Next Article in Journal
Time Course of Priming Effect of TF Inducers on Synergistic TF Expression and Intra-Cellular Gap Formation of Human Vascular Endothelial Cells via the Extrinsic Coagulation Cascade
Previous Article in Journal
A Variant in TBCD Associated with Motoneuronopathy and Corpus Callosum Hypoplasia: A Case Report
Previous Article in Special Issue
Salt Stress—Regulation of Root Water Uptake in a Whole-Plant and Diurnal Context
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Integrated Transcriptomic and Metabolomic Analyses Uncover the Differential Mechanism in Saline–Alkaline Tolerance between Indica and Japonica Rice at the Seedling Stage

1
State Key Laboratory of Hybrid Rice, Key Laboratory for Research and Utilization of Heterosis in Indica Rice, Ministry of Agriculture, College of Life Sciences, Wuhan University, Wuhan 430072, China
2
Lushan Botanical Garden, Chinese Academy of Sciences, Jiujiang 332900, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(15), 12387; https://doi.org/10.3390/ijms241512387
Submission received: 8 June 2023 / Revised: 26 July 2023 / Accepted: 29 July 2023 / Published: 3 August 2023
(This article belongs to the Special Issue Molecular Regulation of Salinity Tolerance in Plants)

Abstract

:
Saline–alkaline stress is one of the major damages that severely affects rice (Oryza sativa L.) growth and grain yield; however, the mechanism of the tolerance remains largely unknown in rice. Herein, we comparatively investigated the transcriptome and metabolome of two contrasting rice subspecies genotypes, Luohui 9 (abbreviation for Chao2R under study, O. sativa ssp. indica, saline–alkaline-sensitive) and RPY geng (O. sativa ssp. japonica, saline–alkaline-tolerant), to identify the main pathways and important factors related to saline–alkaline tolerance. Transcriptome analysis showed that 68 genes involved in fatty acid, amino acid (such as phenylalanine and tryptophan), phenylpropanoid biosynthesis, energy metabolism (such as Glycolysis and TCA cycle), as well as signal transduction (such as hormone and MAPK signaling) were identified to be specifically upregulated in RPY geng under saline–alkaline conditions, implying that a series of cascade changes from these genes promotes saline–alkaline stress tolerance. The transcriptome changes observed in RPY geng were in high accordance with the specifically accumulation of metabolites, consisting mainly of 14 phenolic acids, 8 alkaloids, and 19 lipids based on the combination analysis of transcriptome and metabolome. Moreover, some genes involved in signal transduction as hub genes, such as PR5, FLS2, BRI1, and NAC, may participate in the saline–alkaline stress response of RPY geng by modulating key genes involved in fatty acid, phenylpropanoid biosynthesis, amino acid metabolism, and glycolysis metabolic pathways based on the gene co-expression network analysis. The present research results not only provide important insights for understanding the mechanism underlying of rice saline–alkaline tolerance at the transcriptome and metabolome levels but also provide key candidate target genes for further enhancing rice saline–alkaline stress tolerance.

1. Introduction

Soil saline–alkalization is a widely spread abiotic stress that is emerging as a major limiting factor of global crop growth and production [1]. Soil saline–alkalization is characterized by both high salinity and high alkalinity (often pH > 8.0) due to the enhanced accumulation of two major natural alkaline salts, sodium carbonate (Na2CO3) and sodium hydrogen carbonate (NaHCO3) [2,3,4]. So far, extensive efforts have been conducted to dissect the underlying molecular and physiological mechanisms by which plants respond and adapt to neutral salt stress (e.g., NaCl and Na2SO4), especially in the rice model plant species (Oryza sativa L.) and Arabidopsis thaliana [1,5,6,7,8]. However, fewer studies have been devoted to the elucidation of the plants’ response and adaption to saline–alkaline stress. In contrast to neutral salt–stress–encountering problems related to osmotic stress and ionic toxicity [1,6,7,8], saline–alkaline stress is more complex, which often causes combined damages of osmotic stress, ionic toxicity, oxidative stress, and high pH stress [3,4], indicating that saline–alkaline stress will result in more severe symptoms to plants in the inhibition of growth and the antioxidant system, as well as in an imbalance of ionic homeostasis and osmotic adjustment, etc. [9,10]. For instance, a low concentration of NaHCO3 (<10 mM) resulted in plant leaf chlorosis and lethal symptoms, while a more high–fold concentration of NaCl (200 mM) was required to cause similar symptoms [11]. Notably, high salinity and high alkalinity often occur simultaneously in nature soils, and more than 900 million hectares of lands (>7% of the world’s total land area) are suffering from saline–alkalization all over the world. Soil saline–alkalization has become a major factor constraining the sustainable development of global agriculture. Therefore, there is an urgent need for the elucidation of the plants’ response and adaptation to saline–alkaline stress, the understanding of which can potentially provide important insights for cultivating and breeding crops capable of growth in saline–alkaline farmlands.
Saline–alkaline stress in the soil is characterized by high salt and high pH due to the hydrolysis of Na2CO3 and NaHCO3. High salinity is commonly known to result in a physiological water shortage on plants due to the effect of high Na+ concentrations in the soil, which then leads to stress pathways, including osmotic stress, ionic stress, and secondary stress, especially oxidative stress, in plants [1]. High pH in saline–alkaline soil can also reduce the uptake of many essential nutrients and the Na+ exclusion, thus inducing nutrient stress on plants [12]. All these result in plants suffering from osmotic stress, ionic toxicity, oxidative stress, and high pH stress simultaneously, ultimately severely destroying the cell membrane structure and its integrity, disturbing the cell pH and osmotic and ionic balances, inactivating the enzyme activity, imposing nutrient stress upon plants, etc. [3,4]. In order to cope with the adversity of saline–alkaline stress, plants have evolved multiple adaptive strategies, including mainly the accumulation of osmotic adjustment substances, such as soluble proteins and sugars, to maintain the intracellular water potential [3,4]; the excretion and isolation of Na+ toxicity via transporters or channel proteins or proton pumps, such as salt overly sensitives (SOSs) and high–affinity K+ transporters (HKTs) and H+–ATPase to maintain intracellular ionic homeostasis [13,14,15]; the synthesis of antioxidants (e.g., flavonoids, alkaloids, vitamins, phenolic acids, and mannitol) and antioxidant enzymes (e.g., superoxide dismutase (SOD), peroxidase (POD), and catalase (CAT)) to reduce the oxidative damage of reactive oxygen species (ROS) [3,4]; as well as the secretion of organic anions and H+ to maintain intracellular pH stability via acidifying the rhizosphere [15,16]. During these adaptive strategies of the plants in response to saline–alkaline stress, many genes involved mainly in ionic homeostasis, osmotic regulation, ROS-scavenging activity, as well as signal transduction are induced to express [3,17]. Previous studies have demonstrated that external saline–alkaline conditions are perceived by plants based on a series of corresponding gene changes from a complicated network of signal transduction pathways, including hormone signal transduction pathways, protein kinase pathways (e.g., mitogen–activated protein kinases (MAPKs), Ca2+–dependent protein kinases (CDPKs), and calcineurin–B–like interacting protein kinases (CIPKs)), and SOS pathways [17,18,19]. For instance, activated SOS2, as an intermediate hub in the SOS pathways, was responsible for the Na+ efflux from cytoplasm into vacuole to maintain the intracellular ion balance through phosphorylating and activating the NHX (Na+/H+ antiporter) [20,21,22]. MAPKs and receptor-like protein kinases (RLKs) played a vital role in contributing to the tolerance for plant adaption to salt stress by the modulation of ion and ROS homeostasis [19,23,24]. In addition, the concentration changes in hormones regulated by the expression of hormone-related genes were also a major factor for plant adaptation to saline–alkaline stress, such as abscisic acid (ABA), jasmonic acid (JA), brassinosteroid (BR), gibberellin (GA), and auxin (IAA) [25,26,27]. Furthermore, transcription factors (TFs), such as NACs, WRKYs, bZIPs, and bHLHs, that served as a bridge between upstream stimulus signals and their corresponding downstream–associated genes played important roles for the plants’ adaptation to saline–alkaline stress [18,28]. Moreover, the genes related to resistance, including those that are osmoprotectant–related, such as glutathione S–transferase (GST) and glutamate decarboxylase (GAD); antioxidant–related, such as glutathione peroxidase (GPX) and pathogenesis-related proteins (PRs); as well as transporter-related, such as NHXs and HKTs, were also often induced to express to regulate the intracellular balance of ions, pH, and osmotic potential in a direct way [3,4]. These findings indicate that the transcriptional regulation underlying the plants’ response and adaptation to saline–alkaline stress is a very complicated process regulated by the involvement of multiple genes in multiple metabolic pathways; nevertheless, the mechanism behind this remains largely unknown.
Rice (O. sativa L.), as one of the most important staple crops, is considered to be a saline–alkaline-sensitive glycophytic species [29]. In nature, high salt and high pH are often two coexisting abiotic stresses. However, in contrast to neutral salt stress, fewer studies have been conducted to understand the mechanisms of the response and adaptation to saline–alkaline stress in rice. In China, the total area of saline–alkaline soil is nearly 1 × 108 hectares, accounting for approximately 10% of the total land area, which is distributed mainly in northwestern, northeastern, and coastal areas [30], and more seriously, soil saline–alkalization in these regions is continuing to expand annually at a rate of approximately 1 × 106 hectares due to ecological environment deterioration, the improper application of chemical fertilizers, and poor irrigation practices [31]. Currently, approximately 20% of the total rice is planted in the saline–alkaline farmlands, and soil saline–alkalization is emerging as one of the major constraints of rice production in China [32]. Previous studies have demonstrated that different genotypes within a species show extensive variations of tolerance to saline–alkaline stress [13,33,34,35,36,37,38]. These earlier findings have focused mainly on an individual variety, identical subspecies, or their derived populations based on the analysis of physiological, hormonal, transcriptomic, or metabolomic aspects. Few research studies have been conducted to investigate the different responses to saline–alkaline stress between indica and japonica rice cultivars based on the integrated analysis of the phenotype, physiology, transcriptome, and metabolome. To our knowledge, there are marked differences in many important agronomic traits and environmental stresses (including biotic and abiotic stresses) between indica and japonica rice genotypes due to their difference in genomic structure [39,40,41,42,43,44,45,46]. It is, therefore, urgent to dissect the different underlying mechanisms by which indica and japonica rice cultivars respond and adapt to saline–alkaline stress, the understanding of which will provide the potential for cultivating the new rice cultivars with strong saline–alkaline tolerance via indicajaponica subspecies hybrids [47]. In our studies, we screened and assessed the saline–alkaline tolerance of seven rice cultivars consisting of four indica cultivars (93-11, Chao2R, Fuli, and Baihuamiao) and three japonica cultivars (Nipponbare, ZH11, and RPY geng). Among them, we found that indica cultivars were more sensitive to saline–alkaline stress than japonica cultivars. Intriguingly, hybrid F1 progenies from the crossing of Chao2R (O. sativa ssp. indica) and RPY geng (O. sativa ssp. japonica) showed obvious heterosis in terms of their saline–alkaline tolerance (unpublished data). The two rice genotypes, Chao2R and RPY geng, were, therefore, used to explore the underlying mechanisms by which rice responds and adapts to saline–alkaline stress at the transcriptome and metabolome levels. These findings will contribute to our understanding for breeding rice with a strong saline–alkaline tolerance.

2. Results

2.1. Chao2R and RPY Geng Showed Distinct Tolerance under Saline–Alkaline Treatment

Given that sodium carbonate (Na2CO3) and bicarbonate (NaHCO3) are the major alkaline salts in natural saline–alkaline environments, 0.3% (w/v) mixtures of alkaline salts (Na2CO3:NaHCO3 = 1:3) are added into the Yoshida solution to mimic the stress conditions that occur in saline–alkaline soils with a high pH value (9.35) and high Na+ concentration (approx. 40.84 mM). To compare the difference in the saline–alkaline tolerance of Chao2R (O. sativa ssp. indica) and RPY geng (O. sativa ssp. japonica), we calculated the final survival rate of both rice subspecies genotypes after exposure to saline–alkaline conditions. As shown in Figure 1, the survival rate (SR) of RPY geng (77%) was significantly higher than that of Chao2R (45%). In addition, shoot height, relative growth rate, and shoot biomass (including fresh weight and dry weight) of RPY geng seedlings were significantly higher than those of Chao2R seedlings in both the control and saline–alkaline solutions (Figure 1A and Figure S1 and Table 1). Notably, shoot growth in both genotypes was observed to be reduced after the saline–alkaline treatment, and the magnitude of the reduction in shoot growth was greater in Chao2R than in RPY geng seedlings (Table 1 and Figure S1). These results demonstrate that RPY geng had a stronger tolerance to saline–alkaline stress than did Chao2R at the seedlings.
To adapt to the stress associated with the saline–alkaline soils, plants have evolved a series of strategies to cope with it, such as morphological adaption and physiological adaption (e.g., accumulation of osmoprotectants, antioxidant enzymes, and antioxidants) [3,4]. To explore the difference in the physiological response of both rice subspecies’ genotypes (Chao2R and RPY geng) under saline–alkaline conditions, we then compared the effects of saline–alkaline stress on the physiological adaption of the two genotypes. Under untreated control conditions, the proline (Pro), soluble sugar (SS), and malondialdehyde (MDA) concentrations and the total antioxidant capacity (T-AOC) of the two genotypes were comparable; the activity of superoxide dismutase (SOD) and catalase (CAT) was significantly higher in RPY geng than in Chao2R, while Chao2R had a higher peroxidase (POD) activity than did RPY geng (Table 1 and Figure S1). Compared with the untreated control conditions, the saline–alkaline stress was found to result in notable increases in Pro, SS, MDA, SOD, POD, and T-AOC in both rice seedlings, except for CAT, which had no significant differences between the two genotypes (Table 1 and Figure S1). These changes led to the magnitude of the increase in SS, POD, and T-AOC of the RPY geng seedlings being significantly greater than those of Chao2R, whereas the magnitudes of the increase in SOD, Pro, and MDA of the Chao2R seedlings were markedly greater than those of RPY geng. Among them, SS was significantly higher in RPY geng than in Chao2R, while Pro was markedly higher in Chao2R than in RPY geng, indicating that there was a significant difference between Chao2R and RPY geng in terms of the accumulation of osmoprotectants under saline–alkaline stress conditions. The production of MDA derived from lipid peroxidation can reflect the levels of cell membrane damage after exposure to environmental stresses [48]. We found the MDA concentration and its magnitude of the increase were markedly higher in Chao2R than in RPY geng, which may suggest that the degree of cell membrane damage of Chao2R was more serious than that of RPY geng under the saline–alkaline stress conditions. POD and SOD are two key antioxidant enzymes that help plants eliminate reactive oxygen species (ROS) and facilitate tolerance to environment stresses [49]. In this study, we found the activity of POD and SOD in RPY geng were significantly higher than those of Chao2R, which may indicate that RPY geng had a stronger ability to remove ROS than did Chao2R. As expected, T-AOC was markedly higher in the RPY geng than in the Chao2R seedlings with the saline–alkaline solution, which was in line with the result that RPY geng was more tolerant than Chao2R under saline–alkaline stress conditions. These results were further verified using the Pearson correlation tests between phenotypic traits and physiological characteristics: SR was significantly positively correlated with SOD and T-AOC (Figure S2)

