Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Next Article in Journal
Structural Characteristics and Properties of the RNA-Binding Protein hnRNPK at Multiple Physical States
Previous Article in Journal
Editing the RR-TZF Gene Subfamily in Rice Uncovers Potential Risks of CRISPR/Cas9 for Targeted Genetic Modification
Previous Article in Special Issue
A Maize Calmodulin-like 3 Gene Positively Regulates Drought Tolerance in Maize and Arabidopsis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insights into the Role of GhTAT2 Genes in Tyrosine Metabolism and Drought Stress Tolerance in Cotton

by
Teame Gereziher Mehari
1,
Jungfeng Tang
1,
Haijing Gu
1,
Hui Fang
1,
Jinlei Han
1,
Jie Zheng
2,
Fang Liu
2,
Kai Wang
1,
Dengbing Yao
1,* and
Baohua Wang
1,*
1
School of Life Sciences, Nantong University, Nantong 226019, China
2
State Key Laboratory of Cotton Bio-Breeding and Integrated Utilization, Institute of Cotton Research of Chinese Academy of Agricultural Sciences, Anyang 455000, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(3), 1355; https://doi.org/10.3390/ijms26031355
Submission received: 6 January 2025 / Revised: 2 February 2025 / Accepted: 3 February 2025 / Published: 5 February 2025

Abstract

:
Gossypium hirsutum is a key fiber crop that is sensitive to environmental factors, particularly drought stress, which can reduce boll size, increase flower shedding, and impair photosynthesis. The aminotransferase (AT) gene is essential for abiotic stress tolerance. A total of 3 Gossypium species were analyzed via genome-wide analysis, and the results unveiled 103 genes in G. hirsutum, 47 in G. arboreum, and 53 in G. raimondii. Phylogenetic analysis, gene structure examination, motif analysis, subcellular localization prediction, and promoter analysis revealed that the GhAT genes can be classified into five main categories and play key roles in abiotic stress tolerance. Using RNA-seq expression and KEGG enrichment analysis of GhTAT2, a coexpression network was established, followed by RT-qPCR analysis to identify hub genes. The RT-qPCR results revealed that the genes Gh_A13G1261, Gh_D13G1562, Gh_D10G1155, Gh_A10G1320, and Gh_D06G1003 were significantly upregulated in the leaf and root samples following drought stress treatment, with Gh_A13G1261 identified as the hub gene. The GhTAT2 genes were considerably enriched for tyrosine, cysteine, methionine, and phenylalanine metabolism and isoquinoline alkaloid, tyrosine, tryptophan, tropane, piperidine, and pyridine alkaloid biosynthesis. Under drought stress, KEGG enrichment analysis manifested significant upregulation of amino acids such as L-DOPA, L-alanine, L-serine, L-homoserine, L-methionine, and L-cysteine, whereas metabolites such as maleic acid, p-coumaric acid, quinic acid, vanillin, and hyoscyamine were significantly downregulated. Silencing the GhTAT2 gene significantly affected the shoot and root fresh weights of the plants compared with those of the wild-type plants under drought conditions. RT-qPCR analysis revealed that GhTAT2 expression in VIGS-treated seedlings was lower than that in both wild-type and positive control plants, indicating that silencing GhTAT2 increases sensitivity to drought stress. In summary, this thorough analysis of the gene family lays the groundwork for a detailed study of the GhTAT2 gene members, with a specific focus on their roles and contributions to drought stress tolerance.

1. Introduction

Climate change is a major cause of abiotic and biotic stresses that may adversely affect agricultural productivity to irreversible levels, thereby limiting production growth and jeopardizing sustainable agriculture globally [1]. It is a major cause of biotic and abiotic stresses that negatively affect global agricultural production and productivity [2]. It exacerbates drought conditions around the world, with significant implications for water availability and agricultural systems. It affects the frequency, severity and duration of droughts [3]. Droughts have significant impacts on agricultural production and food security, but global trends in how droughts affect agricultural production remain poorly understood. With more severe droughts expected due to climate change, assessing the vulnerability of agricultural production to these phenomena has become an important area of research [4]. This is the main reason for decreased crop yields caused by abiotic factors globally. This issue leads to food shortages and poses significant challenges for small-scale farmers, who struggle to grow enough grain during periods of low and unpredictable rainfall [5]. Drought stress affects crops differently depending on the growth stage, with vegetative stage stress reducing light interception and grain-filling stage stress causing leaf senescence [6]. This significantly impacts cotton production by reducing growth, photosynthesis, and yield [7].
Cotton is a globally significant crop that is cultivated on more than 30 million hectares, with production exceeding 70 million tons of seed cotton [8]. Gossypium hirsutum, commonly known as upland cotton, is an important fiber crop that is sensitive to various environmental factors. Drought stress is a major concern in cotton cultivation, leading to reduced boll size, increased flower shedding, and impaired photosynthesis. It also causes oxidative damage through the production of reactive oxygen species (ROS), which can be mitigated by natural defense enzymes [9]. Cotton responds to water deficit through various mechanisms, including stomatal closure, osmoregulation, and accumulation of plant growth regulators. At the cellular level, drought stress triggers the overproduction of reactive oxygen species (ROS) and activates signaling pathways involving mitogen-activated protein kinases, Ca2+, and hormones [10]. These responses lead to the expression of stress-related transcription factors and genes, particularly those involved in ROS scavenging and ABA signaling [11]. To mitigate the effects of drought stress, researchers have suggested strategies such as developing drought-tolerant cultivars, applying nutrients and osmoprotectants, and utilizing plant growth-promoting rhizobacteria [10]. Additionally, identifying drought tolerance traits through QTL analysis and transgenic approaches may increase cotton resilience to water scarcity [11].
Aminotransferase genes are vital for improving stress tolerance in crops through the regulation of proline synthesis [12]. Classes of the aminotransferase family are involved in the metabolism of various amino acids, including aromatic amino acid aminotransferases (AAA-ATs) [13], tyrosine aminotransferase (TAT) [14], alanine aminotransferase (ALT) [15], aspartate aminotransferase (AST) [16], and histidinol-phosphate aminotransferase (HisC) [17]. Tyrosine aminotransferase (TAT) is essential for the growth and development of plants. The methyl jasmonate (MeJA)-induced TAT gene reacts to a variety of abiotic stressors. The TAT activity of poplar roots is increased by drought and low nitrogen stress, which alters the amount of RA [18]. TATs are essential for tyrosine metabolism and degradation, with TAT1 and TAT2 exhibiting both distinct and overlapping functions in A. thaliana. The majority of current research indicates that AtTAT1 aids in plant survival in the dark and is involved in tyrosine catabolism [19]. The first enzyme involved in the metabolic breakdown of tyrosine is tyrosine aminotransferase (TATN), which is crucial for tyrosine detoxification and assisting the body in fending off oxidative damage [20]. The TAT gene plays a crucial role in the synthesis of tocopherols in Arabidopsis. The study revealed that knockout mutants presented a substantial reduction in total TAT activity, leading to a significant accumulation of tyrosine and a marked decrease in tocopherol levels. These findings indicate that TAT is important for the utilization of tyrosine in metabolic pathways that produce tocopherols, which are vital for plant health and have significant nutritional value [21].
Multiple studies have highlighted the role of TAT genes across various crop species. SmTAT3-2 is involved in phenolic acid biosynthesis in Salvia miltiorrhiza [22], whereas the overexpression of MdTAT2 improves drought and osmotic stress tolerance in Malus domestica [13]. In Arabidopsis, AtTAT1 and AtTAT2 are linked to tyrosine metabolism and degradation, supporting plant survival [19,23]. Additionally, the AccTATN gene helps Apis cerana cerana respond to pesticide and heavy metal stress [20]. In this study, the GhTAT2 gene from G. hirsutum was cloned for gene knockdown, and the expression patterns of this gene family were analyzed. Recombinant GhTAT2 proteins were subjected to drought stress tests in cotton seedlings, and their role in tyrosine metabolism and other metabolic pathways was examined. This is the first functional characterization of TAT2 in G. hirsutum related to stress tolerance. These findings provide valuable insights into the role of TAT2 genes in the stress response and tyrosine metabolic pathways in cotton.

