Next Article in Journal
Development of High-Quality Nuclei Isolation to Study Plant Root–Microbe Interaction for Single-Nuclei Transcriptomic Sequencing in Soybean
Next Article in Special Issue
Contribution of Duplicated Nucleotide-Binding Leucine-Rich Repeat (NLR) Genes to Wheat Disease Resistance
Previous Article in Journal
Two Congeneric Shrubs from the Atacama Desert Show Different Physiological Strategies That Improve Water Use Efficiency under a Simulated Heat Wave
Previous Article in Special Issue
Insights into the Bioinformatics and Transcriptional Analysis of the Elongator Complexes (ELPs) Gene Family of Wheat: TaELPs Contribute to Wheat Abiotic Stress Tolerance and Leaf Senescence
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Duplicate Genes Contribute to Variability in Abiotic Stress Resistance in Allopolyploid Wheat

1
State Key Laboratory of Crop Stress Biology for Arid Areas, College of Life Science, Northwest A&F University, Yangling 712100, China
2
State Key Laboratory of Crop Stress Biology for Arid Areas, College of Agronomy, Northwest A&F University, Yangling 712100, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Plants 2023, 12(13), 2465; https://doi.org/10.3390/plants12132465
Submission received: 5 May 2023 / Revised: 25 May 2023 / Accepted: 26 May 2023 / Published: 28 June 2023

Abstract

:
Gene duplication is a universal biological phenomenon that drives genomic variation and diversity, plays a crucial role in plant evolution, and contributes to innovations in genetic engineering and crop development. Duplicated genes participate in the emergence of novel functionality, such as adaptability to new or more severe abiotic stress resistance. Future crop research will benefit from advanced, mechanistic understanding of the effects of gene duplication, especially in the development and deployment of high-performance, stress-resistant, elite wheat lines. In this review, we summarize the current knowledge of gene duplication in wheat, including the principle of gene duplication and its effects on gene function, the diversity of duplicated genes, and how they have functionally diverged. Then, we discuss how duplicated genes contribute to abiotic stress response and the mechanisms of duplication. Finally, we have a future prospects section that discusses the direction of future efforts in the short term regarding the elucidation of replication and retention mechanisms of repetitive genes related to abiotic stress response in wheat, excellent gene function research, and practical applications.

1. Introduction

Diversity and phenotypic variation in plants do not arise exclusively from the emergence of completely novel and diverse genes; in many plants, diversity may instead arise primarily through gene duplication and adaptive specialization of preexisting genes. Gene duplication is an evolutionary process in which genetic diversity and new functions are generated via whole-genome duplication (WGD) events or smaller-scale, single-gene or single-base duplications [1,2], the occurrence of which in a gene results in two genes that cannot be functionally distinguished from each other. In the evolutionary history of plants, animals, and fungi, gene duplication events have occurred ubiquitously across kingdoms [3,4,5,6]. In comparison with other eukaryotic genomes, plants typically have a higher rate of evolution, which entails continuous increases in their genomic diversity [7,8]. Recently, previous whole-genome duplication events that led to novel functions, such as biotic or abiotic stress tolerance, higher grain weight and quality, or phenotypic changes in flowering time or plant height, were detected in several domesticated crops, including wheat (Triticum aestivum), cotton (Gossypium hirsutum), and soybean (Glycine max).
If a gene duplication results in two copies with a sufficiently similar nucleotide sequence, then these two genes should also share functional overlap, and considerable research efforts have focused on examining the evolutionary mechanisms through which both genes are retained. Currently, there are several models that could explain the retention of duplicated genes. In eukaryotes, most duplicated genes that are retained become non-functional due to disruption or loss-of-function mutations in regulatory elements; alternatively, ancestral functions are sometimes partitioned among the duplicated genes. However, neither case leads to the evolution of new functions, and this process is thus termed “nonfunctionalization” [9]. In addition, duplicated gene retention is controlled by other evolutionary mechanisms, such as neofunctionalization, functional specialization by sub-functionalization, or dosage amplification [9]. Interestingly, despite the different effects of these various evolutionary mechanisms, duplicated genes can still retain some level of functional redundancy [10], which can increase the plasticity of a genome or enhance the adaptability of a species to changing environments [11].
Plants cannot escape from stress-inducing environmental conditions through movement and, thus, the strong selection pressure imposed by these conditions increases the likelihood and frequency of gaining or losing stress-responsive genes. In varying stress conditions, stress-sensing and stress-response mechanisms should be able to evolve rapidly as needed to adjust to new stimuli and therefore require continual innovation in their genetic basis. Duplicate genes have been proposed to serve as the main source of evolutionary novelty and a possible source of functional innovation. Harsh weather events or climatic conditions are becoming increasingly frequent due to global climate change and the world now faces the prospect of food insecurity.
Bread wheat (AABBDD) is one of the three major staple crops, globally, and future wheat yields will be critical to human survival. Allohexaploid wheat contains three related subgenomes, presumably with triplicate copies of the large majority of genes, resulting in highly complicated gene interaction networks due to the presence of these multiple genomes in each cell [12,13,14]. Gene and genome duplications have likely provided significant contributions to the morphological complexity of plants [15], and these duplication events may also influence the physiological complexity of stress responses. For example, the presence or absence of HPT2 (a low-temperature-responsive gene in wild barley) in the pangenome of barley (diploid), wild emmer (tetraploid), and bread wheat (hexaploid) lines appears to be related to gene dosage constraint and environmental adaptation [16]. Evidence in previous studies at least partially supports that environmental stresses, such as drought, cold, heat, and salt, contribute to the retention of duplicated genes [17,18,19]. Herein, we review recent advances in our understanding of the role of duplicate genes in plant resistance to abiotic stress.

2. Diversity and Divergence of Duplicate Genes Involved in Abiotic Stress Resistance

Distinct from most other eukaryotes, plants cannot move to avoid unfavorable conditions, and thus their genomes tend to evolve at faster rates under these selective conditions, resulting in higher genomic diversity that may enable survival in stressful environments [7]. Plant genomes typically contain a high number of duplicate genes, with 65% of annotated genes, on average, having a duplicate copy in the genome [17]. Several genomic sequencing analyses in plants suggest that whole-genome duplication has occurred several times during the past 200 million years of angiosperm evolution [20,21,22]. In other eukaryotes, such as humans and budding yeast, the most recent whole-genome duplication events occurred roughly 450 million years ago (Mya) and 200 Mya, respectively [6,23,24]. On the other hand, some plant species include both diploid and polyploid individuals [25]. In those species, duplicated genes can increase genetic variability, contributing to increased complexity. spatio-temporal transcriptomic plasticity, and the higher adaptability of polyploids to environmental stress [26,27].
Wheat genomes are relatively large compared with other major cereal crops; diploid einkorn wheat is ~5.0 gigabases (Gb) [28,29], tetraploid emmer is >10 Gb [30], and exaploidy bread wheat is ~17 Gb [31]. Despite the complex composition of the wheat genome, more than 80% of each genome is made up of repetitive DNA sequences [31]. In addition, these large, heterozygous exaploidy wheat genomes contain a remarkable diversity of genetic variations, providing a background conducive to gene duplication, functional/phenotypic variation, and evolutionary conservation. Diversity in duplicated genes is inextricably linked to differences in the wheat genome and differences in genetic diversity among subgenomes [32,33]. Although a majority of wheat genes have three copies distributed across the A, B, and D subgenomes, studies have shown that the degree to which vegetative traits are inherited differs among wheat subgenomes, with trans-acting variants, for example, being more genetically diverse in genomes A and B than in D [34,35,36,37]. This variability in the inheritance of different gene copies suggests that genomic interactions could play an important role in regulating genes selected for domestication and improvement of agricultural traits and stress resistance in wheat.
Receptor-like protein kinases (RLKs), the largest gene family in plants, play critical roles in the regulation of plant developmental processes, signaling transmission, and stress resistance [38,39,40]. In recent years, the RLK family of wheat has been identified, and collinearity events and tandem gene clusters results suggested that polyploidization and tandem duplication events contributed to the RLK member expansions of T. aestivum [38]. Among them are lectin receptor kinases (LEC-RLK) [41], leucine-rich repeat receptor-like kinases (LRR-RLKs) [42,43], Cysteine-rich receptor-like kinases (CRKs) [44], thaumatin-like proteins and thaumatin-like kinases (TLPs) [45], and proline-rich extensin-like receptor protein kinases (PERKs) [46,47]. Gene replication events were identified in these families, and some of these replication gene pairs had differential expression data under abiotic stress, such as TaTLP14-A1/B1/B5 being up-regulated in expression under osmotic stress, while TaTLP14-B4 exhibited the opposite [45]. Enzymatic antioxidants are an important stress-responsive class of proteins that scavenge excess reactive oxygen species (ROS) in the presence of cofactors such as copper and zinc ions [48]. Several studies identified and reported duplication in antioxidant genes, including catalase (CAT) [49,50], superoxide dismutase (SOD) [51,52,53], ascorbate peroxidase (APX) [54], glutathione peroxidase (GPX) [55,56], peroxidase (POD) [57], and glutathione reductase (GR) genes in bread wheat [58]. For example, TaCAT3-A1 and TaCAT3-A2 were found to be clustered into tandem duplication event regions, while the number of cis-elements in the promoter of TaCAT3-A2 was more than TaCAT3-A1; moreover, TaCAT3-A1/A2 contained cis-elements associated with cold response, but not exist in other subgenome copies TaCAT3-B and TaCAT3-D [50], which suggests that there may be differences in the cold response of TaCAT3 homologous gene groups.
In plants, various transporters were reported related to the transport of Ca, Na, and other important molecules during stress response to maintain ion homeostasis in the plant cell [59,60,61]. Hyperosmolality-gated calcium-permeable channels (OSCAs) [62,63], boron transporters (BOR) [64,65], mechanosensitive channels of small conductance-like (MSL) genes [66], Ca2+/cation antiporters (CaCAs) [67], cation proton antiporters (CPAs) [68], P-type II Ca2+ ATPases [69], and thaumatin-like protein kinases (TLPKs) in the bread wheat gene family were identified and analyzed for their expression pattern under abiotic stress [70]. For example, the gene pairs TaMSL4-A1 and TaMSL7-A in MSL were both drought-induced, but the expression of TaMSL4-A1 was inhibited under heat stress and in the early stage of salt stress, while the expression of TaMSL7-A was up-regulated under heat stress and the under early salt stress, which indicates the response of paraline homologous genes was different under different abiotic stresses; the expression was also diverse and began to express and function under different levels of stress [63]. The cation proton antiporter (CPA) superfamily, including K+ efflux antiporter (KEA) and cation/H+ exchanger (CHX) family proteins as well as the Na+/H+ exchanger (NHX), TaNHX4-B.1, and TaNHX4-B.4, facilitated differential drought, salt, and heat stress tolerance to Escherichia coli [64]. A similar phenomenon occurs with transcription factors, in which orthologous or paralogue genes that are produced via gene duplication exhibit identical or opposite expression patterns when subjected to abiotic stress, including NAC [71,72,73,74], DREB [75], Hsf [76,77,78,79], MYB [80,81], bZIP [82,83], WRKY [84,85], AP2/ERF [86], GRF [87], and the homeobox genes HD-Zip [88], TALE [89,90], ZF-HD [91], and WOX [92,93]. Taking TaTAIL as an example, Rathour et al. systemically identified and analyzed TAIL family members in wheat, including gene and protein structural properties, phylogeny, and expression patterns. Gene duplication events were identified, including gene pairs TaTALE8-4A3 and TaTALE8-4A1 produced via fragment replication and five tandem duplicate gene pairs such as TaTALE1-1A2 and TaTALE1-1A1. The cis-acting elements of these repeating gene promoters and their expression data were different under heat and salt stress, indicating that the response of repeating genes to abiotic stress was diverse [89].
The Introduction of diversity in duplicate genes can be grouped into two main categories based on the type of alterations in DNA, that is, changes in gene structure or epigenetic modifications. Changes in DNA sequence mainly impact diversity and variability at the transcriptional and post-transcriptional levels, as well as in the translated protein or post-translational modifications [94,95,96,97,98,99]. At the transcriptional level, differences in cis-acting elements can result in differences in response to the same stimulus between homologous genes, while changes in promoter region binding sites for trans-acting factors can trigger differences in expression between gene copies [34]. Previous research has shown that stress-responsive plant genes are retained at higher rates than nonfunctionalized duplicate genes, especially transcription factors and signal transduction proteins. For instance, the expression levels of 7 out of 25 TaADF genes (TaADF13/16/17/18/20/21/22) were significantly affected by cold or freezing treatment, while overexpression of TaADF16 enhanced tolerance to freezing in wheat plants [100]. Additionally, a number of other transcription factor genes have been reported to respond to and regulate drought stress tolerance, including TaOPF29a, TaDrAp1/2, TaFDL2-1A, TaSNAC4-3D, TaMpc1-D4, TaGT2L1D, TaSNAC4-3A, TaWRKY1-2D, TaNFYC-A7, and TaERF-6-3A [73,74,85,86,101,102,103,104,105]. Besides, genome-wide association studies (GWAS) have identified sequence variations in homologous genes that could increase drought tolerance in wheat, such as TaNAC071-A, TaDTG6-B, and TaSNAC8-A [71,72,75].
Increased diversity and divergence of duplicated genes is also linked to neofunctionalization or sub-functionalization in abiotic stress response, such as heat stress transcription factors (HSFs). HSFs are among the most important TFs in plant response to heat stress, but some HSFs also respond to drought or salt stress. For example, overexpressing TaHsaA2d or TaHsaA6f in wheat not only increases heat tolerance but also drought and salt stress, respectively [76,106], while overexpression of TaMYB344, TaAIDFa, or TaAREB3 enhances tolerance to drought, heat, or salt stress in transgenic lines [80,107,108]. In post-transcriptional regulation, changes in the DNA sequence can introduce variable shear events, such as splice site disruption, which can lead to changes in target gene transcripts that consequently affect the structure and function of the translated protein [98,109,110]. These results illustrate how the accumulation of homoeologs with biased expression patterns can affect stress tolerance. In particular, changes in the DNA sequence between homologous TFs may impact their DNA binding and transcriptional regulatory activities, such as a TaDTG6-B gain-of-function allele that improves drought tolerance in wheat [75].
Changes in the epigenetic modification of duplicated DNA sequences induced by environmental stimuli can also affect heritable variation in gene expression [111,112,113]; such modifications include DNA methylation, histone methylation and acetylation, or modifications to mRNAs or non-coding RNAs [114,115,116,117,118,119]. It should also be noted that the epigenetic modification landscape across the three subgenomes of hexaploid wheat may contribute the predominant regulation to gene dosage for some genes. For example, DNA methylation and acetylation modifications modulate the expression of TaCYP81D5, which contributes to both seedling- and reproductive-stage salt tolerance in bread wheat [120].
The combination of these factors, along with the complexity of the wheat genome, provides rich potential for variation among duplicated genes, from the DNA to protein levels, especially in TFs that function as the major regulators of abiotic and biotic stress response [121]. These altered TF homoeologs can thus participate in sophisticated and versatile regulatory networks that facilitate adaptability and maintenance of homeostasis for essential biological processes in wheat in the face of highly variable and extreme climatic conditions. Genetic diversity causes structural and functional phenotypic differences, and diversity among duplicated genes considerably enriches the genetic diversity of the wheat genome. The resulting complex regulatory networks can further improve the plasticity of signal transduction processes to some extent, and also provide resources for environmental adaptation in both natural evolution and artificial selection processes.

