. The alkali tolerance of broomcorn millet (Panicum miliaceum L.) at the germination and seedling stage: The case of 296 broomcorn millet genotypes. Frontiers in Plant Science. 2021;12:711429. DOI: 10.3389/fpls.2021.711429
  16. 16. Lu X, Chen G, Ma L, Zhang C, Yan H, Bao J, et al. Integrated transcriptome and metabolome analysis reveals antioxidant machinery in grapevine exposed to salt and alkali stress. Physiologia Plantarum. 2023;175(3):e13950. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/ppl.13950
  17. 17. Zhang C, Meng W, Wang Y, Zhou Y, Wang S, Qi F, et al. Comparative analysis of physiological, hormonal and transcriptomic responses reveal mechanisms of saline-alkali tolerance in autotetraploid rice (Oryza sativa L.). International Journal of Molecular Sciences. 2022;23(24):16146. Available from: https://www.mdpi.com/1422-0067/23/24/16146
  18. 18. Geng G, Li R, Stevanato P, Lv C, Lu Z, Yu L, et al. Physiological and transcriptome analysis of sugar beet reveals different mechanisms of response to neutral salt and alkaline salt stresses. Frontiers in Plant Science. 2020;11:571864. DOI: 10.3389/fpls.2020.571864
  19. 19. Ma Q, Wang H, Wu E, Yuan Y, Feng Y, Zhao L, et al. Comprehensive physiological, transcriptomic, and metabolomic analysis of the response of Panicum miliaceum L. roots to alkaline stress. Land Degradation and Development. 2023;34(10):2912-2930. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/ldr.4656
  20. 20. Singh D, Singh CK, Taunk J, Gaikwad K, Singh V, Sanwal SK, et al. Linking genome wide RNA sequencing with physio-biochemical and cytological responses to catalogue key genes and metabolic pathways for alkalinity stress tolerance in lentil (Lens culinaris Medikus). BMC Plant Biology. 2022;22(1):1-17. Available from: https://bmcplantbiol.biomedcentral.com/articles/10.1186/s12870-022-03489-w
  21. 21. Wang W, Pang J, Zhang F, Sun L, Yang L, Zhao Y, et al. Integrated transcriptomics and metabolomics analysis to characterize alkali stress responses in canola (Brassica napus L.). Plant Physiology and Biochemistry. 2021;166:605-620. DOI: 10.1016/j.plaphy.2021.06.021
  22. 22. Ulas F, Aydın A, Ulas A, Yetisir H. The efficacy of grafting on alkali stressed watermelon cultivars under hydroponic conditions. Gesunde Pflanzen. 2021;73(3):345-357. Available from: https://link.springer.com/article/10.1007/s10343-021-00559-1
  23. 23. Machado RMA, Serralheiro RP. Soil salinity: Effect on vegetable crop growth. Management practices to prevent and mitigate soil salinization. Horticulturae. 2017;3(2):30. Available from: https://www.mdpi.com/2311-7524/3/2/30/htm
  24. 24. Nandita K, Kundu M, Rakshit R, Nahakpam S. Effects of foliar application of micronutrients on growth, yield and quality of sweet orange (Citrus sinensis (l.) Osbeck). Bangladesh Journal of Botany. 2022;51(1):57-63. Available from: https://www.banglajol.info/index.php/BJB/article/view/58821
  25. 25. Ghodrati-Tazangi MJ, Samani RB, Tavallali V, Alizadeh A, Honarvar M. Responses of Hyssopus officinalis to bicarbonate stress and foliar application of green synthesized zinc nano-complex formed on Medicago sativa extract. Scientia Horticulturae. 2023;319:112197. DOI: 10.1016/j.scienta.2023.1121197
  26. 26. He K, Xu Y, He G, Zhao X, Wang C, Li S, et al. Combined application of acidic biochar and fertilizer synergistically enhances miscanthus productivity in coastal saline-alkaline soil. Science of the Total Environment. 2023;893:164811. DOI: 10.1016/j.scitotenv.2023.164811
  27. 27. Mishra SK, Misra S, Dixit VK, Kar S, Chauhan PS. Ochrobactrum sp. NBRISH6 inoculation enhances Zea mays productivity, mitigating soil alkalinity and plant immune response. Current Microbiology. 2023;80(10):1-13. Available from: https://link.springer.com/article/10.1007/s00284-023-03441-7