2. Results

2.1. Genome-Wide Identification of Aminotransferase Genes from Cotton Species

The 3 cotton species included a total of 203 genes, with 103 genes in G. hirsutum (47 in the GhAt subgenome and 53 in the GhDt subgenome) and 47 and 53 genes in the diploid species G. arboreum and G. raimondii, respectively. The A genome species (G. arboreum) and D genome species (G. raimondii) are regarded as progenitors of cultivated G. hirsutum [24]. The lengths of the TAT2 gene proteins ranged from 1068 to 3261 bp in the GhAt subgenome of G. hirsutum, 333 to 3378 bp in the GhDt subgenome, 537 to 3360 bp in G. arboreum, and 570 to 4083 bp in G. raimondii. The protein lengths ranged from 355 to 1086 aa in the G. hirsutum GhAt subgenome, 53 to 1125 aa in the GhDt subgenome, 178 to 1119 aa in G. arboreum, and 170 to 1084 aa in G. raimondii. The molecular weights varied from 38.734 to 121.045 kDa in the GhAt subgenome, 5.688 to 123.409 kDa in the GhDt subgenome, 20.782 to 122.993 kDa in G. arboreum, and 18.705 to 121.023 kDa in G. raimondii. The isoelectric points ranged from 5.03 to 9.495 in the GhAt subgenome, 4.324 to 10.748 in the GhDt subgenome, 5.418 to 9.575 in G. arboreum, and 4.826 to 10.43 in G. raimondii (Supplementary File S1).

2.2. Phylogenetic Tree Classification of Aminotransferase Proteins

This study systematically searched for aminotransferase genes in the genomes of diploid and tetraploid cotton, Arabidopsis, and tobacco via BLAST and identified 103, 47, 53, 28, and 34 aminotransferase genes in G. hirsutum, G. arboreum, G. raimondii, T. cacao, and A. thaliana, respectively. This study focused primarily on tetraploid G. hirsutum to explore the origin and evolution of polyploidy by analyzing aminotransferase genes. Additionally, comparisons with two other cotton species, A. thaliana and T. cacao, which are close relatives of upland cotton, were performed to elucidate their evolutionary history. To trace the evolutionary history of aminotransferase genes, we constructed a phylogenetic tree via protein sequences from multiple alignments of 103 G. hirsutum, 47 G. arboreum, 53 G. raimondii, 28 T. cacao, and 34 A. thaliana crop species. The alignment was performed with CLUSTALX (version 2.0) software, and an unrooted phylogenetic tree was generated via maximum likelihood (ML) via the online program iTOL (version 6) [25]. Among the 3 Gossypium species, the aminotransferase tree formed 5 clusters: 80 proteins encoding aromatic amino acid aminotransferases, 15 encoding tyrosine aminotransferases, 20 encoding alanine aminotransferases, 60 encoding aspartate aminotransferases, and 28 encoding histidine phosphate aminotransferases (Figure 1A). Similarly, of the 5 species, including A. thaliana and T. cacao, 94 proteins clustered in group I, 23 in group II, 26 in group III, 79 in group IV, and 43 in group V, each represented with different colors (Figure 1B).

2.3. Chromosomal Location, Gene Architecture and Conserved Domain Distribution

Chromosomal mapping of AT genes from G. hirsutum (AD)1 and two diploid cotton species, G. arboreum (AA) and G. raimondii (DD), was performed to investigate their structural and evolutionary dynamics related to genomic distribution. Among the 103 G. hirsutum genes, 43 were found on 12 chromosomes of the At subgenome (GhAt), with chromosomes 05 and 01 having the largest and lowest numbers of genes, with 7 and 1 genes, respectively. A total of 53 aminotransferase genes were irregularly distributed across 12 chromosomes in GhDt. The highest and lowest number of genes, 07 and 02, are found on chromosomes 05 and 01, respectively. Neither the GhAt nor the GhDt subgenomes contained any genes on chromosome number four. In both subgenomes, four GhAt genes and three GhDt genes were identified as scaffolds (Figure 2A,B,E). Similarly, in G. arboreum, 47 aminotransferase genes are dispersed at random among the A-genome’s 12 chromosomes, excluding chromosome 04. The greatest number of aminotransferase genes are found on chromosome A05 (8), and the fewest are found on chromosome A02 (1). Except for chromosome D12, all 12 chromosomes of the D-genome in G. raimondii contain 53 aminotransferase genes. D09 has the most genes (9), whereas D02 and D04 have the fewest (2 genes). Neither of the diploid species had scaffolds (Figure 2C,D).
A total of 203 aminotransferase genes were classified into 3 subgroups according to their exon and intron counts. The intron-exon structure of aminotransferase genes was visualized to assess the evolutionary impact on the conservation or divergence of introns and exon numbers in cotton. The results revealed that exon counts ranged from 2 to 15 in the GhAt subgenome of G. hirsutum, 2 to 26 in the GhDt subgenome of G. hirsutum, 2 to 25 in G. arboreum, and 3 to 15 in the aminotransferase gene family of G. raimondii (Figure S1) The arrangement of exons and introns illuminates the evolutionary relationships among different gene family members. Notably, closely related genes within the same phylogenetic clade presented greater structural similarity in terms of intron and exon numbers than did those in other clades. A positive correlation was observed between phylogeny and exon-intron structure. Additionally, a one-intron-less structure was identified in the scaffold region of Gh_Sca206957G01. The absence of introns in these aminotransferase genes suggests a lower likelihood of alternative splicing.
Moreover, all cotton aminotransferase protein sequences were analyzed via MEME to identify coding motifs, revealing that aminotransferase proteins contain between one and ten conserved motifs. Among them, in the GhAt subgenome, motifs 4 and 7; in the GhDt subgenome, motifs 1, 4 and 5; in G. arboreum, motifs 1, 4 and 7; and in G. raimondii, motifs 4 and 5 remained almost conserved in all the proteins across the three species of cotton (Figure 3A–D). Specifically, we identified genes with no motif in the Gh_D10G0019 gene of the GhDt subgenome and the Gorai.009G362800, Gorai.002G090700, Gorai.011G002400 and Gorai.011G091200 genes in the G. raimondii genome (Figure 3D). Furthermore, all aminotransferase genes presented a conserved protein motif distribution pattern. For example, motif number 4 was observed in almost all aminotransferase proteins. Overall, aminotransferase genes show a strong ancestral link in their gene structure and phylogeny, maintaining consistent patterns of gene structure and protein motif distribution across subfamilies.

2.4. Analysis of Cis Regulatory Elements and Their Subcellular Localization

Analysis of the cis-acting regulatory elements in the promoter regions revealed that aminotransferase genes play various roles. ABRE, ARE, MBS, and MYB elements are associated with abscisic acid responsiveness, drought inducibility, and anaerobic induction. Elements such as the G-Box, GATA motif, LTR, and MRE are associated with light and low-temperature responsiveness. Similarly, motifs such as GARE, TGACG, TCA, and TGA are involved in hormone signaling related to auxin, gibberellin, MeJA, and salicylic acid (Figure 4A–D).
According to other prediction analyses, the three Gossypium species are expected to be present in various cellular compartments, including the chloroplast, cytoplasm, cytoskeleton, endoplasmic reticulum, extracellular space, Golgi apparatus, mitochondria, nucleus, peroxisome, plastids, and vacuole. Most genes are located predominantly in the chloroplast, cytoplasm, and nucleus, with fewer genes scattered in the other compartments (Figure 5A–D). The results indicated that most aminotransferase gene family members are highly expressed in the chloroplast, cytoplasm, and nucleus of plant cells, suggesting their significant role in food synthesis through photosynthesis, tolerance against stress, and regulation of cell functions.

2.5. Coexpression Network Analysis of Aminotransferases in G. hirsutum

A total of 25 genes from the aminotransferase gene family involved in various metabolic pathways were used to analyze gene coexpression networks via RNA-seq data from the leaf and root tissues of G. hirsutum (Figure 6A,B). The Cytohubba program in Cytoscape software (version 3.10.2) was used to analyze the regulatory network of aminotransferase gene coexpression, with a p value of 0.95 used as the threshold. Key genes were selected from leaf and root tissues on the basis of the range of topological coefficients associated with each node. Gh_A13G1261, Gh_D13G1562 and Gh_D10G1155 were identified as hub genes from both the leaf and root tissue network analyses (Figure 6C,D).

2.6. Identification of Drought-Responsive Candidate Genes

We selected 12 DEGs from the aminotransferase family involved in metabolite pathway enrichment for real-time quantitative validation. After drought stress treatment, Gh_A13G1261, Gh_D13G1562, Gh_D10G1155, Gh_A10G1320, and Gh_D06G1003 were significantly upregulated in both materials, whereas Gh_A03G1375 and Gh_A13G0469 were significantly downregulated in both tissues. The expression of Gh_A07G0174 and Gh_D06G1069 did not significantly change in root tissue, whereas that of Gh_D02G1624 did significantly change in leaf tissue after drought stress. Gh_A13G1261, a member of the tyrosine aminotransferase class (GhTAT2), was chosen as the candidate gene for functional characterization (Figure 7). These results indicate that drought stress induces unique gene expression trends in cotton seedlings with different drought tolerances, influencing various pathways involved in the drought response and accounting for the observed tolerance variations. The RT-qPCR results for these genes aligned with the transcriptome expression trends. This shows a strong correlation between the transcriptome sequencing and RT-qPCR results, confirming the reliability and effectiveness of the data. In conclusion, the AT gene can rapidly adjust its expression through transcription within 48 h of drought stress to increase drought tolerance in cotton.