3. Contribution of Duplicate Genes to Abiotic Stress Resistance

Transcriptional regulation underlies all biological activities in plants, and especially provides a sophisticated set of mechanisms for reacting to changes in the external environment. Duplicated genes typically exhibit distinct patterns of expression, and this regulatory divergence has been proposed to serve as the prelude to functional differentiation among duplicate genes. Thus, duplicate genes are a major source of potential functional innovations in plant response mechanisms to abiotic stress stimuli [120,122,123,124]. Differentiation in expression patterns among duplicated genes is considered a precursor to functional differentiation of genes, because functional divergence may occur long after duplication, whereas changes in expression patterns might begin relatively soon (or immediately) after the duplication event [125]. Although duplicated genes (i.e., homologs and paralogs) may share a high degree of structural or functional similarity, differences in their response to environmental stress can be related to variations in cis-acting regions [126], transcription factor binding sites, and/or methylation status [127]. In particular, the gain of additional or different cis-regulatory elements in the promoter regions of MIKC-type MADS-box [128]; CHY zinc-finger and RING finger [129]; basic leucine zipper [130]; abscisic acid-, stress- and ripening-induced (ASR) [131]; APETALA2/Ethylene-Responsive Factor (AP2/ERF) [132,133]; E3 ubiquitin ligase [134]; or LATERAL ORGAN BOUNDARIES DOMAIN (LBD) [135] genes results in enhanced abiotic stress tolerance in different crops.
The evolution of TATA boxes in duplicated genes may also provide some clues regarding the interrelationship of environmental stress, divergent expression patterns, and the conservation of duplicated genes [136]. Moreover, de novo functionalization is closely correlated with the retention of duplicated genes that inherit little or none of the original function [137]. For instance, CPK7 and CPK12, which are duplicated wheat genes that together comprise the CDPK family, are located on chromosomes 2B and 5A, respectively. The promoters of CPK7 and CPK12 have different cis-acting elements and their expression levels vary. Notably, TaCPK7 is expressed in response to drought, salt, cold, and hydrogen peroxide, whereas TaCPK12 is only expressed in response to ABA treatment. This functional divergence has complementary or amplifying effects on the TaCPK7-dependent stress signaling network in wheat [138,139].
Previous studies have demonstrated that cis-acting elements or trans-acting factors can also cause differences in the expression level of duplicated genes [34]. For instance, heat shock transcription factors (HSFs) upregulate the transcription of target-gene-encoding heat shock proteins (HSPs) via recognition and binding to promoter region heat shock elements (HSEs) in response to heat stress [77,140]. Atypical HSEs, containing mismatched nucleotides at specific positions, are more sensitive/responsive to heat stress and are expressed at higher levels than typical HSEs [96]. Thus, the distribution of diverse HSE motifs in the promoter regions of duplicate HSP genes (homologous or tandem repeats), combined with the effects of the HSE sequence variation, make the TaHSF-TaHSP module essential for heat stress regulation and adaptation in wheat [141]. In addition, similar studies have reported finding homolog expression bias between different copies of the same gene and functional differences due to specific variants, such as CLPB (Caseinolytic Protease B), SKP (S-phase kinase-associated protein 1), ALDH (Aldehyde dehydrogenase), NAC (NAM/ATAF1/2/CUC2), and SOS1 (Salt Overly Sensitive 1) in wheat [142,143,144,145,146,147]. These above results imply that differential regulation of duplicated genes can augment a plant’s ability to respond to environmental abiotic stress challenges.
Phenotypic and functional divergence is facilitated by changes in protein coding sequence between duplicated gene pairs. During gene duplication, structural alterations in regulatory areas, such as protein coding regions and small RNA binding sites, may also occur. Furthermore, gain or loss mutations in exons, introns, pseudo-exons, or indels have been shown to occur more frequently in duplicated genes than in single-copy homologs [148]. Numerous studies have reported finding single-nucleotide alterations or base insertions and deletions that alter the stress response of homologous genes [149,150,151,152]. For example, GWAS identified a favorable allele of TaDTG6, TaDTG6-B Del574, that harbored a frameshift mutation due to deletion of a 26-bp DNA fragment in the coding region. Strong protein and DNA sequence interaction properties of the encoded DREB protein allowed it to bind to the DRE/CRT cis-acting element, upregulate downstream gene expression, and ultimately enhance wheat drought tolerance [75]. Similarly, a variation in the region of subgenome A, but not subgenome B, near TaSnRK2.8 and a single-nucleotide polymorphism (SNP) in the TaSnRK2.8-A-C region conferred a stronger drought-tolerant response in wheat, accompanied by significantly greater seedling biomass and water-soluble carbohydrate contents [150]. Another recent study revealed that TaWD40-4B.1, located in the main drought tolerance QTL, qDT4B, along with an early termination codon generated through a nonsense mutation in TaWD40-4B, was highly correlated with drought resistance in a natural wheat population. These findings showed that wheat carrying TaWD40-4B.1C had significantly higher drought resistance than wheat carrying TaWD40-4B.1T [151]. In addition, microRNAs (miRNAs) also facilitate the control of gene expression in plant species primarily through detecting and cleaving particular regions on target genes to modify their function. Changes in miRNA binding sites and miRNA precursors of duplicated genes may also have different expression patterns and functions [153,154,155,156].
Gene duplication can play an essential role in preserving the integrity of a genetic system while also mitigating the effects of the surrounding environment on that system. That is, if one copy is inactivated through a mutation, other copies can still perform the original function to compensate for potential damage due to the inactivation [157,158]. This functional redundancy is commonly found in polyploid plants, and research has shown that most partially homologous gene copies in wheat are co-expressed. Multiple studies indicate that partially homologous gene copies can be directed by the same regulatory network [32,158,159], and that this functional redundancy between homologous genes is greatly expanded in heteropolyploid wheat. In one typical example, all five of the TaCYP81D tandem repeat genes, generated via lateral doubling within the same subgenome, were experimentally linked to salt tolerance in wheat. Among them, TaCYP81D5 was shown to potentially influence the ZAT12-mediated ROS signaling pathway in wheat [120]. By contrast, tetraploid wheat expressing the cyp81d5-aabb mutant showed no obvious difference from the wild type in salt tolerance, although CYP81D2 and CYP81D4 expression was significantly higher in the mutant than in the wild type. These results implied that the presence of other copies in this gene cluster could make up for the absence of one copy. This functional redundancy enables wheat plants to better withstand unfavorable natural genetic variations and, thus, maintain salt tolerance [112].
The above data show that duplicated genes serve as a crucial source for the development of new defenses against abiotic challenges in wheat, and that these defenses let plants continue to evolve in response to environmental stresses that might have been too severe for their ancestors. In addition to the above cases, a list of duplicate genes is provided in Table 1, including homologous and paralogous genes with structural and functional differences that contribute to wheat response to abiotic stress. A few such examples include drought-responsive TaHVA1, TaRAV, and TaNAC [72,73,160,161]; salt-responsive TaCHYR, TaWRKY75, and TaKNOX11 [84,90,129]; cold-responsive TaEXPA, TaICE, and TaAREB [108,162,163]; and heat-responsive TaMYB, TabZIP, and TaHAG [80,82,83,164,165].

4. Molecular Mechanisms of Abiotic Stress Resistance by Duplicated Genes

Duplicate genes are produced through a variety of mechanisms (Figure 1), and the rapid doubling of a genome during polyploidization can trigger large-scale genomic alterations, such as chromosomal rearrangements, gene inversions, and gene loss, in addition to generating a large number of duplicate genes. Other work suggests that polyploidy may be correlated with enhanced stress tolerance and higher reproductive fitness under stress conditions based on evidence that polyploids typically have a broader geographic distribution than diploid relatives [196,197,198,199,200]. Furthermore, crops that have undergone polyploidization are more prevalently cultivated. For instance, the tetraploid wild emmer wheat that is grown today, Triticum turgidum ssp. dicoccoides (2n = 4x = 28, BBAA), originated from the diploid wheat, Triticum urartu (2n = 2x = 14, AA), and its close relative, Aegilops speltoides (2n = 2x = 14, BB). Triticum aestivum L (2n = 6x = 42, BBAADD) was created through crossing wild emmer wheat with the diploid, Aegilops tauschii (2n = 2x = 14, DD) [201]. In the complete published genome of wheat cultivar, Chinese Spring, three copies can be found for more than half of the genes, and the copies are evenly distributed among the three homologous A, B, and D subgenomes [31]. These duplicated genes are referred to as homoeologs, and they share a high degree of sequence similarity as well as functional conservation and redundancy [31].
Although most of these homologous genes show synergistic expression patterns in the wheat population, a few homologous genes show negatively correlated patterns of expression. This inverse relationship is potentially due to regulatory effects of genetic variations, indicated by a gradual shift in the expression profiles of the same homologs from a positive to a negative correlation concomitant with an increasing number of SNPs in the region adjacent to one of the homologs [34]. Other research has shown that typical allohexaploid species have higher salt tolerance than their tetraploid wheat progenitors [164]. Thus, WGDs appear to improve the organismal capacity for adaptation to environmental challenges through introducing new genetic features and increasingly complex intragenomic network interactions [202,203]. Some recent work has uncovered another duplication phenomenon, distinct from WGD, related to significant genomic enrichment with TEs (transposable elements) [204,205,206,207,208], and there are important distinctions in functional enrichment and retention between genes produced via recent duplication and those produced via WGD [205,207]. In addition to WGD events, the pool of duplicate genes has also been considerably increased by numerous, small-scale subgenomic duplication events, including tandem duplications [1], segmental duplications [209], DNA-based transposition [208], and retrotransposon-mediated duplications [209]. Importantly, in each whole-genome or partial-chromosome-segment duplication event, a portion of genes are eliminated while another genomic fragment carrying duplicated genes is preserved to participate in further evolution.
The ability of crops to adapt to severe climatic conditions can depend heavily on copy number variations (CNVs). The FR-2 locus (Frost Resistance-2) has been linked to cold tolerance, which is consistent with a set of CBFs (C-repeat binding factors) that were found to regulate pathways involved in cold-climate domestication and cold tolerance [210,211]. Variability in CBF gene expression due to CNVs in the corresponding locus have been correlated with a cold-tolerant phenotype [210,211,212,213,214,215]. In particular, differences in gene coding sequence have been identified that directly impact phenotype, such as sequence variations in the CBF12 binding domains between winter and spring wheat accessions that modulate CBF12 binding activity at target loci [216] and enhanced cold tolerance in winter cereals with a high CBF copy number compared to that in single-copy spring cereals. In addition, CBF13 appears to have undergone pseudogenization in spring barley, based on the prevalence of sporadic nonsense codons, whereas its coding sequence remains intact in winter barley [212]. Other cold tolerance loci have been identified in wheat that are also functionally linked to CNVs in CBF genes [213,217].
Transposable elements (TEs) are small sequences, typically of viral origin, that mobilize to random locations throughout the genome in a “cut-and-paste” fashion (i.e., excised and reinserted), in the case of DNA transposons, or in a “copy-and-paste” fashion (i.e., RNAs encoding these elements are reverse transcribed and integrated into new sites, leaving the original), in the case of retrotransposons. Duplication can occur if a gene is co-replicated or co-transcribed/reverse transcribed with a TE, then integrated back into the genome at a new site, sometimes disrupting genes at the integration site [218]. Transposons thus mediate two main genomic effects: (1) they mobilize randomly throughout the genome, disrupting genes; and (2) they generate copies of the sequence in proximity to their integration site, resulting in new genes, pseudogenes, and cis-regulatory elements. Alterations in environmental conditions such as temperature, light, and water availability can impact transposon activation, in addition to internal factors that contribute to genomic instability. In a recent study examining the genomes of eleven crops, transposons accounted for 22% to 85% of the total genome content [219]. The wheat genome contains a multitude of TEs, which comprise ~85% of the sequence. For example, TE insertions in the promoter regions of Vrn1 homologs resulted in several loss-of-function mutations related to cold domestication [220]. Thus, the discovery, characterization, and use of TEs in crops, especially wheat, which necessarily entails an in-depth understanding of their role in gene duplication and the functions of the duplicated genes, could serve as a promising direction for future crop breeding.

5. Prospects

Gene duplications are mainly produced via polyploidization and WGD events [160,163]. In cereal crops, polyploidization increases the phenotypic diversity of polyploid species, improves the fitness of fixed hybrids, and enriches genetic diversity through altering genome structure [221,222]. Crops that can adapt to new habitats are needed to ensure stable, long-term food production as soil and climate conditions deteriorate globally. Hence, a key strategy for removing genetic barriers to crop improvement is to engineer the polyploid evolutionary process through utilizing distant crosses between related species [223,224]. In addition, it is necessary to identify potential stress-responsive allelic variations through dissecting the mechanisms underpinning transcriptional regulation and functional differentiation of duplicated genes in crops. Transgenic methods have been used to introduce superior genes into crops which confer high-yield potential but are accompanied by poor stress resistance. However, stress-tolerant phenotypes may be engineered through identifying differences in copy number, gene structure, expression level, and protein coding sequences of stress-responsive genes between tolerant and sensitive accessions. Despite the potential benefits of having several copies of stress response genes in the genome, genetic redundancy can also confound screening for distinct phenotypes. Gene-centric approaches, such as those utilizing CRISPR-Cas systems, could be used to achieve targeted improvements, reducing time and potentially increasing success rates. These approaches have been successful in model plants as well as in polyploid wheat, enabling functional genetic analysis in complex polyploid genomic backgrounds [225,226].
In summary, with natural selection continuing to serve as the cornerstone of evolutionary theory, we must expedite a deeper understanding of the origins, evolutionary mechanisms, and genetic basis of domestication that have culminated in modern wheat (Figure 2). Utilizing the genetic diversity and exploiting differences in wheat and its wild relatives is the foundation of germplasm innovation for the foreseeable future. At the same time, the threats posed by the increasing frequency and severity of global climate change warrant ongoing, extensive study of the genetic basis of traits that enable crop adaptation to environmental change. This advanced understanding will provide sound theoretical support for developing high-performance, stress-resilient germplasm and genetic resources for molecular breeding. Finally, the phenotypic characterization of alleles that enable continued high crop performance under variable or extreme field conditions is also essential for the effective deployment of stress-responsive genes identified in the laboratory.

