Next Article in Journal
Optimal Planting Density and Nutrient Application of Soybeans: A Case Study in Northeastern China
Previous Article in Journal
Barley Genetic Resources: Advancing Conservation and Applications for Breeding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genomic and Transcriptomic Approaches to Developing Abiotic Stress-Resilient Crops

by
Saravanappriyan Kamali
1 and
Amarjeet Singh
1,2,*
1
National Institute of Plant Genome Research, New Delhi 110067, India
2
School of Life Sciences, Jawaharlal Nehru University (JNU), New Delhi 110067, India
*
Author to whom correspondence should be addressed.
Agronomy 2023, 13(12), 2903; https://doi.org/10.3390/agronomy13122903
Submission received: 8 October 2023 / Revised: 9 November 2023 / Accepted: 15 November 2023 / Published: 26 November 2023
(This article belongs to the Special Issue The Application of Genomics Methods for Crop Improvement)

Abstract

:
In the realm of agriculture, a pressing concern remains the abiotic stresses, such as temperature fluctuation, drought, soil salinity, and heavy metal contamination. These adverse growth conditions hamper crop yields and global food security. In this review, we present a comprehensive examination of the recent advancements in utilizing genomics and transcriptomics, tools to enhance crop resilience against these stress factors. Genomics aids in the identification of genes responsive to stress, unravels regulatory networks, and pinpoints genetic variations linked to stress tolerance. Concurrently, transcriptomics sheds light on the intricate dynamics of gene expression during stress conditions, unearthing novel stress-responsive genes and signaling pathways. This wealth of knowledge shapes the development of stress-tolerant crop varieties, achieved through conventional breeding programs and state-of-the-art genetic engineering and gene editing techniques like CRISPR-Cas9. Moreover, the integration of diverse omics data and functional genomics tools empowers precise manipulation of crop genomes to fortify their stress resilience. In summary, the integration of genomics and transcriptomics holds substantial promise in elucidating the molecular mechanisms behind crop stress tolerance, offering a path towards sustainable agriculture and safeguarding food security amidst shifting environmental challenges.

1. Introduction

The Food and Agriculture Organization (FAO) of the United Nations reported in its statistical yearbook 2022 that hunger is still on the rise with 770 million people undernourished in the year 2021. In every continent on the earth, there is moderate to severe level of food insecurity. Between 2000 and 2020, agricultural land declined by 134 million ha and water stress is prevalent in most countries [1]. Considering these circumstances, the primary goal must be ensuring nourishment for the growing population. Unfortunately, crop production seems to have reached a plateau and is even decreasing in recent times, primarily because of climate change and the scarcity of arable land for cultivation. In particular, abiotic stresses like temperature extremes (such as heat or cold stress), water availability (drought or flooding), salinity, nutrient deficiency or excess, heavy metal toxicity, pollution, radiation, and other physical or chemical factors account for significant crop yield losses every year all over the world [2]. Abiotic stress can disrupt various physiological and biochemical processes within plants, leading to reduced growth, development, and productivity. It can also make plants more susceptible to diseases, pests, and other biotic stresses [3]. Hence, enhancing agricultural productivity and sustainability holds immense significance on a global scale. There is an urgent requirement for crop improvement to guarantee food security for the world’s population [4]. The goal of crop improvement is to develop improved plant varieties that exhibit desirable traits such as increased yield, disease resistance, tolerance to environmental stresses, improved nutritional content, and better post-harvest characteristics [5]. During the green revolution, biotechnology tools were not widely employed as the field of biotechnology was still in its infancy. The full potential of modern biotechnology, including genetic engineering and genomics was realized later [6]. Since then, these technologies have played a significant role in crop improvement. Multiple “omics” approaches are extensively used for crop improvement. The progress made in next-generation sequencing (NGS) has opened avenues for the development of new omics fields, including genomics, transcriptomics, proteomics, metabolomics, ionomics, etc. [7]). Especially, genomics and transcriptomics play crucial roles in crop improvement by providing valuable insights into the genetic makeup and gene expression patterns of crucial genes and gene families [7]. In this review, we will focus on genomics and transcriptomics approaches employed in crop improvement against abiotic stresses.

2. Genomics for Understanding Abiotic Stress Response in Plants

Genomics focuses on exploring and studying the genomes, i.e., the complete set of DNA or genetic material within an organism, with the help of a range of techniques and approaches [8]. It involves analyzing and interpreting the structure, function, and evolution of genomes. Genomics has been broadly classified into functional, structural, and comparative genomics, based on its methodologies and outcomes [9]. In this section, the three classes of genomics, their methodologies, and their application in crop improvement for abiotic stress tolerance will be discussed.