2.7. Metabolite Pathway Enrichment Analysis

KEGG enrichment analysis was conducted on G. hirsutum races to identify key and significantly enriched metabolites that are vital for drought stress tolerance. The GhTAT2 gene is involved in and regulates various metabolic pathways, namely, tyrosine metabolism; cysteine and methionine metabolism; isoquinoline alkaloid biosynthesis; phenylalanine metabolism; phenylalanine, tyrosine and tryptophan biosynthesis; and tropane, piperidine and pyridine alkaloid biosynthesis (Figure 8A–F). In cysteine and methionine metabolism, L-alanine, L-serine, L-homoserine, and L-methionine amino acids were upregulated, whereas homocystine was downregulated (Figure 8A). In tyrosine metabolism, phenylalanine, tyrosine and tryptophan biosynthesis amino acids, such as L-phenylalanine, L-DOPA, N-methyltyramine and succinic acid, were significantly upregulated, whereas maleic acid, p-coumaric acid, quinic acid, fructose 1-phosphate, phenylpyruvic acid and anthranilate were significantly downregulated (Figure 8C,D). In phenylalanine metabolism, isoquinoline alkaloid biosynthesis and tropane, piperidine and pyridine alkaloid biosynthesis metabolites such as 2-hydroxycinnamic acid, L-phenylalanine, benzoic acid, putrescine and L-DOPA were significantly upregulated, whereas papaverine, p-coumaric acid, vanillin, hyoscyamine and nicotine were downregulated after drought stress treatment (Figure 8B,F). A list of the metabolite data used for the enrichment analysis is provided in Supplementary File S2.

2.8. VIGS Agroinfiltration, Drought Stress Treatment and Relative Expression Analysis

To assess the role of the GhTAT2 gene (Gh_A13G1261) in drought stress tolerance, GhTAT2 gene silencing was performed on cotton seedlings. The experimental setup included TRV2:00, TRV2:GhTAT2, and TRV2:CLA1, with TRV2:CLA1 exhibiting an albino phenotype in leaf tissues from the second week onward (Figure 9A). Significant differences (p < 0.05) in morphological traits were noted among VIGS-treated plants and positive controls under drought conditions. Although plant height and root length remained similar, the VIGS-treated plants presented significantly lower shoot and root fresh weights than the positive controls did (Figure 9B,C). Samples were collected from the leaves and roots of TRV2:00 and TRV2:GhTAT2 plants to investigate the role of the GhTAT2 gene under drought conditions.

3. Discussion

Cotton (Gossypium hirsutum L.) is a commercially valuable fiber crop grown worldwide in various climates, with increasing demand driven by its use in the textile and oil industries [26]. This versatile crop is vulnerable to biotic and abiotic stresses, especially high temperatures and drought, which can considerably affect yield and quality [27]. Drought stress is a pressing global issue that significantly impacts cotton production. There is an increasing need to identify or breed drought-tolerant varieties to ensure sustainable cotton farming. This stress hinders cotton growth and development by altering metabolic pathways, lowering photosynthesis, and causing responses such as stomatal closure. It decreases photosynthesis and reduces the supply of photosynthates, which results in boll shedding and a low lint yield [10,28]. It impacts several physiological processes in plants, including photosynthesis, stomatal regulation, root–shoot growth ratio, leaf area expansion, transpiration, and osmoregulation. This stress triggers sensing and signaling pathways that activate various parallel responses, including physiological, molecular, and biochemical mechanisms [29,30]. Drought stress at the cellular level triggers excessive production of reactive oxygen species, mitogen-activated protein kinases, Ca2+, and hormone-mediated signaling. It also activates transcription factors involved in both abscisic acid-dependent and abscisic acid-independent stress signaling pathways in cotton. Cotton plants adapt to drought through mechanisms such as the accumulation of reactive oxygen species (ROS) and the activation of stress-responsive transcription factors [7,10].
This study identified 103, 47, 53, 28, and 34 aminotransferase genes in G. hirsutum, G. arboreum, G. raimondii, A. thaliana, and T. cacao, respectively. The number of genes in cottons were greater than Arabidopsis, and tobacco. This is because of the polyploidy nature of cotton species due to interspecific hybridization. There are diploid and tetraploid sets of chromosomes in cotton, while Arabidopsis has only two sets. This extra genetic evolution in cotton gives it more genes for various functions [31]. The phylogenetic analysis grouped the aminotransferase class into 5 different classes, with 94, 23, 26, 79 and 43 aminotransferase proteins in the I, II, III, IV, and V groups, respectively. In the Jinjiang oyster, 18 aminotransferase class I and II genes were identified, and their expression varied under salinity stress [32]. Aminotransferases are categorized into five classes: I, II, III, IV, and V. Classes I and II include alanine aminotransferase, aromatic amino acid aminotransferase, aspartate aminotransferase, and histidine phosphate aminotransferase [32,33]. Aminotransferases are essential enzymes responsible for amino acid metabolism and transfer [33]. The aminotransferase gene family plays crucial roles in amino acid metabolism and osmotic regulation across various organisms [32]. Across species, aminotransferases are involved in crucial cellular processes, including osmotic pressure regulation, nitrogen metabolism, and stress response, highlighting their importance in cellular adaptation and homeostasis [32,34].
In the subcellular localization prediction analysis, most genes were found to be located in the chloroplast, cytoplasm, or nucleus. These cell components play crucial roles in photosynthesis, regulating cell functions, and enhancing stress tolerance. During stress, chloroplasts synthesize essential compounds such as amino acids, vitamins, phytohormones, lipids, nucleotides, and secondary metabolites [35]. Similarly, cytoplasmic stress granules and cytosolic pH homeostasis act as signaling hubs that influence cell viability and stress recovery and are essential for normal growth and stress responses in plants [36,37]. In addition, nucleolar proteins are essential for stress adaptation, affecting plant growth and tolerance to environmental stresses [38]. Analysis of cis-acting elements in the promoter regions of aminotransferase genes revealed several important motifs, including ABREs, AREs, MBSs, MYBs, G-boxes, GATA, LTRs, MREs, GAREs, TGACGs, TCAs, and TGATAs, which are associated with stress tolerance, light regulation, and hormone signaling. Various studies have identified motifs involved in hormonal regulation and phytohormonal responses, including ABREs, GARE motifs, GATA transcription factors [39,40,41], G-box, GT1 motifs, TCT motifs, MREs, and GATA motifs, which regulate development and responses to abiotic stresses [42]. These findings indicate that TAT2 is important for plant growth and abiotic stress resistance.
The aminotransferase gene plays a crucial role in abiotic stress tolerance, as revealed by KEGG enrichment analysis of pathways such as tyrosine metabolism, cysteine and methionine metabolism, and phenylalanine metabolism. Key metabolites, including N-methyltyramine, succinic acid, L-alanine, L-serine, L-homoserine, L-cysteine, L-methionine, 2-hydroxycinnamic acid, benzoic acid, and L-phenylalanine, were upregulated, indicating that aminotransferases are involved in drought stress tolerance. Tyrosine aminotransferases play a role in the oxidative stress response through m-tyrosine metabolism in Caenorhabditis elegans [43]. In Vigna radiata, tyrosine and lysine enhance growth under cadmium stress, highlighting their importance in plant nitrogen metabolism [44]. Additionally, tyrosine metabolism may improve drought tolerance by affecting carbon and nitrogen metabolism in okra [45]. The tyrosine metabolism pathway initiates the production of various structurally diverse natural compounds in plants, including tocopherols, betalains, salidroside, plastoquinone, ubiquinone, and benzylisoquinoline alkaloids. Notably, tyrosine-derived metabolites, tocopherols, ubiquinone and plastoquinone are vital for plant survival [46].
Plants generate several l-tyrosine (Tyr)-derived compounds essential for their adaptation, as well as having pharmaceutical and nutritional significance for human health. TAT catalyzes the reversible reaction between Tyr and 4-hydroxyphenylpyruvate, serving as the entry point for the biosynthesis of various natural products and the degradation of Tyr for energy and nutrient recycling [19]. Similar TAT genes have been found in other plants, including Salvia miltiorrhiza. In this species, three SmTAT genes exhibit various expression patterns and react to methyl jasmonate stimuli [47]. Methionine is crucial in the oxidative stress response and functions as a scavenger of reactive oxygen species [48]. It is central to the interconversion of sulfur-containing amino acids, with cysteine contributing to oxygen tolerance [49]. Both methionine and cysteine are highly sensitive to various reactive oxygen species, highlighting their antioxidant properties [50]. Furthermore, phenylalanine metabolism plays a crucial role in mitigating the negative effects of drought in Brassica campestris [51], cold stress in tartary buckwheat landraces [35], and heat stress in Cucumis sativus [52]. The enhanced production of aromatic amino acids in tobacco plants results in a significant increase in the levels of phenylpropanoid metabolites. This, in turn, contributes to improved tolerance against stresses that the plants may encounter [53]. Similarly, TAT enzymes are involved in various metabolic pathways, like amino acid metabolism, vitamin biosynthesis and secondary metabolites [33].
VIGS analyses revealed phenotype differences in TRV2:00 and TRV2:GhTAT2 before and after drought stress treatment, highlighting its potential role in cotton drought tolerance (Figure 10A,B). RT-qPCR analysis disclosed that GhTAT2 expression in VIGS-treated plants was lower than that in positive control plants, indicating that TRV2:GhTAT2 is sensitive to drought stress. Additionally, the morphological traits of VIGS-treated seedlings were notably different from those of TRV:00-treated seedlings. Silencing the GhTAT2 gene reduced both shoot and root fresh weight. Furthermore, after drought, the relative expression of the GhTAT2 gene significantly increased in the leaves and roots of positive control plants but notably decreased in both tissues of VIGS-treated plants (Figure 10C,D). Previous reports indicate that the TAT2 gene enhances stress responses in various crop species. In Populus simonii, TAT2 genes improve drought tolerance and low nitrogen tolerance [18]. The overexpression of MdTAT2 in Malus domestica enhances resistance to osmotic and drought stress [13], whereas the overexpression of AccTATN in Apis cerana cerana increases heavy metal stress tolerance and antioxidant capacity [20]. An ornithine δ-aminotransferase gene OsOAT gene, plays a crucial role in the metabolism of proline and arginine. This gene has been found to enhance drought tolerance as well as provide protection against oxidative stress in rice plants [54].