  28. 28. Ma Q, Wu E, Wang H, Feng Y, Zhao L, Feng B. Comparative transcriptomic and metabolomic study reveal that exogenous 24-epiandrosterone mitigate alkaline stress in broomcorn millet (Panicum miliaceum L.) via regulating photosynthesis and antioxidant capacity. GCB Bioenergy. 2023;15(4):494-507. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/gcbb.13032
  29. 29. Liu X, Xie X, Zheng C, Wei L, Li X, Jin Y, et al. RNAi-mediated suppression of the abscisic acid catabolism gene OsABA8ox1 increases abscisic acid content and tolerance to saline–alkaline stress in rice (Oryza sativa L.). Crop Journal. 2022;10(2):354-367. DOI: 10.1016/j.cj.2021.06.011
  30. 30. Wang F, Xu H, Zhang L, Shi Y, Song Y, Wang X, et al. The lipoxygenase OsLOX10 affects seed longevity and resistance to saline-alkaline stress during rice seedlings. Plant Molecular Biology. 2023;111(4–5):415-428. Available from: https://link.springer.com/article/10.1007/s11103-023-01334-8
  31. 31. Jin T, An J, Xu H, Chen J, Pan L, Zhao R, et al. A soybean sodium/hydrogen exchanger GmNHX6 confers plant alkaline salt tolerance by regulating Na+/K+ homeostasis. Frontiers in Plant Science. 2022;13:938635. DOI: 10.3389/fpls.2022.938635
  32. 32. Zhao K, Wu Y. Effects of Zn deficiency and bicarbonate on the growth and photosynthetic characteristics of four plant species. PLoS One. 2017;12(1):e0169812. Available from: https://journals.plos.org/plosone/article?id=10.1371/journal.pone.0169812
  33. 33. Tsai HH, Schmidt W. The enigma of environmental pH sensing in plants. Nature Plants. 2021;7(2):106-115. Available from: https://www.nature.com/articles/s41477-020-00831-8
  34. 34. Sposito G. The chemical composition of soils. In: Sposito G, editor. The Chemistry of Soils. New York, United States of America: Oxford University Press Inc.; 2008. pp. 2-24. Available from: https://books.google.com/books/about/The_Chemistry_of_Soils.html?hl=es&id=XCJnDAAAQBAJ
  35. 35. Tavakkoli E, Rengasamy P, Smith E, Mcdonald GK. The effect of cation-anion interactions on soil pH and solubility of organic carbon. European Journal of Soil Science. 2015;66(6):1054-1062. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/ejss.12294
  36. 36. Boyd CE, Tucker CS, Somridhivej B. Alkalinity and hardness: Critical but elusive concepts in aquaculture. Journal of the World Aquaculture Society. 2016;47(1):6-41. Available from: https://www.researchgate.net/publication/292185663
  37. 37. Roosta HR. Interaction between water alkalinity and nutrient solution pH on the vegetative growth, chlorophyll fluorescence and leaf magnesium, iron, manganese, and zinc concentrations in lettuce. Journal of Plant Nutrition. 2011;34(5):717-731. Available from: https://www.tandfonline.com/doi/abs/10.1080/01904167.2011.540687
  38. 38. Jiao X, Li Y, Zhang X, Liu C, Liang W, Li C, et al. Exogenous dopamine application promotes alkali tolerance of apple seedlings. Plants. 2019;8(12):580. Available from: https://www.mdpi.com/2223-7747/8/12/580/htm
  39. 39. Yang CW, Wang P, Li CY, Shi DC, Wang DL. Comparison of effects of salt and alkali stresses on the growth and photosynthesis of wheat. Photosynthetica. 2008;46(1):107-114. Available from: https://link.springer.com/article/10.1007/s11099-008-0018-8
  40. 40. Roosta HR, Manzari Tavakkoli M, Hamidpour M. Comparison of different soilless media for growing gerbera under alkalinity stress condition. Journal of Plant Nutrition. 2016;39(8):1063-1073. Available from: https://www.tandfonline.com/doi/abs/10.1080/01904167.2015.1109112
  41. 41. Brautigan DJ, Rengasamy P, Chittleborough DJ. Amelioration of alkaline phytotoxicity by lowering soil pH. Crop and Pasture Science. 2014;65(12):1278-1287. Available from: https://www.publish.csiro.au/cp/CP13435
  42. 42. Nandita K, Kundu M, Nahakpam S, Rakshit R. Micronutrients spray to combat granulation and improve fruit quality of sweet orange [Citrus sinensis (L.) Osbeck] cv. Mosambi under nontraditional citrus growing track. Journal of Plant Nutrition. 2023;46(17):4207-4223. Available from: https://www.tandfonline.com/doi/abs/10.1080/01904167.2023.2224384