Author Contributions

H.M. designed the framework of the article; L.D. and Z.M. wrote the manuscript and drew the figures; H.M. supervised the overall process. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China (grant no. 2022YFD1200202), the Science Foundation for Distinguished Young Scholars of Shaanxi Province (grant no. 2023-JC-JQ-20), and Chinese Universities Scientific Fund (grant no. 2452022347).

Acknowledgments

We apologize to colleagues whose work we were unable to cite due to space and reference limitations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Flagel, L.E.; Wendel, J.F. Gene duplication and evolutionary novelty in plants. New Phytol. 2009, 183, 557–564. [Google Scholar] [CrossRef]
  2. Wang, Y.; Wang, X.; Paterson, A.H. Genome and gene duplications and gene expression divergence: A view from plants. Ann. N. Y. Acad. Sci. 2012, 1256, 1–14. [Google Scholar] [CrossRef]
  3. Jaillon, O.; Aury, J.-M.; Brunet, F.; Petit, J.-L.; Stange-Thomann, N.; Mauceli, E.; Bouneau, L.; Fischer, C.; Ozouf-Costaz, C.; Bernot, A.; et al. Genome duplication in the teleost fish Tetraodon nigroviridis reveals the early vertebrate proto-karyotype. Nature 2004, 431, 946–957. [Google Scholar] [CrossRef] [Green Version]
  4. Aury, J.-M.; Jaillon, O.; Duret, L.; Noel, B.; Jubin, C.; Porcel, B.M.; Ségurens, B.; Daubin, V.; Anthouard, V.; Aiach, N.; et al. Global trends of whole-genome duplications revealed by the ciliate Paramecium tetraurelia. Nature 2006, 444, 171–178. [Google Scholar] [CrossRef] [Green Version]
  5. Wolfe, K.H.; Shields, D.C. Molecular evidence for an ancient duplication of the entire yeast genome. Nature 1997, 387, 708–713. [Google Scholar] [CrossRef]
  6. Kellis, M.; Blrren, B.W.; Lander, E.S. Proof and evolutionary analysis of ancient genome duplication in the yeast Saccharomyces cerevisiae. Nature 2004, 428, 617–624. [Google Scholar] [CrossRef]
  7. Kejnovsky, E.; Leitch, I.J.; Leitch, A.R. Contrasting evolutionary dynamics between angiosperm and mammalian genomes. Trends Ecol. Evol. 2009, 24, 572–582. [Google Scholar] [CrossRef]
  8. Murat, F.; Van De Peer, Y.; Salse, J. Decoding Plant and Animal Genome Plasticity from Differential Paleo-Evolutionary Patterns and Processes. Genome Biol. Evol. 2012, 4, 917–928. [Google Scholar] [CrossRef] [Green Version]
  9. Force, A.; Lynch, M.; Pickett, F.B.; Amores, A.; Yan, Y.-L.; Postlethwait, J. Preservation of Duplicate Genes by Complementary, Degenerative Mutations. Genetics 1999, 151, 1531–1545. [Google Scholar] [CrossRef]
  10. Kuzmin, E.; VanderSluis, B.; Ba, A.N.N.; Wang, W.; Koch, E.N.; Usaj, M.; Khmelinskii, A.; Usaj, M.M.; van Leeuwen, J.; Kraus, O.; et al. Exploring whole-genome duplicate gene retention with complex genetic interaction analysis. Science 2020, 368, eaaz5667. [Google Scholar] [CrossRef]
  11. Magadum, S.; Banerjee, U.; Murugan, P.; Gangapur, D.; Ravikesavan, R. Gene duplication as a major force in evolution. J. Genet. 2013, 92, 155–161. [Google Scholar] [CrossRef]
  12. Matsuoka, Y. Evolution of Polyploid Triticum Wheats under Cultivation: The Role of Domestication, Natural Hybridization and Allopolyploid Speciation in their Diversification. Plant Cell Physiol. 2011, 52, 750–764. [Google Scholar] [CrossRef] [Green Version]
  13. Chen, Z.J. Genetic and Epigenetic Mechanisms for Gene Expression and Phenotypic Variation in Plant Polyploids. Annu. Rev. Plant Biol. 2007, 58, 377–406. [Google Scholar] [CrossRef] [Green Version]
  14. Li, A.-L.; Geng, S.-F.; Zhang, L.-Q.; Liu, D.-C.; Mao, L. Making the Bread: Insights from Newly Synthesized Allohexaploid Wheat. Mol. Plant 2015, 8, 847–859. [Google Scholar] [CrossRef] [Green Version]
  15. Freeling, M.; Thomas, B.C. Gene-balanced duplications, like tetraploidy, provide predictable drive to increase morphological complexity. Genome Res. 2006, 16, 805–814. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, D.; Wang, Y.; Long, W.; Niu, M.; Zhao, Z.; Teng, X.; Zhu, X.; Zhu, J.; Hao, Y.; Wang, Y.; et al. SGD1, a key enzyme in tocopherol biosynthesis, is essential for plant development and cold tolerance in rice. Plant Sci. 2017, 260, 90–100. [Google Scholar] [CrossRef]
  17. Panchy, N.; Lehti-Shiu, M.; Shiu, S.-H. Evolution of Gene Duplication in Plants. Plant Physiol. 2016, 171, 2294–2316. [Google Scholar] [CrossRef] [Green Version]
  18. Dong, S.; Adams, K.L. Differential contributions to the transcriptome of duplicated genes in response to abiotic stresses in natural and synthetic polyploids. New Phytol. 2011, 190, 1045–1057. [Google Scholar] [CrossRef]
  19. Liu, Z.; Adams, K.L. Expression Partitioning between Genes Duplicated by Polyploidy under Abiotic Stress and during Organ Development. Curr. Biol. 2007, 17, 1669–1674. [Google Scholar] [CrossRef] [Green Version]
  20. Myburg, A.A.; Grattapaglia, D.; Tuskan, G.A.; Hellsten, U.; Hayes, R.D.; Grimwood, J.; Jenkins, J.; Lindquist, E.; Tice, H.; Bauer, D.; et al. The genome of Eucalyptus grandis. Nature 2014, 510, 356–362. [Google Scholar] [CrossRef] [Green Version]
  21. Lee, T.-H.; Tang, H.; Wang, X.; Paterson, A.H. PGDD: A database of gene and genome duplication in plants. Nucleic Acids Res. 2012, 41, 1152–1158. [Google Scholar] [CrossRef] [PubMed]
  22. Lu, F.; Lipka, A.E.; Glaubitz, J.; Elshire, R.; Cherney, J.H.; Casler, M.D.; Buckler, E.S.; Costich, D.E. Switchgrass Genomic Diversity, Ploidy, and Evolution: Novel Insights from a Network-Based SNP Discovery Protocol. PLoS Genet. 2013, 9, e1003215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Panopoulou, G.; Hennig, S.; Groth, D.; Krause, A.; Poustka, A.J.; Herwig, R.; Vingron, M.; Lehrach, H. New Evidence for Genome-Wide Duplications at the Origin of Vertebrates Using an Amphioxus Gene Set and Completed Animal Genomes. Genome Res. 2003, 13, 1056–1066. [Google Scholar] [CrossRef] [Green Version]
  24. Dehal, P.; Boore, J.L. Two Rounds of Whole Genome Duplication in the Ancestral Vertebrate. PLoS Biol. 2005, 3, e314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Greilhuber, J.; Dolezel, J.; Wendel, J.F. The incidence of polyploidy in natural plant populations: Major patterns and evolutionary processes. In Plant Genome Diversity; Spinger: Berlin/Heidelberg, Germany, 2013; Chapter 16; pp. 255–276. [Google Scholar]
  26. Scarpino, S.V.; Levin, D.A.; Meyers, L.A. Polyploid formation shapes flowering plant diversity. Am. Nat. Devoted Concept. Unification Biol. Sci. 2014, 184, 456–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Diallo, A.M.; Nielsen, L.R.; Kjær, E.D.; Petersen, K.K.; Ræbild, A. Polyploidy can Confer Superiority to West African Acacia senegal (L.) Willd. Trees. Front. Plant Sci. 2016, 7, 821. [Google Scholar] [CrossRef] [Green Version]
  28. Ling, H.-Q.; Ma, B.; Shi, X.; Liu, H.; Dong, L.; Sun, H.; Cao, Y.; Gao, Q.; Zheng, S.; Li, Y.; et al. Genome sequence of the progenitor of wheat A subgenome Triticum urartu. Nature 2018, 557, 424–428. [Google Scholar] [CrossRef] [Green Version]
  29. Luo, M.-C.; Gu, Y.Q.; Puiu, D.; Wang, H.; Twardziok, S.O.; Deal, K.R.; Huo, N.; Zhu, T.; Wang, L.; Wang, Y.; et al. Genome sequence of the progenitor of the wheat D genome Aegilops tauschii. Nature 2017, 551, 498–502. [Google Scholar] [CrossRef] [Green Version]
  30. Maccaferri, M.; Harris, N.S.; Twardziok, S.O.; Pasam, R.K.; Gundlach, H.; Spannagl, M.; Ormanbekova, D.; Lux, T.; Prade, V.M.; Milner, S.G.; et al. Durum wheat genome highlights past domestication signatures and future improvement targets. Nat. Genet. 2019, 51, 885–895. [Google Scholar] [CrossRef] [Green Version]
  31. International Wheat Genome Sequencing Consortium (IWGSC). Shifting the limits in wheat research and breeding using a fully annotated reference genome. Science 2018, 361, eaar7191. [Google Scholar] [CrossRef] [Green Version]
  32. Jordan, K.W.; Wang, S.; Lun, Y.; Gardiner, L.J.; MacLachlan, R.; Hucl, P.; Wiebe, K.; Wong, D.; Forrest, K.L.; IWGS-Consortium; et al. A haplotype map of allohexaploid wheat reveals distinct patterns of selection on homoeologous genomes. Genome Biol. 2015, 16, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Wang, X.; Wang, R.; Ma, C.; Shi, X.; Liu, Z.; Wang, Z.; Sun, Q.; Cao, J.; Xu, S. Massive expansion and differential evolution of small heat shock proteins with wheat (Triticum aestivum L.) polyploidization. Sci. Rep. 2017, 7, 2581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. He, F.; Wang, W.; Rutter, W.B.; Jordan, K.W.; Ren, J.; Taagen, E.; DeWitt, N.; Sehgal, D.; Sukumaran, S.; Dreisigacker, S.; et al. Genomic variants affecting homoeologous gene expression dosage contribute to agronomic trait variation in allopolyploid wheat. Nat. Commun. 2022, 13, 826. [Google Scholar] [CrossRef]
  35. He, F.; Pasam, R.; Shi, F.; Kant, S.; Keeble-Gagnere, G.; Kay, P.; Forrest, K.; Fritz, A.; Hucl, P.; Wiebe, K.; et al. Exome sequencing highlights the role of wild-relative introgression in shaping the adaptive landscape of the wheat genome. Nat. Genet. 2019, 51, 896–904. [Google Scholar] [CrossRef] [PubMed]
  36. Akhunov, E.D.; Akhunova, A.R.; Anderson, O.D.; Anderson, J.A.; Blake, N.; Clegg, M.T.; Coleman-Derr, D.; Conley, E.J.; Crossman, C.C.; Deal, K.R.; et al. Nucleotide diversity maps reveal variation in diversity among wheat genomes and chromosomes. BMC Genom. 2010, 14, 702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Dvorak, J.; Akhunov, E.D.; Akhunov, A.R.; Deal, K.R.; Luo, M.-C. Molecular Characterization of a Diagnostic DNA Marker for Domesticated Tetraploid Wheat Provides Evidence for Gene Flow from Wild Tetraploid Wheat to Hexaploid Wheat. Mol. Biol. Evol. 2006, 23, 1386–1396. [Google Scholar] [CrossRef] [Green Version]
  38. Yan, J.; Su, P.; Meng, X.; Liu, P. Phylogeny of the plant receptor-like kinase (RLK) gene family and expression analysis of wheat RLK genes in response to biotic and abiotic stresses. BMC Genom. 2023, 24, 224. [Google Scholar] [CrossRef]
  39. 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]
  40. Dievart, A.; Gottin, C.; Périn, C.; Ranwez, V.; Chantret, N. Origin and Diversity of Plant Receptor-Like Kinases. Annu. Rev. Plant Biol. 2020, 71, 131–156. [Google Scholar] [CrossRef] [Green Version]
  41. Shumayla Sharma, S.; Pandey, A.K.; Singh, K.; Upadhyay, S.K. Molecular characterization and global expression analysis of lectin receptor kinases in bread wheat (Triticum aestivum). PLoS ONE 2016, 11, e0153925. [Google Scholar] [CrossRef] [Green Version]
  42. Shumayla Sharma, S.; Kumar, R.; Mendu, V.; Singh, K.; Upadhyay, S.K. Genomic dissection and expression profiling revealed functional divergence in Triticum aestivum leucine rich repeat receptor like kinases (TaLRRKs). Front. Plant Sci. 2016, 7, 1374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Soltabayeva, A.; Dauletova, N.; Serik, S.; Sandybek, M.; Omondi, J.O.; Kurmanbayeva, A.; Srivastava, S. Receptor-like Kinases (LRR-RLKs) in Response of Plants to Biotic and Abiotic Stresses. Plants 2022, 11, 2660. [Google Scholar] [CrossRef]
  44. Shumayla; Tyagi, S.; Sharma, A.; Singh, K.; Upadhyay, S.K. Genomic dissection and transcriptional profiling of Cysteine-rich receptor-like kinases in five cereals and functional characterization of TaCRK68-A. Int. J. Biol. Macromol. 2019, 134, 316–329. [Google Scholar] [CrossRef]
  45. Sharma, A.; Sharma, H.; Rajput, R.; Pandey, A.; Upadhyay, S.K. Molecular Characterization Revealed the Role of Thaumatin-Like Proteins of Bread Wheat in Stress Response. Front. Plant Sci. 2022, 12, 807448. [Google Scholar] [CrossRef]