2.2. Transcriptomic Analysis of Chao2R and RPY Geng in Response to Saline–Alkaline Stress

Three independent samples of both rice subspecies’ genotypes, Chao2R (indica, saline–alkaline-sensitive) and RPY geng (japonica, saline–alkaline-tolerant), were performed from the saline–alkaline treatment sample and the untreated control sample for RNA-sequencing to determine the genes responsible for the saline–alkaline stress tolerance. In total, approximately 745.07 million raw sequencing reads were generated from the BGISEQ sequencing of the 12 samples (Table S1). Each sample covered raw reads ranging from 47.05 to 99.75 million, with an average of 62.09 million. After adapter sequences and low-quality reads were removed, the remaining clean reads of each sample ranged from 44.80 to 95.90 million, with an average of 59.41 million. Then, the 12 samples generated, in total, approximately 712.98 million clean sequencing reads and 106.94 Gb of clean bases (exceeding 6 Gb of each sample), respectively. At least 96.67% and 91.44% bases of the clean reads separately had a quality score ≥ Q20 (less than or equal to the 1% incorrect rate) and ≥ Q30 (less than or equal to the 1‰ incorrect rate), which indicated the high quality of the clean reads in each sample. Here, 92.65–96.76% of the clean reads were mapped to the rice reference genome, with an average mapping rate of 94.60% by using HISAT v2.1.0 software. There were 30,950, 29,344, 30,393, and 30,054 predicted genes of the clean reads mapped in the rice reference genome in Chao2R (Chao2R with untreated control conditions), Chao2RA (Chao2R treated with saline–alkaline conditions for 72 h), RPY (RPY geng with untreated control conditions), and RPYA (RPY geng treated with saline–alkaline conditions for 72 h), respectively. Therefore, the results indicated that more than half of the rice reference genes were expressed in each sample, and these genes were sufficient for the following differential gene expression analysis: The global gene expression levels were similar among the three biological replicates of each treatment (Figure 2A), which suggested good repeatability in each treatment. Principal component analysis (PCA) was performed to evaluate the overall differences of all detected genes between the two rice genotypes under the different treatments. Chao2R and RPY geng could be clearly divided into different regions based on the first principal component (PC1) and PC2, which highlighted the great influence of saline–alkaline stress on the change in the detected genes in the two rice genotypes (Figure 2B). PC1 (35.05%) and PC2 (23.87%) together explained 58.92% of the total variation. In addition, biological replicates from the same treatment were obviously clustered together, which further indicated good repeatability in each treatment.
Differential expression genes (DEGs) in this study were designated as those genes that showed at least a twofold change (the absolute value of the log2 (fold change with FPKM) 1) and were determined to be significantly different (false discovery rate (FDR) < 0.05) under saline–alkaline stress compared with the control. RPY geng (10,117) had more DEGs than did Chao2R (8,623), and more genes were downregulated than upregulated in both rice genotypes (Figure 2C). It was also apparent that RPY geng had more specific DEGs than did Chao2R in either the downregulated or the upregulated fraction (Figure 2D), revealing an apparent difference between the two rice subspecies in response to the saline–alkaline stress at the transcriptional level. Combined with the data of the survival rates (Figure 1B), this could reasonably suggest that RPY geng was more tolerant to saline–alkaline stress than was Chao2R due to its more rapid response to the saline–alkaline stress at the level of transcription.
To further determine the saline–alkaline responsive pathways of the two rice genotypes, both the Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) approaches were used to investigate the functional enrichment and annotation analysis of the DEGs. Both approaches showed that Chao2R and RPY geng had modified their metabolism pathways after exposure to the saline–alkaline treatment. In the GO analysis (Figure 3 and Table S2), the GO category biology process ‘response to salt stress’ and ‘response to osmotic stress’ were enriched in the upregulated DEGs of both Chao2R and RPY geng. The lipid metabolic pathway was especially enriched in the upregulated DEGs of RPY geng, including the biology process categories of ‘fatty acid catabolic process’, ‘fatty acid β-oxidation’, ‘fatty acid oxidation’, ‘lipid oxidation’, ‘cellular lipid catabolic process’ and ‘lipid modification’. However, the GO terms of the upregulated DEGs of Chao2R were associated mainly with the chitin-related process and water-responsive processes, including the categories of ‘response to chitin’, ‘chitinase activity’, ‘chitin binding’, ‘chitin metabolic process’, ‘chitin catabolic process’, ‘response to water’, ‘response to water deprivation’, and ‘positive regulation of response to water deprivation’. The findings implied a different saline–alkaline responsive expression mode of the two rice subspecies’ genotypes. In contrast, the downregulated DEGs of both Chao2R and RPY geng were enriched mainly in ‘cell structure’, ‘cell cycle process’, and ‘photosynthesis process’. It could be inferred that both Chao2R and RPY geng may have transferred the energy requirements of the cell structure, the cell cycle process, and the photosynthesis process to produce greater metabolites, such as lipids and chitins. In the KEGG analysis, similar conclusions could be drawn (Figure S3 and Table S3). Several pathways were enriched in the upregulated DEGs of both Chao2R and RPY geng, including the categories of ‘α-linolenic acid metabolism’, ‘butanoate metabolism’, ‘propanoate metabolism’, ‘galactose metabolism’, ‘valine, leucine, and isoleucine degradation’, ‘β-alanine metabolism’, as well as ‘arginine and proline metabolism’. Notably, the plant signal transduction pathway was also enriched in the upregulation of both Chao2R and RPY geng, with the KEGG pathway ‘plant hormone signal transduction’ upregulated in RPY geng and ‘MAPK signaling pathway-plant’ upregulated in Chao2R. The lipid and amino acid metabolic pathways were specifically enriched in the upregulated DEGs of RPY geng, including the KEGG pathways of ‘fatty acid degradation’, ‘glycerolipid metabolism’, ‘biosynthesis of unsaturated fatty acids’, ’phenylalanine metabolism’, ‘tyrosin metabolism’, and ‘tryptophan metabolism’. In contrast, many pathways were enriched in the downregulated DEGs of both Chao2R and RPY geng genotypes, such as ‘ribosome’, ‘photosynthesis’, ‘aminoacyl-tRNA biosynthesis’, ‘photosynthesis-antenna proteins’, and ‘glycine, serine, and threonine metabolism’. In all, the findings of the functional enrichment and annotation of the DEGs indicated that Chao2R and RPY geng modified the expression of multiple genes involved in many pathways, including lipid metabolic, amino acid and organic acid metabolic, terpenoid biosynthesis, plant signal transduction, chitin-related, as well as salt and osmotic stress responsive, which may suggest their involvement in rice saline–alkaline stress resistance. The genes related to lipid metabolism, phenylalanine metabolism, tryptophan metabolism, tyrosine metabolism, signal transduction, and sugar metabolism that were especially enhanced by the saline–alkaline stress in RPY geng may play exceptional roles in building up its resistance against saline–alkaline stress.

2.3. Metabolomic Analysis of Chao2R and RPY Geng in Response to Saline–Alkaline Stress

To determine how metabolites respond to saline–alkaline stress, the 12 samples taken from both rice genotypes (Chao2R and RPY geng) during saline–alkaline treatment and untreated control were next subjected to wide-targeted metabolic profiling using UPLC-ESI-MS/MS. Here, a total of 801 metabolites were identified and annotated among the 12 samples, which could be divided into 13 different categories, including 74 amino acids and their derivatives (9.24%), 131 lipids (16.35%), 49 nucleotides and their derivatives (6.12%), 78 organic acids (9.74%), 42 saccharides and alcohols (5.24%), 18 vitamins (2.25%), 73 alkaloids (9.11%), 189 flavonoids (23.60%), 18 types of lignans and coumarins (2.25%), 108 phenolic acids (13.48%), 7 tannins (0.87%), 6 terpenoids (0.74%), and 8 types of other compounds (1.00%) (Figure 4A). The overall metabolic profile of Chao2R and RPY geng in each treatment could be visualized based on hierarchical cluster analysis (HCA) (Figure 4B). The heat map showed that the three biological replicates from the same group were clustered together, which suggested a high repeatability and a robust correlation between replicates. Flavonoids, lipids, phenolic acids, organic acids, amino acids and their derivatives, and alkaloids were six dominant metabolites based on the type and proportion of the metabolites. PCA was subsequently used to evaluate the overall differences between Chao2R and RPY geng under the different treatments at the metabolic level. Chao2R and RPY geng could be obviously separated into distinct regions by PC1 and PC2, indicating that the saline–alkaline conditions could have a great impact on the change of the metabolites in both rice subspecies’ genotypes (Figure 4C). The PCA results of all identified metabolites showed that the first two PCs explained 64.98% of the total variation. Similar to gene expression (Figure 2B), all the biological replicates of metabolomics samples were also grouped together, indicating the high repeatability and good homogeneity among the replicates.
Both rice subspecies’ genotypes showed a distinctive saline–alkaline responsive accumulation pattern at the metabolic level. Here, significantly differentially accumulated metabolites (DAMs) between the groups in this study were defined as those with at least a twofold alteration (|log2FC (fold change)| 1) and a variable importance in projection (VIP) 1. In all, 560 DAMs were identified between Chao2R and RPY geng under the saline–alkaline treatment in comparison with the control condition, including mainly flavonoids (24.46%), lipids (16.96%), phenolic acids (13.21%), alkaloids (10.36%), organic acids (10.36%), and amino acids and their derivatives (9.46%) (Figure S4). More DAMs were observed in RPY geng (446) than in Chao2R (392); more metabolites were downregulated than upregulated in Chao2R, while more metabolites were upregulated than downregulated in RPY geng (Figure 4D and Figure S4A). There were 278 DAMs common in Chao2R and RPY geng, with the DAMs specifically accumulated in both rice subspecies’ genotypes: 114 in Chao2R and 168 in RPY geng (Figure S4B). In particular, RPY geng had more unique DAMs than did Chao2R in the upregulated fraction, while more specific DAMs were detected in Chao2R than RPY geng in the downregulated fraction (Figure 4E).
The DAMs identified in both rice genotypes were further investigated using the KEGG pathway database. The KEGG analysis revealed that Chao2R more so than RPY geng had modified their metabolism after exposure to the saline–alkaline conditions. Notably, completely different pathways were specifically enriched in the DAMs between Chao2R and RPY geng, respectively (Figure S5). Five pathways were significantly enriched in the 169 DAMs of RPY geng, including ‘tryptophan metabolism’, ‘glucosinolate biosynthesis’, ‘aminoacyl-tRNA biosynthesis’, ‘phenylalanine metabolism’, and ‘tropane, piperidine, and pyridine alkaloid biosynthesis’, while the significant categories among 9 pathways in the 137 DAMs of Chao2R included ‘pyruvate metabolism’, ‘valine, leucine, and isoleucine degradation’, ‘2-oxocarboxylic acid metabolism’, ‘citrate cycle (TCA cycle)’, ‘propanoate metabolism’, ‘carbon metabolism‘, ‘biosynthesis of secondary metabolites ‘, ‘valine, leucine, and isoleucine biosynthesis’, and ‘biosynthesis of amino acids’. To observe more intuitively the variations occurring among these different metabolites in Chao2R and RPY geng during the saline–alkaline conditions and the untreated control, a heat map of the 560 different metabolites was obtained (Figure S6). Among them, 108 metabolites increased significantly in both Chao2R and RPY geng after exposure to the saline–alkaline treatment when compared with their levels in the untreated control (Figure 4E and Figure S6), including 25 alkaloids, 23 amino acids and their derivatives, 21 lipids, and 18 organic acids, indicating these metabolites were most involved in the rice’s saline–alkaline tolerance. The 229 metabolites that induced specifically in RPY geng consisted of lipids (56), flavonoids (48), phenolic acids (40), and alkaloids (21), etc., and this could be an indication of their exceptional functional involvement in the saline–alkaline tolerance of RPY geng. In contrast, those 86 metabolites that increased uniquely in Chao2R were flavonoids (39), organic acids (13), phenolic acids (10), etc. Subsequently, 560 DAMs were classified, and the changes in the various types of metabolites were observed between Chao2R and RPY geng after exposure to the saline–alkaline conditions in comparison with the control. The results showed that many types of metabolites displayed distinct changes and trends in both rice genotypes, such as lignans and coumarins, nucleotides and their derivatives, vitamins, lipids, alkaloids, as well as phenolic acids (Figure 4F and Figure S7). It is noteworthy that the content and alteration of lipids, alkaloids, and phenolic acids in RPY geng were significantly higher than those in Chao2R. In contrast, higher contents of organic acids, tannins, terpenoids, as well as saccharides and alcohols were detected in Chao2R than in RPY geng. Based on the above results, a clear difference was observed between the two rice subspecies’ genotypes in response to the saline–alkaline stress at the metabolic level. Together with the survival data (Figure 1B), it could be speculated that RPY geng was more tolerant to the saline–alkaline stress than Chao2R by rapidly and specifically increasing the metabolites under the saline–alkaline stress treatment, especially lipids, alkaloids, as well as phenolic acids.

2.4. Mining the Potential Important Candidate Genes and Metabolites for Saline–Alkaline Tolerance

In order to explore further the molecular mechanisms of RPY geng in regulating and adapting to salt–alkaline stress more than Chao2R, an integrated systemic KEGG mapping analysis was constructed based on the differences in the metabolites and genes. A total of 77 metabolic pathways displayed alterations in the rice in response to the saline–alkaline stress, and 1504 genes and 169 metabolites were mapped to these pathways (Table S4). As shown in Figure 5, these metabolites and genes were involved mainly in the pathways related to amino acid metabolism, energy metabolism (Glycolysis and TCA cycle), fatty acid metabolism, alkaloid biosynthesis, phenylpropanoid biosynthesis, as well as plant signal transduction pathway (hormone signal transduction and MAPK signaling pathways).
Considering all the results of the transcriptome and metabolome, we then focused on DEGs and DAMs which showed the opposite regulation by the saline–alkaline condition in RPY geng and Chao2R to narrow down the number of potential important candidate genes and metabolites that were specifically regulated by the saline–alkaline stress. We speculated that these genes and metabolites may play different roles in the RPY geng and Chao2R responses to the saline–alkaline stress. Based on the criteria, we finally screened 68 genes and 53 metabolites (including 19 lipids, 14 phenolic acids, 8 alkaloids, 4 lignans, 3 vitamins, 3 amino acid, and 1 coumarin), those most involved in the rice saline–alkaline tolerance (Figure 6A,B). Amongst these genes and metabolites, most were specifically upregulated by the saline–alkaline stress in RPY geng. The functional correlation of these 68 genes in response to the saline–alkaline stress was firstly tested by constructing a co-expression network using Cytoscape v3.8.2 software (Pearson correlation coefficient (PCC) ≥ 0.9, p < 0.01). Based on gene connectivity, it was found that some genes involved in the plant signal transduction pathway were identified as hub genes, such as PR5 (pathogenesis-related protein, Os02g12510), WRKY (Os08g0235800 and Os07g0111400), SPCH/bHLH (Os02g0691500), NAC (Os02g0285900), FLS2 (Os04g0349700), LUG (Os02g0813800), BRI1 (protein brassinosteroid-insensitive 1, Os02g0154200), as well as TGA/bZIP (Os05g0443900), and these hub genes were found to be highly co-expressed with genes involved mainly in the amino sugar and nucleotide sugar metabolism (especially Os08g0237200 and Os03g0268400), glutathione metabolism (Os06g0185900 and Os10g0525500), phenylpropanoid biosynthesis (especially Os08g0543400, Os04g0473900, and Os01g0591300), ubiquitin-mediated proteolysis (especially Os05g0182500, Os09g0272900, and Os05g0193900), fatty acid metabolism (especially Os03g0108500, Os06g0300000, Os08g0184800, Os12g0559200, and Os12g0102100), amino acid metabolism (especially Os04g0107500 and Os04g0447800), glycolysis (especially Os10g0204400 and Os07g0197100), etc. (Figure 6C). Therefore, we may speculate that these genes related to the plant signal transduction pathway play important roles in regulating the key genes’ involvement in sugar metabolism, amino acid metabolism, energy metabolism, fatty acid metabolism, phenylpropanoid biosynthesis, etc., and a series of these genes’ cascade correlations may accelerate and drive the higher saline–alkaline tolerance of RPY geng compared with that of Chao2R.
In order to validate the reliability of the transcriptome datasets under study, twenty potential candidate genes related to the plant signal transduction pathway, sugar metabolism, glutathione metabolism, phenylpropanoid biosynthesis, fatty acid metabolism, amino acid metabolism, as well as glycolysis were selected for study by quantitative reverse transcription-polymerase chain reaction (qRT-PCR), with gene-specific primers in all samples of Chao2R and RPY geng (Table S5). All twenty genes showed a largely consistent expression pattern between the qRT-PCR results and the RNA-seq data (R2 = 0.9573, p < 0.0001) (Figure S8), revealing that the RNA-seq dataset under study was highly adequate for further analysis.
To reveal the correlation between the 68 genes and 53 metabolites involved in the rice saline–alkaline stress tolerance, an interactive network between these genes and the metabolites was constructed using Cytoscape v3.8.2 software based on the screening criteria of the absolute value of PCC ≥ 0.9 and p < 0.01. As shown in Figure 6D, there are 52 metabolites (except for N-Acetylornithine) correlated with 67 genes (except for Os01g0952800). For instance, 5 genes (especially Os08g0543400, Os04g0473900, and Os01g0591300) related to phenylpropanoid biosynthesis were positively correlated with 48 metabolic compounds, including 18 lipids, 13 phenolic acids, 8 alkaloids, 4 lignans, 3 vitamins, 1 coumarin, and 1 amino acid; there were 12 genes (especially Os04g0146800, Os06g0300000, Os12g0559200, and Os03g0423300) involved in fatty acid metabolism that were positively related with 50 metabolites, including 19 lipids, 13 phenolic acids, 9 alkaloids, 4 lignans, 3 vitamins, 1 coumarin, and 1 amino acid; another 5 genes (especially Os08g0237200 and Os11g0204600) related to amino sugar and nucleotide sugar metabolism were positively correlated with 46 metabolic compounds; there were 6 genes (especially Os10g0213100, Os10g0155400, and Os04g0447800) mapped to amino acid metabolism that were positively related with 40 metabolites; another 5 genes (especially Os06g0145600 and Os02g0734400) assigned to cutin, suberine, and wax biosynthesis were positively correlated with 35 metabolic compounds; we found 3 genes (especially Os07g0197100 and Os10g0204400) involved in glycolysis that were positively related with 32 metabolites; there were 3 genes (especially Os05g0182500 and Os09g0272900) related to ubiquitin-mediated proteolysis that were positively correlated with 42 metabolites; lastly, 2 genes (Os10g0525500 and Os06g0185900) involved in glutathione metabolism were positively related with 30 metabolites. Notably, 19 genes assigned to the plant signal transduction pathway were also positively correlated with 50 metabolic compounds, including 19 lipids, 13 phenolic acids, 9 alkaloids, 4 lignans, 3 vitamins, 1 coumarin, and 1 amino acid, indicating that the involvement of these genes (especially Os02g12510, Os04g0349700, Os06g0306600, Os08g0235800, Os02g0154200, Os05g0443900, Os07g0111400, and Os12g0609200) in the plant signal transduction pathway may have played important roles in positively regulating the biosynthesis of these 50 metabolites.