2.1. Functional Genomics

Functional studies of genomes readily produce information that is applicable to crop improvement. Functional genomics involves the functional characterization of genes and their interactions with other genes in a regulatory network [10]. Functional genomics includes different approaches to identify gene functions, such as sequence- or hybridization-based methodologies, gene inactivation or editing-based approaches, and gene overexpression [10]. These different approaches will be discussed in detail here.

2.1.1. Gene Inactivation and Editing Approaches for Functional Analysis of Abiotic Stress-Related Genes

RNAi and VIGS

RNAi technology can be used for gene inactivation and functional studies [11]. It involves the introduction of small interfering RNA (siRNA) or short hairpin RNA (shRNA) molecules into the cell. These molecules bind to the target mRNA, leading to its degradation or inhibition of its protein translation. Through the utilization of RNAi techniques, the roles of various genes in abiotic stress responses have been uncovered across plant species. In soybean, overexpression of the Tubby-like protein gene GmTLP8 enhanced the plant’s resilience to drought and salt stress, while its suppression decreased tolerance [12]. Similarly, in rice, overexpression of the Aux/IAA gene OsIAA20 and the nuclear export receptor OsXPO1 improved plant tolerance to drought and salt stress, but reduced plant height and seed-setting rate in the case of OsXPO1. Conversely, suppressing these genes by RNAi decreased plant resilience to stress and induced developmental defects [13]. In tobacco, suppression of the APETALA2/ethylene response factor (AP2/ERF) gene NtRAV-4 enhanced root development, leaf photosynthetic ability, and drought tolerance [14].
Virus-induced gene silencing (VIGS) is a transient induction of RNAi by using modified viral vectors for plant functional genomics. RNAi and VIGS have been used for improvement of several traits in crop plants. For example, VIGS of CaWRKY40a in pepper enhances resistance against Xanthomonas campestris infection [15]. In wheat VIGS of Diketone Metabolism-PKS (DMP), -Hydrolase (DMH) affects β-Diketone biosynthesis and results in increased glaucousness which is associated with yield [16]. In rice, both the General Control Non-derepressible 5 (GCN5) and Adenosine Deaminase2 (ADA2) RNAi lines produced fewer crown roots and showed reduced primary root length and shoot height compared with the wild type [17]. These examples show that both RNAi and VIGS are useful tools in improving several important traits in crops through gene silencing [18]. Antisense oligonucleotides (ASO) are short synthetic nucleic acid sequences that are designed to be complementary to the target gene’s mRNA. Upon binding to the mRNA, ASOs can prevent mRNA translation or promote degradation of the target mRNA. ASOs can be chemically modified to enhance stability and specificity, and they are being explored as therapeutic agents for various genetic diseases [19]. Conditional gene knockout techniques enable the inactivation of genes in a specific tissue, at a particular developmental stage, or upon induction by external stimuli. This can be achieved using Cre-loxP or Flp-FRT systems by flanking the target gene with loxP sites. When Cre recombinase is expressed under the control of a specific promoter (e.g., tissue-specific promoters or inducible promoters), it mediates the excision of the target gene between the loxP sites, resulting in gene inactivation [20]. Flp recombinase recognizes FRT sites and catalyzes recombination between them, resulting in DNA excision, inversion, or rearrangement, depending on the orientation and arrangement of the FRT sites [21].