  46. Shumayla Mendu, V.; Singh, K.; Upadhyay, S.K. Insight into the roles of proline-rich extensin-like receptor protein kinases of bread wheat (Triticum aestivum L.). Life 2022, 12, 941. [Google Scholar] [CrossRef]
  47. Kesawat, M.S.; Kherawat, B.S.; Singh, A.; Dey, P.; Routray, S.; Mohapatra, C.; Saha, D.; Ram, C.; Siddique, K.H.M.; Kumar, A.; et al. Genome-Wide Analysis and Characterization of the Proline-Rich Extensin-like Receptor Kinases (PERKs) Gene Family Reveals Their Role in Different Developmental Stages and Stress Conditions in Wheat (Triticum aestivum L.). Plants 2022, 11, 496. [Google Scholar] [CrossRef] [PubMed]
  48. Nadarajah, K.K. ROS Homeostasis in Abiotic Stress Tolerance in Plants. Int. J. Mol. Sci. 2020, 21, 5208. [Google Scholar] [CrossRef] [PubMed]
  49. Tyagi, S.; Shumayla; Madhu; Singh, K.; Upadhyay, S.K. Molecular characterization revealed the role of catalases under abiotic and arsenic stress in bread wheat (Triticum aestivum L.). J. Hazard. Mater. 2021, 403, 123585. [Google Scholar] [CrossRef]
  50. Zhang, Y.; Zheng, L.; Yun, L.; Ji, L.; Li, G.; Ji, M.; Shi, Y.; Zheng, X. Catalase (CAT) gene family in wheat (Triticum aestivum L.): Evolution, expression pattern and function analysis. Int. J. Mol. Sci. 2022, 23, 542. [Google Scholar] [CrossRef]
  51. Tyagi, S.; Sharma, S.; Taneja, M.; Kumar, R.; Sembi, J.K.; Upadhyay, S.K. Superoxide dismutases in bread wheat (Triticum aestivum L.): Comprehensive characterization and expression analysis during development and, biotic and abiotic stresses. Agri. Gene 2017, 6, 1–13. [Google Scholar] [CrossRef]
  52. Wang, M.; Zhao, X.; Xiao, Z.; Yin, X.; Xing, T.; Xia, G. A wheat superoxide dismutase gene TaSOD2 enhances salt resistance through modulating redox homeostasis by promoting NADPH oxidase activity. Plant Mol. Biol. 2016, 91, 115–130. [Google Scholar] [CrossRef] [PubMed]
  53. Jiang, W.; Yang, L.; He, Y.; Zhang, H.; Li, W.; Chen, H.; Ma, D.; Yin, J. Genome-wide identification and transcriptional expression analysis of superoxide dismutase (SOD) family in wheat (Triticum aestivum). PeerJ 2019, 7, e8062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Tyagi, S.; Shumayla Verma, P.C.; Singh, K.; Upadhyay, S.K. Molecular characterization of ascorbate peroxidase (APX) and APX-related (APX-R) genes in Triticum aestivum L. Genomics 2020, 112, 4208–4223. [Google Scholar] [CrossRef] [PubMed]
  55. Tyagi, S.; Himani; Sembi, J.K.; Upadhyay, S.K. Gene architecture and expression analyses provide insights into the role of glutathione peroxidases (GPXs) in bread wheat (Triticum aestivum L.). J. Plant Physiol. 2018, 223, 19–31. [Google Scholar] [CrossRef]
  56. Zhai, C.-Z.; Zhao, L.; Yin, L.-J.; Chen, M.; Wang, Q.-Y.; Li, L.-C.; Xu, Z.-S.; Ma, Y.-Z. Two Wheat Glutathione Peroxidase Genes Whose Products Are Located in Chloroplasts Improve Salt and H2O2 Tolerances in Arabidopsis. PLoS ONE 2013, 8, e73989. [Google Scholar] [CrossRef] [Green Version]
  57. Su, P.; Yan, J.; Li, W.; Wang, L.; Zhao, J.; Ma, X.; Li, A.; Wang, H.; Kong, L. A member of wheat class III peroxidase gene family, TaPRX-2A, enhanced the tolerance of salt stress. BMC Plant Biol. 2020, 20, 392. [Google Scholar] [CrossRef]
  58. Madhu; Kaur, A.; Tyagi, S.; Shumayla; Singh, K.; Upadhyay, S.K. Exploration of glutathione reductase for abiotic stress response in bread wheat (Triticum aestivum L.). Plant Cell Rep. 2022, 41, 639–654. [Google Scholar] [CrossRef]
  59. Hamamoto, S.; Horie, T.; Hauser, F.; Deinlein, U.; Schroeder, J.I.; Uozumi, N. HKT transporters mediate salt stress resistance in plants: From structure and function to the field. Curr. Opin. Biotechnol. 2015, 32, 113–120. [Google Scholar] [CrossRef] [Green Version]
  60. Saddhe, A.A.; Mishra, A.K.; Kumar, K. Molecular insights into the role of plant transporters in salt stress response. Physiol. Plant. 2021, 173, 1481–1494. [Google Scholar] [CrossRef]
  61. Gill, R.A.; Ahmar, S.; Ali, B.; Saleem, M.H.; Khan, M.U.; Zhou, W.; Liu, S. The Role of Membrane Transporters in Plant Growth and Development, and Abiotic Stress Tolerance. Int. J. Mol. Sci. 2021, 22, 12792. [Google Scholar] [CrossRef]
  62. Tong, K.; Wu, X.; He, L.; Qiu, S.; Liu, S.; Cai, L.; Rao, S.; Chen, J. Genome-Wide Identification and Expression Profile of OSCA Gene Family Members in Triticum aestivum L. Int. J. Mol. Sci. 2021, 23, 469. [Google Scholar] [CrossRef] [PubMed]
  63. Kaur, A.; Sharma, A.; Madhu; Dixit, S.; Singh, K.; Upadhyay, S.K. OSCA Genes in Bread Wheat: Molecular Characterization, Expression Profiling, and Interaction Analyses Indicated Their Diverse Roles during Development and Stress Response. Int. J. Mol. Sci. 2022, 23, 14867. [Google Scholar] [CrossRef]
  64. Wang, Y.; Niu, Z.; Hu, X.; Wu, X.; Yang, Z.; Hao, C.; Zhou, M.; Yang, S.; Dong, N.; Liu, M.; et al. Molecular characterization of the genome-wide BOR transporter family and their responses to boron conditions in common wheat (Triticum aestivum L.). Front. Plant Sci. 2022, 13, 997915. [Google Scholar] [CrossRef] [PubMed]
  65. Sharma, H.; Sharma, A.; Rajput, R.; Sidhu, S.; Dhillon, H.; Verma, P.C.; Pandey, A.; Upadhyay, S.K. Molecular characterization, evolutionary analysis, and expression profiling of BOR genes in important cereals. Plants 2022, 11, 911. [Google Scholar] [CrossRef] [PubMed]
  66. Kaur, A.; Taneja, M.; Tyagi, S.; Sharma, A.; Singh, K.; Upadhyay, S.K. Genome-wide characterization and expression analysis suggested diverse functions of the mechanosensitive channel of small conductance-like (MSL) genes in cereal crops. Sci. Rep. 2020, 10, 16583. [Google Scholar] [CrossRef]
  67. Taneja, M.; Tyagi, S.; Sharma, S.; Upadhyay, S.K. Ca2+/Cation Antiporters (CaCA): Identification, Characterization and Expression Profiling in Bread Wheat (Triticum aestivum L.). Front. Plant Sci. 2016, 7, 1775. [Google Scholar] [CrossRef] [Green Version]
  68. Sharma, H.; Taneja, M.; Upadhyay, S.K. Identification, characterization and expression profiling of cation-proton antiporter superfamily in Triticum aestivum L. and functional analysis of TaNHX4-B. Genomics 2020, 112, 356–370. [Google Scholar] [CrossRef]
  69. Taneja, M.; Upadhyay, S.K. Molecular characterization and differential expression suggested diverse functions of P-type II Ca2+ATPases in Triticum aestivum L. BMC Genom. 2018, 19, 389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Sharma, A.; Shumayla Tyagi, S.; Alok, A.; Singh, K.; Upadhyay, S.K. Thaumatin-like protein kinases: Molecular characterization and transcriptional profiling in five cereal crops. Plant Sci. 2020, 290, 110317. [Google Scholar] [CrossRef]
  71. Mao, H.; Li, S.; Wang, Z.; Cheng, X.; Li, F.; Mei, F.; Chen, N.; Kang, Z. Regulatory changes in TaSNAC8-6A are associated with drought tolerance in wheat seedlings. Plant Biotechnol. J. 2020, 18, 1078–1092. [Google Scholar] [CrossRef] [Green Version]
  72. Mao, H.; Li, S.; Chen, B.; Jian, C.; Mei, F.; Zhang, Y.; Li, F.; Chen, N.; Li, T.; Du, L.; et al. Variation in cis-regulation of a NAC transcription factor contributes to drought tolerance in wheat. Mol. Plant 2022, 15, 276–292. [Google Scholar] [CrossRef] [PubMed]
  73. Ma, J.; Zhang, M.; Lv, W.; Tang, X.; Zhao, D.; Wang, L.; Li, C.; Jiang, L. Overexpression of TaSNAC4-3D in Common Wheat (Triticum aestivum L.) Negatively Regulates Drought Tolerance. Front. Plant Sci. 2022, 13, 945272. [Google Scholar] [CrossRef]
  74. Mei, F.; Chen, B.; Li, F.; Zhang, Y.; Kang, Z.; Wang, X.; Mao, H. Overexpression of the wheat NAC transcription factor TaSNAC4-3A gene confers drought tolerance in transgenic Arabidopsis. Plant Physiol. Biochem. 2021, 160, 37–50. [Google Scholar] [CrossRef]
  75. Mei, F.; Chen, B.; Du, L.; Li, S.; Zhu, D.; Chen, N.; Zhang, Y.; Li, F.; Wang, Z.; Cheng, X.; et al. A gain-of-function allele of a DREB transcription factor gene ameliorates drought tolerance in wheat. Plant Cell 2022, 34, 4472–4494. [Google Scholar] [CrossRef] [PubMed]
  76. Bi, H.; Zhao, Y.; Li, H.; Liu, W. Wheat Heat Shock Factor TaHsfA6f Increases ABA Levels and Enhances Tolerance to Multiple Abiotic Stresses in Transgenic Plants. Int. J. Mol. Sci. 2020, 21, 3121. [Google Scholar] [CrossRef] [PubMed]
  77. Bi, H.; Miao, J.; He, J.; Chen, Q.; Qian, J.; Li, H.; Xu, Y.; Ma, D.; Zhao, Y.; Tian, X.; et al. Characterization of the Wheat Heat Shock Factor TaHsfA2e-5D Conferring Heat and Drought Tolerance in Arabidopsis. Int. J. Mol. Sci. 2022, 23, 2784. [Google Scholar] [CrossRef] [PubMed]
  78. Meena, S.; Samtani, H.; Khurana, P. Elucidating the functional role of heat stress transcription factor A6b (TaHsfA6b) in linking heat stress response and the unfolded protein response in wheat. Plant Mol. Biol. 2022, 108, 621–634. [Google Scholar] [CrossRef]
  79. Kalaipandian, S.; Powell, J.; Karunakaran, A.; Stiller, J.; Adkins, S.; Kage, U.; Kazan, K.; Fleury, D. Transcriptome Analysis of Heat Shock Factor C2a Over-Expressing Wheat Roots Reveals Ferroptosis-like Cell Death in Heat Stress Recovery. Int. J. Mol. Sci. 2023, 24, 3099. [Google Scholar] [CrossRef]
  80. Wei, Q.; Chen, R.; Wei, X.; Liu, Y.; Zhao, S.; Yin, X.; Xie, T. Genome-wide identification of R2R3-MYB family in wheat and functional characteristics of the abiotic stress responsive gene TaMYB344. BMC Genom. 2020, 21, 792. [Google Scholar] [CrossRef]
  81. Zhang, L.; Zhao, G.; Xia, C.; Jia, J.; Liu, X.; Kong, X. Overexpression of a wheat MYB transcription factor gene, TaMYB56-B, enhances tolerances to freezing and salt stresses in transgenic Arabidopsis. Gene 2012, 505, 100–107. [Google Scholar] [CrossRef]
  82. Geng, X.; Zang, X.; Li, H.; Liu, Z.; Zhao, A.; Liu, J.; Peng, H.; Yao, Y.; Hu, Z.; Ni, Z.; et al. Unconventional splicing of wheat TabZIP60 confers heat tolerance in transgenic Arabidopsis. Plant Sci. 2018, 274, 252–260. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, L.; Zhang, L.; Xia, C.; Gao, L.; Hao, C.; Zhao, G.; Jia, J.; Kong, X. A Novel Wheat C-bZIP Gene, TabZIP14-B, Participates in Salt and Freezing Tolerance in Transgenic Plants. Front. Plant Sci. 2017, 9, 710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ye, H.; Qiao, L.; Guo, H.; Guo, L.; Ren, F.; Bai, J.; Wang, Y. Genome-Wide Identification of Wheat WRKY Gene Family Reveals That TaWRKY75-A Is Referred to Drought and Salt Resistances. Front. Plant Sci. 2021, 12, 663118. [Google Scholar] [CrossRef] [PubMed]
  85. Yu, Y.; Song, T.; Wang, Y.; Zhang, M.; Li, N.; Yu, M.; Zhang, S.; Zhou, H.; Guo, S.; Bu, Y.; et al. The wheat WRKY transcription factor TaWRKY1-2D confers drought resistance in transgenic Arabidopsis and wheat (Triticum aestivum L.). Int. J. Biol. Macromol. 2023, 226, 1203–1217. [Google Scholar] [CrossRef]
  86. Yu, Y.; Yu, M.; Zhang, S.; Song, T.; Zhang, M.; Zhou, H.; Wang, Y.; Xiang, J.; Zhang, X. Transcriptomic identification of wheat AP2/ERF transcription factors and functional characterization of TaERF-6-3A in response to drought and salinity stresses. Int. J. Mol. Sci. 2022, 23, 3272. [Google Scholar] [CrossRef]
  87. Shao, W.; Chen, W.; Zhu, X.; Zhou, X.; Jin, Y.; Zhan, C.; Liu, G.; Liu, X.; Ma, D.; Qiao, Y. Genome-Wide Identification and Characterization of Wheat 14-3-3 Genes Unravels the Role of TaGRF6-A in Salt Stress Tolerance by Binding MYB Transcription Factor. Int. J. Mol. Sci. 2021, 22, 1904. [Google Scholar] [CrossRef]