3. Discussion

Rice (O. sativa L.), as one of the most important staple crops, is a saline–alkaline-sensitive crop. Currently, approximately 20% of the total rice is planted in saline–alkaline soil, and soil saline–alkalization has become one of the major constraints of rice production [32]. In contrast to neutral salinity, fewer related studies have been conducted on saline–alkaline stress. However, high salt concentration and high pH are often two coexisting abiotic stresses in nature. Therefore, there is an urgent need to dissect the mechanisms regarding how rice regulates and adapts to this saline–alkaline stress, the understanding of which can potentially increase the rice saline–alkaline tolerance and exploit adequately the saline–alkaline farmlands. Sodium carbonate (Na2CO3) and bicarbonate (NaHCO3) are the two major salts existing in natural saline–alkaline farmlands, which can cause the synergistic damages of high Na+ toxicity, osmotic stress, and pH stress on rice [2,3,4]. Here, we have shown that RPY geng (japonica) is a more saline–alkaline-tolerant genotype than Chao2R (indica), as evidenced by a lower magnitude of reductions in shoot height and shoot biomass, lower magnitude of increase in MDA, as well as more marked increases in SR, SS, POD, SOD, and T-AOC (Figure 1 and Figure S1 and Table 1), indicating that an in-depth investigation of the different mechanisms between indica Chao2R and japonica RPY geng responses to saline–alkali stress might provide a new insight for improving rice saline–alkali tolerance.
Given the deleterious effects of saline–alkaline stress on rice, previous some studies have been conducted to investigate the potential mechanisms by which rice regulates and adapts to saline–alkaline stress conditions. For instance, japonica ‘Nipponbare’ was used to alleviate the harmful effect of the saline–alkaline stress using magnetized water [50]; japonica ‘Tongxi926′ was used to dissect the effects of long-term high saline–alkaline stress on rice [36]; indica ’93-11′ and its polyploidy were used to explore the mechanisms of saline–alkali tolerance [51]; two japonica cultivars ‘dongdao-4′ and ‘jigeng-88′ were used to study their differences in tolerance to saline–alkaline stress [16,35,37]; and two indica cultivars ‘FL478′ and ‘IR29′ were used to reveal their differences under saline–alkaline stress by regulating Na+ transport [13]. These earlier studies have focused mainly on an individual variety, identical subspecies, or their derived populations based on the analysis of physiological, hormonal, transcriptomic, or metabolomic aspects. In contrast, fewer studies have been conducted to explore the different responses to saline–alkaline stress between indica and japonica rice cultivars based on the comprehensive phenotypic, physiological, transcriptomic, and metabolomics analyses. As we know, there are marked differences in many important agronomic traits and environmental stresses (including biotic and abiotic stresses) between indica and japonica rice cultivars due to the difference in their genomic structures [39,40,41,42,43,44,45,46]. In order to take full advantage of the differences between indica Chao2R (saline–alkaline-sensitive genotype) and japonica RPY geng (saline–alkaline-tolerant genotype) for breeding new rice cultivars with higher saline–alkaline stress tolerance via the indicajaponica subspecific hybrids, there is first an urgent need to explore the different mechanisms by which they regulate and adapt to saline–alkaline stress at the transcriptional and metabolomic levels. In our study, the lipid metabolic pathway, the plant signal transduction pathway, as well as the secondary metabolic biosynthesis (including phenolic acid and alkaloid) that were specifically enriched by the saline–alkaline stress conditions in RPY geng may play important roles in building up its resistance against saline–alkaline stress according to the GO and KEGG aspects (Figure 3, Figure 5 and Figure S4), suggesting that RPY geng was equipped with a more efficient mechanism by coupling with multiple biological processes for adapting to the saline–alkaline conditions. Recently, QTL mapping and candidate gene analysis for rice saline–alkaline tolerance has been explored based on the genome-wide association study (GWAS) and linkage-mapping approaches using bi-parental populations (such as recombinant inbred lines (RILs)) or natural populations [29,33,38]. For example, linkage mapping and a GWAS analysis had been used to detect 3 candidate genes (Os11g37300, Os11g37320, and Os11g37390) within a major QTL (qAT11) on chromosome 11 for alkali tolerance in Japonica rice using 184 RILs and 295 japonica rice varieties [38]. Similarly, one candidate gene (OsHKT1;1) within a major QTL (RNC4) significantly associated with root Na+ content and root Na+/K+ ratio was identified for regulating the distinct salt tolerance between indica accessions and japonica accessions by GWAS [52]. Li et al. (2019) used 295 japonica rice varieties to evaluate the alkalinity tolerance at the seedling stage via GWAS, and Os03g26210 (OsIRO3) within a common QTL on chromosome 3 associated with the three phenotypes of alkalinity tolerance was identified for improving the alkalinity tolerance in rice [33]. These results suggest that QTL mapping and GWAS in bi-parental populations and natural populations are powerful strategies for dissecting the genetic variations and identifying the candidate genes in saline–alkaline tolerance among the rice germplasms. Therefore, major QTLs or candidate genes related to the saline–alkaline-tolerant differences between indica ‘Chao2R’ and japonica ‘RPY geng’ will be expected to be more finely mapped using the comprehensive analysis of the phenotypical, metabolomics, transcriptomic, and genomic datasets based on the linkage-mapping populations consisting of 270 derived RILs from a cross between Chao2R and RPY geng in our lab [39]. Notably, QTL mapping in bi-parental populations is insufficient to reveal the genetic architecture and identify the loci and candidate genes in rice saline–alkaline tolerance due to the difficulty in the acquisition of effective phenotypic data [29,53]. Having effective phenotypic data is a prerequisite for the detection of QTLs or genes for the saline–alkaline-tolerance trait. In contrast to the QTL linkage analysis in bi-parental segregating populations, GWAS is a more powerful tool for gaining insight into the valuable natural variations in trait-associated loci or genes, thus detecting multiple alleles at the same site based on high-density variations in the natural populations [33]. In order to improve the accuracy and breadth of major QTL loci, it is expected that mining the candidate genes or QTLs significantly associated with saline–alkaline tolerance in japonica ‘RPY geng’ will be carried out in the future through the combination analysis of bi-parental QTL mapping and GWAS.
Saline–alkaline tolerance is a very complicated quantitative trait controlled by multiple genes in the rice [29]. External saline–alkaline stress can induce expression changes of a series of cascade-related genes that are involved mainly in ionic homeostasis, osmotic regulation, ROS-scavenging activity, as well as signal transduction [3,4]. Herein, we ultimately screened the 68 genes that are most involved in the rice saline–alkaline tolerance based on an integrated analysis of transcriptome and metabolome (Figure 6A). Notably, the largest portion of these genes was most enriched in the plant signal transduction pathways. Previous studies have demonstrated that external saline–alkaline conditions were firstly perceived by plants based on a complicated network of signal transduction pathways, including hormone signal transduction pathways (e.g., ABA and IAA), protein kinase pathways (e.g., MAPKs, CDPKs, and CIPKs), and salt overly sensitive (SOSs) pathways [6,7,8,17]. The SOS pathway, as the first signal transduction pathway established in plants during saline–alkaline stress, was responsible for the Na+ efflux in the cells [14]. The activated SOS2 that acted as an intermediate hub in the SOS pathways could activate the NHX (Na+/H+ antiporter) by phosphorylation to maintain the intracellular ion balance through the efflux of Na+ from the cytoplasm into the vacuole [20,21,22]. NHX-encoding reverse transport protein/channel ion could enhance the plant salt tolerance through the modulation of a high Na+ efflux rate and K+/Na+ ratio by inducing the expression of SKOR (stellar K+ outward rectifier), SOS1 (salt overly sensitive 1), and AKT1 (Arabidopsis K+ transporter 1) [54,55,56]. Plants could also respond to the saline–alkaline stress through the regulation of protein kinases consisting of mitogen-activated protein kinases (MAPKs) [19,23,57] and receptor-like protein kinases (RLKs) [24,58], and overexpressing these family members in transgenic plants significantly enhanced the tolerance for plant adaption to salt stress by modulating the ion and ROS homeostasis. In addition, alterations in the hormones are an indispensable factor for plant adaption to growth and development during signal transduction pathway involvement in response to the saline–alkaline stress [25,26,27]. For instance, the involvement of BRI1 (brassinosteroid-insensitive 1) in the brassinosteroid (BR) signaling pathway was induced to express by salinity or alkalinity stress, and its overexpression in transgenic plants was more tolerant to salinity or alkalinity [59,60]; GID1 (gibberellin-insensitive dwarf 1) that could function as a soluble GA receptor [61], was found to be involved in the response to salt or alkali stresses by modulating the endogenous GA levels [37]; ABCGs that functioned as an ABA exporter or importer participated in the plants’ adaption to salt stress by cooperating to transport ABA and Na+ in this process [62]; Aux/IAAs, as one of primary protein components associated with the auxin signaling pathway, mediated the distribution and accumulation of IAA in plant roots in response to salt or alkali stresses [63,64]. Studies have also demonstrated that the transcript levels of genes involved in JA biosynthesis and its signal pathway were significantly upregulated in the plant’s response to high salinity or alkalinity stress, such as AOC (allene oxide cyclase), AOS (allene oxide synthase), LOX (lipoxygenase), MFP (multifunctional protein), OPR (12-oxophytodienoate reductase), and JAZ (jasmonate ZIM-domain protein), and their overexpressions in transgenic plants were more tolerant to salinity or alkalinity [65,66,67,68]. In this study, genes encoding the protein homologs of NHX (Os06g0318500), SKOR (Os01g0189100), MKK (Os02g0694900), RLK (Os01g0137700), BRI1s (Os02g0154200, Os11g0618300, Os06g0544100, and Os08g0521200), GID1s (Os08g0475400 and Os06g0306600), ABCG11 (Os12g0411700), Aux/IAA (Os03g0797800), as well as JAZ (Os12g0609200) were rapidly and specifically induced by saline–alkaline stress in RPY geng, suggesting that the plant signal transduction pathways may play vital roles in rice for saline–alkaline stress tolerance.
Transcription factors act as a bridge between upstream stimulus signals and a series of their corresponding downstream–associated genes along the signal transduction pathway in response to the saline–alkaline stress [18,28]. For example, NACs encoding the plant-specific transcription factor family were reported to be highly linked to the responses to salinity and alkalinity stress through the adjustment of the intracellular ion homeostasis, osmotic potential, and the ROS-scavenging capability [69,70], which may participate in the ABA- or ethylene-dependent signal transduction pathway [1,71,72]. WRKY was induced to express during salinity or alkalinity, and the overexpression of WRKY in transgenic plants contributed to the salt or alkali tolerance by enhancing the contents of the osmotic adjustment compounds (such as proline and soluble proteins) and reducing the ROS levels [73,74]. Recent studies have shown that salt or alkali stress could trigger the transcript levels of bZIP transcription factors, and bZIP-overexpressing plants increased the salinity or alkalinity resistance by maintaining an ion homeostasis, contributing to the ROS detoxification and reducing the MDA accumulation based on an ABA-dependent signal transduction pathway [75,76]. In addition, a study has proven that the bHLH transcription factor functioned positively in salt tolerance through its effects on the intracellular ion homeostasis and proline biosynthesis by directly activating the expression of the SOS1 and P5CS (pyrroline-5-carboxylate acid synthetase) gene homologs [77]. In this study, several transcription factors, such as NAC (Os02g0285900), WRKYs (Os07g0111400 and Os08g0235800), bZIP (Os05g0443900), as well as bHLHs (Os01g0952800 and Os02g0691500), were significantly and uniquely upregulated in the RPY geng response to saline–alkaline stress, suggesting these transcription factors could be involved in the saline–alkaline tolerance of rice by mediating their corresponding related pathways.
Given the adversity associated with saline–alkaline stress, the genes related mainly to ionic homeostasis, osmotic regulation, and ROS-scavenging activity were specifically induced to maintain an intracellular balance of ions, pH, and osmotic potential in a direct way [3,4]. Herein, the transcriptome analysis showed that some candidate genes were specifically upregulated in RPY geng under saline–alkaline stress, including three-, five-, six-, five-, twelve-, five-, two-, and three-gene homologs related to glycolysis/gluconeogenesis metabolism (ATP, ADH, and HK), amino sugar and nucleotide sugar metabolism (e.g., MPGs), biosynthesis of amino acids (e.g., GAD), phenylpropanoid biosynthesis (e.g., PAL and 4CL), fatty acid metabolism (e.g., PLA2, ACS, MFP2, OPR, LOX, and ACX), cutin, suberine, and wax biosynthesis (e.g., POGs, WAX2, and CER1), glutathione metabolism (GST and GPX), as well as ubiquitin-mediated proteolysis (e.g., UBC and SKP1), respectively (Figure 6A). Previous studies revealed that these genes played particular roles in the response and adaptation to the saline or alkaline stress conditions. For instance, glucose, as a signal, regulated the plant growth, development and stress response, and hexokinases (HKs) functioned in the sugar-mediated sensing and signaling pathways to regulate and adapt to the salinity tolerance by interacting with and phosphorylating NHX1 [78,79,80]. Alcohol dehydrogenase (ADH) could be induced to express in the plants’ response to the salinity stress, and the ADH-overexpressing plants could contribute to the interconversion of aldehydes to alcohols; then, this occurrence might trigger the production of green leaf volatiles (GLVs), thereby increasing the expression levels of multiple stress-related genes and the accumulation of callose depositions and soluble sugars, ultimately conferring enhanced resistance to the salinity stress of the plants [81,82]. ATP, as one of the key salt-signaling network molecules, played important roles in the plant adaption to salinity stress through triggering ethylene to maintain K+/Na+ homeostasis by activating and strengthening the H2O2 and Ca2+ signaling pathways [83,84]. Recent new candidate genes encoding mannose-1-phosphate guanyl transferases (MPGs) were identified by the functional screening of a cDNA library from a salt-tolerant rice, and the overexpression of MPG1 in the transgenic lines played a particular role in enhancing the salinity tolerance [85]. In addition, glutamate decarboxylase (GAD) was involved in the γ-aminobutyric acid (GABA) biosynthesis, and the increasing endogenous GABA levels alleviated the plants’ effects of salt damage through modulating the Na+ uptake, increasing the amino acid level, as well as strengthening the antioxidant metabolism to decrease the ROS and MDA contents [86,87]. In addition, several key genes (e.g., PALs and 4CLs) involved in the phenylpropanoid biosynthesis may have contributed to the salinity or alkalinity tolerance by triggering the ROS-scavenging activity according to recent comparative omics analyses [88,89]. Moreover, fatty acids and their derivatives participated in the plant tolerance to stress response together with their roles in lipid-dependent signaling cascades [90,91]. The involvement of acyl-coenzyme A oxidases (ACXs) in the β-oxidation played an important role in the plant growth, development, and stress response, and the overexpression of the ACX gene in the plants enhanced their drought and salinity tolerance by reducing the experienced oxidative stress [92]. Long-chain acyl-CoA synthetases (ACSs) participated in plant wax biosynthesis, and overexpressing ACS transgenic plants were more tolerant to salt stress through their accumulation of cuticular wax together with reduced water loss rates and epidermal permeability [93,94]. Similarly, Eceriferum1 (CER1) and Eceriferum3 (WAX2) also played pivotal roles in cuticular wax biosynthesis, and the transgenic lines of CER1- and WAX2-overexpression affected the plants’ response to the biotic and abiotic stresses by altering the composition and ultrastructure of the cuticular waxes [95,96]. Furthermore, given that saline–alkaline stress often results in the generation and accumulation of ROS, the plants had evolved a set of scavenging systems to maintain the intracellular ROS balance consisting of antioxidant enzymes, such as peroxidase (POD), catalase (CAT), glutathione peroxidase (GPX), and glutathione S-transferase (GST) [3,4]. GST-overexpressing plants induced the accumulation of proline and soluble sugar together with the expression of three ROS detoxification-related genes (CAT, POD, and SOD), thereby enhancing the activity of CAT, POD, and SOD, which, in turn, reduced the content of ROS and MDA, ultimately promoting the resistance to saline–alkaline stress [97]. Ubiquitination played vital roles in the plant stress responses, and the overexpressed UBC gene encoding a ubiquitin-conjugating enzyme contributed to the salt and drought stress tolerance by adjusting the transcription level of the genes related to the oxidative stress responsive, osmolyte synthesis, and ion homeostasis in the plants [98]. Similarly, S-phase kinase-associated protein 1 (SKP1), as one of the main important components of the SKP1/Cullin/F-box (SCF) complex, was involved in multiple functions, including control of the cell cycle progression, signal transduction, and transcriptional regulation. The overexpression of SKP1 conferred salinity tolerance by enhancing the expression of the genes’ involvement in plant reproduction, growth, and stress response [99]. Additionally, Os02g12510 and Os03g0160600 were identified as novel gene homologs of PR5 (pathogenesis-related protein 5) and 5PTase4 (phosphatidylinositol 5-phosphatases 4) under study, respectively. The PRs were regarded as a group of heterogeneous proteins encoded by the genes that were rapidly regulated by environmental factors consisting of pathogenic infections, abiotic stresses, and hormones (e.g., SA, JA, ABA, brassinosteroid, and ethylene), suggesting that they played important roles in protecting the plants against certain stress responses (e.g., pathogens, salt, and drought) and developmental processes [100,101]. 5PTases were key components of the intracellular vesicle-trafficking system involved in plant salt tolerance through the coordination of the ROS production, endocytosis, calcium influx, and the induction of stress-responsive genes [102,103]. In all, from these transcriptome findings under study, it can be speculated that the molecular mechanism of saline–alkaline tolerance in rice is a very complicated processes controlled by multiple genes’ involvement in the plant signal transduction, sugar and fatty acid metabolism, amino acid and secondary metabolite biosynthesis, ubiquitin-mediated proteolysis, etc., suggesting that the mechanism behind the saline–alkaline resistance in rice may be the result of a series of cascade regulations of the genes’ involvement in multiple metabolic pathways. In particular, the transcription factors NAC (Os02g0285900) and WRKYs (Os07g0111400 and Os08g0235800) were highly co-expressed with the genes related to plant signal transduction (PR5 (Os02g12510), BRI1 (Os02g0154200), and FLS2 (Os04g0349700)), secondary metabolite biosynthesis (HCT (Os08g0543400) and 4CL (Os04g0473900)), glutathione metabolism (GPX (Os06g0185900)), ubiquitin-mediated proteolysis (UBC (Os05g0182500) and Os09g0272900 (RPS2)), fatty acid metabolism (Os03g0108500 (SMO1-1)), biosynthesis of amino acids (Os04g0107500 (HPR)), as well as sugar metabolism (Os08g0237200 (MPG)), indicating that the three transcription factors may have potential functions in regulating these metabolism-related genes in response to saline–alkaline stress (Figure 6C). However, it is worth further investigating this issue to obtain a full understanding of the mechanism behind these gene regulations in response to saline–alkaline tolerance.
The metabolome analysis results in this study showed that more specific DAMs were observed in RPY geng than in Chao2R in the upregulated fraction, suggesting a clear apparent difference between the two rice genotypes in response to the saline–alkaline stress at the metabolic level, especially lipids, alkaloids, and phenolic acids (Figure 4). Lipids, as one of the main constituents of the plasma membrane, play crucial roles in plant resistance to stress responses through modulating the fluidity of the cell membranes, along with their roles in lipid-dependent signaling cascades [90,91,104]. It has been reported that lysophosphatidylcholine (lysoPC) and lysophosphatidylethanolamine (LysoPE), acting as the major components of lipids, are rapidly induced by various biotic or abiotic stress [105,106], in addition to the accumulation of unsaturated fatty acids contributing to the cell membrane fluidity [104]. In this study, lipids (19) were the largest part of the metabolites that were specifically upregulated by the saline–alkaline stress in RPY geng (saline–alkaline-tolerant genotype) (Figure 6B), such as 5 free fatty acids, 5 phosphatidylcholines (1 PC and 4 lysoPCs), and 5 LysoPEs, suggesting these lipids’ involvement in the RPY geng tolerance to saline–alkaline stress by enhancing the fluidity of the plasma membrane. Increasing reports suggest that phenolic compounds and their derivatives (collectively termed salicylic acids, ferulic acids, caffeic acids, etc.), as endogenously active secondary metabolites, play crucial roles in the plants’ defense against various abiotic and biotic stresses due to their roles in antioxidant and cell wall remodeling, together with their roles in improving the metabolic activities [107,108]. Moreover, ferulic acids and caffeic acids, also as the downstream metabolites of the TCA cycle, are involved in the process of energy metabolism [91]. In our findings, 14 phenolic acids (including salicylic acids, ferulic acids, caffeic acids, and their corresponding derivatives) were also specifically induced in the saline–alkaline-tolerant genotypes (RPY geng), suggesting that these phenolic compounds may participate in the saline–alkaline resistance of rice by maintaining ROS homeostasis and enhancing the activity of energy metabolism. Recent studies revealed that polyamines (including mainly putrescine, spermidine, and spermine) had potential roles in protecting the plants from various abiotic stresses through the modulation of hormones, ROS, membrane stability, amino acid and carbon metabolism, and ion channels [109,110]. Among them, putrescine, as a free radical scavenger, was supposed to elevate the membrane stability through altering the fatty acid level and to protect the membrane from oxidative stress damage through regulating the antioxidant system under abiotic stress [109]. Spermidine, acting as a stress-protecting compound, has been shown to alleviate plant tolerance towards various abiotic stresses through elevating the activities of the antioxidants or stabilizing the structural integrity of the photosynthetic apparatus [111,112]. Herein, nine alkaloids, consisting mostly of putrescine, spermidine, and their related derivatives, were also specifically accumulated by the saline–alkaline stress in RPY geng, suggesting that these alkaloid compounds may be involved in the saline–alkaline tolerance of rice through alleviating oxidative stress damage and stabilizing the structural integrity of the membrane or photosynthetic apparatus. Additionally, 11 other metabolites were also observed to be specifically upregulated in RPY geng under saline–alkaline stress, including 3 amino acids and their derivatives, 4 lignans, 1 coumarin, and 3 vitamins. Previous studies revealed that amino acids participated in the response to various abiotic stresses by serving as the precursor molecule for the synthesis of secondary metabolites and protein and functioning as effective antioxidants for ROS detoxification, in addition to their roles in osmotic adjustment [113]. Coumarins and lignans, as a group of phenylpropanoids, had important functions in improving the tolerance to multiple stresses for the plant’s interaction with its environment through the modulation of the cell wall, ROS scavenging capacity, or semiochemical role [114,115]. Plant-derived vitamins were essential for the metabolism in the plant’s response to its environment because of their role as enzymatic cofactors, redox chemistry, and important antioxidants [116]. Overall, based on the metabolome analysis results, it can be reasonably inferred that the capacities of osmotic potential, energy metabolism, plasma membrane fluidity, ROS scavenging, as well as total antioxidant activity, are the main factors driving the difference between RPY geng and Chao2R in response to saline–alkaline stress.
Based on comprehensive phenotype, physiology, transcriptome, and metabolome analyses of RPY geng (a saline–alkaline-tolerant genotype) and Chao2R (a saline–alkaline-sensitive genotype), we propose a hypothetical model of rice saline–alkaline adaption: the external saline–alkaline condition is firstly perceived through a complicated network of signal transduction pathways, including hormone signaling and MAPK cascade pathways; then, it activates the downstream transcription factors, thereby activating these metabolic processes consisting mainly of lipids, alkaloids, and phenolic acids by inducing the expression of the saline–alkaline-responsive genes directly, ultimately promoting saline–alkaline tolerance by increasing the capacities of osmotic potential, energy metabolism, plasma membrane fluidity, ROS scavenging, as well as total antioxidant (Figure 7).