  43. 43. Dawood MFA, Sohag AAM, Tahjib-Ul-Arif M, Abdel Latef AAH. Hydrogen sulfide priming can enhance the tolerance of artichoke seedlings to individual and combined saline-alkaline and aniline stresses. Plant Physiology and Biochemistry. 2021;159:347-362. DOI: 10.1016/j.plaphy.2020.12.034
  44. 44. Hitti Y, MacPherson S, Lefsrud M. Separate effects of sodium on germination in saline–sodic and alkaline forms at different concentrations. Plants. 2023;12(6):1234. Available from: https://www.mdpi.com/2223-7747/12/6/1234/htm
  45. 45. Liu D, Ma Y, Rui M, Lv X, Chen R, Chen X, et al. Is high ph the key factor of alkali stress on plant growth and physiology? A case study with wheat (Triticum aestivum L.) seedlings. Agronomy. 2022;12(8):1820. Available from: https://www.mdpi.com/2073-4395/12/8/1820/htm
  46. 46. Wang Y, Wang J, Guo D, Zhang H, Che Y, Li Y, et al. Physiological and comparative transcriptome analysis of leaf response and physiological adaption to saline alkali stress across pH values in alfalfa (Medicago sativa). Plant Physiology and Biochemistry. 2021;167:140-152. DOI: 10.1016/j.plaphy.2021.07.040
  47. 47. Malekzadeh Shamsabad MR, Esmaeilizadeh M, Roosta HR, Dąbrowski P, Telesiński A, Kalaji HM. Supplemental light application can improve the growth and development of strawberry plants under salinity and alkalinity stress conditions. Scientific Reports. 2022;12(1):1-13. Available from: https://www.nature.com/articles/s41598-022-12925-8
  48. 48. Gill PK. Effect of various abiotic stresses on the growth , soluble sugars and water relations of Sorghum seedlings grown in light and darkness. Bulgarian Journal of Plant Physiology. 2014;27(1–2):72-84
  49. 49. Deng C, Zhang G, Pan X. Photosynthetic responses in reed (Phragmites australis (CAV.) TRIN. ex Steud.) seedlings induced by different salinity-alkalinity and nitrogen levels. Journal of Agricultural Science and Technology. 2011;13(5):687-699. Available from: http://jast.modares.ac.ir/article-23-1590-en.html
  50. 50. Yang CW, Xu HH, Wang LL, Liu J, Shi DC, Wang DL. Comparative effects of salt-stress and alkali-stress on the growth, photosynthesis, solute accumulation, and ion balance of barley plants. Photosynthetica. 2009;47(1):79-86. Available from: https://link.springer.com/article/10.1007/s11099-009-0013-8
  51. 51. Scheiner JJ, Labay A, Kamas J. Rootstocks improve blanc Du Bois vine performance and fruit quality on alkaline soil. Catalyst: Discovery into Practice. 2020;4(2):63-73. Available from: https://www.ajevonline.org/content/4/Suppl_2/63
  52. 52. Roura S, Davidovich L, Del Valle C. Quality loss in minimally processed swiss chard related to amount of damaged area. LWT. 2000;33(1):53-59. DOI: 10.1006/fstl.1999.0615
  53. 53. Wang Y, Feng Y, Yan M, Yu J, Zhou X, Bao J, et al. Effect of saline–alkali stress on sugar metabolism of jujube fruit. Horticulturae. 2022;8(6):474. Available from: https://www.mdpi.com/2311-7524/8/6/474/htm
  54. 54. Rasheed R, Ashraf MA, Ahmad SJN, Parveen N, Hussain I, Bashir R. Taurine regulates ROS metabolism, osmotic adjustment, and nutrient uptake to lessen the effects of alkaline stress on Trifolium alexandrinum L. plants. South African Journal of Botany. 2022;148:482-498. DOI: 10.1016/j.sajb.2022.05.023
  55. 55. Mourad AMI, Farghly KA, Börner A, Moursi YS. Candidate genes controlling alkaline-saline tolerance in two different growing stages of wheat life cycle. Plant and Soil. 2023;493:283-307. Available from: https://link.springer.com/article/10.1007/s11104-023-06232-y
  56. 56. Navarro-León E, Grazioso A, Atero-Calvo S, Rios JJ, Esposito S, Blasco B. Evaluation of the alkalinity stress tolerance of three Brassica rapa CAX1 TILLING mutants. Plant Physiology and Biochemistry. 2023;198:107712. DOI: 10.1016/j.plaphy.2023.107712