  88. Yue, H.; Shu, D.; Wang, M.; Xing, G.; Zhan, H.; Du, X.; Song, W.; Nie, X. Genome-Wide Identification and Expression Analysis of the HD-Zip Gene Family in Wheat (Triticum aestivum L.). Genes 2018, 9, 70. [Google Scholar] [CrossRef] [Green Version]
  89. Rathour, M.; Shumayla; Alok, A.; Upadhyay, S.K. Investigation of Roles of TaTALE Genes during Development and Stress Response in Bread Wheat. Plants 2022, 11, 587. [Google Scholar] [CrossRef]
  90. Han, Y.; Zhang, L.; Yan, L.; Xiong, X.; Wang, W.; Zhang, X.H.; Min, D.H. Genome-wide analysis of TALE superfamily in Triticum aestivum reveals TaKNOX11-A is involved in abiotic stress response. BMC Genom. 2022, 23, 89. [Google Scholar] [CrossRef]
  91. Liu, H.; Yang, Y.; Zhang, L. Zinc finger-homeodomain transcriptional factors (ZF-HDs) in wheat (Triticum aestivum L.): Identification, evolution, expression analysis and response to abiotic stresses. Plants 2021, 10, 593. [Google Scholar] [CrossRef]
  92. Li, Z.; Liu, D.; Xia, Y.; Li, Z.; Jing, D.; Du, J.; Niu, N.; Ma, S.; Wang, J.; Song, Y.; et al. Identification of the WUSCHEL-Related Homeobox (WOX) Gene Family, and Interaction and Functional Analysis of TaWOX9 and TaWUS in Wheat. Int. J. Mol. Sci. 2020, 21, 1581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Rathour, M.; Sharma, A.; Kaur, A.; Upadhyay, S.K. Genome-wide characterization and expression and co-expression analysis suggested diverse functions of WOX genes in bread wheat. Heliyon 2020, 6, e05762. [Google Scholar] [CrossRef] [PubMed]
  94. Hu, Z.; Yu, Y.; Wang, R.; Yao, Y.; Peng, H.; Ni, Z.; Sun, Q. Expression divergence of TaMBD2 homoeologous genes encoding methyl CpG-binding domain proteins in wheat (Triticum aestivum L.). Gene 2011, 471, 13–18. [Google Scholar] [CrossRef] [PubMed]
  95. Derakhshani, B.; Ayalew, H.; Mishina, K.; Tanaka, T.; Kawahara, Y.; Jafary, H.; Oono, Y. Comparative Analysis of Root Transcriptome Reveals Candidate Genes and Expression Divergence of Homoeologous Genes in Response to Water Stress in Wheat. Plants 2020, 9, 596. [Google Scholar] [CrossRef]
  96. Zhao, P.; Javed, S.; Shi, X.; Wu, B.; Zhang, D.; Xu, S.; Wang, X. Varying Architecture of Heat Shock Elements Contributes to Distinct Magnitudes of Target Gene Expression and Diverged Biological Pathways in Heat Stress Response of Bread Wheat. Front. Genet. 2020, 11, 30. [Google Scholar] [CrossRef] [Green Version]
  97. Yu, G.; Hou, W.; Du, X.; Wang, L.; Wu, H.; Zhao, L.; Kong, L.; Wang, H. Identification of wheat non-specific lipid transfer proteins involved in chilling tolerance. Plant Cell Rep. 2014, 33, 1757–1766. [Google Scholar] [CrossRef] [PubMed]
  98. Yu, K.; Feng, M.; Yang, G.; Sun, L.; Qin, Z.; Cao, J.; Wen, J.; Li, H.; Zhou, Y.; Chen, X.; et al. Changes in Alternative Splicing in Response to Domestication and Polyploidization in Wheat. Plant Physiol. 2020, 184, 1955–1968. [Google Scholar] [CrossRef]
  99. Ma, Z.; Li, M.; Zhang, H.; Zhao, B.; Liu, Z.; Duan, S.; Meng, X.; Li, G.; Guo, X. Alternative Splicing of TaHsfA2-7 Is Involved in the Improvement of Thermotolerance in Wheat. Int. J. Mol. Sci. 2023, 24, 1014. [Google Scholar] [CrossRef]
  100. Xu, K.; Zhao, Y.; Zhao, S.; Liu, H.; Wang, W.; Zhang, S.; Yang, X. Genome-Wide Identification and Low Temperature Responsive Pattern of Actin Depolymerizing Factor (ADF) Gene Family in Wheat (Triticum aestivum L.). Front. Plant Sci. 2021, 12, 618984. [Google Scholar] [CrossRef]
  101. Zotova, L.; Shamambaeva, N.; Lethola, K.; Alharthi, B.; Vavilova, V.; Smolenskaya, S.E.; Goncharov, N.P.; Kurishbayev, A.; Jatayev, S.; Gupta, N.K.; et al. TaDrAp1 and TaDrAp2, Partner Genes of a Transcription Repressor, Coordinate Plant Development and Drought Tolerance in Spelt and Bread Wheat. Int. J. Mol. Sci. 2020, 21, 8296. [Google Scholar] [CrossRef]
  102. Wang, B.; Li, L.; Liu, M.; Peng, D.; Wei, A.; Hou, B.; Lei, Y.; Li, X. TaFDL2-1A confers drought stress tolerance by promoting ABA biosynthesis, ABA responses, and ROS scavenging in transgenic wheat. Plant J. 2022, 112, 722–737. [Google Scholar] [CrossRef]
  103. Li, X.; Tang, Y.; Li, H.; Luo, W.; Zhou, C.; Zhang, L.; Lv, J. A wheat R2R3 MYB gene TaMpc1-D4 negatively regulates drought tolerance in transgenic Arabidopsis and wheat. Plant Sci. 2020, 299, 110613. [Google Scholar] [CrossRef]
  104. Zheng, X.; Liu, H.; Ji, H.; Wang, Y.; Dong, B.; Qiao, Y.; Liu, M.; Li, X. The Wheat GT Factor TaGT2L1D Negatively Regulates Drought Tolerance and Plant Development. Sci. Rep. 2016, 6, 27042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zotova, L.; Kurishbayev, A.; Jatayev, S.; Khassanova, G.; Zhubatkanov, A.; Serikbay, D.; Sereda, S.; Sereda, T.; Shvidchenko, V.; Lopato, S.; et al. Genes encoding transcription factors TaDREB5 and TaNFYC-A7 are differentially expressed in leaves of bread wheat in response to drought, dehydration and ABA. Front Plant Sci. 2018, 9, 1441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Chauhan, H.; Khurana, N.; Agarwal, P.; Khurana, J.P.; Khurana, P. A Seed Preferential Heat Shock Transcription Factor from Wheat Provides Abiotic Stress Tolerance and Yield Enhancement in Transgenic Arabidopsis under Heat Stress Environment. PLoS ONE 2013, 8, e79577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Xu, Z.-S.; Ni, Z.-Y.; Liu, L.; Nie, L.-N.; Li, L.-C.; Chen, M.; Ma, Y.-Z. Characterization of the TaAIDFa gene encoding a CRT/DRE-binding factor responsive to drought, high-salt, and cold stress in wheat. Mol. Genet. Genom. 2008, 280, 497–508. [Google Scholar] [CrossRef] [PubMed]
  108. Wang, J.Y.; Li, Q.; Mao, X.G.; Li, A.; Jing, R.L. Wheat transcription factor TaAREB3 participates in drought and freezing tolerances in Arabidopsis. Int. J. Biol. Sci. 2016, 12, 257–269. [Google Scholar] [CrossRef] [Green Version]
  109. Egawa, C.; Kobayashi, F.; Ishibashi, M.; Nakamura, T.; Nakamura, C.; Takumi, S. Differential regulation of transcript accumulation and alternative splicing of a DREB2 homolog under abiotic stress conditions in common wheat. Genes Genet. Syst. 2006, 81, 77–91. [Google Scholar] [CrossRef] [Green Version]
  110. Xu, S.; Dong, Q.; Deng, M.; Lin, D.; Xiao, J.; Cheng, P.; Xing, L.; Niu, Y.; Gao, C.; Zhang, W.; et al. The vernalization-induced long non-coding RNA VAS functions with the transcription factor TaRF2b to promote TaVRN1 expression for flowering in hexaploid wheat. Mol. Plant 2021, 14, 1525–1538. [Google Scholar] [CrossRef]
  111. Lv, Z.; Li, Z.; Wang, M.; Zhao, F.; Zhang, W.; Li, C.; Gong, L.; Zhang, Y.; Mason, A.S.; Liu, B. Conservation and trans-regulation of histone modification in the A and B subgenomes of polyploid wheat during domestication and ploidy transition. BMC Biol. 2021, 19, 42. [Google Scholar] [CrossRef]
  112. Kong, L.; Liu, Y.; Wang, X.; Chang, C. Insight into the Role of Epigenetic Processes in Abiotic and Biotic Stress Response in Wheat and Barley. Int. J. Mol. Sci. 2020, 21, 1480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Guo, X.; Han, F. Asymmetric Epigenetic Modification and Elimination of rDNA Sequences by Polyploidization in Wheat. Plant Cell 2014, 26, 4311–4327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. S Alotaibi, S.; El-Shehawi, A.M.; M Elseehy, M. Heat shock proteins expression is regulated by promoter cpg methylation/demethylation under heat stress in wheat varieties. Pak. J. Biol. Sci. 2020, 23, 1310–1320. [Google Scholar] [CrossRef]
  115. Khan, A.R.; Enjalbert, J.; Marsollier, A.-C.; Rousselet, A.; Goldringer, I.; Vitte, C. Vernalization treatment induces site-specific DNA hypermethylation at the VERNALIZATION-A1 (VRN-A1) locus in hexaploid winter wheat. BMC Plant Biol. 2013, 13, 209. [Google Scholar] [CrossRef] [Green Version]
  116. Fei, Y.; Xue, Y.; Du, P.; Yang, S.; Deng, X. Expression analysis and promoter methylation under osmotic and salinity stress of TaGAPC1 in wheat (Triticum aestivum L). Protoplasma 2017, 254, 987–996. [Google Scholar] [CrossRef]
  117. Duan, H.; Li, J.; Zhu, Y.; Jia, W.; Wang, H.; Jiang, L.; Zhou, Y. Responsive changes of DNA methylation in wheat (Triticum aestivum) under water deficit. Sci. Rep. 2020, 13, 7938. [Google Scholar] [CrossRef]
  118. Ma, H.; Lin, J.; Mei, F.; Mao, H.; Li, Q.Q. Differential alternative polyadenylation of homoeologous genes of allohexaploid wheat ABD subgenomes during drought stress response. Plant J. 2023, 114, 499–518. [Google Scholar] [CrossRef] [PubMed]
  119. Wu, X.; Gong, F.; Cao, D.; Hu, X.; Wang, W. Advances in crop proteomics: PTMs of proteins under abiotic stress. Proteomics 2016, 16, 847–865. [Google Scholar] [CrossRef] [PubMed]
  120. Wang, M.; Yuan, J.; Qin, L.; Shi, W.; Xia, G.; Liu, S.; Meng, W. Ta CYP 81D5 one member in a wheat cytochrome P450 gene cluster, confers salinity tolerance via reactive oxygen species scavenging. Plant Biotechnol. J. 2020, 18, 791–804. [Google Scholar] [CrossRef] [Green Version]
  121. Foyer, C.H.; Rasool, B.; Davey, J.W.; Hancock, R.D. Cross-tolerance to biotic and abiotic stresses in plants: A focus on resistance to aphid infestation. J. Exp. Bot. 2016, 67, 2025–2037. [Google Scholar] [CrossRef]
  122. Nemchenko, A.; Kunze, S.; Feussner, I.; Kolomiets, M. Duplicate maize 13-lipoxygenase genes are differentially regulated by circadian rhythm, cold stress, wounding, pathogen infection, and hormonal treatments. J. Exp. Bot. 2006, 57, 3767–3779. [Google Scholar] [CrossRef] [PubMed]
  123. Ye, L.; Wang, B.; Zhang, W.; Shan, H.; Kong, H. Gains and Losses of Cis-regulatory elements led to divergence of the Arabidopsis APETALA1 and CAULIFLOWER duplicate genes in the time, space, and level of expression and regulation of one paralog by the other. Plant Physiol. 2016, 171, 1055–1069. [Google Scholar] [PubMed] [Green Version]
  124. Hu, C.; Lin, S.-Y.; Chi, W.-T.; Charng, Y.-Y. Recent Gene Duplication and Subfunctionalization Produced a Mitochondrial GrpE, the Nucleotide Exchange Factor of the Hsp70 Complex, Specialized in Thermotolerance to Chronic Heat Stress in Arabidopsis. Plant Physiol. 2012, 158, 747–758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Arsovski, A.A.; Pradinuk, J.; Guo, X.Q.; Wang, S.; Adams, K.L. Evolution of cis-regulatory elements and regulatory networks in duplicated genes of Arabidopsis thaliana. Plant Physiol. 2015, 169, 2982–2991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Oh, D.-H.; Hong, H.; Lee, S.Y.; Yun, D.-J.; Bohnert, H.J.; Dassanayake, M. Genome Structures and Transcriptomes Signify Niche Adaptation for the Multiple-Ion-Tolerant Extremophyte Schrenkiella parvula. Plant Physiol. 2014, 164, 2123–2138. [Google Scholar] [CrossRef] [Green Version]
  127. Wang, J.; Marowsky, N.C.; Fan, C. Divergence of Gene Body DNA Methylation and Evolution of Plant Duplicate Genes. PLoS ONE 2014, 9, e110357. [Google Scholar] [CrossRef] [Green Version]
  128. Schilling, S.; Kennedy, A.; Pan, S.R.; Jermiin, L.S.; Melzer, R. Genome-wide analysis of MIKC-type MADS-box genes in wheat: Pervasive duplications may have facilitated adaptation to different environmental conditions. Cold Spring Harb. Lab. 2019, 255, 511–529. [Google Scholar]
  129. Liu, H.; Yang, W.; Zhao, X.; Kang, G.; Na Li, N.; Xu, H. Genome-wide analysis and functional characterization of CHYR gene family associated with abiotic stress tolerance in bread wheat (Triticum aestivum L.). BMC Plant Biol. 2022, 22, 204. [Google Scholar] [CrossRef]