4. Materials and Methods

4.1. Plant Materials, Growth, Treatment, and Sample Collection

The plant materials used under study include Luohui 9 (short for Chao2R, Oryza sativa ssp. indica) and RPY geng (Oryza sativa ssp. japonica). Seeds were firstly germinated on 1/4 MS medium containing 0.7% agar at 28 °C for 2 days. The germinated seedlings were then transferred to 96-well plate black polypropylene containers (127 mm × 87 mm × 114 mm) with Yoshida solution in the Wuhan University greenhouse under 14 h light (200 µmol m−2 s−1, 28 °C)/10 h dark (22 °C), with approximately 70% relative humidity. At 15 days after sowing (15 DAS), half the seedlings were transferred to Yoshida solution supplemented with 0.3% (w/v) mixtures of alkaline salts (Na2CO3:NaHCO3 = 1:3) to mimic the saline–alkaline stress conditions with a high pH value (9.35) and high Na+ concentration (approx. 40.84 mM) for 72 h. The remaining seedlings grown under original Yoshida solution were used as an untreated control (CK). Shoot samples were collected in the light after treatment (between 6 and 8 h), rapidly frozen in liquid N2, and stored at −80 °C prior to analysis. Three independent biological replicates (at least fifteen seedlings/each replicate) were applied for RNA and metabolite extractions.
To evaluate the saline–alkaline stress tolerance of Chao2R and RPY geng, 15 DAS seedlings were transferred to Yoshida solution containing 0.3% (w/v) mixtures of alkaline salts (Na2CO3:NaHCO3 = 1:3) for 10 days. After saline–alkaline treatment, the seedlings were cultured in the control Yoshida solution for one week recovery, and then the final survival rate of the seedlings were calculated on the two rice genotypes.

4.2. Measurements of Phenotypic and Physiological Traits

Images of the rice seedlings (at least 10 plants) from treatment and control groups were taken with a Nikon D7100 camera (Nikon, Tokyo, Japan), and shoot height was then analyzed as the distance from the base of the shoot of the seedling to the tip of the tallest leaf blade of each plant using ImageJ 1.50i software (National Institutes of Health, Bethesda, MD, USA). The harvested fresh rice seedlings (excluding roots) were immediately measured as shoot fresh weight. For shoot dry weight (DW) measurement, these harvested fresh samples were oven-dried for 1 h at 105 °C and for 5 days at 80 °C until they reached constant mass. Relative growth rates (RGRs) between 6 and 25 DAS were calculated as (ln(DWn+1) − ln(DWn))/(tn+1 − tn) from mean shoot biomass upon control and saline–alkaline treatments. To analyze the physiological traits of the shoot of the rice seedlings with and without saline–alkaline stress treatment, the following physiological parameters were measured according to the manufacturer’s protocols of their corresponding assay kits (Suzhou Comin Biotechnology Co., Ltd., Suzhou, China): the content of proline (Pro), soluble sugar (SS), and malondialdehyde (MDA), the activity of superoxide dismutase (SOD), peroxidase (POD), and catalase (CAT), as well as the total antioxidant capacity (T-AOC). Briefly, the proline content measurement was as follows, according to the previous study with some modification [117]: proline of shoot samples (0.1 g frozen weight) was extracted with 1 mL 3% sulfosalicylic acid at 90 °C for 10 min. After centrifugation at 10,000× g for 10 min, the cooling supernatant was treated with 0.5 mL acetic acid and 0.5 mL 2.5% ninhydrin solution in a boiling water bath for 30min, then 1 mL toluene was added, and the absorbance was determined at 520 nm. The soluble sugar content was performed based on the previous report with minor modifications [118]: soluble sugar content of shoot samples (0.1 g frozen weight) was extracted with 1 mL distilled water at 95 °C for 10 min, and then the cooling homogenate was centrifuged at 8000× g for 10 min, followed by the supernatant being filled to 10 mL with distilled water. The 0.2 mL supernatant reacted with 0.1 mL 2% anthrone solution, 1 mL 98% sulfuric acid, and 0.2 mL distilled water in a water bath at 95 °C for 10 min. After cooling, the absorbance of the mixture was measured at 620 nm. The malondialdehyde (MDA) content was measured by the thiobarbituric acid (TBA) reaction method as described by Dhindsa et al. (1981) with some modifications [119]: shoot samples (0.1 g frozen weight) were homogenized in 1 mL of 5% cold trichloroacetic acid (TCA) and centrifuged at 8000× g for 10 min at 4 °C. The 0.1 mL supernatant was mixed with 0.3 mL of 0.67% TBA in a water bath at 95 °C for 30 min and then immediately cooled on ice. After centrifugation at 10,000× g for 10 min, the absorbance of supernatant was determined at 532 nm and 600 nm, respectively. To assay the activity of antioxidant enzymes, including superoxide dismutase (SOD), peroxidase (POD), and catalase (CAT), shoot samples (0.1 g frozen weight) were homogenized in 1 mL of 50 mM cold phosphate buffer for extraction of antioxidant enzymes. After centrifugation at 8000× g for 10 min at 4 °C, the supernatant was subsequently used for measurement of antioxidant enzyme activity. For SOD activity assay, 90 μL supernatant reacted with 0.936 mL nitro blue tetrazolium (NBT) reaction solution for 30 min, and then the absorbance of the reaction mixture was determined at 560 nm. One unit of SOD activity was assayed by measuring its ability to inhibit 50% of NBT photoreduction at 560 nm as previously described [119]. For POD activity assay, 50 μL supernatant was added to 0.95 mL guaiacol reaction solution containing 20 mM phosphate buffer, 20 mM guaiacol, and 40 mM H2O2, and then changes of the absorbance of the reaction buffer were read at 470 nm every 1 min [120]. One unit of POD activity was defined as the amount of enzyme that would increase 0.01 absorbance unit at 470 nm min−1 g−1 fresh weight. For CAT activity assay, 35 μL supernatant was added to 1 mL reaction buffer containing 15 mM phosphate buffer and 15 mM H2O2. Reaction was initiated immediately once enzyme extract was added, followed by reading changes of the absorbance of the reaction buffer at 240 nm every 1 min [119]. One unit of CAT activity was defined as the amount of enzyme that would decrease 1nmol H2O2 min−1 g−1 fresh weight. For total antioxidant capacity (T-AOC) assay, shoot samples (0.1 g frozen weight) were homogenized in 1 mL of 50 mM cold phosphate solution. After centrifugation at 10,000× g for 10 min at 4 °C, 50 μL supernatant was mixed and reacted with 950 μL ferric-reducing ability of plasma (FRAP) solution for 20 min, and then the absorbance of the mixture was recorded at 593 nm using a spectrophotometer [121].