4. Materials and Methods

4.1. Aminotransferase Gene Identification in Cotton

Cotton species genome data for G. hirsutum (NAU assembly), G. arboreum (CRI assembly) and G. raimondii (JGI assembly) were retrieved from CottonFGD (https://cottonfgd.org/, accessed on 12 December 2024) [55]. The protein sequences of Theobroma cacao (Ver 1.1) and Arabidopsis thaliana (TAIR 10) were obtained from the Phytozome database (https://phytozome-next.jgi.doe.gov/, accessed on 12 December 2024) [56]. To identify aminotransferase proteins, Pfam (https://pfam.xfam.org/, accessed on 18 December 2024) and HMMER (version 3.4) software were used with default settings to obtain aminotransferases containing the Pfam PF00155 domain, which probably belongs to the aminotransferase gene family. Incomplete and repeated sequences with an e value greater than 1 × 10−5 were removed [57]. We analyzed various gene features, protein statistics, and transcript characteristics of cotton aminotransferase genes, including lengths of CDSs, proteins and genes, molecular weights (MWs), isoelectric points (PIs), exon/intron lengths, and start and end codons, via CottonFGD [55].

4.2. Protein Sequence Alignment and Phylogenetic Tree Construction

The ClustalX tool (Ver. 2.1) was used to fully align all of the aminotransferase protein sequences identified from the five species and perform bootstrap N-J tree analysis with the default settings [58]. A Newick file was produced after manual modification in MEGA 7.0 software via the maximum likelihood approach with 1000 bootstrap repetitions [59]. The Newick file from MEGA 7.0 was used to construct the final tree via the iTOL online tool [25]. Comparative phylogenetic analysis was also performed to identify the evolutionary relationships of cotton aminotransferase with A. thaliana and T. cacao. All relevant list of proteins associated with aminotransferases are given in Supplementary File S3.

4.3. Chromosomal Mapping, Gene Structure and Conserved Domain Analysis

General feature format (GFF3) files and genome assembly sequences for the three cotton species were obtained from CottonFGD [55]. TBtools software [60] was used to visualize the chromosomal locations of aminotransferase genes in the cotton genomes. Gene structure analysis involved the retrieval of detailed coding sequence (CDS) and genomic sequences from CottonFGD. The intron/exon arrangement of the cotton aminotransferase genes was illustrated via the gene structure display server GSDS 2.0 (https://gsds.gao-lab.org/, accessed on 21 December 2024) [61]. Similarly, using protein sequences as inputs on the MEME Suite 5.5.7 online server (https://meme-suite.org/meme/, accessed on 21 December 2024), we identified the highly conserved motifs of aminotransferase proteins [62].

4.4. Promoter Region Analysis and Prediction of Subcellular Localization

We utilized the PlantCARE (https://bioinformatics.psb.ugent.be/webtools/plantcare/html/, accessed on 21 December 2024) webserver to identify cis-acting regulatory elements in the promoter regions of aminotransferase genes in Gossypium species [63]. The 2000 bp promoter sequences of aminotransferases from the upstream region were sourced from the CottonFGD database [55]. Protein sequences were used to determine the subcellular localization of aminotransferase proteins from Gossypium species via the WoLF PSORT online prediction tool (https://wolfpsort.hgc.jp/, accessed on 21 December 2024) [64].

4.5. Expression Profiling, Coexpression Network and RT-qPCR Analysis

RNA-seq data (NCBI accession number: PRJNA663204) were used to examine the relative expression patterns of aminotransferases under drought stress for 0 h, 24 h and 48 h for each treatment and two tissues, namely, roots and leaves, from three cotton races [65]. To verify the sequencing results, the identified candidate genes were analyzed via real-time quantitative PCR. The primers for the target genes (Supplementary File S4) were designed via NCBI Primer Blast (https://www.ncbi.nlm.nih.gov/tools/primer-blast/, accessed on 12 December 2024). Cytoscape software (version 3.10.2) was utilized for coexpression network analysis of leaf and root tissue FPKM values from RNA-seq data. The top 25 genes associated with various metabolic pathways were imported into Cytoscape based on their correlation values. The FPKM values of leaf and root tissues at 0 h, 24 h, and 48 h under drought stress were used to determine the correlations between G. hirsutum genes via a correlation calculation tool with default settings [66]. RT-qPCR was performed with the ChamQ SYBR qPCR Master Mix (LowROX Premixed) kit on a SLAN-96S real-time PCR system (Hongshi Medical Technology, Shanghai, China). The 20 μL RT-qPCR mixture comprised 10 μL of ChamQ SYBR qPCR Master Mix (Vazyme Biotech, Nanjing, China), 0.4 μL of each forward and reverse primer, 2 μL of cDNA template, and 7.2 μL of dd water. GhUBQ served as the internal reference, and the results were analyzed via the 2−ΔΔCT method [67]. A heatmap was generated via TBtools software [60] with FPKM values for relative expression analysis, incorporating three biological and technical replicates.

4.6. Cotton Seedlings and Growth Conditions

The experiment took place in seedling growth chamber at Nantong University’s School of Life Sciences. For VIGS experiment, CRI variety, a G. hirsutum drought-tolerant cotton variety released by the Institute of cotton research, Chinese Academy of Agricultural Sciences (ICR, CAAS) was used [68]. Seedlings were planted in pots containing equal parts vermiculite and humus and were grown under the same conditions. A total of 12 pot cotton seedlings for the TRV2:00, TRV2:GhTAT2 and TRV2:CLA1 treatment sets were used for the experiment. When the seedlings grew to three true leaves, drought stress treatment was applied by adding the nutrient solution with PEG-6000 [69].

4.7. VIGS Experiment for Drought Stress Tolerance

The VIGS experiment used the G. hirsutum variety “CRI-12” to study the role of GhTAT2 genes in drought stress tolerance and tissue-specific expression profiling via gene silencing genetic transformation. The functional analysis focused on significantly upregulated expression of the GhTAT2 gene (Gh_A13G1261) in G. hirsutum. The candidate GhTAT2 with a 271 bp fragment was amplified via specific forward and reverse primers. The amplified primers were subsequently cloned and inserted into the vector plasmid tobacco rattle virus (pTRV2) via XbaI and BamHI enzymes. A. tumefaciens “GV3101” served as the recombinant vector carrier via the freeze-thaw method, with TRV2:GhTAT2 as the silenced gene, TRV2:CLA1 as the positive control and TRV2:00 as the negative control [70]. The cotton seedlings’ cotyledons underwent agroinfiltration when the cotyledons were fully expanded and before the first true leaf had just emerged. After infiltration, the plants were kept in darkness for 24 h before being transferred to a growth chamber at 25 °C with a 16/8 h light/dark cycle. Two weeks post-infection, plants transformed with the positive control gene CLA1 presented an albino phenotype in their newly emerged leaves. 15% PEG-6000 was used to induce drought stress at the three-leaf stage, and leaf and root samples were collected at 0, 24 and 48 h in liquid nitrogen and stored at −80 °C for expression analysis in three replications from before to after treatment [71].