  130. Liang, Y.; Xia, J.; Jiang, Y.; Bao, Y.; Chen, H.; Wang, D.; Zhang, D.; Yu, J.; Cang, J. Genome-Wide Identification and Analysis of bZIP Gene Family and Resistance of TaABI5 (TabZIP96) under Freezing Stress in Wheat (Triticum aestivum). Int. J. Mol. Sci. 2022, 23, 2351. [Google Scholar] [CrossRef]
  131. Li, H.; Guan, H.; Zhuo, Q.; Wang, Z.; Li, S.; Si, J.; Zhang, B.; Feng, B.; Kong, L.-A.; Wang, F.; et al. Genome-wide characterization of the abscisic acid-, stress- and ripening-induced (ASR) gene family in wheat (Triticum aestivum L.). Biol. Res. 2020, 24, 23. [Google Scholar] [CrossRef]
  132. Magar, M.; Liu, H.; Yan, G. Genome-wide analysis of AP2/ERF superfamily genes in contrasting wheat genotypes reveals heat stress-related candidate genes. Front. Plant Sci. 2022, 13, 853086. [Google Scholar] [CrossRef]
  133. Zhang, J.; Liao, J.; Ling, Q.; Xi, Y.; Qian, Y. Genome-wide identification and expression profiling analysis of maize AP2/ERF superfamily genes reveal essential roles in abiotic stress tolerance. BMC Genom. 2022, 23, 125. [Google Scholar] [CrossRef] [PubMed]
  134. Sharma, B.; Saxena, H.; Negi, H. Genome-wide analysis of HECT E3 ubiquitin ligase gene family in Solanum lycopersicum. Sci. Rep. 2021, 11, 15891. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, Z.; Zhang, R.; Cheng, Y.; Lei, P.; Song, W.; Zheng, W.; Nie, X. Genome-Wide Identification, Evolution, and Expression Analysis of LBD Transcription Factor Family in Bread Wheat (Triticum aestivum L.). Front. Plant Sci. 2021, 12, 721253. [Google Scholar] [CrossRef] [PubMed]
  136. Zou, Y.; Huang, W.; Gu, Z.; Gu, X. Predominant Gain of Promoter TATA Box after Gene Duplication Associated with Stress Responses. Mol. Biol. Evol. 2011, 28, 2893–2904. [Google Scholar] [CrossRef] [Green Version]
  137. Zou, C.; Lehti-Shiu, M.; Michael, T.; Shin-Han, S.; Copenhaver, G.P. Evolution of stress-regulated gene expression in duplicate genes of Arabidopsis thaliana. PLoS Genet. 2009, 5, e1000581. [Google Scholar] [CrossRef] [Green Version]
  138. Li, A.-L.; Zhu, Y.-F.; Tan, X.-M.; Wang, X.; Wei, B.; Guo, H.-Z.; Zhang, Z.-L.; Chen, X.-B.; Zhao, G.-Y.; Kong, X.-Y.; et al. Evolutionary and functional study of the CDPK gene family in wheat (Triticum aestivum L.). Plant Mol. Biol. 2008, 66, 429–443. [Google Scholar] [CrossRef]
  139. Geng, S.; Zhao, Y.; Tang, L.; Zhang, R.; Sun, M.; Guo, H.; Kong, X.; Li, A.; Mao, L. Molecular evolution of two duplicated CDPK genes CPK7 and CPK12 in grass species: A case study in wheat (Triticum aestivum L.). Gene 2011, 475, 94–103. [Google Scholar] [CrossRef]
  140. Efeoglu, B.; Terzioglu, S. Varying patterns of protein synthesis in bread wheat during heat shock. Acta Biol. Hung. 2007, 58, 93–104. [Google Scholar] [CrossRef]
  141. Lu, Y.; Zhao, P.; Zhang, A.; Wang, J.; Ha, M. Genome-Wide Analysis of HSP70s in Hexaploid Wheat: Tandem Duplication, Heat Response, and Regulation. Cells 2022, 11, 818. [Google Scholar] [CrossRef]
  142. Erdayani, E.; Nagarajan, R.; Grant, N.P.; Gill, K.S. Genome-wide analysis of the HSP101/CLPB gene family for heat tolerance in hexaploid wheat. Sci. Rep. 2020, 10, 3948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. El Beji, I.H.; Mouzeyar, S.; Bouzidi, M.-F.; Roche, J. Expansion and Functional Diversification of SKP1-Like Genes in Wheat (Triticum aestivum L.). Int. J. Mol. Sci. 2019, 20, 3295. [Google Scholar] [CrossRef] [Green Version]
  144. Chen, J.; Wei, B.; Li, G.; Fan, R.; Zhong, Y.; Wang, X.; Zhang, X. TraeALDH7B1-5A, encoding aldehyde dehydrogenase 7 in wheat, confers improved drought tolerance in Arabidopsis. Planta 2015, 242, 137–151. [Google Scholar] [CrossRef]
  145. Ma, J.; Yuan, M.; Sun, B.; Zhang, D.; Zhang, J.; Li, C.; Shao, Y.; Liu, W.; Jiang, L. Evolutionary Divergence and Biased Expression of NAC Transcription Factors in Hexaploid Bread Wheat (Triticum aestivum L.). Plants 2021, 10, 382. [Google Scholar] [CrossRef]
  146. Jiang, W.; Pan, R.; Buitrago, S.; Wu, C.; Abou-Elwafa, S.F.; Xu, Y.; Zhang, W. Conservation and divergence of the TaSOS1 gene family in salt stress response in wheat (Triticum aestivum L.). Physiol. Mol. Biol. Plants 2021, 27, 1245–1260. [Google Scholar] [CrossRef]
  147. Zheng, M.; Li, J.; Zeng, C.; Liu, X.; Chu, W.; Lin, J.; Wang, F.; Wang, W.; Guo, W.; Xin, M.; et al. Subgenome-biased expression and functional diversification of a Na+/H+ antiporter homoeologs in salt tolerance of polyploid wheat. Front. Plant Sci. 2022, 13, 1072009. [Google Scholar] [CrossRef] [PubMed]
  148. Xu, G.; Guo, C.; Shan, H.; Kong, H. Divergence of duplicate genes in exon-intron structure. Proc. Natl. Acad. Sci. USA 2012, 109, 1187–1192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Ali, Z.; Park, H.C.; Ali, A.; Oh, D.H.; Aman, R.; Kropornicka, A.; Hong, H.; Choi, W.; Chung, W.S.; Kim, W.Y. TsHKT1;2, a HKT1 homolog from the extremophile Arabidopsis relative Thellungiella salsuginea, shows K specificity in the presence of NaCl. Plant Physiol. 2012, 158, 1463–1474. [Google Scholar] [CrossRef] [Green Version]
  150. Zhang, H.; Mao, X.; Zhang, J.; Chang, X.; Jing, R. Single-nucleotide polymorphisms and association analysis of drought-resistance gene TaSnRK2.8 in common wheat. Plant Physiol. Biochem. 2013, 70, 174–181. [Google Scholar] [CrossRef] [Green Version]
  151. Tian, G.; Wang, S.; Wu, J.; Wang, Y.; Wang, X.; Liu, S.; Han, D.; Xia, G.; Wang, M. Allelic variation of TaWD40-4B.1 contributes to drought tolerance by modulating catalase activity in wheat. Nat. Commun. 2023, 14, 1200. [Google Scholar] [CrossRef]
  152. Mondini, L.; Nachit, M.M.; Pagnotta, M.A. Allelic variants in durum wheat (Triticum turgidum L. var. durum) DREB genes conferring tolerance to abiotic stresses. Mol. Genet. Genom. 2015, 290, 531–544. [Google Scholar] [CrossRef] [PubMed]
  153. Bologna, N.; Voinnet, O. The diversity, biogenesis, and activities of endogenous silencing small RNAs in Arabidopsis. Annu. Rev. Plant Biol. 2014, 65, 473–503. [Google Scholar] [CrossRef] [PubMed]
  154. Wang, S.; Adams, K.L. Duplicate Gene Divergence by Changes in MicroRNA Binding Sites in Arabidopsis and Brassica. Genome Biol. Evol. 2015, 7, 646–655. [Google Scholar] [CrossRef] [Green Version]
  155. Paterson, A.H.; Bowers, J.E.; Bruggmann, R.; Dubchak, I.; Grimwood, J.; Gundlach, H.; Haberer, G.; Hellsten, U.; Mitros, T.; Poliakov, A.; et al. The Sorghum bicolor genome and the diversification of grasses. Nature 2009, 457, 551–556. [Google Scholar] [CrossRef] [Green Version]
  156. Cheng, X.; He, Q.; Tang, S.; Wang, H.; Zhang, X.; Lv, M.; Liu, H.; Gao, Q.; Zhou, Y.; Wang, Q.; et al. The miR172/IDS1 signaling module confers salt tolerance through maintaining ROS homeostasis in cereal crops. New Phytol. 2021, 230, 1017–1033. [Google Scholar] [CrossRef]
  157. Chapman, B.A.; Bowers, J.E.; Feltus, F.A.; Paterson, A.H. Buffering of crucial functions by paleologous duplicated genes may contribute cyclicality to angiosperm genome duplication. Proc. Natl. Acad. Sci. USA 2006, 103, 2730–2735. [Google Scholar] [CrossRef] [Green Version]
  158. Akhunova, A.R.; Matniyazov, R.T.; Liang, H.; Akhunov, E.D. Homoeolog-specific transcriptional bias in allopolyploid wheat. BMC Genom. 2010, 11, 505–516. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Kan, J.; Cai, Y.; Cheng, C.; Jiang, C.; Jin, Y.; Yang, P. Simultaneous editing of host factor gene TaPDIL5-1 homoeoalleles confers wheat yellow mosaic virus resistance in hexaploid wheat. New Phytol. 2022, 234, 340–344. [Google Scholar] [CrossRef]
  160. Samtani, H.; Sharma, A.; Khurana, P. Overexpression of HVA1 Enhances Drought and Heat Stress Tolerance in Triticum aestivum Doubled Haploid Plants. Cells 2022, 11, 912. [Google Scholar] [CrossRef]
  161. Karami, M.; Fatahi, N.; Lohrasebi, T.; Razavi, K. RAV transcription factor regulatory function in response to salt stress in two Iranian wheat landraces. J. Plant Res. 2022, 135, 121–136. [Google Scholar] [CrossRef]
  162. Badawi, M.; Reddy, Y.V.; Agharbaoui, Z.; Tominaga, Y.; Danyluk, J.; Sarhan, F.; Houde, M. Structure and Functional Analysis of Wheat ICE (Inducer of CBF Expression) Genes. Plant Cell Physiol. 2008, 49, 1237–1249. [Google Scholar] [CrossRef] [Green Version]
  163. Peng, L.N.; Xu, Y.Q.; Wang, X.; Feng, X.; Zhao, Q.Q.; Feng, S.S.; Zhao, Z.Y.; Hu, B.Z.; Li, F.L.; Xu, Y.; et al. Overexpression of paralogues of the wheat expansin gene TaEXPA8 improves low-temperature tolerance in Arabidopsis. Plant Biol. 2019, 21, 1119–1131. [Google Scholar] [CrossRef]
  164. Zheng, M.; Lin, J.; Liu, X.; Chu, W.; Li, J.; Gao, Y.; An, K.; Song, W.; Xin, M.; Yao, Y.; et al. Histone acetyltransferase TaHAG1 acts as a crucial regulator to strengthen salt tolerance of hexaploid wheat. Plant Physiol. 2021, 186, 1951–1969. [Google Scholar] [CrossRef] [PubMed]
  165. Lin, J.; Song, N.; Liu, D.; Liu, X.; Chu, W.; Li, J.; Chang, S.; Liu, Z.; Chen, Y.; Yang, Q.; et al. Histone acetyltransferase TaHAG1 interacts with TaNACL to promote heat stress tolerance in wheat. Plant Biotechnol. J. 2022, 20, 1645–1647. [Google Scholar] [CrossRef] [PubMed]
  166. Chandra, A.K.; Jha, S.K.; Agarwal, P.; Mallick, N.; Niranjana, M. Vinod Leaf rolling in bread wheat (Triticum aestivum L.) is controlled by the upregulation of a pair of closely linked/duplicate zinc finger homeodomain class transcription factors during moisture stress conditions. Front. Plant Sci. 2022, 13, 1038881. [Google Scholar] [CrossRef]
  167. Soltész, A.; Smedley, M.; Vashegyi, I.; Galiba, G.; Harwood, W.; Vágújfalvi, A. Transgenic barley lines prove the involvement of TaCBF14 and TaCBF15 in the cold acclimation process and in frost tolerance. J. Exp. Bot. 2013, 64, 1849–1862. [Google Scholar] [CrossRef] [Green Version]
  168. Ma, H.-Z.; Liu, G.-Q.; Li, C.-W.; Kang, G.-Z.; Guo, T.-C. Identification of the TaBTF3 gene in wheat (Triticum aestivum L.) and the effect of its silencing on wheat chloroplast, mitochondria and mesophyll cell development. Biochem. Biophys. Res. Commun. 2012, 426, 608–614. [Google Scholar] [CrossRef]
  169. Li, L.; Peng, Z.; Mao, X.; Wang, J.; Li, C.; Chang, X.; Jing, R. Genetic insights into natural variation underlying salt tolerance in wheat. J. Exp. Bot. 2020, 72, 1135–1150. [Google Scholar] [CrossRef]
  170. Wang, D.; Cao, Z.; Wang, W.; Zhu, W.; Hao, X.; Fang, Z.; Liu, S.; Wang, X.; Zhao, C.; Tang, Y. Genome-Wide Characterization of OFP Family Genes in Wheat (Triticum aestivum L.) Reveals That TaOPF29a-A Promotes Drought Tolerance. BioMed Res. Int. 2020, 2020, 9708324. [Google Scholar] [CrossRef]
  171. Saidi, M.N.; Mergby, D.; Souibgui, A.; Yacoubi, I. Overexpression of durum wheat NAC transcription factor TtNTL3A promotes early flowering and increases multiple stress tolerance in transgenic Arabidopsis. Plant Physiol. Biochem. 2022, 192, 1–9. [Google Scholar] [CrossRef] [PubMed]
  172. Borjigin, C.; Schilling, R.K.; Bose, J.; Hrmova, M.; Qiu, J.; Wege, S.; Situmorang, A.; Byrt, C.; Brien, C.; Berger, B.; et al. A single nucleotide substitution in TaHKT1;5-D controls shoot Na+ accumulation in bread wheat. Plant Cell Environ. 2020, 43, 2158–2171. [Google Scholar] [CrossRef] [PubMed]
  173. Byrt, C.S.; Xu, B.; Krishnan, M.; Lightfoot, D.J.; Athman, A.; Jacobs, A.K.; Watson-Haigh, N.S.; Plett, D.; Munns, R.; Tester, M.; et al. The Na(+) transporter, TaHKT1;5-D, limits shoot Na(+) accumulation in bread wheat. Plant J. 2014, 80, 516–526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Genc, Y.; Oldach, K.; Verbyla, A.P.; Lott, G.; Hassan, M.; Tester, M.; Wallwork, H.; McDonald, G.K. Sodium exclusion QTL associated with improved seedling growth in bread wheat under salinity stress. Theor. Appl. Genet. 2010, 121, 877–894. [Google Scholar] [CrossRef]