4.3. RNA Sequencing (RNA-seq) and Data Analysis

To investigate the transcription levels’ changes of Chao2 and RPY geng under saline–alkaline stress conditions, RNA extraction and cDNA library construction of 12 shoots of Chao2 and RPY geng from three independent biological replicates for each genotype with and without saline–alkaline stress treatment were used for RNA-seq on the BGISEQ sequencing platform by Metware Biotechnology Co., Ltd. (Wuhan, China). Total RNA was extracted from the shoots using ethanol precipitation protocol and CTAB-PBIOZOL reagent following the manufacturer’s instructions. RNA integrity was monitored using a Nano Drop and Agilent 2100 bioanalyzer (Thermo Fisher Scientific, Waltham, MA, USA). Messenger RNA (mRNA) was purified from total RNA using oligo (dT- attached magnetic beads. Short fragments of mRNA were conducted by using fragment buffer at appropriate temperature. First-strand cDNA was synthesized using the mRNA fragments as templates under random hexamer primer and M-MuLV Reverse Transcriptase (RNase H). Subsequently, second-strand cDNA synthesis was performed by using buffer, dNTPs, RNase H, and DNA polymerase I. Double-stranded cDNA were purified and processed via AMPure XP beads for end repair with addition of A-tailing mix and RNA index adapters. The cDNA fragments were purified with AMPure XP system (Beckman Coulter, Beverly, CA, USA) and enriched by PCR amplification. The final transcriptome libraries were constructed from the double-stranded PCR products, which were heated, denatured, and circularized by the splint oligo sequence, and library products’ quality was checked using the Agilent Bioanalyzer 2100 system. The single-strand circle DNA (ssCir DNA) was formatted as the final library. The final library was amplified with phi29 to make DNA nanoball (DNB) which consisted of more than 300 copies of one molecule. DNBs were loaded into the patterned nanoarray, and single-end 50-base reads were produced for sequencing analysis on BGISEQ MGISEQ-2000RS platform (BGI, Shenzhen, China). The obtained original raw data (raw reads) were cleaned using fastp v 0.19.3 by removing reads containing adapter, reads containing the ploy-N content at more than 10%, and low-quality sequences (Q ≤ 20, the number of low-quality bases contained in more than 50% in a read). The obtained clean reads were mapped to the rice reference genome available at ftp://ftp.ensemblgenomes.org/pub/plants/release-50/fasta/oryza_sativa/dna/ (accessed on 2 June 2021) using HISAT v2.1.0 tool soft, and the new genes were predicted using StringTie v1.3.4d software. In order to eliminate the influence of different genes in lengths and sequencing discrepancies, quantification of gene expression levels was estimated by the fragments per kilobase of exon per million mapped fragments (FPKM) method using featureCounts v1.6.2 software. The false discovery rate (FDR) was used as the corrected p–value threshold in multiple tests using the Benjamini–Hochberg method. In this study, the FDR ≤ 0.05 and |log2 fold change with FPKM| ≥ 1 were used as the cutoff values for significant differences in gene expression between groups using DESeq2 v1.22.1. To obtain more information of the differentially expressed genes (DEGs), the enrichment analyses consisting of Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) were carried out by following the hypergeometric test, with an FDR ≤ 0.05 as a threshold for significance.

4.4. Quantitative Reverse Transcription Polymerase Chain Reaction (qRT-PCR) Analysis

In order to validate the accuracy of transcriptome data, the expression of 20 selected genes, including 10 randomly selected genes and ten DEGs, were determined by using quantitative real-time reverse transcription polymerase chain reaction (qRT-PCR) assay. Primers of these 20 selected genes for qRT-PCR are listed in Table S5. For quantification of these genes, samples of Chao2 and RPY geng used were the same as for RNA-seq. Total RNA extraction and first-strand cDNA synthesis and qRT-PCR reaction (10 µL), and the relative expression levels of 20 query genes were performed as previously described [122].

4.5. Metabolomics and Data Analysis

The samples for RNA-seq were next used for wide–targeted metabolic profiling using an ultra-performance liquid chromatography electrospray ionization tandem mass spectrometry (UPLC-ESI-MS/MS) (UPLC, Nexera X2, Shimadzu, Kyoto, Japan; MS, 4500 Q TRAP, Applied Biosystems, Waltham, MA, USA) platform by Metware Biotechnology Co., Ltd. according to the standard instructions. Briefly, shoot samples were freeze-dried using a vacuum freeze-dryer (Scientz-100F) and crushed into a powder using a mixer mill (MM 400, Retsch, Germany) with a zirconia bead for 1.5 min at 30 Hz. Lyophilized powder (100 mg) was weighed and extracted overnight at 4 °C with 1.2 mL of 70% methanol solution. After centrifugation at 12,000× g rpm for 10 min, the supernatants of the samples were filtrated using 0.22 μm pore size (ANPEL, Shanghai, China) for the next UPLC-MS/MS analysis. During the conditions of UPLC analysis, the sample effluents were alternatively connected to an ESI-triple quadrupole-linear ion trap (QTRAP)-MS. Linear ion trap (LIT) and triple quadrupole (QQQ) scans were acquired for a UPLC-MS/MS system, which was equipped with an ESI Turbo Ion-Spray interface and controlled by Analyst 1.6.3 software (AB Sciex, Framingham, MA, USA), with the operation of positive and negative ion modes. Qualitative identification of metabolites under study was obtained according to the source of public database (e.g., MassBank, KNAPSAcK, and METLIN) and private MVDB database (Metware Biotechnology Co., Ltd.). After the scaled unit variance of data, unsupervised principal component analysis (PCA) was conducted with the statistics function ‘prcomp’ in R package. The hierarchical cluster analysis (HCA) and Pearson correlation coefficients (PCCs) were both presented by the ‘pheatmap’ within R, and PCCs were calculated using the ‘cor’ function in R package. The values of variable importance in projection (VIP) were obtained from the result of orthogonal partial least squares discriminant analysis (OPLS-DA) using ‘MetaboAnalystR’ within R package. The VIP ≥ 1 and |log2 fold change (FC)| ≥ 1 were set as the threshold to determine significantly differentially accumulated metabolites (DAMs) between groups. To obtain more functions of identified metabolites, these metabolites were annotated with the KEGG compound database and were next mapped to the KEGG pathway database. Metabolite sets enrichment analyses (MSEAs) were carried out with a hypergeometric test’s p-values ≤ 0.05 as the cutoff value for significance.

4.6. Statistical Analyses

All data were analyzed by two-way ANOVA using SPSS17.0 or GraphPad Prism 8. Significant differences between the multiple treatment groups were evaluated using Student’s t-test (p < 0.05). All heatmaps of expression levels in the study were carried out by TBtools [123] or R package. A co-expression network of candidate genes as well as an interactive network between candidate genes and metabolites were carried out using Cytoscape v3.8.2 software by following the screening criteria of the absolute value of |PCC| ≥ 0.9 and p < 0.01.

5. Conclusions

In this study, we explored a potential molecular mechanism difference in salt–alkaline tolerance between indica Chao2R and japonica RPY geng at the seedling stage through the investigation of changes in genes and metabolites by integrating phenotype, physiology, transcriptome, and metabolome analyses. Compared with Chao2R, RPY geng is more inclined to activate these metabolic processes, including mainly lipids, alkaloids, and phenolic acids, through a series of cascade regulations of these genes mainly involved in plant signal transduction pathways (hormone signaling and MAPK cascade pathways), such as PR, FLS2, VIP1, BRI1, AUX/IAA, JAZ, and GID1; transcription factors, such as NAC and WRKYs; and saline–alkaline-responsive genes, such as 4CL, GPX, GST, MPG, HK, GAD, ACX, LOX, and ACS, so as to promote saline–alkaline stress tolerance by improving the abilities of ROS scavenging, osmotic potential, energy metabolism, and plasma membrane fluidity, etc. The findings of this study will be valuable for providing a mechanistic understanding of the rice saline–alkaline tolerance at the transcriptomic and metabolomic levels.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241512387/s1.

Author Contributions

Conceptualization, validation, writing—review and editing, and methodology, J.W. (Jianyong Wang) and Y.L.; funding acquisition, project administration, resources, supervision, Y.L.; portion of software, visualization, and data curation, K.H., J.W. (Jien Wang), Z.G., S.L. and X.D.; data curation, formal analysis, investigation, visualization, software, and writing—original draft preparation, J.W. (Jianyong Wang) All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported from the National Key Research and Development Program of China (No. 2016YFD0100400) and the National Special Key Project for Transgenic Breeding, China (No. 2016ZX08001001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data of the RNA-seq from the next generation sequence under study were deposited into the National Center for Biotechnology Information (NCBI) Sequence Reads Archive (SRA) database under accession number PRJNA961320.