  57. 57. Cao X, Sun L, Wang W, Zhang F. Exogenous calcium application mediates K+ and Na+ homeostasis of different salt-tolerant rapeseed varieties under NaHCO3 stress. Plant Growth Regulation. 2023;102:367-378. Available from: https://link.springer.com/article/10.1007/s10725-023-01066-1
  58. 58. Wang W, Zhang F, Sun L, Yang L, Yang Y, Wang Y, et al. Alkaline salt inhibits seed germination and seedling growth of canola more than neutral salt. Frontiers in Plant Science. 2022;13:814755. DOI: 10.3389/fpls.2022.814755
  59. 59. Chen L, Zhao R, Yu J, Gu J, Li Y, Chen W, et al. Functional analysis of plasma membrane H+-ATPases in response to alkaline stress in blueberry. Scientia Horticulturae. 2022;306:111453. DOI: 10.1016/j.scienta.2022.111453
  60. 60. Han P, Li S, Yao K, Geng H, Liu J, Wang Y, et al. Integrated metabolomic and transcriptomic strategies to reveal adaptive mechanisms in castor plant during germination stage under alkali stress. Environmental and Experimental Botany. 2022;203:105031. DOI: 10.1016/j.envexpbot.2022.105031
  61. 61. Zhang L, Qian Q, Song S. Combating alkaline stress with alkaline tolerance 1, which encodes a conserved Gγ protein in multiple crops. Plant Communications. 2023;4(3):100603. DOI: 10.1016/j.xplc.2023.100603
  62. 62. Zaib-un-Nisa MX, Anwar S, Chen C, Jin X, Yu L, et al. Strigolactone enhances alkaline tolerance in soybean seeds germination by altering expression profiles of ABA biosynthetic and signaling genes. Journal of Plant Biology. 2023;66(4):373-381. Available from: https://link.springer.com/article/10.1007/s12374-022-09357-2
  63. 63. Li J, Jin Y, Liu Z, Ba T, Wang W, Xu S. Morphology and metabolism of storage substances contribution to alkali stress responses in two contrasting barley cultivars during germination stage. Journal of Seed Science. 2023;45:e202345030. Available from: https://www.scielo.br/j/jss/a/nQKBBGLq8Lq5zbFrRzTsZdP/?lang=en
  64. 64. Fatima A, Hussain S, Hussain S, Ali B, Ashraf U, Zulfiqar U, et al. Differential morphophysiological, biochemical, and molecular responses of maize hybrids to salinity and alkalinity stresses. Agronomy. 2021;11(6):1150. Available from: https://www.mdpi.com/2073-4395/11/6/1150/htm
  65. 65. Xu X, Zhang J, Yan B, Wei Y, Ge S, Li J, et al. The adjustment of membrane lipid metabolism pathways in maize roots under saline–alkaline stress. Frontiers in Plant Science. 2021;12:635327. DOI: 10.3389/fpls.2021.635327
  66. 66. Wu GQ, Li H, Zhu YH, Li SJ. Comparative physiological response of sainfoin (Onobrychis viciaefolia) seedlings to alkaline and saline-alkaline stress. The Journal of Animal and Plant Sciences. 2020;31(4):2021. DOI: 10.36899/JAPS.2021.4.0299
  67. 67. Liang X, Wang Y, Li Y, An W, He X, Chen Y, et al. Widely-targeted metabolic profiling in Lycium barbarum fruits under salt-alkaline stress uncovers mechanism of salinity tolerance. Molecules. 2022;27(5):1564. Available from: https://www.mdpi.com/1420-3049/27/5/1564/htm
  68. 68. Zhang H, Liu XL, Zhang RX, Yuan HY, Wang MM, Yang HY, et al. Root damage under alkaline stress is associated with reactive oxygen species accumulation in rice (Oryza sativa L.). Frontiers in Plant Science. 2017;8:288138. DOI: 10.3389/fpls.2017.01580
  69. 69. Huihui Z, Xin L, Yupeng G, Mabo L, Yue W, Meijun A, et al. Physiological and proteomic responses of reactive oxygen species metabolism and antioxidant machinery in mulberry (Morus alba L.) seedling leaves to NaCl and NaHCO3 stress. Ecotoxicology and Environmental Safety. 2020;193:110259. DOI: 10.1016/j.ecoenv.2020.110259
  70. 70. Sriramachandrasekharan MV, Priya NG, Manivannan R, Prakash M. Ameliorative role of silicon on osmoprotectants, antioxidant enzymes and growth of maize grown under alkaline stress. Silicon. 2022;14(12):6577-6585. Available from: https://link.springer.com/article/10.1007/s12633-021-01434-4