4.8. Statistical and Graphic Analysis

The data analysis was performed using one-way analysis of variance (ANOVA) at 5% probability. The significant differences between treatment means were estimated through t-test at the 5% and 1% (p ≤ 0.05 and p ≤ 0.01) confidence levels. GraphPad Prism (version 8.4.3) was used to display significant differences. The data are presented as the means ± SDs from three separate experiments, with * and ** indicating significance at p ≤ 0.05 and p ≤ 0.01, respectively [72].

5. Conclusions

A total of 203 aminotransferase genes were identified across 3 Gossypium species, with 103 in G. hirsutum, 47 in G. arboreum, and 53 in G. raimondii, through gene family analysis. Genome-wide identification, KEGG enrichment analysis, VIGS, and RT-qPCR profiling suggest that the GhTAT2 gene cluster plays a role in drought stress tolerance through its involvement in tyrosine, cysteine and methionine, and phenylalanine metabolism. Silencing of GhTAT2 via VIGS and subsequent RT-qPCR revealed downregulation and reduced relative expression under drought stress. GhTAT2 has significant potential for drought stress tolerance in cotton. Thus, further gene transformation and editing are required to enhance our understanding of their roles in drought stress tolerance and metabolic mechanisms at the genetic and molecular levels.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms26031355/s1.

Author Contributions

Conceptualization and writing—original draft preparation and visualization, T.G.M. and J.T.; methodology and data curation, H.G.; software and formal analysis, H.F. and J.H.; validation and resources, J.Z. and F.L.; investigation and writing—review and editing, K.W.; supervision, project administration and funding acquisition, D.Y. and B.W. All authors have read and agreed to the published version of the manuscript.

Funding

We appreciate financial help from the Key Research and Development Project of Jiangsu Province, China (Modern Agriculture, BE2022364); the State Key Laboratory of Cotton Bio-breeding and Integrated Utilization Open Fund (CB2024A06).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within this article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

CLACloroplastos alterados
CottonFGDCotton functional genomics database
TATTyrosine aminotransferases
TRVTobacco rattle virus
VIGSVirus-induced gene silencing