  175. Genc, Y.; Taylor, J.; Rongala, J.; Oldach, K. A Major Locus for Chloride Accumulation on Chromosome 5A in Bread Wheat. PLoS ONE 2014, 9, e98845. [Google Scholar] [CrossRef]
  176. Ben Amar, S.; Brini, F.; Sentenac, H.; Masmoudi, K.; Véry, A.-A. Functional characterization in Xenopus oocytes of Na+ transport systems from durum wheat reveals diversity among two HKT1;4 transporters. J. Exp. Bot. 2013, 65, 213–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Pang, F.; Niu, J.; Solanki, M.K.; Nosheen, S.; Liu, Z.; Wang, Z. PHD-finger family genes in wheat (Triticum aestivum L.): Evolutionary conservatism, functional diversification, and active expression in abiotic stress. Front. Plant Sci. 2022, 13, 1016831. [Google Scholar] [CrossRef]
  178. Li, S.; Yu, S.; Zhang, Y.; Zhu, D.; Li, F.; Chen, B.; Mei, F.; Du, L.; Ding, L.; Chen, L.; et al. Genome-wide association study revealed TaHXK3-2A as a candidate gene controlling stomatal index in wheat seedlings. Plant Cell Environ. 2022, 45, 2306–2323. [Google Scholar] [CrossRef] [PubMed]
  179. Liu, X.; Fu, L.; Qin, P.; Sun, Y.; Liu, J.; Wang, X. Overexpression of the wheat trehalose 6-phosphate synthase 11 gene enhances cold tolerance in Arabidopsis thaliana. Gene 2019, 710, 210–217. [Google Scholar] [CrossRef]
  180. Tian, Y.; Peng, K.; Bao, Y.; Zhang, D.; Meng, J.; Wang, D.; Wang, X.; Cang, J. Glucose-6-phosphate dehydrogenase and 6-phosphogluconate dehydrogenase genes of winter wheat enhance the cold tolerance of transgenic Arabidopsis. Plant Physiol. Biochem. 2021, 161, 86–97. [Google Scholar] [CrossRef]
  181. Lu, Y.; Zhao, P.; Zhang, A.; Ma, L.; Xu, S.; Wang, X. Alternative Splicing Diversified the Heat Response and Evolutionary Strategy of Conserved Heat Shock Protein 90s in Hexaploid Wheat (Triticum aestivum L.). Front. Genet. 2020, 11, 577897. [Google Scholar] [CrossRef]
  182. Zang, X.; Geng, X.; Wang, F.; Liu, Z.; Zhang, L.; Zhao, Y.; Tian, X.; Ni, Z.; Yao, Y.; Xin, M.; et al. Overexpression of wheat ferritin gene TaFER-5B enhances tolerance to heat stress and other abiotic stresses associated with the ROS scavenging. BMC Plant Biol. 2017, 17, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Ru, J.N.; Hou, Z.H.; Zheng, L.; Zhao, Q.; Wang, F.Z.; Chen, J.; Zhou, Y.B.; Chen, M.; Ma, Y.Z.; Xi, Y.J.; et al. Genome-wide analysis of DEAD-box RNA helicase family in wheat (Triticum aestivum) and functional identification of TaDEAD-box57 in abiotic stress responses. Front. Plant Sci. 2021, 12, 797276. [Google Scholar] [CrossRef] [PubMed]
  184. He, J.; Li, C.; Hu, N.; Zhu, Y.; He, Z.; Sun, Y.; Wang, Z.; Wang, Y. ECERIFERUM1-6A is required for the synthesis of cuticular wax alkanes and promotes drought tolerance in wheat. Plant Physiol. 2022, 190, 1640–1657. [Google Scholar] [CrossRef] [PubMed]
  185. Li, T.; Sun, Y.; Liu, T.; Wu, H.; An, P.; Shui, Z.; Wang, J.; Zhu, Y.; Li, C.; Wang, Y.; et al. TaCER1-1A is involved in cuticular wax alkane biosynthesis in hexaploid wheat and responds to plant abiotic stresses. Plant Cell Environ. 2019, 42, 3077–3091. [Google Scholar] [CrossRef] [PubMed]
  186. Mao, H.; Jian, C.; Cheng, X.; Chen, B.; Mei, F.; Li, F.; Zhang, Y.; Li, S.; Du, L.; Li, T.; et al. The wheat ABA receptor gene TaPYL1-1B contributes to drought tolerance and grain yield by increasing water-use efficiency. Plant Biotechnol. J. 2021, 20, 846–861. [Google Scholar] [CrossRef] [PubMed]
  187. Iehisa, J.C.; Matsuura, T.; Mori, I.C.; Yokota, H.; Kobayashi, F.; Takumi, S. Identification of quantitative trait loci for abscisic acid responsiveness in the D-genome of hexaploid wheat. J. Plant Physiol. 2014, 171, 830–841. [Google Scholar] [CrossRef] [PubMed]
  188. Xu, X.; Zhang, L.; Zhao, W.; Fu, L.; Han, Y.; Wang, K.; Yan, L.; Li, Y.; Zhang, X.-H.; Min, D.-H. Genome-wide analysis of the serine carboxypeptidase-like protein family in Triticum aestivum reveals TaSCPL184-6D is involved in abiotic stress response. BMC Genom. 2021, 22, 350. [Google Scholar] [CrossRef]
  189. Gahlaut, V.; Samtani, H.; Khurana, P. Genome-wide identification and expression profiling of cytosine-5 DNA methyltransferases during drought and heat stress in wheat (Triticum aestivum). Genomics 2020, 112, 4796–4807. [Google Scholar] [CrossRef]
  190. Wang, X.; Pan, C.; Long, J.; Bai, S.; Yao, M.; Chen, J.; Sun, G.; Fan, Y.; Wang, Z.; Liu, F.; et al. Genome-wide identification of the jumonji C domain- containing histone demethylase gene family in wheat and their expression analysis under drought stress. Front. Plant Sci. 2022, 13, 987257. [Google Scholar] [CrossRef]
  191. Park, Y.C.; Jang, C.S. Molecular dissection of two homoeologous wheat genes encoding RING H2-type E3 ligases: TaSIRFP-3A and TaSIRFP-3B. Planta 2020, 252, 26. [Google Scholar] [CrossRef]
  192. Kim, J.H.; Kim, M.S.; Kim, D.Y.; Amoah, J.N.; Seo, Y.W. Molecular Characterization of U-box E3 Ubiquitin Ligases (TaPUB2 and TaPUB3) Involved in the Positive Regulation of Drought Stress Response in Arabidopsis. Int. J. Mol. Sci. 2021, 22, 13658. [Google Scholar] [CrossRef] [PubMed]
  193. Smirnova, O.G.; Pshenichnikova, T.A. The relationship between the genetic status of the Vrn-1 locus and the size of the root system in bread wheat (Triticum aestivum L.). Vavilov J. Genet. Breed. 2022, 25, 805–811. [Google Scholar] [CrossRef] [PubMed]
  194. Zheng, B.; Ben Biddulph, B.; Li, D.; Kuchel, H.; Chapman, S. Quantification of the effects of VRN1 and Ppd-D1 to predict spring wheat (Triticum aestivum) heading time across diverse environments. J. Exp. Bot. 2013, 64, 3747–3761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Lu, P.-P.; Zheng, W.-J.; Wang, C.-T.; Shi, W.-Y.; Fu, J.-D.; Chen, M.; Chen, J.; Zhou, Y.-B.; Xi, Y.-J.; Xu, Z.-S. Wheat Bax Inhibitor-1 interacts with TaFKBP62 and mediates response to heat stress. BMC Plant Biol. 2018, 18, 259. [Google Scholar] [CrossRef] [Green Version]
  196. Huang, S.; Ding, J.; Deng, D.; Tang, W.; Sun, H.; Liu, D.; Zhang, L.; Niu, X.; Zhang, X.; Meng, M.; et al. Draft genome of the kiwifruit Actinidia chinensis. Nat. Commun. 2013, 4, 2640. [Google Scholar] [CrossRef] [Green Version]
  197. Ramsey, J.; Ramsey, T.S. Ecological studies of polyploidy in the 100 years following its discovery. Philos. Trans. R. Soc. B Biol. Sci. 2014, 369, 20130352. [Google Scholar] [CrossRef]
  198. Lynch, M.; Conery, J.S. The evolutionary fate and consequences of duplicate genes. Science 2000, 290, 1151–1155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Qiao, X.; Li, Q.; Yin, H.; Qi, K.; Li, L.; Wang, R.; Zhang, S.; Paterson, A.H. Gene duplication and evolution in recurring polyploidization–diploidization cycles in plants. Genome Biol. 2019, 20, 1–23. [Google Scholar] [CrossRef] [Green Version]
  200. Song, K.; Lu, P.; Tang, K.; Osborn, T.C. Rapid genome change in synthetic polyploids of Brassica and its implications for polyploid evolution. Proc. Natl. Acad. Sci. USA 1995, 92, 7719–7723. [Google Scholar] [CrossRef] [Green Version]
  201. Marcussen, T.; Sandve, S.R.; Heier, L.; Spannagl, M.; Pfeifer, M.; Jakobsen, K.S.; Wulff, B.B.H.; Steuernagel, B.; Mayer, K.F.X.; Olsen, O.-A. Ancient hybridizations among the ancestral genomes of bread wheat. Science 2014, 345, 1250092. [Google Scholar] [CrossRef]
  202. Hegarty, M.J.; Hiscock, S.J. Genomic Clues to the Evolutionary Success of Polyploid Plants. Curr. Biol. 2008, 18, R435–R444. [Google Scholar] [CrossRef] [Green Version]
  203. Salman-Minkov, A.; Sabath, N.; Mayrose, I. Whole-genome duplication as a key factor in crop domestication. Nat. Plants 2016, 2, 16115. [Google Scholar] [CrossRef]
  204. Zhao, G.; Zou, C.; Li, K.; Wang, K.; Li, T.; Gao, L.; Zhang, X.; Wang, H.; Yang, Z.; Liu, X.; et al. The Aegilops tauschii genome reveals multiple impacts of transposons. Nat. Plants 2017, 3, 946–955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Freeling, M. Bias in Plant Gene Content Following Different Sorts of Duplication: Tandem, Whole-Genome, Segmental, or by Transposition. Annu. Rev. Plant Biol. 2009, 60, 433–453. [Google Scholar] [CrossRef] [PubMed]
  206. Wang, X.; Yan, X.; Hu, Y.; Qin, L.; Wang, D.; Jia, J.; Jiao, Y. A recent burst of gene duplications in Triticeae. Plant Commun. 2022, 3, 100268. [Google Scholar] [CrossRef] [PubMed]
  207. Yi-Tzu, K.; Chao, Y.T.; Chen, W.C.; Ming-Che, S.; Chang, S.B. Segmental and tandem chromosome duplications led to divergent evolution of the chalcone synthase gene family in Phalaenopsis orchids. Ann. Bot. 2018, 1, 69–77. [Google Scholar]
  208. Vicient, C.M.; Casacuberta, J.M. Impact of transposable elements on polyploid plant genomes. Ann. Bot. 2017, 120, 195–207. [Google Scholar] [CrossRef] [Green Version]
  209. Kim, S.; Park, J.; Yeom, S.-I.; Kim, Y.-M.; Seo, E.; Kim, K.-T.; Kim, M.-S.; Lee, J.M.; Cheong, K.; Shin, H.-S.; et al. New reference genome sequences of hot pepper reveal the massive evolution of plant disease-resistance genes by retroduplication. Genome Biol. 2017, 18, 210. [Google Scholar] [CrossRef] [Green Version]
  210. Francia, E.; Barabaschi, D.; Tondelli, A.; Laidò, G.; Rizza, F.; Stanca, A.M.; Busconi, M.; Fogher, C.; Stockinger, E.J.; Pecchioni, N. Fine mapping of a HvCBF gene cluster at the frost resistance locus Fr-H2 in barley. Theor. Appl. Genet. 2007, 115, 1083–1091. [Google Scholar] [CrossRef]
  211. Båga, M.; Chodaparambil, S.V.; Limin, A.E.; Pecar, M.; Fowler, D.B.; Chibbar, R.N. Identification of quantitative trait loci and associated candidate genes for low-temperature tolerance in cold-hardy winter wheat. Funct. Integr. Genom. 2007, 7, 53–68. [Google Scholar] [CrossRef]
  212. Knox, A.K.; Dhillon, T.; Cheng, H.; Tondelli, A.; Pecchioni, N.; Stockinger, E.J. CBF gene copy number variation at Frost Resistance-2 is associated with levels of freezing tolerance in temperate-climate cereals. Theor. Appl. Genet. 2010, 121, 21–35. [Google Scholar] [CrossRef] [PubMed]
  213. Zhu, J.; Pearce, S.; Burke, A.; See, D.R.; Skinner, D.Z.; Dubcovsky, J.; Garland-Campbell, K. Copy number and haplotype variation at the VRN-A1 and central FR-A2 loci are associated with frost tolerance in hexaploid wheat. Theor. Appl. Genet. 2014, 127, 1183–1197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Miller, A.K.; Galiba, G.; Dubcovsky, J. A cluster of 11 CBF transcription factors is located at the frost tolerance locus Fr-A m 2 in Triticum monococcum. Mol. Genet. Genom. 2006, 275, 193–203. [Google Scholar] [CrossRef] [PubMed]
  215. Zhang, X.; Fowler, S.; Cheng, H.; Lou YRhee, S.; Stockinger, E.; Thomashow, M. Freezing-sensitive tomato has a functional CBF cold response pathway, but a CBF regulon that differs from that of freezing-tolerant Arabidopsis. Plant J. Cell Mol. Biol. 2004, 39, 905–919. [Google Scholar] [CrossRef]
  216. Knox, A.K.; Li, C.; Vágújfalvi, A.; Galiba, G.; Stockinger, E.J.; Dubcovsky, J. Identification of candidate CBF genes for the frost tolerance locus Fr-A m 2 in Triticum monococcum. Plant Mol. Biol. 2008, 67, 257–270. [Google Scholar] [CrossRef]
  217. Dixon, L.; Karsai IKiss, T.; Adamski, N.; Liu, Z.; Ding, Y.; Allard, V.; Boden, S.; Griffiths, S. VERNALIZATION1 controls developmental responses of winter wheat under high ambient temperatures. Development 2019, 146, dev172684. [Google Scholar] [CrossRef] [Green Version]