  71. 71. Todevska D, Kovacevik B, Kostadinvik-Velickovska S, Markova-Ruzdik N, Mihajlov L. Evaluation of the quality traits of red cherry tomato varieties grown in alkaline soil. Zemdirbyste. 2023;110(2):157-164. DOI: 10.13080/z-a.2023.110.019
  72. 72. Andrioli RJ. Adaptive mechanisms of tall wheatgrass to salinity and alkalinity stress. Grass and Forage Science. 2023;78(1):23-36. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/gfs.12585
  73. 73. Fu D, Zhang Z, Wallrad L, Wang Z, Höller S, Ju CF, et al. Ca2+-dependent phosphorylation of NRAMP1 by CPK21 and CPK23 facilitates manganese uptake and homeostasis in Arabidopsis. Proceedings of the National Academy of Sciences of the United States of America. 2022;119(40):e2204574119. Available from: https://www.pnas.org/doi/abs/10.1073/pnas.2204574119
  74. 74. Müller B, Sterner VAC, Papp L, May Z, Orlóci L, Gyuricza C, et al. Alkaline salt tolerance of the biomass plant Arundo donax. Agronomy. 2022;12(7):1589. Available from: https://www.mdpi.com/2073-4395/12/7/1589/htm
  75. 75. Xu Y, Tao S, Zhu Y, Zhang Q, Li P, Wang H, et al. Identification of alkaline salt tolerance genes in Brassica napus L. by transcriptome analysis. Genes. 2022;13(8):1493. Available from: https://www.mdpi.com/2073-4425/13/8/1493/htm
  76. 76. Wang N, Fan X, Lin Y, Li Z, Wang Y, Zhou Y, et al. Alkaline stress induces different physiological, hormonal and gene expression responses in diploid and autotetraploid Rice. International Journal of Molecular Sciences. 2022;23(10):5561. Available from: https://www.mdpi.com/1422-0067/23/10/5561/htm
  77. 77. Pérez-Martín L, Busoms S, Tolrà R, Poschenrieder C. Transcriptomics reveals fast changes in salicylate and jasmonate signaling pathways in shoots of carbonate-tolerant Arabidopsis thaliana under bicarbonate exposure. International Journal of Molecular Sciences. 2021;22(3):1-24. Available from: https://www.mdpi.com/1422-0067/22/3/1226/htm
  78. 78. Ma Q, Wang H, Wu E, Zhang H, Feng Y, Feng B. Widely targeted metabolomic analysis revealed the effects of alkaline stress on nonvolatile and volatile metabolites in broomcorn millet grains. Food Research International. 2023;171:113066. DOI: 10.1016/j.foodres.2023.113066
  79. 79. Geng G, Wang G, Stevanato P, Lv C, Wang Q, Yu L, et al. Physiological and proteomic analysis of different molecular mechanisms of sugar beet response to acidic and alkaline pH environment. Frontiers in Plant Science. 2021;12:682799. DOI: 10.3389/fpls.2021.682799
  80. 80. Dong S, Pang W, Liu Z, Li H, Zhang K, Cong L, et al. Transcriptome analysis of leaf senescence regulation under alkaline stress in Medicago truncatula. Frontiers in Plant Science. 2022;13:881456. DOI: 10.3389/fpls.2022.881456
  81. 81. Singh L, Coronejo S, Pruthi R, Chapagain S, Subudhi PK. Integration of QTL mapping and whole genome sequencing identifies candidate genes for alkalinity tolerance in Rice (Oryza sativa). International Journal of Molecular Sciences. 2022;23(19):11791. Available from: https://www.mdpi.com/1422-0067/23/19/11791/htm
  82. 82. Jin T, Shan Z, Zhou S, Yang Q, Gai J, Li Y. GmDNAJC7 from soybean is involved in plant tolerance to alkaline-salt, salt, and drought stresses. Agronomy. 2022;12(6):1419. Available from: https://www.mdpi.com/2073-4395/12/6/1419/htm
  83. 83. Zhang H, Yu F, Xie P, Sun S, Qiao X, Tang S, et al. A Gγ protein regulates alkaline sensitivity in crops. Science. 2023;379(6638):eade8416. Available from: https://www.science.org/doi/10.1126/science.ade8416
  84. 84. Sun W, Zhang H, Yang S, Liu L, Xie P, Li J, et al. Genetic modification of Gγ subunit AT1 enhances salt-alkali tolerance in main graminaceous crops. National Science Review. 2023;10(6):nwad075. DOI: 10.1093/nsr/nwad075
  85. 85. Wang P, Ma JF. Knockout of a gene encoding a Gγ protein boosts alkaline tolerance in cereal crops. aBIOTECH. 2023;4(2):180-183. Available from: https://link.springer.com/article/10.1007/s42994-023-00106-8