References

  1. Cao, M.; Narayanan, M.; Shi, X.; Chen, X.; Li, Z.; Ma, Y. Optimistic contributions of plant growth-promoting bacteria for sustainable agriculture and climate stress alleviation. Environ. Res. 2023, 217, 114924. [Google Scholar] [CrossRef] [PubMed]
  2. Raza, A.; Razzaq, A.; Mehmood, S.S.; Zou, X.; Zhang, X.; Lv, Y.; Xu, J. Impact of climate change on crops adaptation and strategies to tackle its outcome: A review. Plants 2019, 8, 34. [Google Scholar] [CrossRef] [PubMed]
  3. Naumann, G.; Alfieri, L.; Wyser, K.; Mentaschi, L.; Betts, R.A.; Carrao, H.; Spinoni, J.; Vogt, J.; Feyen, L. Global changes in drought conditions under different levels of warming. Geophys. Res. Lett. 2018, 45, 3285–3296. [Google Scholar] [CrossRef]
  4. Kim, W.; Iizumi, T.; Nishimori, M. Global patterns of crop production losses associated with droughts from 1983 to 2009. J. Appl. Meteorol. Climatol. 2019, 58, 1233–1244. [Google Scholar] [CrossRef]
  5. Begna, T. Effects of drought stress on crop production and productivity. Int. J. Agric. Sci. 2020, 6, 34–43. [Google Scholar]
  6. Geneti, T.Z. Review on the effect of moisture or rain fall on crop production. Civ. Environ. Res. 2019, 11, 1–7. [Google Scholar]
  7. Zafar, S.; Afzal, H.; Ijaz, A.; Mahmood, A.; Ayub, A.; Nayab, A.; Hussain, S.; Maqsood, U.-H.; Sabir, M.A.; Zulfiqar, U. Cotton and drought stress: An updated overview for improving stress tolerance. S. Afr. J. Bot. 2023, 161, 258–268. [Google Scholar] [CrossRef]
  8. Jabran, K.; Ul-Allah, S.; Chauhan, B.S.; Bakhsh, A. An introduction to global production trends and uses, history and evolution, and genetic and biotechnological improvements in cotton. Cotton Prod. 2019, 1–22. [Google Scholar] [CrossRef]
  9. Bhutta, M.; Qureshi, M.; Shabaz, M.; Mehmood Ahmad, M.A.; Saif-ul-Malook, S.-u.-M.; Qurban Ali, Q.A. Oxidative damage caused by Reactive oxygen species under drought stress in Gossypium hirsutum. Life Sci J. 2015, 12, 51–59. [Google Scholar]
  10. Meshram, J.; Singh, S.; Raghavendra, K.; Waghmare, V. Chapter 6—Drought stress tolerance in cotton: Progress and perspectives. In Climate Change and Crop Stress; Shanker, A.K., Shanker, C., Anand, A., Maheswari, M., Eds.; Academic Press: Cambridge, MA, USA, 2022; pp. 135–169. [Google Scholar]
  11. Ullah, A.; Sun, H.; Yang, X.; Zhang, X. Drought coping strategies in cotton: Increased crop per drop. Plant Biotechnol. J. 2017, 15, 271–284. [Google Scholar] [CrossRef]
  12. Anwar, A.; Wang, K.; Wang, J.; Shi, L.; Du, L.; Ye, X. Expression of Arabidopsis Ornithine Aminotransferase (AtOAT) encoded gene enhances multiple abiotic stress tolerances in wheat. Plant Cell Rep. 2021, 40, 1155–1170. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, H.; Dong, Q.; Duan, D.; Zhao, S.; Li, M.; van Nocker, S.; Ma, F.; Mao, K. Comprehensive genomic analysis of the Tyrosine Aminotransferase (TAT) genes in apple (Malus domestica) allows the identification of MdTAT2 conferring tolerance to drought and osmotic stresses in plants. Plant Physiol. Biochem. 2018, 133, 81–91. [Google Scholar] [CrossRef] [PubMed]
  14. Mehere, P.; Han, Q.; Lemkul, J.A.; Vavricka, C.J.; Robinson, H.; Bevan, D.R.; Li, J. Tyrosine aminotransferase: Biochemical and structural properties and molecular dynamics simulations. Protein Cell. 2010, 1, 1023–1032. [Google Scholar] [CrossRef] [PubMed]
  15. Senior, J.R. Alanine aminotransferase: A clinical and regulatory tool for detecting liver injury-past, present, and future. Clin. Pharmacol. Ther. 2012, 92, 332–339. [Google Scholar] [CrossRef] [PubMed]
  16. Ndrepepa, G.; Holdenrieder, S.; Cassese, S.; Xhepa, E.; Fusaro, M.; Laugwitz, K.L.; Schunkert, H.; Kastrati, A. Aspartate aminotransferase and mortality in patients with ischemic heart disease. Nutr. Metab. Cardiovasc. Dis. 2020, 30, 2335–2342. [Google Scholar] [CrossRef]
  17. Marienhagen, J.; Sandalova, T.; Sahm, H.; Eggeling, L.; Schneider, G. Insights into the structural basis of substrate recognition by histidinol-phosphate aminotransferase from Corynebacterium glutamicum. Acta Crystallogr. D Biol. Crystallogr. 2008, 64, 675–685. [Google Scholar] [CrossRef]
  18. Zhang, C.; Meng, S.; Li, M.; Zhao, Z. Transcriptomic insight into nitrogen uptake and metabolism of Populus simonii in response to drought and low nitrogen stresses. Tree Physiol. 2018, 38, 1672–1684. [Google Scholar] [CrossRef]
  19. Wang, M.; Toda, K.; Block, A.; Maeda, H.A. TAT1 and TAT2 tyrosine aminotransferases have both distinct and shared functions in tyrosine metabolism and degradation in Arabidopsis thaliana. J. Biol. Chem. 2019, 294, 3563–3576. [Google Scholar] [CrossRef]
  20. Sun, Y.; Niu, X.; Huang, Y.; Wang, L.; Liu, Z.; Guo, X.; Xu, B.; Wang, C. Role of the tyrosine aminotransferase AccTATN gene in the response to pesticide and heavy metal stress in Apis cerana cerana. Pestic. Biochem. Physiol. 2023, 191, 105372. [Google Scholar] [CrossRef]
  21. Riewe, D.; Koohi, M.; Lisec, J.; Pfeiffer, M.; Lippmann, R.; Schmeichel, J.; Willmitzer, L.; Altmann, T. A tyrosine aminotransferase involved in tocopherol synthesis in Arabidopsis. Plant J. 2012, 71, 850–859. [Google Scholar] [CrossRef]
  22. Zhong, M.; Zhang, L.; Yu, H.; Liao, J.; Jiang, Y.; Chai, S.; Yang, R.; Wang, L.; Deng, X.; Zhang, S. Identification and characterization of a novel tyrosine aminotransferase gene (SmTAT3-2) promotes the biosynthesis of phenolic acids in Salvia miltiorrhiza Bunge. Int. J. Biol. Macromol. 2024, 254, 127858. [Google Scholar] [CrossRef] [PubMed]
  23. Zhu, F.; Alseekh, S.; Koper, K.; Tong, H.; Nikoloski, Z.; Naake, T.; Liu, H.; Yan, J.; Brotman, Y.; Wen, W. Genome-wide association of the metabolic shifts underpinning dark-induced senescence in Arabidopsis. Plant Cell 2022, 34, 557–578. [Google Scholar] [CrossRef] [PubMed]
  24. Li, F.; Fan, G.; Wang, K.; Sun, F.; Yuan, Y.; Song, G.; Li, Q.; Ma, Z.; Lu, C.; Zou, C.; et al. Genome sequence of the cultivated cotton Gossypium arboreum. Nat Genet. 2014, 46, 567–572. [Google Scholar] [CrossRef] [PubMed]
  25. Letunic, I.; Bork, P. Interactive Tree of Life (iTOL) v6: Recent updates to the phylogenetic tree display and annotation tool. Nucleic Acids Res. 2024, 52, W78–W82. [Google Scholar] [CrossRef]
  26. Hassan, A.; Ijaz, M.; Sattar, A.; Sher, A.; Rasheed, I.; Saleem, M.Z.; Hussain, I. Abiotic stress tolerance in cotton. In Advances in Cotton Research; IntechOpen: London, UK, 2020. [Google Scholar] [CrossRef]
  27. Khan, A.; Pan, X.; Najeeb, U.; Tan, D.K.Y.; Fahad, S.; Zahoor, R.; Luo, H. Coping with drought: Stress and adaptive mechanisms, and management through cultural and molecular alternatives in cotton as vital constituents for plant stress resilience and fitness. Biol. Res. 2018, 51. [Google Scholar] [CrossRef]
  28. Rehman, M.; Bakhsh, A.; Zubair, M.; Rehmani, M.I.A.; Shahzad, A.; Nayab, S.; Khan, M.; Anum, W.; Akhtar, R.; Kanwal, N. Effects of water stress on cotton (Gossypium spp.) plants and productivity. Egypt. J. Agron 2021, 43, 307–315. [Google Scholar] [CrossRef]
  29. Anjum, S.A.; Xie, X.Y.; Wang, L.C.; Saleem, M.F.; Man, C.; Lei, W. Morphological, physiological and biochemical responses of plants to drought stress. Afr. J. Agric. Res. 2011, 6, 2026–2032. [Google Scholar]
  30. Sourour, A.; Afef, O.; Mounir, R.; Mongi, B.Y. A review: Morphological, physiological, biochemical and molecular plant responses to water deficit stress. Int. J. Eng. Sci. 2017, 6, 1–4. [Google Scholar] [CrossRef]
  31. Renny-Byfield, S.; Wendel, J.F. Doubling down on genomes: Polyploidy and crop plants. Am. J. Bot. 2014, 101, 1711–1725. [Google Scholar] [CrossRef]
  32. Ge, G.; Liu, Z.; Yu, T.; Zhou, L.; Sun, X.; Li, Z.; Zheng, Y. Aminotransferase Class I and II Gene Family in the Jinjiang Oyster (Crassostrea ariakensis): Genomewide Identification, Phylogenetic Analysis and Expression Profiles after Salinity Stress. Fishes 2023, 8, 459. [Google Scholar] [CrossRef]
  33. Koper, K.; Han, S.-W.; Pastor, D.C.; Yoshikuni, Y.; Maeda, H.A. Evolutionary origin and functional diversification of aminotransferases. J. Biol. Chem. 2022, 298. [Google Scholar] [CrossRef] [PubMed]
  34. Liepman, A.H.; Olsen, L.J. Genomic analysis of aminotransferases in Arabidopsis thaliana. CRC Crit. Rev. Plant Sci. 2004, 23, 73–89. [Google Scholar] [CrossRef]
  35. Song, Y.; Feng, J.; Liu, D.; Long, C. Different phenylalanine pathway responses to cold stress based on metabolomics and transcriptomics in Tartary buckwheat landraces. J. Agri. Food Chem. 2022, 70, 687–698. [Google Scholar] [CrossRef] [PubMed]
  36. Mahboubi, H.; Stochaj, U. Cytoplasmic stress granules: Dynamic modulators of cell signaling and disease. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 884–895. [Google Scholar] [CrossRef]
  37. Zhou, J.-Y.; Hao, D.-L.; Yang, G.-Z. Regulation of cytosolic pH: The contributions of plant plasma membrane H+-ATPases and multiple transporters. Int. J. Mol. Sci. 2021, 22, 12998. [Google Scholar] [CrossRef]
  38. Lee, S.; Seo, Y.E.; Choi, J.; Yan, X.; Kim, T.; Choi, D.; Lee, J.H. Nucleolar actions in plant development and stress responses. Plant Cell Environ. 2024, 47, 5189–5204. [Google Scholar] [CrossRef]
  39. Ain-Ali, Q.-U.; Mushtaq, N.; Amir, R.; Gul, A.; Tahir, M.; Munir, F. Genome-wide promoter analysis, homology modeling and protein interaction network of Dehydration Responsive Element Binding (DREB) gene family in Solanum tuberosum. PLoS ONE 2021, 16, e0261215. [Google Scholar] [CrossRef]
  40. Maqsood, H.; Munir, F.; Amir, R.; Gul, A. Genome-wide identification, comprehensive characterization of transcription factors, cis-regulatory elements, protein homology, and protein interaction network of DREB gene family in Solanum lycopersicum. Front. Plant Sci. 2022, 13, 1031679. [Google Scholar] [CrossRef]
  41. Manzoor, M.A.; Sabir, I.A.; Shah, I.H.; Wang, H.; Yu, Z.; Rasool, F.; Mazhar, M.Z.; Younas, S.; Abdullah, M.; Cai, Y. Comprehensive comparative analysis of the GATA transcription factors in four Rosaceae species and phytohormonal response in Chinese pear (Pyrus bretschneideri) fruit. Int. J. Mol. Sci. 2021, 22, 12492. [Google Scholar] [CrossRef]
  42. Ahmar, S.; Zolkiewicz, K.; Gruszka, D. Analyses of genes encoding the Glycogen Synthase Kinases in rice and Arabidopsis reveal mechanisms which regulate their expression during development and responses to abiotic stresses. Plant Sci. 2023, 332, 111724. [Google Scholar] [CrossRef]
  43. Ipson, B.R.; Green, R.A.; Wilson, J.T.; Watson, J.N.; Faull, K.F.; Fisher, A.L. Tyrosine aminotransferase is involved in the oxidative stress response by metabolizing meta-tyrosine in Caenorhabditis elegans. J. Biol. Chem. 2019, 294, 9536–9554. [Google Scholar] [CrossRef] [PubMed]
  44. Mahmood, S.; Wahid, A.; Azeem, M.; Zafar, S.; Bashir, R.; Bajwa, M.O.S.; Ali, S. Tyrosine or lysine priming modulated phenolic metabolism and improved cadmium stress tolerance in mung bean (Vigna radiata L.). S. Afr. J. Bot. 2022, 149, 397–406. [Google Scholar] [CrossRef]
  45. Wang, J.; Shi, D.; Bai, Y.; Zhang, T.; Wu, Y.; Liu, Z.; Jiang, L.; Ye, L.; Peng, Z.; Yuan, H. Comprehensive proteomic and metabolomic analysis uncover the response of okra to drought stress. Peer J. 2022, 10, e14312. [Google Scholar] [CrossRef] [PubMed]
  46. Xu, J.-J.; Fang, X.; Li, C.-Y.; Yang, L.; Chen, X.-Y. General and specialized tyrosine metabolism pathways in plants. Abiotech 2020, 1, 97–105. [Google Scholar] [CrossRef]
  47. Dong, S.; Wang, L.; Qin, H.; Zhan, H.; Wang, D.; Cao, X. Expression Patterns and Functional Analysis of Three SmTAT Genes Encoding Tyrosine Aminotransferases in Salvia miltiorrhiza. Int. J. Mol. Sci. 2023, 24, 15575. [Google Scholar] [CrossRef]
  48. Campbell, K.; Vowinckel, J.; Keller, M.A.; Ralser, M. Methionine metabolism alters oxidative stress resistance via the pentose phosphate pathway. Antioxid. Redox. Signal. 2016, 24, 543–547. [Google Scholar] [CrossRef]
  49. Qiao, Y.; Liu, G.; Leng, C.; Zhang, Y.; Lv, X.; Chen, H.; Sun, J.; Feng, Z. Metabolic profiles of cysteine, methionine, glutamate, glutamine, arginine, aspartate, asparagine, alanine and glutathione in Streptococcus thermophilus during pH-controlled batch fermentations. Sci. Rep. 2018, 8, 12441. [Google Scholar] [CrossRef]
  50. Bin, P.; Huang, R.; Zhou, X. Oxidation resistance of the sulfur amino acids: Methionine and cysteine. BioMed Res. Int. 2017, 2017, 9584932. [Google Scholar] [CrossRef]
  51. Ramzan, T.; Shahbaz, M.; Maqsood, M.F.; Zulfiqar, U.; Saman, R.U.; Lili, N.; Irshad, M.; Maqsood, S.; Haider, A.; Shahzad, B. Phenylalanine supply alleviates the drought stress in mustard (Brassica campestris) by modulating plant growth, photosynthesis, and antioxidant defense system. Plant Physiol. Biochem. 2023, 201, 107828. [Google Scholar] [CrossRef]