Acknowledgments

We gratefully acknowledge the anonymous referees for their valuable advice and constructive comments on this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, Y.; Guo, Y. Elucidating the molecular mechanisms mediating plant salt-stress responses. New Phytol. 2018, 217, 523–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Cao, Y.; Zhang, M.; Liang, X.; Li, F.; Shi, Y.; Yang, X.; Jiang, C. Natural variation of an EF-hand Ca2+-binding-protein coding gene confers saline-alkaline tolerance in maize. Nat. Commun. 2020, 11, 186. [Google Scholar] [CrossRef] [Green Version]
  3. Fang, S.; Hou, X.; Liang, X. Response mechanisms of plants under saline-alkali stress. Front. Plant Sci. 2021, 12, 667458. [Google Scholar] [CrossRef]
  4. Rao, Y.; Peng, T.; Xue, S. Mechanisms of plant saline-alkaline tolerance. J. Plant Physiol. 2023, 281, 153916. [Google Scholar] [CrossRef]
  5. Ismail, A.M.; Horie, T. Genomics, physiology, and molecular breeding approaches for improving salt tolerance. Annu. Rev. Plant Biol. 2017, 68, 405–434. [Google Scholar] [CrossRef] [Green Version]
  6. Deinlein, U.; Stephan, A.B.; Horie, T.; Luo, W.; Xu, G.; Schroeder, J.I. Plant salt-tolerance mechanisms. Trends Plant Sci. 2014, 19, 371–379. [Google Scholar] [CrossRef] [Green Version]
  7. Munns, R.; Tester, M. Mechanisms of salinity tolerance. Annu. Rev. Plant Biol. 2008, 59, 651–681. [Google Scholar] [CrossRef] [Green Version]
  8. Van Zelm, E.; Zhang, Y.; Testerink, C. Salt tolerance mechanisms of plants. Annu. Rev. Plant Biol. 2020, 71, 403–433. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, X.; Ren, H.; Wei, Z.; Wang, Y.; Ren, W. Effects of neutral salt and alkali on ion distributions in the roots, shoots, and leaves of two alfalfa cultivars with differing degrees of salt tolerance. J. Integr. Agric. 2017, 16, 1800–1807. [Google Scholar] [CrossRef] [Green Version]
  10. Chen, Y.; Li, Y.; Sun, P.; Chen, G.; Xin, J. Interactive effects of salt and alkali stresses on growth, physiological responses and nutrient (N, P) removal performance of Ruppia maritima. Ecol. Eng. 2017, 104, 177–183. [Google Scholar] [CrossRef]
  11. Guan, Q.; Liao, X.; He, M.; Li, X.; Wang, Z.; Ma, H.; Yu, S.; Liu, S. Tolerance analysis of chloroplast OsCu/Zn-SOD overexpressing rice under NaCl and NaHCO3 stress. PLoS ONE 2017, 12, e0186052. [Google Scholar] [CrossRef] [Green Version]
  12. Zhang, H.; Yu, F.; Xie, P.; Sun, S.; Qiao, X.; Tang, S.; Chen, C.; Yang, S.; Mei, C.; Yang, D.; et al. A Gγ protein regulates alkaline sensitivity in crops. Science 2023, 379, eade8416. [Google Scholar] [CrossRef] [PubMed]
  13. Chuamnakthong, S.; Nampei, M.; Ueda, A. Characterization of Na+ exclusion mechanism in rice under saline-alkaline stress conditions. Plant Sci. 2019, 287, 110171. [Google Scholar] [CrossRef]
  14. Ji, H.; Pardo, J.M.; Batelli, G.; Van Oosten, M.J.; Bressan, R.A.; Li, X. The Salt Overly Sensitive (SOS) pathway: Established and emerging roles. Mol. Plant 2013, 6, 275–286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Cosse, M.; Seidel, T. Plant proton pumps and cytosolic pH-homeostasis. Front. Plant Sci. 2021, 12, 672873. [Google Scholar] [CrossRef]
  16. Li, Q.; Yang, A.; Zhang, W.H. Efficient acquisition of iron confers greater tolerance to saline-alkaline stress in rice (Oryza sativa L.). J. Exp. Bot. 2016, 67, 6431–6444. [Google Scholar] [CrossRef] [Green Version]
  17. Zhu, J.K. Abiotic stress signaling and responses in plants. Cell 2016, 167, 313–324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Shah, W.H.; Rasool, A.; Saleem, S.; Mushtaq, N.U.; Tahir, I.; Hakeem, K.R.; Rehman, R.U. Understanding the integrated pathways and mechanisms of transporters, protein kinases, and transcription factors in plants under salt stress. Int. J. Genom. 2021, 2021, 5578727. [Google Scholar] [CrossRef] [PubMed]
  19. Zhang, M.; Zhang, S. Mitogen-activated protein kinase cascades in plant signaling. J. Integr. Plant Biol. 2022, 64, 301–341. [Google Scholar] [CrossRef]
  20. Zhu, M.; Shabala, L.; Cuin, T.A.; Huang, X.; Zhou, M.; Munns, R.; Shabala, S. Nax loci affect SOS1-like Na+/H+ exchanger expression and activity in wheat. J. Exp. Bot. 2016, 67, 835–844. [Google Scholar] [CrossRef] [Green Version]
  21. Cuin, T.A.; Bose, J.; Stefano, G.; Jha, D.; Tester, M.; Mancuso, S.; Shabala, S. Assessing the role of root plasma membrane and tonoplast Na+/H+ exchangers in salinity tolerance in wheat: In planta quantification methods. Plant Cell Environ. 2011, 34, 947–961. [Google Scholar] [CrossRef]
  22. El Mahi, H.; Pérez-Hormaeche, J.; De Luca, A.; Villalta, I.; Espartero, J.; Gámez-Arjona, F.; Fernández, J.L.; Bundó, M.; Mendoza, I.; Mieulet, D.; et al. A critical role of sodium flux via the plasma membrane Na+/H+ exchanger SOS1 in the salt tolerance of rice. Plant Physiol. 2019, 180, 1046–1065. [Google Scholar] [CrossRef] [Green Version]
  23. Chen, X.; Ding, Y.; Yang, Y.; Song, C.; Wang, B.; Yang, S.; Guo, Y.; Gong, Z. Protein kinases in plant responses to drought, salt, and cold stress. J. Integr. Plant Biol. 2021, 63, 53–78. [Google Scholar] [CrossRef] [PubMed]
  24. Ye, Y.; Ding, Y.; Jiang, Q.; Wang, F.; Sun, J.; Zhu, C. The role of receptor-like protein kinases (RLKs) in abiotic stress response in plants. Plant Cell Rep. 2017, 36, 235–242. [Google Scholar] [CrossRef] [PubMed]
  25. Ciura, J.; Kruk, J. Phytohormones as targets for improving plant productivity and stress tolerance. J. Plant Physiol. 2018, 229, 32–40. [Google Scholar] [CrossRef]
  26. Bari, R.; Jones, J.D. Role of plant hormones in plant defence responses. Plant Mol. Biol. 2009, 69, 473–488. [Google Scholar] [CrossRef] [PubMed]
  27. Verma, V.; Ravindran, P.; Kumar, P.P. Plant hormone-mediated regulation of stress responses. BMC Plant Biol. 2016, 16, 86. [Google Scholar] [CrossRef] [Green Version]
  28. Signor, S.A.; Nuzhdin, S.V. The evolution of gene expression in cis and trans. Trends Genet. 2018, 34, 532–544. [Google Scholar] [CrossRef]
  29. Ganapati, R.K.; Naveed, S.A.; Zafar, S.; Wang, W.; Xu, J. Saline-alkali tolerance in rice: Physiological response, molecular mechanism, and QTL identification and application to breeding. Rice Sci. 2022, 29, 412–434. [Google Scholar]
  30. Liu, J.; Wan, J.; Yin, X.; Gu, X.; Yin, H.; Yu, M.; Luan, Y.; Li, R.; Mi, T.; Li, J. Progress and prospect of developing salt and alkali tolerant rice using hybrid rice technology in China. Plant Breed. 2023, 142, 1–10. [Google Scholar] [CrossRef]
  31. Xu, Z.; Shao, T.; Lv, Z.; Yue, Y.; Liu, A.; Long, X.; Zhou, Z.; Gao, X.; Rengel, Z. The mechanisms of improving coastal saline soils by planting rice. Sci. Total Environ. 2020, 703, 135529. [Google Scholar] [CrossRef]
  32. Li, J.; Pu, L.; Han, M.; Zhu, M.; Zhang, R.; Xiang, Y. Soil salinization research in China: Advances and prospects. J. Geogr. Sci. 2014, 24, 943–960. [Google Scholar] [CrossRef]
  33. Li, N.; Zheng, H.; Cui, J.; Wang, J.; Liu, H.; Sun, J.; Liu, T.; Zhao, H.; Lai, Y.; Zou, D. Genome-wide association study and candidate gene analysis of alkalinity tolerance in japonica rice germplasm at the seedling stage. Rice 2019, 12, 24. [Google Scholar] [CrossRef]
  34. Li, N.; Liu, H.; Sun, J.; Zheng, H.; Wang, J.; Yang, L.; Zhao, H.; Zou, D. Transcriptome analysis of two contrasting rice cultivars during alkaline stress. Sci. Rep. 2018, 8, 9586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Li, Q.; Ma, C.; Tai, H.; Qiu, H.; Yang, A. Comparative transcriptome analysis of two rice genotypes differing in their tolerance to saline-alkaline stress. PLoS ONE 2020, 15, e0243112. [Google Scholar] [CrossRef] [PubMed]
  36. Qian, G.; Wang, M.; Wang, X.; Liu, K.; Li, Y.; Bu, Y.; Li, L. Integrated transcriptome and metabolome analysis of rice leaves response to high saline–alkali stress. Int. J. Mol. Sci. 2023, 24, 4062. [Google Scholar] [PubMed]
  37. Li, Q.; Yang, A.; Zhang, W. Higher endogenous bioactive gibberellins and α-amylase activity confer greater tolerance of rice seed germination to saline-alkaline stress. Environ. Exp. Bot. 2019, 162, 357–363. [Google Scholar] [CrossRef]
  38. Li, X.; Zheng, H.; Wu, W.; Liu, H.; Wang, J.; Jia, Y.; Li, J.; Yang, L.; Lei, L.; Zou, D.; et al. QTL mapping and candidate gene analysis for alkali tolerance in japonica rice at the bud stage based on linkage mapping and genome-wide association study. Rice 2020, 13, 48. [Google Scholar] [CrossRef]
  39. Kong, W.; Deng, X.; Yang, J.; Zhang, C.; Sun, T.; Ji, W.; Zhong, H.; Fu, X.; Li, Y. High-resolution bin-based linkage mapping uncovers the genetic architecture and heterosis-related loci of plant height in indica-japonica derived populations. Plant J. 2022, 110, 814–827. [Google Scholar]
  40. Kong, W.; Sun, T.; Zhang, C.; Deng, X.; Li, Y. Comparative transcriptome analysis reveals the mechanisms underlying differences in salt tolerance between indica and japonica rice at seedling stage. Front. Plant Sci. 2021, 12, 725436. [Google Scholar]
  41. Umemoto, T.; Yano, M.; Satoh, H.; Shomura, A.; Nakamura, Y. Mapping of a gene responsible for the difference in amylopectin structure between japonica-type and indica-type rice varieties. Theor. Appl. Genet. 2002, 104, 1–8. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, Q.; Liang, Z.; Cui, X.; Ji, C.; Li, Y.; Zhang, P.; Liu, J.; Riaz, A.; Yao, P.; Liu, M.; et al. N6-Methyladenine DNA methylation in japonica and indica rice genomes and its association with gene expression, plant development, and stress responses. Mol. Plant 2018, 11, 1492–1508. [Google Scholar] [CrossRef] [Green Version]
  43. Shang, J.; Tao, Y.; Chen, X.; Zou, Y.; Lei, C.; Wang, J.; Li, X.; Zhao, X.; Zhang, M.; Lu, Z.; et al. Identification of a new rice blast resistance gene, Pid3, by genomewide comparison of paired nucleotide-binding site--leucine-rich repeat genes and their pseudogene alleles between the two sequenced rice genomes. Genetics 2009, 182, 1303–1311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Lv, Y.; Guo, Z.; Li, X.; Ye, H.; Li, X.; Xiong, L. New insights into the genetic basis of natural chilling and cold shock tolerance in rice by genome-wide association analysis. Plant Cell Environ. 2016, 39, 556–570. [Google Scholar] [PubMed] [Green Version]
  45. Hori, K.; Suzuki, K.; Ishikawa, H.; Nonoue, Y.; Nagata, K.; Fukuoka, S.; Tanaka, J. Genomic regions involved in differences in eating and cooking quality other than Wx and Alk genes between indica and japonica rice cultivars. Rice 2021, 14, 8. [Google Scholar] [CrossRef] [PubMed]
  46. Sun, H.; Qian, Q.; Wu, K.; Luo, J.; Wang, S.; Zhang, C.; Ma, Y.; Liu, Q.; Huang, X.; Yuan, Q.; et al. Heterotrimeric G proteins regulate nitrogen-use efficiency in rice. Nat. Genet. 2014, 46, 652–656. [Google Scholar] [CrossRef]
  47. Zhang, G. Prospects of utilization of inter-subspecific heterosis between indica and japonica rice. J. Integr. Agric. 2020, 19, 1–10. [Google Scholar] [CrossRef]
  48. Tsikas, D. Assessment of lipid peroxidation by measuring malondialdehyde (MDA) and relatives in biological samples: Analytical and biological challenges. Anal. Biochem. 2017, 524, 13–30. [Google Scholar]
  49. Sharma, P.; Jha, A.B.; Dubey, R.S.; Pessarakli, M. Reactive oxygen species, oxidative damage, and antioxidative defense mechanism in plants under stressful conditions. J. Bot. 2012, 2012, 1–26. [Google Scholar]
  50. Ma, C.; Li, Q.; Song, Z.; Su, L.; Tao, W.; Zhou, B.; Wang, Q. Irrigation with magnetized water alleviates the harmful effect of saline-alkaline stress on rice seedlings. Int. J. Mol. Sci. 2022, 23, 10048. [Google Scholar] [CrossRef]
  51. Zhang, C.; Meng, W.; Wang, Y.; Zhou, Y.; Wang, S.; Qi, F.; Wang, N.; Ma, J. Comparative analysis of physiological, hormonal and transcriptomic responses reveal mechanisms of saline-alkali tolerance in autotetraploid rice (Oryza sativa L.). Int. J. Mol. Sci. 2022, 23, 16146. [Google Scholar] [CrossRef]
  52. Campbell, M.T.; Bandillo, N.; Al Shiblawi, F.R.A.; Sharma, S.; Liu, K.; Du, Q.; Schmitz, A.J.; Zhang, C.; Véry, A.A.; Lorenz, A.J.; et al. Allelic variants of OsHKT1;1 underlie the divergence between indica and japonica subspecies of rice (Oryza sativa) for root sodium content. PLoS Genet. 2017, 13, e1006823. [Google Scholar]
  53. Mir, R.R.; Reynolds, M.; Pinto, F.; Khan, M.A.; Bhat, M.A. High-throughput phenotyping for crop improvement in the genomics era. Plant Sci. 2019, 282, 60–72. [Google Scholar] [PubMed]
  54. Sun, T.; Ma, N.; Wang, C.; Fan, H.; Wang, M.; Zhang, J.; Cao, J.; Wang, D. A golgi-localized sodium/hydrogen exchanger positively regulates salt tolerance by maintaining higher K+/Na+ ratio in soybean. Front. Plant Sci. 2021, 12, 638340. [Google Scholar] [PubMed]
  55. Sun, T.; Fan, L.; Yang, J.; Cao, R.; Yang, C.; Zhang, J.; Wang, D. A Glycine max sodium/hydrogen exchanger enhances salt tolerance through maintaining higher Na+ efflux rate and K+/Na+ ratio in Arabidopsis. BMC Plant Biol. 2019, 19, 469. [Google Scholar]
  56. Zou, Q.; Xu, H.; Yang, G.; Yu, L.; Jiang, H.; Mao, Z.; Hu, J.; Zhang, Z.; Wang, N.; Chen, X. MdbHLH106-like transcription factor enhances apple salt tolerance by upregulating MdNHX1 expression. Plant Cell Tissue Organ Cult. 2021, 145, 333–345. [Google Scholar] [CrossRef]
  57. Rehman, N.; Khan, M.R.; Abbas, Z.; Rafique, R.S.; Zaynab, M.; Qasim, M.; Noor, S.; Inam, S.; Ali, G.M. Functional characterization of Mitogen-Activated Protein Kinase Kinase (MAPKK) gene in Halophytic Salicornia europaea against salt stress. Environ. Exp. Bot. 2020, 171, 103934. [Google Scholar] [CrossRef]
  58. Zhou, H.; Xiao, F.; Zheng, Y.; Liu, G.; Zhuang, Y.; Wang, Z.; Zhang, Y.; He, J.; Fu, C.; Lin, H. PAMP-INDUCED SECRETED PEPTIDE 3 modulates salt tolerance through RECEPTOR-LIKE KINASE 7 in plants. Plant Cell 2022, 34, 927–944. [Google Scholar] [CrossRef]
  59. Chen, J.; Li, X.; Ye, X.; Guo, P.; Hu, Z.; Qi, G.; Cui, F.; Liu, S. An S-ribonuclease binding protein EBS1 and brassinolide signaling are specifically required for Arabidopsis tolerance to bicarbonate. J. Exp. Bot. 2021, 72, 1449–1459. [Google Scholar]
  60. Belkhadir, Y.; Jaillais, Y.; Epple, P.; Balsemão-Pires, E.; Dangl, J.L.; Chory, J. Brassinosteroids modulate the efficiency of plant immune responses to microbe-associated molecular patterns. Proc. Natl. Acad. Sci. USA 2012, 109, 297–302. [Google Scholar] [CrossRef]
  61. Ueguchi-Tanaka, M.; Ashikari, M.; Nakajima, M.; Itoh, H.; Katoh, E.; Kobayashi, M.; Chow, T.Y.; Hsing, Y.I.; Kitano, H.; Yamaguchi, I.; et al. GIBBERELLIN INSENSITIVE DWARF1 encodes a soluble receptor for gibberellin. Nature 2005, 437, 693–698. [Google Scholar] [CrossRef]
  62. Gräfe, K.; Schmitt, L. The ABC transporter G subfamily in Arabidopsis thaliana. J. Exp. Bot. 2021, 72, 92–106. [Google Scholar] [PubMed]