  86. 86. Li Z, An M, Hong D, Chang D, Wang K, Fan H. Transcriptomic and metabolomic analyses reveal the differential regulatory mechanisms of compound material on the responses of Brassica campestris to saline and alkaline stresses. Frontiers in Plant Science. 2022;13:820540. DOI: 10.3389/fpls.2022.820540
  87. 87. Seraj RGM, Mohammadi M, Shahbazi M, Tohidfar M. Co-expression network analysis of soybean transcriptome identify hub genes under saline-alkaline and water deficit stress. Iranian Journal of Biotechnology. 2022;20(4):110-121. Available from: https://www.ijbiotech.com/article_157961.html
  88. 88. Zhang P, Duo T, Wang F, Zhang X, Yang Z, Hu G. De novo transcriptome in roots of switchgrass (Panicum virgatum L.) reveals gene expression dynamic and act network under alkaline salt stress. BMC Genomics. 2021;22(1):82. Available from: https://bmcgenomics.biomedcentral.com/articles/10.1186/s12864-021-07368-w
  89. 89. Fan Y, Lu X, Chen X, Wang J, Wang D, Wang S, et al. Cotton transcriptome analysis reveals novel biological pathways that eliminate reactive oxygen species (ROS) under sodium bicarbonate (NaHCO3) alkaline stress. Genomics. 2021;113(3):1157-1169. DOI: 10.1016/j.ygeno.2021.02.022
  90. 90. Ma Q, Yuan Y, Wu E, Wang H, Dang K, Feng Y, et al. Endogenous bioactive gibberellin/abscisic acids and enzyme activity synergistically promote the phytoremediation of alkaline soil by broomcorn millet (Panicum miliaceum L.). Journal of Environmental Management. 2022;305:114362. DOI: 10.1016/j.jenvman.2021.114362
  91. 91. Rui C, Chen X, Xu N, Wang J, Zhang H, Li S, et al. Identification and structure analysis of KCS family genes suggest their reponding to regulate fiber development in long-staple cotton under salt-alkaline stress. Frontiers in Genetics. 2022;13:812449. DOI: 10.3389/fgene.2022.812449
  92. 92. Wang W, Pang J, Zhang F, Sun L, Yang L, Fu T, et al. Transcriptome analysis reveals key molecular pathways in response to alkaline salt stress in canola (Brassica napus L.) roots. Journal of Plant Growth Regulation. 2023;42(5):3111-3127. Available from: https://link.springer.com/article/10.1007/s00344-022-10774-3
  93. 93. Han H, Qu Y, Wang Y, Zhang Z, Geng Y, Li Y, et al. Transcriptome and small rna sequencing reveals the basis of response to salinity, alkalinity and hypertonia in quinoa (Chenopodium quinoa Willd.). International Journal of Molecular Sciences. 2023;24(14):11789. Available from: https://www.mdpi.com/1422-0067/24/14/11789/htm
  94. 94. Rehman O, Uzair M, Chao H, Khan MR, Chen M. Decoding RNA editing sites through transcriptome analysis in rice under alkaline stress. Frontiers in Plant Science. 2022;13:892729. DOI: 10.3389/fpls.2022.892729
  95. 95. Zhang B, Liu Z, Zhou R, Cheng P, Li H, Wang Z, et al. Genome-wide analysis of soybean DnaJA-family genes and functional characterization of GmDnaJA6 responses to saline and alkaline stress. Crop Journal. 2023;11(4):1230-1241. DOI: 10.1016/j.cj.2023.06.005
  96. 96. Xu Y, Magwanga RO, Yang X, Jin D, Cai X, Hou Y, et al. Genetic regulatory networks for salt-alkali stress in Gossypium hirsutum with differing morphological characteristics. BMC Genomics. 2020;21(1):1-19. Available from: https://bmcgenomics.biomedcentral.com/articles/10.1186/s12864-019-6375-9
  97. 97. Rehman OU, Uzair M, Farooq MS, Saleem B, Attacha S, Attia KA, et al. Comprehensive insights into the regulatory mechanisms of lncRNA in alkaline-salt stress tolerance in rice. Molecular Biology Reports. 2023;50(9):7381-7392. Available from: https://link.springer.com/article/10.1007/s11033-023-08648-2
  98. 98. Wang N, Wang Y, Wang C, Leng Z, Qi F, Wang S, et al. Evaluating the differential response of transcription factors in diploid versus autotetraploid Rice leaves subjected to diverse saline–alkali stresses. Genes. 2023;14(6):1151. Available from: https://www.mdpi.com/2073-4425/14/6/1151/htm