  218. Freeling, M.; Lyons, E.; Pedersen, B.; Alam, M.; Ming, R.; Lisch, D. Many or most genes in Arabidopsis transposed after the origin of the order Brassicales. Genome Res. 2008, 18, 1924–1937. [Google Scholar] [CrossRef] [Green Version]
  219. Morrell, P.L.; Buckler, E.S.; Ross-Ibarra, J. Crop genomics: Advances and applications. Nat. Rev. Genet. 2011, 13, 85–96. [Google Scholar] [CrossRef]
  220. Golovnina, K.; Kondratenko, E.; Blinov, A.; Goncharov, N. Molecular characterization of vernalization loci VRN1 in wild and cultivated wheats. BMC Plant Biol. 2010, 10, 168. [Google Scholar] [CrossRef] [Green Version]
  221. Yu, H.; Lin, T.; Meng, X.; Du, H.; Zhang, J.; Liu, G.; Chen, M.; Jing, Y.; Kou, L.; Li, X.; et al. A route to de novo domestication of wild allotetraploid rice. Cell 2021, 184, 1156–1170.e14. [Google Scholar] [CrossRef]
  222. Renny-Byfield, S.; Wendel, J.F. Doubling down on genomes: Polyploidy and crop plants. Am. J. Bot. 2014, 101, 1711–1725. [Google Scholar] [CrossRef] [Green Version]
  223. Sayed, H.I. Diversity of salt tolerance in a germplasm collection of wheat (Triticum spp.). Theor. Appl. Genet. 1985, 69, 651–657. [Google Scholar] [CrossRef]
  224. Budak, H.; Kantar, M.; Kurtoglu, K.Y. Drought Tolerance in Modern and Wild Wheat. Sci. World J. 2013, 2013, 548246. [Google Scholar] [CrossRef] [Green Version]
  225. Wang, Y.; Cheng, X.; Shan, Q.; Zhang, Y.; Liu, J.; Gao, C.; Qiu, J.-L. Simultaneous editing of three homoeoalleles in hexaploid bread wheat confers heritable resistance to powdery mildew. Nat. Biotechnol. 2014, 32, 947–951. [Google Scholar] [CrossRef]
  226. Shan, Q.; Wang, Y.; Li, J.; Zhang, Y.; Chen, K.; Liang, Z.; Zhang, K.; Liu, J.; Xi, J.J.; Qiu, J.-L.; et al. Targeted genome modification of crop plants using a CRISPR-Cas system. Nat. Biotechnol. 2013, 31, 686–688. [Google Scholar] [CrossRef]
Figure 1. Mechanisms of gene duplication. (A) Whole-genome duplication (WGD) or polyploidy to produce duplicate genes. (B) Tandem duplication: duplication of a gene via unequal crossing-over between similar alleles. A chromosomal region within 200-kb containing two or more genes is defined as a tandem duplication event. (C) Segmental duplication: multiple genes through polyploidy followed by chromosome rearrangements. (D) Transposon-mediated duplication. (E) Retroduplication. mRNA, which has been transcribed and cleaved, goes through a reverse transcription process to form cDNA, which is then randomly inserted into a chromosome to form a new duplication gene.
Figure 1. Mechanisms of gene duplication. (A) Whole-genome duplication (WGD) or polyploidy to produce duplicate genes. (B) Tandem duplication: duplication of a gene via unequal crossing-over between similar alleles. A chromosomal region within 200-kb containing two or more genes is defined as a tandem duplication event. (C) Segmental duplication: multiple genes through polyploidy followed by chromosome rearrangements. (D) Transposon-mediated duplication. (E) Retroduplication. mRNA, which has been transcribed and cleaved, goes through a reverse transcription process to form cDNA, which is then randomly inserted into a chromosome to form a new duplication gene.
Plants 12 02465 g001
Figure 2. Prospects of abiotic stress-related duplicates in wheat. Making full use of the rich germplasm resources of the Triticeae, identification of abiotic stress-resistant varieties of bread wheat and the mining of superior genes in ancestral wild-type and closely related species. Explore the replication, retention, and evolutionary mechanisms of repetitive genes in wheat, as well as the diversity and differentiation of the functions and mechanisms of action of these genes in the abiotic response process, and rationalize the use of repetitive genes to balance their effects in stress resistance and wheat yield to achieve the desired results of stress resistance and yield. Make full use of multi-omics analysis platforms such as bioinformatics analysis and phenome in pursuit of application of excellent genes from greenhouse to field production. This proposal provides a potential precision-breeding method for improved abiotic resistance of wheat.
Figure 2. Prospects of abiotic stress-related duplicates in wheat. Making full use of the rich germplasm resources of the Triticeae, identification of abiotic stress-resistant varieties of bread wheat and the mining of superior genes in ancestral wild-type and closely related species. Explore the replication, retention, and evolutionary mechanisms of repetitive genes in wheat, as well as the diversity and differentiation of the functions and mechanisms of action of these genes in the abiotic response process, and rationalize the use of repetitive genes to balance their effects in stress resistance and wheat yield to achieve the desired results of stress resistance and yield. Make full use of multi-omics analysis platforms such as bioinformatics analysis and phenome in pursuit of application of excellent genes from greenhouse to field production. This proposal provides a potential precision-breeding method for improved abiotic resistance of wheat.
Plants 12 02465 g002
Table 1. Recent examples of duplicate genes in wheat.
Table 1. Recent examples of duplicate genes in wheat.
Gene NameGene ProductEnvironmental ConditionReference
Transcription factorTaSNAC8-6ANAC transcription factorsdrought stress[71]
TaNAC071-ANAC transcription factorsdrought stress[72]
TaSNAC4-3DNAC transcription factorsdrought stress[73]
TaSNAC4-3ANAC transcription factorsdrought stress[74]
TaDTG6-BDehydration-responsive element-binding proteindrought stress[75]
TaHsfA6fHeat shock factorheat stress[76]
TaHsfA2e-5DHeat shock transcription factordrought and heat stress[77]
TaHsfA6b-4DHeat shock transcription factorheat stress[78]
TaHsfC2aHeat shock factorheat stress[79]
TaMYB344MYB transcription factorsdrought, heat, and salt stress[80]
TaMYB56-BMYB transcription factorsfreezing and salt stress[81]
TabZIP60Basic leucine zipper proteinsheat stress[82]
TabZIP14-BbZIP transcription factorssalt and freezing stress[83]
TaWRKY75-AWRKY domain proteinsalt stress[84]
TaWRKY1-2DWRKY transcription factorsdrought stress[85]
TaERF-6-3AAP2/ERF transcription factorsdrought and salt stress[86]
TaGRF6-AGeneral regulatory factorssalt stress[87]
TaDrAp1, TaDrAp2Down-regulator associated proteindrought stress[101]
TaFDL2-1AbZIP transcription factordrought stress[102]
TaMpc1-D4MYB transcription factorsdrought stress[103]
TaGT2L1Dtrihelix transcription factorsdrought stress[104]
TaNFYC-A7Recombinant Nuclear Transcription Factordrought stress[105]
TaAIDFαCRT/DRE-binding factorcold stress[107]
TaAREB3ABA-responsive element-binding proteinscold stress[108]
TaWD40-4B.1WD40 transcription factorsdrought stress[151]
TaRAV4 and TaRAV5RAV (related to ABI3/VP1) transcription factordrought stress[161]
TaZHD1 and TaZHD10Zinc finger homeodomain class transcription factorsdrought stress[166]
TaCBF14 and TaCBF15C-repeat/DREB binding factorscold stress[167]
TaBTF3Basic transcription factor 3cold stress[168]
TaRN2ASYMMETRIC LEAVES2 (AS2)/LATERAL ORGAN BOUNDARIES (LOB) domain transcription factorheat stress[169]
TaOPF29aOVATE family proteinsdrought stress[170]
TtNTL3ANAC transcription factorsdrought and salt stress[171]
Cytoprotective protein/enzymeTaCAT3Catalasecold stress[49,50]
TaSOD2Superoxide dismutasessalt stress[51,52,53]
TaAPX-RAscorbate peroxidasedrought and salt stress[54]
TaGPXGlutathione peroxidase genessalt stress[55,56]
TaPRX-2APeroxidase gene familysalt stress[57]
TaGR2-B1Glutathione reductasesalt stress[58]
TaADF16Actin depolymerizing factorcold stress[100]
TaHVA1Group 3 Late Embryogenesis Abundant proteindrought and heat stress[160]
TaEXPA8Expansin proteincold stress[163]
TransportersTaOSCAsHyperosmolality-gated calcium-permeable channelsdrought, salt, heat stress[62,63]
TaBORsBOR transporter familydrought, salt, heat stress[64]
TaMSLMechanosensitive channel of small conductance-likedrought, salt, heat stress[66]
TaCaCACa2+/cation antiportersdrought, salt, heat stress[67]
TaNHX4-B.1 and TaNHX4-B.4Cation proton antiporterdrought, salt, heat stress[68]
TaACAs and TaECAsP-type II Ca2+ATPasesdrought, salt, heat stress[69]
TaSOS1Na+/H+ antiportersalt stress[146,147]
TaHKT1;5-D, TmHKT1;5- ANa+ transportersalt stress[172,173]
HKT1;4Na⁺ transportersalt stress[174]
TaCLC; TaCCCChloride channel; cation chloride co-transportersalt stress[175]
TdHKT1;4Na⁺ transportersalt stress[176]
Homeobox genesTaHD-ZipHD-Zip gene familysalt and drought stress[88]
TaKNOX11-ATALE superfamily proteindrought, salt stress[89,90]
TaZF-HDZinc Finger-Homeodomain Transcriptional Factorsdrought, salt, and cold stress[91]
TaWUS and TaWOX14WUSCHEL-Related Homeoboxdrought, salt, heat stress[92,93]
TaPHDPlant homeodomain (PHD) transcription factorscold, drought, and heat stress[177]
Metabolism-related enzymeTaLTPIb.1, TaLTPIb.5, and TaLTPIdNon-specific lipid transfer proteinscold stress[97]
TaCYP81D5Cytochrome P450 proteinsalt stress[120]
TaHSP70sHeat shock proteinheat stress[141]
TraeALDH7B1-5AAldehyde dehydrogenasedrought stress[144]
TaHXK3-2AHexokinasedrought stress[178]
TaTPS11Trehalose 6-phosphate synthasecold stress[179]
TaG6PDHGlucose-6-phosphate dehydrogenasecold stress[180]
TaHSP90sHeat shock proteinheat stress[181]
TaFER-5BFerritinheat stress[182]
TaDEAD-box57-3BDEAD-box RNA Helicasedrought and salt stress[183]
TaCER1-6A,TaCER1-1AAlkane biosynthesis genedrought stress[184,185]
Cell signaling protein/enzymesTaCPK7 and TaCPK12Calcium-dependent protein kinasesdrought stress[139]
TaRN1Serine/threonine protein kinasesalt stress[169]
TaPYL1ABA receptordrought stress[186]
CYCB2, CDKA1B2-type cyclin in mitotic; cyclin-dependent kinasesdrought stress[187]
TaSCPL184-6DSerine carboxypeptidase-like proteinsalt stress[188]
Receptor like protein kinaseTaLRRKsLeucine-rich repeat kinaseheat and drought, and salt[42,43]
TaCRK68-ACysteine-rich receptor-like kinasesheat, drought, cold and salt stress[44]
TaTLPsThaumatin-like protein kinasesheat and drought, and salt[45]
TaPERKsProline-Rich Extensin-like Receptor Kinasesheat stress[46,47]
Epigenetic regulation genesTaMBD2Methyl CpG-binding domain proteinscold stress[94]
TaHAG1Histone acetyltransferaseheat and salt stress[164,165]
TaCMTCytosine-5 DNA methyltransferasesdrought, heat stress[189]
Tr-7A-JMJ1,Tr-1B-JMJ3Histone demethylasedrought stress[190]
TaSIRFP-3A,TaSIRFP-3BRING-HC-type E3 ligasescold stress[191]
TaPUB2/TaPUB3U-box E3 ubiquitin ligasedrought stress[192]
Other stress response genesTaCHYR2.1, TaCHYR9.2, TaCHYR11.1CHY zinc-finger and RING finger proteinsalt stress[129]
TaICE41 and TaICE87Inducer of CBF expressioncold stress[162]
Vrn-B1, Vrn-D3Vernalization genesdrought and heat stress[193,194]
TaBI-1.1Bax Inhibitorheat stress[195]
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

Du, L.; Ma, Z.; Mao, H. Duplicate Genes Contribute to Variability in Abiotic Stress Resistance in Allopolyploid Wheat. Plants 2023, 12, 2465. https://doi.org/10.3390/plants12132465

AMA Style

Du L, Ma Z, Mao H. Duplicate Genes Contribute to Variability in Abiotic Stress Resistance in Allopolyploid Wheat. Plants. 2023; 12(13):2465. https://doi.org/10.3390/plants12132465

Chicago/Turabian Style

Du, Linying, Zhenbing Ma, and Hude Mao. 2023. "Duplicate Genes Contribute to Variability in Abiotic Stress Resistance in Allopolyploid Wheat" Plants 12, no. 13: 2465. https://doi.org/10.3390/plants12132465

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