  52. Amjad, M.; Wang, Y.; Han, S.; Haider, M.Z.; Sami, A.; Batool, A.; Shafiq, M.; Ali, Q.; Dong, J.; Sabir, I.A. Genome wide identification of phenylalanine ammonia-lyase (PAL) gene family in Cucumis sativus (cucumber) against abiotic stress. BMC Genom. Data 2024, 25, 76. [Google Scholar] [CrossRef]
  53. Oliva, M.; Guy, A.; Galili, G.; Dor, E.; Schweitzer, R.; Amir, R.; Hacham, Y. Enhanced Production of Aromatic Amino Acids in Tobacco Plants Leads to Increased Phenylpropanoid Metabolites and Tolerance to Stresses. Front. Plant Sci. 2020, 11, 604349. [Google Scholar] [CrossRef] [PubMed]
  54. You, J.; Hu, H.; Xiong, L. An ornithine δ-aminotransferase gene OsOAT confers drought and oxidative stress tolerance in rice. Plant Sci. 2012, 197, 59–69. [Google Scholar] [CrossRef] [PubMed]
  55. Zhu, T.; Liang, C.; Meng, Z.; Sun, G.; Meng, Z.; Guo, S.; Zhang, R. CottonFGD: An integrated functional genomics database for cotton. BMC Plant Biol. 2017, 17, 101. [Google Scholar] [CrossRef] [PubMed]
  56. Goodstein, D.M.; Shu, S.; Howson, R.; Neupane, R.; Hayes, R.D.; Fazo, J.; Mitros, T.; Dirks, W.; Hellsten, U.; Putnam, N.; et al. Phytozome: A comparative platform for green plant genomics. Nucleic Acids Res. 2012, 40, D1178–D1186. [Google Scholar] [CrossRef] [PubMed]
  57. Finn, R.D.; Clements, J.; Eddy, S.R. HMMER web server: Interactive sequence similarity searching. Nucleic acids Res. 2011, 39, W29–W37. [Google Scholar] [CrossRef]
  58. Larkin, M.A.; Blackshields, G.; Brown, N.P.; Chenna, R.; McGettigan, P.A.; McWilliam, H.; Valentin, F.; Wallace, I.M.; Wilm, A.; Lopez, R. Clustal W and Clustal X version 2.0. Bioinformatics 2007, 23, 2947–2948. [Google Scholar] [CrossRef]
  59. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef]
  60. Chen, C.; Wu, Y.; Li, J.; Wang, X.; Zeng, Z.; Xu, J.; Liu, Y.; Feng, J.; Chen, H.; He, Y. TBtools-II: A “one for all, all for one” bioinformatics platform for biological big-data mining. Mol. Plant 2023, 16, 1733–1742. [Google Scholar] [CrossRef]
  61. Hu, B.; Jin, J.; Guo, A.-Y.; Zhang, H.; Luo, J.; Gao, G. GSDS 2.0: An upgraded gene feature visualization server. Bioinformatics 2015, 31, 1296–1297. [Google Scholar] [CrossRef]
  62. Bailey, T.L.; Boden, M.; Buske, F.A.; Frith, M.; Grant, C.E.; Clementi, L.; Ren, J.; Li, W.W.; Noble, W.S. MEME SUITE: Tools for motif discovery and searching. Nucleic Acids Res. 2009, 37, W202–W208. [Google Scholar] [CrossRef]
  63. Lescot, M.; Déhais, P.; Thijs, G.; Marchal, K.; Moreau, Y.; Van de Peer, Y.; Rouzé, P.; Rombauts, S. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002, 30, 325–327. [Google Scholar] [CrossRef] [PubMed]
  64. Horton, P.; Park, K.-J.; Obayashi, T.; Fujita, N.; Harada, H.; Adams-Collier, C.; Nakai, K. WoLF PSORT: Protein localization predictor. Nucleic Acids Res. 2007, 35, W585–W587. [Google Scholar] [CrossRef] [PubMed]
  65. Mehari, T.G.; Xu, Y.; Umer, M.J.; Shiraku, M.L.; Hou, Y.; Wang, Y.; Yu, S.; Zhang, X.; Wang, K.; Cai, X. Multiomics-based identification and functional characterization of Gh_A06G1257 proves its potential role in drought stress tolerance in Gossypium hirsutum. Front. Plant Sci. 2021, 12, 746771. [Google Scholar] [CrossRef] [PubMed]
  66. Shannon, P.; Markiel, A.; Ozier, O.; Baliga, N.S.; Wang, J.T.; Ramage, D.; Amin, N.; Schwikowski, B.; Ideker, T. Cytoscape: A software environment for integrated models of biomolecular interaction networks. Genome Res. 2003, 13, 2498–2504. [Google Scholar] [CrossRef]
  67. Schmittgen, T.D.; Livak, K.J. Analyzing real-time PCR data by the comparative CT method. Nat. Protoc. 2008, 3, 1101–1108. [Google Scholar] [CrossRef]
  68. Xu, Y.; Magwanga, R.O.; Yang, X.; Jin, D.; Cai, X.; Hou, Y.; Wei, Y.; Zhou, Z.; Wang, K.; Liu, F. Genetic regulatory networks for salt-alkali stress in Gossypium hirsutum with differing morphological characteristics. BMC Genom. 2020, 21, 15. [Google Scholar] [CrossRef]
  69. Mehari, T.G.; Xu, Y.; Magwanga, R.O.; Umer, M.J.; Shiraku, M.L.; Hou, Y.; Wang, Y.; Wang, K.; Cai, X.; Zhou, Z.; et al. Identification and functional characterization of Gh_D01G0514 (GhNAC072) transcription factor in response to drought stress tolerance in cotton. Plant Physiol. Biochem. 2021, 166, 361–375. [Google Scholar] [CrossRef]
  70. Feng, W.; Guo, L.; Fang, H.; Mehari, T.G.; Gu, H.; Wu, Y.; Jia, M.; Han, J.; Guo, Q.; Xu, Z. Transcriptomic profiling reveals salt-responsive long non-coding RNAs and their putative target genes for improving salt tolerance in upland cotton (Gossypium hirsutum). Ind. Crop. Prod. 2024, 216, 118744. [Google Scholar] [CrossRef]
  71. Muhammad Tajo, S.; Pan, Z.; He, S.; Chen, B.; Km, Y.; Mahmood, T.; Bello Sadau, S.; Shahid Iqbal, M.; Gereziher, T.; Suleiman Abubakar, U. Characterization of WOX genes revealed drought tolerance, callus induction, and tissue regeneration in Gossypium hirsutum. Front. Genet. 2022, 13, 928055. [Google Scholar] [CrossRef]
  72. Kim, T.K. T test as a parametric statistic. Korean J. Anesthesiol. 2015, 68, 540–546. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic tree analysis of aminotransferase gene family members. (A) Tree constructed using the protein sequences of G. hirsutum, G. arboreum and G. raimondii. (B) Tree constructed using the protein sequences of G. hirsutum, G. arboreum, G. raimondii, A. thaliana and T. cacao. Different colors represent different clusters, and AT stands for aminotransferase.
Figure 1. Phylogenetic tree analysis of aminotransferase gene family members. (A) Tree constructed using the protein sequences of G. hirsutum, G. arboreum and G. raimondii. (B) Tree constructed using the protein sequences of G. hirsutum, G. arboreum, G. raimondii, A. thaliana and T. cacao. Different colors represent different clusters, and AT stands for aminotransferase.
Ijms 26 01355 g001
Figure 2. Chromosome mapping of aminotransferase genes. (A) G. hirsutum GhAt subgenome; (B) G. hirsutum GhDt subgenome; (C) G. arboreum; (D) G. raimondii; (E) genes located in scaffold regions.
Figure 2. Chromosome mapping of aminotransferase genes. (A) G. hirsutum GhAt subgenome; (B) G. hirsutum GhDt subgenome; (C) G. arboreum; (D) G. raimondii; (E) genes located in scaffold regions.
Ijms 26 01355 g002
Figure 3. Phylogenetics, gene structure and motif analysis of the aminotransferase genes in cotton species: (A) the G. hirsutum GhAt subgenome (B), the G. hirsutum GhDt subgenome, (C) G. arboreum and (D) G. raimondii. The phylogenetic tree was created with MEGA 7 via the neighbor-joining method and 1000 bootstrap replicates, while conserved motifs in aminotransferase proteins were visualized with TBtools (version 2.154) software. The exon-intron structures of the aminotransferase genes reflect their evolutionary relationships, with yellow circles representing exons and gray lines indicating introns. Each motif is marked by a colored box in the legend, and the lengths of the motifs in each protein are shown proportionally.
Figure 3. Phylogenetics, gene structure and motif analysis of the aminotransferase genes in cotton species: (A) the G. hirsutum GhAt subgenome (B), the G. hirsutum GhDt subgenome, (C) G. arboreum and (D) G. raimondii. The phylogenetic tree was created with MEGA 7 via the neighbor-joining method and 1000 bootstrap replicates, while conserved motifs in aminotransferase proteins were visualized with TBtools (version 2.154) software. The exon-intron structures of the aminotransferase genes reflect their evolutionary relationships, with yellow circles representing exons and gray lines indicating introns. Each motif is marked by a colored box in the legend, and the lengths of the motifs in each protein are shown proportionally.
Ijms 26 01355 g003
Figure 4. Cis-regulatory element analysis of the promoter regions of the aminotransferase gene family. (A) G. hirsutum GhAt subgenome; (B) G. hirsutum GhDt subgenome; (C) G. arboreum; (D) G. raimondii.
Figure 4. Cis-regulatory element analysis of the promoter regions of the aminotransferase gene family. (A) G. hirsutum GhAt subgenome; (B) G. hirsutum GhDt subgenome; (C) G. arboreum; (D) G. raimondii.
Ijms 26 01355 g004
Figure 5. Prediction of the subcellular localization of aminotransferase genes in Gossypium species. (A) G. hirsutum in the GhAt subgenome; (B) G. hirsutum in the GhDt subgenome; (C) G. arboreum; (D) G. raimondii.
Figure 5. Prediction of the subcellular localization of aminotransferase genes in Gossypium species. (A) G. hirsutum in the GhAt subgenome; (B) G. hirsutum in the GhDt subgenome; (C) G. arboreum; (D) G. raimondii.
Ijms 26 01355 g005
Figure 6. Transcriptome and coexpression network analysis of aminotransferase genes. (A) Transcriptome analysis of aminotransferase genes in G. hirsutum leaf tissue; (B) transcriptome analysis of aminotransferase genes in G. hirsutum root tissue; (C) coexpression analysis and candidate gene identification in leaf expression; (D) coexpression analysis and candidate gene identification in root expression. Genes in red have positive correlation values and are considered hub genes.
Figure 6. Transcriptome and coexpression network analysis of aminotransferase genes. (A) Transcriptome analysis of aminotransferase genes in G. hirsutum leaf tissue; (B) transcriptome analysis of aminotransferase genes in G. hirsutum root tissue; (C) coexpression analysis and candidate gene identification in leaf expression; (D) coexpression analysis and candidate gene identification in root expression. Genes in red have positive correlation values and are considered hub genes.
Ijms 26 01355 g006
Figure 7. The relative expression levels of aminotransferase genes after drought stress were analyzed via a t-test, and the results are presented as the means ± SDs from three independent experiments. Significance is indicated by *, **, and *** for p ≤ 0.05, p ≤ 0.01, and p ≤ 0.001, respectively, ns stands for non-significant.
Figure 7. The relative expression levels of aminotransferase genes after drought stress were analyzed via a t-test, and the results are presented as the means ± SDs from three independent experiments. Significance is indicated by *, **, and *** for p ≤ 0.05, p ≤ 0.01, and p ≤ 0.001, respectively, ns stands for non-significant.
Ijms 26 01355 g007
Figure 8. Metabolite enrichment analyses of GhTAT2 genes under drought stress. (A) Cysteine and methionine metabolism; (B) phenylalanine metabolism; (C) tyrosine metabolism; (D) phenylalanine, tyrosine and tryptophan biosynthesis; (E) tropane, piperidine and pyridine alkaloid biosynthesis; (F) isoquinoline alkaloid biosynthesis. The heatmap was generated in TBtools via log2-transformed metabolite expression data.
Figure 8. Metabolite enrichment analyses of GhTAT2 genes under drought stress. (A) Cysteine and methionine metabolism; (B) phenylalanine metabolism; (C) tyrosine metabolism; (D) phenylalanine, tyrosine and tryptophan biosynthesis; (E) tropane, piperidine and pyridine alkaloid biosynthesis; (F) isoquinoline alkaloid biosynthesis. The heatmap was generated in TBtools via log2-transformed metabolite expression data.
Ijms 26 01355 g008
Figure 9. Phenotypic variations in cotton seedlings after agroinfiltration (A) TRV2:00, empty vector, TRV2:GhTAT2, VIGS plant, TRV2:CLA1, positive control, (B) plant height and root length measurement, and (C) shoot fresh weight and root fresh weight measurement. ns stands for non-significant; * indicates a significant difference at p < 0.05.
Figure 9. Phenotypic variations in cotton seedlings after agroinfiltration (A) TRV2:00, empty vector, TRV2:GhTAT2, VIGS plant, TRV2:CLA1, positive control, (B) plant height and root length measurement, and (C) shoot fresh weight and root fresh weight measurement. ns stands for non-significant; * indicates a significant difference at p < 0.05.
Ijms 26 01355 g009
Figure 10. Relative expression analysis of positive control and silenced tissues. (A) Phenotypic images of TRV2:00 and TRV2:GhTAT2 before treatment. (B) Phenotypic images of TRV2:00 and TRV2:GhTAT2 after PEG-6000 treatment. (C) Expression profiling of the empty vector and TRV2:GhTAT2 before drought stress treatment in leaf and root tissues. (D) Expression profiling of the empty vector and TRV2:GhTAT2 after drought stress treatment in leaf and root tissues. Bars with different letters indicate significant differences at * for p ≤ 0.05, ns stands for non-significant.
Figure 10. Relative expression analysis of positive control and silenced tissues. (A) Phenotypic images of TRV2:00 and TRV2:GhTAT2 before treatment. (B) Phenotypic images of TRV2:00 and TRV2:GhTAT2 after PEG-6000 treatment. (C) Expression profiling of the empty vector and TRV2:GhTAT2 before drought stress treatment in leaf and root tissues. (D) Expression profiling of the empty vector and TRV2:GhTAT2 after drought stress treatment in leaf and root tissues. Bars with different letters indicate significant differences at * for p ≤ 0.05, ns stands for non-significant.
Ijms 26 01355 g010
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