  99. 99. Shamsabad MRM, Esmaeilizadeh M, Roosta HR, Dehghani MR, Dąbrowski P, Kalaji HM. The effect of supplementary light on the photosynthetic apparatus of strawberry plants under salinity and alkalinity stress. Scientific Reports. 2022;12(1):13257. Available from: https://www.nature.com/articles/s41598-022-17377-8
  100. 100. Xudong G, Fengju Z, Teng W, Xiaowei X, Xiaohui J, Xing X. Effects of nitrogen and phosphorus addition on growth and leaf nitrogen metabolism of alfalfa in alkaline soil in Yinchuan plain of Hetao Basin. PeerJ. 2022;10:e13261. Available from: https://peerj.com/articles/13261
  101. 101. Zhang P, Yang F, Zhang H, Liu L, Liu X, Chen J, et al. Beneficial effects of biochar-based organic fertilizer on nitrogen assimilation, antioxidant capacities, and photosynthesis of sugar beet (Beta vulgaris L.) under saline-alkaline stress. Agronomy. 2020;10(10):1562. Available from: https://www.mdpi.com/2073-4395/10/10/1562/htm
  102. 102. Rehman B, Hussain S, Zulfiqar A. Zinc sulfate biofortification enhances physio-biochemical attributes and oxidative stress tolerance in rice varieties grown in zinc deficient alkaline soil. South African Journal of Botany. 2023;162:271-281. DOI: 10.1016/j.sajb.2023.09.023
  103. 103. Danaeifar A, Khaleghi E, Zivdar S, Mehdikhanlou K. Physiological, biochemical, and gene expression of Sour Orange (Citrus aurantium L.) to Iron(II)-Arginine Chelate under salinity, alkalinity, and salt–alkali combined stresses. Scientia Horticulturae. 2023;319:112146. DOI: 10.1016/j.scienta.2023.112146
  104. 104. Saleem I, Maqsood MA, Aziz T, Bhatti IA, Jabbar A. Potassium ferrite nano coated dap fertilizer improves wheat (Triticum aestivum L.) growth and yield under alkaline calcareous soil conditions. Pakistan Journal of Agricultural Sciences. 2021;58(2):485-492. DOI: 10.21162/PAKJAS/21.939
  105. 105. Wang X, An M, Wang K, Fan H, Shi J, Chen K. Effects of organic polymer compound material on K+ and Na+ distribution and physiological characteristics of cotton under saline and alkaline stresses. Frontiers in Plant Science. 2021;12:636536. DOI: 10.3389/fpls.2021.636536
  106. 106. Chen Q, Cao X, Li Y, Sun Q, Dai L, Li J, et al. Functional carbon nanodots improve soil quality and tomato tolerance in saline-alkali soils. Science of the Total Environment. 2022;830:154817. DOI: 10.1016/j.scitotenv.2022.154817
  107. 107. El-temsah ME, Abd-Elkrem YM, El-Gabry YA, Abdelkader MA, Morsi NAA, Taha NM, et al. Response of diverse peanut cultivars to nano and conventional calcium forms under alkaline sandy soil. Plants. 2023;12(14):2598. Available from: https://www.mdpi.com/2223-7747/12/14/2598/htm
  108. 108. Li F, Wu Z, Zuo S, Fan L, Wei Z, Ma L, et al. Application of enzymatic hydrolysate of Ulva clathrata as biostimulant improved physiological and metabolic adaptation to salt-alkaline stress in wheat. Journal of Applied Phycology. 2022;34(3):1779-1789. Available from: https://link.springer.com/article/10.1007/s10811-022-02684-4
  109. 109. Dixit VK, Misra S, Mishra SK, Tewari SK, Joshi N, Chauhan PS. Characterization of plant growth-promoting alkalotolerant Alcaligenes and bacillus strains for mitigating the alkaline stress in Zea mays. Antonie van Leeuwenhoek, International Journal of General and Molecular Microbiology. 2020;113(7):889-905. Available from: https://link.springer.com/article/10.1007/s10482-020-01399-1
  110. 110. Youssef SM, El-Serafy RS, Ghanem KZ, Elhakem A, Abdel Aal AA. Foliar spray or soil drench: Microalgae application impacts on soil microbiology, morpho-physiological and biochemical responses, oil and fatty acid profiles of chia plants under alkaline stress. Biology. 2022;11(12):1844. Available from: https://www.mdpi.com/2079-7737/11/12/1844/htm