  63. Xuan, L.; Liang, W.; Li, Y.; Li, M.; Liu, C.; Li, C. Transcriptome analysis reveals the effects of alkali stress on root system architecture and endogenous hormones in apple rootstocks. J. Integr. Agric. 2019, 18, 2264–2271. [Google Scholar]
  64. Li, J.; Xu, H.H.; Liu, W.C.; Zhang, X.W.; Lu, Y.T. Ethylene inhibits root elongation during alkaline stress through AUXIN1 and associated changes in auxin accumulation. Plant Physiol. 2015, 168, 1777–1791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Wu, J.; Cao, J.; Su, M.; Feng, G.; Xu, Y.; Yi, H. Genome-wide comprehensive analysis of transcriptomes and small RNAs offers insights into the molecular mechanism of alkaline stress tolerance in a citrus rootstock. Hortic. Res. 2019, 6, 33. [Google Scholar] [CrossRef] [Green Version]
  66. Wang, F.; Xu, H.; Zhang, L.; Shi, Y.; Song, Y.; Wang, X.; Cai, Q.; He, W.; Xie, H.; Zhang, J. The lipoxygenase OsLOX10 affects seed longevity and resistance to saline-alkaline stress during rice seedlings. Plant Mol. Biol. 2023, 111, 415–428. [Google Scholar]
  67. Zhu, D.; Cai, H.; Luo, X.; Bai, X.; Deyholos, M.K.; Chen, Q.; Chen, C.; Ji, W.; Zhu, Y. Over-expression of a novel JAZ family gene from Glycine soja, increases salt and alkali stress tolerance. Biochem. Biophys. Res. Commun. 2012, 426, 273–279. [Google Scholar] [CrossRef]
  68. Zhao, G.; Song, Y.; Wang, Q.; Yao, D.; Li, D.; Qin, W.; Ge, X.; Yang, Z.; Xu, W.; Su, Z.; et al. Gossypium hirsutum salt tolerance is enhanced by overexpression of G. arboreum JAZ1. Front. Bioeng. Biotechnol. 2020, 8, 157. [Google Scholar] [CrossRef]
  69. Wang, L.; Li, Z.; Lu, M.; Wang, Y. ThNAC13, a NAC transcription factor from Tamarix hispida, confers salt and osmotic stress tolerance to transgenic Tamarix and Arabidopsis. Front. Plant Sci. 2017, 8, 635. [Google Scholar] [CrossRef] [Green Version]
  70. Li, M.; Chen, R.; Jiang, Q.; Sun, X.; Zhang, H.; Hu, Z. GmNAC06, a NAC domain transcription factor enhances salt stress tolerance in soybean. Plant Mol. Biol. 2021, 105, 333–345. [Google Scholar] [CrossRef]
  71. Zhang, L.; Zhang, L.; Xia, C.; Zhao, G.; Jia, J.; Kong, X. The novel wheat transcription factor TaNAC47 enhances multiple abiotic stress tolerances in transgenic plants. Front. Plant Sci. 2015, 6, 1174. [Google Scholar] [CrossRef] [Green Version]
  72. Cao, L.; Yu, Y.; Ding, X.; Zhu, D.; Yang, F.; Liu, B.; Sun, X.; Duan, X.; Yin, K.; Zhu, Y. The Glycine soja NAC transcription factor GsNAC019 mediates the regulation of plant alkaline tolerance and ABA sensitivity. Plant Mol. Biol. 2017, 95, 253–268. [Google Scholar] [CrossRef] [PubMed]
  73. Wang, X.; Ajab, Z.; Liu, C.; Hu, S.; Liu, J.; Guan, Q. Overexpression of transcription factor SlWRKY28 improved the tolerance of Populus davidiana × P. bolleana to alkaline salt stress. BMC Genet. 2020, 21, 103. [Google Scholar] [CrossRef]
  74. Wang, Y.; Jiang, L.; Chen, J.; Tao, L.; An, Y.; Cai, H.; Guo, C. Overexpression of the alfalfa WRKY11 gene enhances salt tolerance in soybean. PLoS ONE 2018, 13, e0192382. [Google Scholar] [CrossRef] [Green Version]
  75. Zhao, S.; Gao, H.; Jia, X.; Li, X.; Mao, K.; Ma, F. The γ-clade HD-Zip I transcription factor MdHB-7 regulates salt tolerance in transgenic apple (Malus domestica). Plant Soil 2021, 463, 509–522. [Google Scholar] [CrossRef]
  76. Qu, D.; Show, P.L.; Miao, X. Transcription factor ChbZIP1 from Alkaliphilic Microalgae Chlorella sp. BLD enhancing alkaline tolerance in transgenic Arabidopsis thaliana. Int. J. Mol. Sci. 2021, 22, 2387. [Google Scholar] [CrossRef]
  77. Zhang, H.; Guo, J.; Chen, X.; Zhou, Y.; Pei, Y.; Chen, L.; Ul Haq, S.; Lu, M.; Gong, H.; Chen, R. Pepper bHLH transcription factor CabHLH035 contributes to salt tolerance by modulating ion homeostasis and proline biosynthesis. Hortic. Res. 2022, 9, uhac203. [Google Scholar] [CrossRef]
  78. Sun, M.H.; Ma, Q.J.; Hu, D.G.; Zhu, X.P.; You, C.X.; Shu, H.R.; Hao, Y.J. The glucose sensor MdHXK1 phosphorylates a tonoplast Na+/H+ exchanger to improve salt tolerance. Plant Physiol. 2018, 176, 2977–2990. [Google Scholar] [CrossRef] [Green Version]
  79. Pérez-Díaz, J.; Batista-Silva, W.; Almada, R.; Medeiros, D.B.; Arrivault, S.; Correa, F.; Bastías, A.; Rojas, P.; Beltrán, M.F.; Pozo, M.F. Prunus Hexokinase 3 genes alter primary C-metabolism and promote drought and salt stress tolerance in Arabidopsis transgenic plants. Sci. Rep. 2021, 11, 7098. [Google Scholar] [CrossRef] [PubMed]
  80. Lugassi, N.; Yadav, B.S.; Egbaria, A.; Wolf, D.; Kelly, G.; Neuhaus, E.; Raveh, E.; Carmi, N.; Granot, D. Expression of Arabidopsis hexokinase in tobacco guard cells increases water-use efficiency and confers tolerance to drought and salt stress. Plants 2019, 8, 613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Yi, S.Y.; Ku, S.S.; Sim, H.J.; Kim, S.K.; Park, J.H.; Lyu, J.I.; So, E.J.; Choi, S.Y.; Kim, J.; Ahn, M.S. An alcohol dehydrogenase gene from Synechocystis sp. confers salt tolerance in transgenic tobacco. Front. Plant Sci. 2017, 8, 1965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Shi, H.; Liu, W.; Yao, Y.; Wei, Y.; Chan, Z. Alcohol dehydrogenase 1 (ADH1) confers both abiotic and biotic stress resistance in Arabidopsis. Plant Sci. 2017, 262, 24–31. [Google Scholar] [CrossRef] [PubMed]
  83. Zhu, W.; Miao, Q.; Sun, D.; Yang, G.; Wu, C.; Huang, J.; Zheng, C. The mitochondrial phosphate transporters modulate plant responses to salt stress via affecting ATP and gibberellin metabolism in Arabidopsis thaliana. PLoS ONE 2012, 7, e43530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Lang, T.; Deng, C.; Yao, J.; Zhang, H.; Wang, Y.; Deng, S. A salt-signaling network involving ethylene, extracellular ATP, hydrogen peroxide, and calcium mediates K+/Na+ homeostasis in Arabidopsis. Int. J. Mol. Sci. 2020, 21, 8683. [Google Scholar] [CrossRef]
  85. Kumar, R.; Mustafiz, A.; Sahoo, K.K.; Sharma, V.; Samanta, S.; Sopory, S.K.; Pareek, A.; Singla-Pareek, S.L. Functional screening of cDNA library from a salt tolerant rice genotype Pokkali identifies mannose-1-phosphate guanyl transferase gene (OsMPG1) as a key member of salinity stress response. Plant Mol. Biol. 2012, 79, 555–568. [Google Scholar] [CrossRef] [PubMed]
  86. Akama, K.; Akihiro, T.; Kitagawa, M.; Takaiwa, F. Rice (Oryza sativa) contains a novel isoform of glutamate decarboxylase that lacks an authentic calmodulin-binding domain at the C-terminus. Biochim. Biophys. Acta 2001, 1522, 143–150. [Google Scholar] [CrossRef]
  87. Wu, X.; Jia, Q.; Ji, S.; Gong, B.; Li, J.; Lü, G.; Gao, H. Gamma-aminobutyric acid (GABA) alleviates salt damage in tomato by modulating Na+ uptake, the GAD gene, amino acid synthesis and reactive oxygen species metabolism. BMC Plant Biol. 2020, 20, 465. [Google Scholar] [CrossRef]
  88. Gan, T.; Lin, Z.; Bao, L.; Hui, T.; Cui, X.; Huang, Y.; Wang, H.; Su, C.; Jiao, F.; Zhang, M. Comparative proteomic analysis of tolerant and sensitive varieties reveals that phenylpropanoid biosynthesis contributes to salt tolerance in mulberry. Int. J. Mol. Sci. 2021, 22, 9402. [Google Scholar] [CrossRef]
  89. Zhu, Y.; Wang, Q.; Wang, Y.; Xu, Y.; Li, J.; Zhao, S.; Wang, D.; Ma, Z.; Yan, F.; Liu, Y. Combined transcriptomic and metabolomic analysis reveals the role of phenylpropanoid biosynthesis pathway in the salt tolerance process of Sophora alopecuroides. Int. J. Mol. Sci. 2021, 22, 2399. [Google Scholar] [CrossRef]
  90. Hou, Q.; Ufer, G.; Bartels, D. Lipid signalling in plant responses to abiotic stress. Plant Cell Environ. 2016, 39, 1029–1048. [Google Scholar] [CrossRef] [Green Version]
  91. Sun, S.; Fang, J.; Lin, M.; Hu, C.; Qi, X.; Chen, J.; Zhong, Y.; Muhammad, A.; Li, Z.; Li, Y. Comparative metabolomic and transcriptomic studies reveal key metabolism pathways contributing to freezing tolerance under cold stress in kiwifruit. Front. Plant Sci. 2021, 12, 628969. [Google Scholar] [CrossRef]
  92. Shiraku, M.L.; Magwanga, R.O.; Cai, X.; Kirungu, J.N.; Xu, Y.; Mehari, T.G.; Hou, Y.; Wang, Y.; Agong, S.G.; Peng, R. Functional characterization of GhACX3 gene reveals its significant role in enhancing drought and salt stress tolerance in cotton. Front. Plant Sci. 2021, 12, 658755. [Google Scholar] [CrossRef]
  93. Zhang, C.L.; Zhang, Y.L.; Hu, X.; Xiao, X.; Wang, G.L.; You, C.X.; Li, Y.Y.; Hao, Y.J. An apple long-chain acyl-CoA synthetase, MdLACS4, induces early flowering and enhances abiotic stress resistance in Arabidopsis. Plant Sci. 2020, 297, 110529. [Google Scholar] [CrossRef] [PubMed]
  94. Zhang, C.L.; Hu, X.; Zhang, Y.L.; Liu, Y.; Wang, G.L.; You, C.X.; Li, Y.Y.; Hao, Y.J. An apple long-chain acyl-CoA synthetase 2 gene enhances plant resistance to abiotic stress by regulating the accumulation of cuticular wax. Tree Physiol. 2020, 40, 1450–1465. [Google Scholar] [CrossRef] [PubMed]
  95. Bourdenx, B.; Bernard, A.; Domergue, F.; Pascal, S.; Léger, A.; Roby, D.; Pervent, M.; Vile, D.; Haslam, R.P.; Napier, J.A.; et al. Overexpression of Arabidopsis ECERIFERUM1 promotes wax very-long-chain alkane biosynthesis and influences plant response to biotic and abiotic stresses. Plant Physiol. 2011, 156, 29–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Wang, W.; Liu, X.; Gai, X.; Ren, J.; Liu, X.; Cai, Y.; Wang, Q.; Ren, H. Cucumis sativus L. WAX2 plays a pivotal role in wax biosynthesis, influencing pollen fertility and plant biotic and abiotic stress responses. Plant Cell Physiol. 2015, 56, 1339–1354. [Google Scholar] [CrossRef] [Green Version]
  97. Du, B.; Zhao, W.; An, Y.; Li, Y.; Zhang, X.; Song, L.; Guo, C. Overexpression of an alfalfa glutathione S-transferase gene improved the saline-alkali tolerance of transgenic tobacco. Biol. Open 2019, 8, bio043505. [Google Scholar] [CrossRef] [Green Version]
  98. Zhou, G.A.; Chang, R.Z.; Qiu, L.J. Overexpression of soybean ubiquitin-conjugating enzyme gene GmUBC2 confers enhanced drought and salt tolerance through modulating abiotic stress-responsive gene expression in Arabidopsis. Plant Mol. Biol. 2010, 72, 357–367. [Google Scholar] [CrossRef] [Green Version]
  99. Hao, Q.; Ren, H.; Zhu, J.; Wang, L.; Huang, S.; Liu, Z.A.; Gao, Z.; Shu, Q. Overexpression of PSK1, a SKP1-like gene homologue, from Paeonia suffruticosa, confers salinity tolerance in Arabidopsis. Plant Cell Rep. 2017, 36, 151–162. [Google Scholar] [CrossRef]
  100. Onishi, M.; Tachi, H.; Kojima, T.; Shiraiwa, M.; Takahara, H. Molecular cloning and characterization of a novel salt-inducible gene encoding an acidic isoform of PR-5 protein in soybean (Glycine max [L.] Merr.). Plant Physiol. Biochem. 2006, 44, 574–580. [Google Scholar] [CrossRef]
  101. Seo, P.J.; Lee, A.K.; Xiang, F.; Park, C.M. Molecular and functional profiling of Arabidopsis pathogenesis-related genes: Insights into their roles in salt response of seed germination. Plant Cell Physiol. 2008, 49, 334–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Golani, Y.; Kaye, Y.; Gilhar, O.; Ercetin, M.; Gillaspy, G.; Levine, A. Inositol polyphosphate phosphatidylinositol 5-phosphatase9 (At5ptase9) controls plant salt tolerance by regulating endocytosis. Mol. Plant 2013, 6, 1781–1794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Kaye, Y.; Golani, Y.; Singer, Y.; Leshem, Y.; Cohen, G.; Ercetin, M.; Gillaspy, G.; Levine, A. Inositol polyphosphate 5-phosphatase 7 regulates the production of reactive oxygen species and salt tolerance in Arabidopsis. Plant Physiol. 2011, 157, 229–241. [Google Scholar] [CrossRef] [Green Version]
  104. Yaeno, T.; Matsuda, O.; Iba, K. Role of chloroplast trienoic fatty acids in plant disease defense responses. Plant J. 2004, 40, 931–941. [Google Scholar] [CrossRef] [PubMed]
  105. Kim, M.J.; Lee, M.Y.; Shon, J.C.; Kwon, Y.S.; Liu, K.H.; Lee, C.H.; Ku, K.M. Untargeted and targeted metabolomics analyses of blackberries—Understanding postharvest red drupelet disorder. Food Chem. 2019, 300, 125169. [Google Scholar] [CrossRef] [PubMed]
  106. Narvaez-Vasquez, J.; Florin-Christensen, J.; Ryan, C.A. Positional specificity of a phospholipase A activity induced by wounding, systemin, and oligosaccharide elicitors in tomato leaves. Plant Cell 1999, 11, 2249–2260. [Google Scholar] [CrossRef] [Green Version]
  107. Shalaby, S.; Horwitz, B.A. Plant phenolic compounds and oxidative stress: Integrated signals in fungal-plant interactions. Curr. Genet. 2015, 61, 347–357. [Google Scholar] [CrossRef]
  108. Sharma, A.; Shahzad, B.; Rehman, A.; Bhardwaj, R.; Landi, M.; Zheng, B. Response of Phenylpropanoid pathway and the role of polyphenols in plants under abiotic stress. Molecules 2019, 24, 2452. [Google Scholar] [CrossRef] [Green Version]
  109. Kou, S.; Chen, L.; Tu, W.; Scossa, F.; Wang, Y.; Liu, J.; Fernie, A.R.; Song, B.; Xie, C. The arginine decarboxylase gene ADC1, associated to the putrescine pathway, plays an important role in potato cold-acclimated freezing tolerance as revealed by transcriptome and metabolome analyses. Plant J. 2018, 96, 1283–1298. [Google Scholar] [CrossRef] [Green Version]
  110. Shi, H.; Chan, Z. Improvement of plant abiotic stress tolerance through modulation of the polyamine pathway. J. Integr. Plant Biol. 2014, 56, 114–121. [Google Scholar] [CrossRef]
  111. Jiang, D.; Hou, J.; Gao, W.; Tong, X.; Li, M.; Chu, X.; Chen, G. Exogenous spermidine alleviates the adverse effects of aluminum toxicity on photosystem II through improved antioxidant system and endogenous polyamine contents. Ecotoxicol. Environ. Saf. 2021, 207, 111265. [Google Scholar] [CrossRef] [PubMed]
  112. Moschou, P.N.; Paschalidis, K.A.; Delis, I.D.; Andriopoulou, A.H.; Lagiotis, G.D.; Yakoumakis, D.I.; Roubelakis-Angelakis, K.A. Spermidine exodus and oxidation in the apoplast induced by abiotic stress is responsible for H2O2 signatures that direct tolerance responses in tobacco. Plant Cell 2008, 20, 1708–1724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Cai, J.; Aharoni, A. Amino acids and their derivatives mediating defense priming and growth tradeoff. Curr. Opin. Plant Biol. 2022, 69, 102288. [Google Scholar] [CrossRef]
  114. Stringlis, I.A.; de Jonge, R.; Pieterse, C.M.J. The age of coumarins in plant-microbe interactions. Plant Cell Physiol. 2019, 60, 1405–1419. [Google Scholar] [CrossRef] [Green Version]
  115. Paniagua, C.; Bilkova, A.; Jackson, P.; Dabravolski, S.; Riber, W.; Didi, V.; Houser, J.; Gigli-Bisceglia, N.; Wimmerova, M.; Budínská, E.; et al. Dirigent proteins in plants: Modulating cell wall metabolism during abiotic and biotic stress exposure. J. Exp. Bot. 2017, 68, 3287–3301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Asensi-Fabado, M.A.; Munné-Bosch, S. Vitamins in plants: Occurrence, biosynthesis and antioxidant function. Trends Plant Sci. 2010, 15, 582–592. [Google Scholar] [CrossRef]
  117. Bates, L.; Waldren, R.A.; Teare, I. Rapid determination of free proline for water-stress studies. Plant Soil 1973, 39, 205–207. [Google Scholar] [CrossRef]