Mehari, T.G.; Tang, J.; Gu, H.; Fang, H.; Han, J.; Zheng, J.; Liu, F.; Wang, K.; Yao, D.; Wang, B. Insights into the Role of GhTAT2 Genes in Tyrosine Metabolism and Drought Stress Tolerance in Cotton. Int. J. Mol. Sci. 2025, 26, 1355. https://doi.org/10.3390/ijms26031355

AMA Style

Mehari TG, Tang J, Gu H, Fang H, Han J, Zheng J, Liu F, Wang K, Yao D, Wang B. Insights into the Role of GhTAT2 Genes in Tyrosine Metabolism and Drought Stress Tolerance in Cotton. International Journal of Molecular Sciences. 2025; 26(3):1355. https://doi.org/10.3390/ijms26031355

Chicago/Turabian Style

Mehari, Teame Gereziher, Jungfeng Tang, Haijing Gu, Hui Fang, Jinlei Han, Jie Zheng, Fang Liu, Kai Wang, Dengbing Yao, and Baohua Wang. 2025. "Insights into the Role of GhTAT2 Genes in Tyrosine Metabolism and Drought Stress Tolerance in Cotton" International Journal of Molecular Sciences 26, no. 3: 1355. https://doi.org/10.3390/ijms26031355

APA Style

Mehari, T. G., Tang, J., Gu, H., Fang, H., Han, J., Zheng, J., Liu, F., Wang, K., Yao, D., & Wang, B. (2025). Insights into the Role of GhTAT2 Genes in Tyrosine Metabolism and Drought Stress Tolerance in Cotton. International Journal of Molecular Sciences, 26(3), 1355. https://doi.org/10.3390/ijms26031355

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