  111. 111. Fu J, Xiao Y, Wang YF, Liu ZH, Zhang YF, Yang KJ. Trichoderma asperellum alters fungal community composition in saline–alkaline soil maize rhizospheres. Soil Science Society of America Journal. 2021;85(4):1091-1104. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/saj2.20245
  112. 112. Cabral-Miramontes JP, Olmedo-Monfil V, Lara-Banda M, Zúñiga-Romo ER, Aréchiga-Carvajal ET. Promotion of plant growth in arid zones by selected Trichoderma spp. strains with adaptation plasticity to alkaline pH. Biology. 2022;11(8):1206. Available from: https://www.mdpi.com/2079-7737/11/8/1206/htm
  113. 113. Farghaly FA, Nafady NA, Abdel-Wahab DA. The efficiency of arbuscular mycorrhiza in increasing tolerance of Triticum aestivum L. to alkaline stress. BMC Plant Biology. 2022;22(1):1-17. Available from: https://bmcplantbiol.biomedcentral.com/articles/10.1186/s12870-022-03790-8
  114. 114. İpek M, Eşitken A. Effects of rhizobacteria on plant growth and fruit quality of blackberry in alkaline soil. Selcuk Journal of Agricultural and Food Sciences. 2022;36(3):387-392. DOI: 10.15316/SJAFS.2022.050
  115. 115. Xu J, Yang J, Xu Z, Zhao D, Hu X. Exogenous spermine-induced expression of SlSPMS gene improves salinity–alkalinity stress tolerance by regulating the antioxidant enzyme system and ion homeostasis in tomato. Plant Physiology and Biochemistry. 2020;157:79-92. DOI: 10.1016/j.plaphy.2020.09.033
  116. 116. Ma C, Yuan S, Xie B, Li Q, Wang Q, Shao M. IAA plays an important role in alkaline stress tolerance by modulating root development and ROS detoxifying systems in rice plants. International Journal of Molecular Sciences. 2022;23(23):14817. Available from: https://www.mdpi.com/1422-0067/23/23/14817/htm
  117. 117. Chang D, Liu H, An M, Hong D, Fan H, Wang K, et al. Integrated transcriptomic and metabolomic analysis of the mechanism of foliar application of hormone-type growth regulator in the improvement of grape (Vitis vinifera L.) coloration in saline-alkaline soil. Plants. 2022;11(16):2115. Available from: https://www.mdpi.com/2223-7747/11/16/2115/htm
  118. 118. Youssef SM, Abdella EMM, Al-Elwany OA, Alshallash KS, Alharbi K, Ibrahim MTS, et al. Integrative application of foliar yeast extract and gibberellic acid improves morpho-physiological responses and nutrient uptake of Solidago virgaurea plant in alkaline soil. Life. 2022;12(9):1405. Available from: https://www.mdpi.com/2075-1729/12/9/1405/htm
  119. 119. Xu Z, Wang J, Zhen W, Sun T, Hu X. Abscisic acid alleviates harmful effect of saline–alkaline stress on tomato seedlings. Plant Physiology and Biochemistry. 2022;175:58-67. DOI: 10.1016/j.plaphy.2022.01.018
  120. 120. Qu D, Show PL, Miao X. Transcription factor chbzip1 from alkaliphilic microalgae Chlorella sp. BLD enhancing alkaline tolerance in transgenic Arabidopsis thaliana. International Journal of Molecular Sciences. 2021;22(5):1-16. Available from: https://www.mdpi.com/1422-0067/22/5/2387/htm
  121. 121. Wei L, Zhang R, Zhang M, Xia G, Liu S. Functional analysis of long non-coding RNAs involved in alkaline stress responses in wheat. Journal of Experimental Botany. 2022;73(16):5698-5714. DOI: 10.1093/jxb/erac211
  122. 122. Cankurt K, İpek M. The effects of some organic compounds on yield and fruit quality in albion strawberry (Fragaria x ananassa Duch) Cultivar. Selcuk Journal of Agricultural and Food Sciences. 2023;37(1):19-124. DOI: 10.15316/SJAFS.2023.003
  123. 123. Sanwal SK, Mann A, Kesh H, Kaur G, Kumar R, Rai AK. Genotype environment interaction analysis for fruit yield in okra (Abelmoschus esculentus L.) under alkaline environments. Indian Journal of Genetics and Plant Breeding. 2021;81(1):101-110. DOI: 10.31742/IJGPB.81.1.11

Written By

Fabián Pérez-Labrada, José Luis Espinoza-Acosta, Daniel Bárcenas-Santana, Elizabeth García-León and Mari Carmen López-Pérez

Submitted: 22 December 2023 Reviewed: 21 February 2024 Published: 24 April 2024