  118. Hansen, J.; Moller, I. Percolation of starch and soluble carbohydrates from plant tissue for quantitative determination with anthrone. Anal. Biochem. 1975, 68, 87–94. [Google Scholar] [CrossRef]
  119. Dhindsa, R.S.; Plumb-Dhindsa, P.; Thorpe, T.A. Leaf senescence: Correlated with increased levels of membrane permeability and lipid peroxidation, and decreased levels of superoxide dismutase and catalase. J. Exp. Bot. 1981, 32, 93–101. [Google Scholar] [CrossRef]
  120. Kochba, J.; Lavee, S.; Spiegel-Roy, P. Differences in peroxidase activity and isoenzymes in embryogenic ane non-embryogenic ‘Shamouti’orange ovular callus lines. Plant Cell Physiol. 1977, 18, 463–467. [Google Scholar] [CrossRef]
  121. Benzie, I.F.; Strain, J.J. The ferric reducing ability of plasma (FRAP) as a measure of “antioxidant power”: The FRAP assay. Anal. Biochem. 1996, 239, 70–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Wang, J.; Jiang, Y.; Sun, T.; Zhang, C.; Liu, X.; Li, Y. Genome-wide classification and evolutionary analysis reveal diverged patterns of chalcone isomerase in plants. Biomolecules 2022, 12, 961. [Google Scholar] [CrossRef] [PubMed]
  123. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools: An integrative toolkit developed for interactive analyses of big biological data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Effects of saline–alkaline stress on Chao2R and RPY geng at the seedlings. (A) Growth phenotype of both rice cultivars with and without saline–alkaline treatment. Scale bars = 3 cm. (B) The final survival rate of the two rice cultivars after 10 days of saline–alkaline treatment followed by 7 days for recovery at normal conditions. Asterisks represent significance between groups by Student’s t-test (*** p < 0.001).
Figure 1. Effects of saline–alkaline stress on Chao2R and RPY geng at the seedlings. (A) Growth phenotype of both rice cultivars with and without saline–alkaline treatment. Scale bars = 3 cm. (B) The final survival rate of the two rice cultivars after 10 days of saline–alkaline treatment followed by 7 days for recovery at normal conditions. Asterisks represent significance between groups by Student’s t-test (*** p < 0.001).
Ijms 24 12387 g001
Figure 2. Comparative transcriptome analysis of Chao2R and RPY geng seedlings under saline–alkaline stress conditions. (A) FPKM box diagram of all tested 12 samples under study. Chao2R and RPY indicate samples of Chao2R and RPY geng under normal conditions, respectively; Chao2RA and RPYA indicate samples of Chao2R and RPY geng treated with saline–alkaline stress for 72 h, respectively. (B) Principal component analysis (PCA) of transcriptome data from the two rice cultivars with and without saline–alkaline stress. (C) Number of differential expression genes (DEGs) in Chao2R and RPY geng after exposure to saline–alkaline condition. 0 and A represent the untreated control (CK) and the saline–alkaline treatment, respectively. (D) Venn diagram of upregulated DEGs (left) and Venn diagram of downregulated DEGs (right) in Chao2R and RPY geng.
Figure 2. Comparative transcriptome analysis of Chao2R and RPY geng seedlings under saline–alkaline stress conditions. (A) FPKM box diagram of all tested 12 samples under study. Chao2R and RPY indicate samples of Chao2R and RPY geng under normal conditions, respectively; Chao2RA and RPYA indicate samples of Chao2R and RPY geng treated with saline–alkaline stress for 72 h, respectively. (B) Principal component analysis (PCA) of transcriptome data from the two rice cultivars with and without saline–alkaline stress. (C) Number of differential expression genes (DEGs) in Chao2R and RPY geng after exposure to saline–alkaline condition. 0 and A represent the untreated control (CK) and the saline–alkaline treatment, respectively. (D) Venn diagram of upregulated DEGs (left) and Venn diagram of downregulated DEGs (right) in Chao2R and RPY geng.
Ijms 24 12387 g002
Figure 3. Top 20 Gene Ontology (GO) enrichment of DEGs from Chao2R and RPY geng under saline–alkaline stress for 72 h in comparison with untreated control (short for 0_vs_A). Upregulated DEGs in Chao2R (A), upregulated DEGs in RPY geng (B), downregulated DEGs in Chao2R (C), downregulated DEGs in RPY geng (D).
Figure 3. Top 20 Gene Ontology (GO) enrichment of DEGs from Chao2R and RPY geng under saline–alkaline stress for 72 h in comparison with untreated control (short for 0_vs_A). Upregulated DEGs in Chao2R (A), upregulated DEGs in RPY geng (B), downregulated DEGs in Chao2R (C), downregulated DEGs in RPY geng (D).
Ijms 24 12387 g003
Figure 4. Comparative metabolomic analysis of Chao2R and RPY geng at the seedlings under saline–alkaline stress treatment. (A) Classification of the identified 801 metabolites under study. (B) Clustering heatmap of all metabolites. Chao2R and RPY indicate samples of Chao2R and RPY geng without saline–alkaline treatment, respectively; Chao2RA and RPYA indicate samples of Chao2R and RPY geng with saline–alkaline stress treatment for 72 h, respectively. (C) PCA of metabolome data from the two rice cultivars with and without saline–alkaline stress treatment. (D) Number of differentially accumulated metabolites (DAMs) in Chao2R and RPY geng after subjected to saline–alkaline condition in comparison with untreated control (short for 0_vs_A). (E) Venn diagram of upregulated DAMs (left) and Venn diagram of downregulated DAMs (right) in Chao2R and RPY geng. (F) Statistics numbers (left) and log2 (fold change) values (right) of DAMs distribution for different types of metabolites in Chao2R and RPY geng treated with salt–alkaline stress in comparison with untreated control. Asterisks represent significance between the groups by Student’s t-test (* p < 0.05 and ** p < 0.01).
Figure 4. Comparative metabolomic analysis of Chao2R and RPY geng at the seedlings under saline–alkaline stress treatment. (A) Classification of the identified 801 metabolites under study. (B) Clustering heatmap of all metabolites. Chao2R and RPY indicate samples of Chao2R and RPY geng without saline–alkaline treatment, respectively; Chao2RA and RPYA indicate samples of Chao2R and RPY geng with saline–alkaline stress treatment for 72 h, respectively. (C) PCA of metabolome data from the two rice cultivars with and without saline–alkaline stress treatment. (D) Number of differentially accumulated metabolites (DAMs) in Chao2R and RPY geng after subjected to saline–alkaline condition in comparison with untreated control (short for 0_vs_A). (E) Venn diagram of upregulated DAMs (left) and Venn diagram of downregulated DAMs (right) in Chao2R and RPY geng. (F) Statistics numbers (left) and log2 (fold change) values (right) of DAMs distribution for different types of metabolites in Chao2R and RPY geng treated with salt–alkaline stress in comparison with untreated control. Asterisks represent significance between the groups by Student’s t-test (* p < 0.05 and ** p < 0.01).
Ijms 24 12387 g004
Figure 5. Main metabolic pathways of Chao2R and RPY geng seedlings in response to saline–alkaline stress. Metabolites and genes were mapped to an integrated systemic metabolic pathway diagram based on combining the KEGG pathways of DAMs and DEGs. The first and second boxes indicate the log2 (fold change) of metabolites in Chao2R and RPY geng, respectively, after subjected to saline–alkaline treatment in contrast to untreated control, in which red indicates upregulation, and blue indicates downregulation. The metabolic pathways involvement in alkaloid, phenolic acid, lipid, amino acid, organic acid, saccharide, and plant signal transduction are marked in purple, green, blue, orange, black, light green, and red, respectively.
Figure 5. Main metabolic pathways of Chao2R and RPY geng seedlings in response to saline–alkaline stress. Metabolites and genes were mapped to an integrated systemic metabolic pathway diagram based on combining the KEGG pathways of DAMs and DEGs. The first and second boxes indicate the log2 (fold change) of metabolites in Chao2R and RPY geng, respectively, after subjected to saline–alkaline treatment in contrast to untreated control, in which red indicates upregulation, and blue indicates downregulation. The metabolic pathways involvement in alkaloid, phenolic acid, lipid, amino acid, organic acid, saccharide, and plant signal transduction are marked in purple, green, blue, orange, black, light green, and red, respectively.
Ijms 24 12387 g005
Figure 6. Potential important candidate genes and metabolites for rice saline–alkaline tolerance. (A) Expression patterns of important candidate genes. The data in the boxes indicate the log2 (fold change) of genes in Chao2R and RPY geng after exposure to saline–alkaline condition compared with untreated control (short for 0_vs_A). The upregulation and downregulation are masked in blue and red, respectively. (B) The expression heatmap showing the log2 (fold change) of important candidate metabolites in Chao2R and RPY geng after exposure to saline–alkaline condition in comparison with untreated control (short for 0_vs_A). Blue and red indicate downregulation and upregulation, respectively. (C) Co-expression network of candidate genes (Pearson correlation coefficient (PCC) ≥ 0.9, p < 0.01). (D) An interactive correlation network of potential important candidate genes and metabolites involved in rice salt–alkaline tolerance (PCC ≥ 0.9, p < 0.01). Red lines indicate the positive correlation, blue solid ellipses indicate metabolites, red solid ellipses indicate genes involved in plant signal transduction pathways, and green solid ellipses indicate genes related to non-signal transduction pathways.
Figure 6. Potential important candidate genes and metabolites for rice saline–alkaline tolerance. (A) Expression patterns of important candidate genes. The data in the boxes indicate the log2 (fold change) of genes in Chao2R and RPY geng after exposure to saline–alkaline condition compared with untreated control (short for 0_vs_A). The upregulation and downregulation are masked in blue and red, respectively. (B) The expression heatmap showing the log2 (fold change) of important candidate metabolites in Chao2R and RPY geng after exposure to saline–alkaline condition in comparison with untreated control (short for 0_vs_A). Blue and red indicate downregulation and upregulation, respectively. (C) Co-expression network of candidate genes (Pearson correlation coefficient (PCC) ≥ 0.9, p < 0.01). (D) An interactive correlation network of potential important candidate genes and metabolites involved in rice salt–alkaline tolerance (PCC ≥ 0.9, p < 0.01). Red lines indicate the positive correlation, blue solid ellipses indicate metabolites, red solid ellipses indicate genes involved in plant signal transduction pathways, and green solid ellipses indicate genes related to non-signal transduction pathways.
Ijms 24 12387 g006
Figure 7. Proposed molecular models for rice saline–alkaline tolerance. Under the saline–alkaline stress, in contrast to Chao2R, RPY geng may be more inclined to active the downstream transcription factors through these genes’ cascade regulation between the plant hormone signaling transduction pathway and the plant MAPK signaling pathway, then activating these metabolic processes, including lipids, alkaloids, and phenolic acids, which ultimately improves the ability of ROS scavenging, osmotic pressure, plasma membrane fluidity, energy metabolism, and total antioxidants.
Figure 7. Proposed molecular models for rice saline–alkaline tolerance. Under the saline–alkaline stress, in contrast to Chao2R, RPY geng may be more inclined to active the downstream transcription factors through these genes’ cascade regulation between the plant hormone signaling transduction pathway and the plant MAPK signaling pathway, then activating these metabolic processes, including lipids, alkaloids, and phenolic acids, which ultimately improves the ability of ROS scavenging, osmotic pressure, plasma membrane fluidity, energy metabolism, and total antioxidants.
Ijms 24 12387 g007
Table 1. Phenotypical and physiological characteristics of Chao2R and RPY geng with and without the saline–alkaline conditions.
Table 1. Phenotypical and physiological characteristics of Chao2R and RPY geng with and without the saline–alkaline conditions.
TraitChao2RRPY Geng
CK0.3%ACK0.3%A
SFW (g)0.235 ± 0.021 a0.198 ± 0.016 a0.456 ± 0.027 c0.308 ± 0.065 b
SDW (g)0.047 ± 0.009 b0.032 ± 0.004 a0.090 ± 0.0045 c0.051 ± 0.008 b
SH (cm)26.388 ± 1.195 a21.511 ± 1.478 b37.436 ± 0.590 c31.979 ± 1.305 d
RGR (g g−1 d−1)0.062 ± 0.007 a0.022 ± 0.003 b0.107 ± 0.002 c0.049 ± 0.004 d
Pro (μg g−1)114.375 ± 5.166 a617.825 ± 31.839 b113.619 ± 6.381 a525.277 ± 10.187 c
SS (mg g−1)9.982 ± 1.259 a11.188 ± 0.404 b10.257 ± 0.150 a,b16.574 ± 0.296 c
MDA (nmol g−1)2.123 ± 0.398 a4.056 ± 0.101 b2.473 ± 0.464 a3.583 ± 0.054 b
SOD (U g−1)1291.054 ± 30.045 a2811.185 ± 118.884 b2432.248 ± 62.447 c3492.370 ± 208.532 d
POD (U g−1)9787.804 ± 80.488 a12,079.593 ± 14.285 b7990.000 ± 130.000 c26,776.920 ± 84.620 d
CAT (nmol min−1 g−1)925.068 ± 121.810 a885.388 ± 58.494 a1689.188 ± 236.332 b1471.082 ± 24.087 b
T-AOC (μmol Trolox g−1)2.721 ± 0.039 a2.817 ± 0.067 a2.633 ± 0.044 a3.939 ± 0.097 b
Data are means ± SE (n ≥ 3). Small letters represent significant difference (p < 0.05) within the same trait. CK represents the untreated control group, 0.3%A represents the saline–alkaline treatment group. SFW: Shoot fresh weight; SDW: Shoot dry weight; SH: Shoot height; RGR: Relative growth rate; Pro: Proline content; SS: Soluble sugar content; MDA: Malondialdehyde content; SOD: Superoxide dismutase; POD: Peroxidase activity; CAT: Catalase activity; T-AOC: Total antioxidant capacity.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, J.; Hu, K.; Wang, J.; Gong, Z.; Li, S.; Deng, X.; Li, Y. Integrated Transcriptomic and Metabolomic Analyses Uncover the Differential Mechanism in Saline–Alkaline Tolerance between Indica and Japonica Rice at the Seedling Stage. Int. J. Mol. Sci. 2023, 24, 12387. https://doi.org/10.3390/ijms241512387

AMA Style

Wang J, Hu K, Wang J, Gong Z, Li S, Deng X, Li Y. Integrated Transcriptomic and Metabolomic Analyses Uncover the Differential Mechanism in Saline–Alkaline Tolerance between Indica and Japonica Rice at the Seedling Stage. International Journal of Molecular Sciences. 2023; 24(15):12387. https://doi.org/10.3390/ijms241512387

Chicago/Turabian Style

Wang, Jianyong, Keke Hu, Jien Wang, Ziyun Gong, Shuangmiao Li, Xiaoxiao Deng, and Yangsheng Li. 2023. "Integrated Transcriptomic and Metabolomic Analyses Uncover the Differential Mechanism in Saline–Alkaline Tolerance between Indica and Japonica Rice at the Seedling Stage" International Journal of Molecular Sciences 24, no. 15: 12387. https://doi.org/10.3390/ijms241512387

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

Article Metrics

Back to TopTop