ZFN, TALEN, and CRISPR-Cas9

To functionally characterize plant genes, genome editing techniques like targeted mutation, INDEL creation, and genomic sequence modifications can be applied [11]. Common genome editing tools include Zinc finger nucleases (ZFNs), transcriptional activator-like effector nucleases (TALENs), and clustered regularly interspaced short palindromic repeat (CRISPR)-Cas9 [11]. ZFNs, for instance, induce double-strand breaks at specific genomic locations, resulting in targeted mutagenesis such as chromosomal deletions, transgene removal, and precise DNA integration [22]. They consist of a non-specific cleavage domain from the FoKI endonuclease fused with custom-designed Cys2-His2 zinc finger proteins, leading to DSB formation. Plant systems employ error-prone non-homologous end joining (NHEJ) for DNA repair (Wada et al., 2022). TALEN- and ZFN-based genome editing for abiotic stress tolerance have not been carried out extensively in crops, yet. But CRISPR-Cas (clustered regularly interspaced short palindromic repeats-CRISPR-associated proteins) has been extensively used in crops for genome editing in this decade. Using CRISPR-Cas9 knockout mutation of miRNAs, the OsMIR408 and OsMIR528 lines were developed. These lines were salt-sensitive, and it has been found that these genes were positive regulators of salt stress tolerance [23]. In the past decade, many salt stress-related genes have been identified, for example, OsRR22, identified as a salt stress-related gene, encodes a 696-amino acid B-type response regulator transcription factor that is involved in cytokinin signaling. CRISPR/Cas9 editing leading to loss of function mutation in the OsRR22 gene results in increased salt tolerance [24]. Many genes have been identified using mutation studies and among them, the drought and salt tolerance (DST) gene was identified as a non-desirable gene, and is present in the genome due to linkage. Genome editing via CRISPR-Cas technology has caused 366 bp deletion and has generated the loss of function mutation of the DST gene. This mutant line shows enhanced leaf water retention under dehydration stress [25]. Hybrid proline-rich proteins (HyPRPs) have been demonstrated to play distinct roles in responses to biotic and abiotic stresses across various plant species. However, in the case of tomato, a specific HyPRP called SlHyPRP1 has been identified as a suppressor of multiple stress responses. Precise elimination of SlHyPRP1 negative-response domain(s) using CRISPR-Cas9 has led to high salinity tolerance at the germination and vegetative stages [26]. The ARGOS8 gene, which acts as a negative regulator in the ethylene response pathway, has been found to induce drought tolerance genes in maize [27]. Although numerous natural genotypes displaying drought resistance have been identified, the expression of ARGOS8 therein has been observed to be very low. To overcome this limitation, CRISPR-Cas9 was used with a less-restrictive constitutive promoter called GOS2 to enhance the expression of ARGOS8 [27]. This led to increased drought tolerance in maize. CRISPR-Cas9-mediated editing of genes encoding two enzymes, poly ADP-ribose polymerase (PARP) and ADP-ribose specific Nudix hydrolase (NUDX), has resulted in increased tolerance to oxidative, drought, and genotoxic stresses in maize [27]. NPR1 (nonexpressor of pathogenesis-related gene 1) serves as a central regulator in the plant defense response against pathogens. While its role in the defense pathway is well understood, its involvement in abiotic stress remains unclear. CRISPR-Cas9-mediated mutagenesis of SlNPR1 in tomatoes has led to reduced drought tolerance, which was accompanied by decreased expression of key drought-related genes, including SlGST, SlDHN, and SlDREB [28]. These findings suggest that NPR1 not only regulates the response to biotic stress but also plays a role in the plant’s response to abiotic stress. Overall, these findings demonstrate the potential of CRISPR-Cas9 technology in crop improvement particularly for abiotic stress tolerance traits. All this information aids in the development of transgenic or targeted mutant lines of crops for abiotic stress tolerance.

TILLING

Targeted induced local lesions in genomes (TILLING) is a useful and high-throughput technique to identify single nucleotide mutations in a specific region of a gene of interest. TILLING methods are generally employed for screening of both phenotypic and genotypic variations in crops under abiotic stresses [29]. Its updated variant EcoTILLING has been used for the identification of natural polymorphisms [30]. EcoTILLING is useful particularly for non-model organisms with limited genetic resources and genomic information. In one study, chemically mutated lines were screened using the TILLING approach to identify variants in membrane transport genes and their response to salt stress [31]. Among 2961 mutant lines, 41 mutants with single nucleotide polymorphisms (SNPs) in nine membrane transporters were discovered. Altered sequences found in the exon region of seven genes, and these seven mutants exhibited salt tolerance. Additionally, five mutants with SNPs, i.e., OsAKT1 (outward rectifying Potassium channel 1), OsHKT6 (high-affinity Potassium transporter 6), OsNSCC2 (nonselective cation channel 2), OsHAK11 (high affinity Potassium transporter 11), and OsSOS1 (salt overly sensitive 1) showed altered gene expression levels [31]. These mutants hold potential for developing salt-tolerant lines. Similarly, chilling-tolerant lines were identified in a TILLING mutant population of Donganbyeo rice cultivar. Comparative transcriptome analysis revealed that chilling stress tolerance was associated with monosaccharide catabolic processes, which provide the necessary energy for cold adaptation in rice [32]. High-temperature stress during grain filling leads to delayed endosperm formation and grain chalkiness. Multi-omics analysis indicated that the downregulation of starch synthesis enzymes and upregulation of α-amylases could be the possible reason behind this [32]. Targeting the TILLING mutants of α-amylase genes may reduce chalkiness in heat-stressed grains without the need for transgenic approaches. TILLING was also used to create three mutations in the Brassica rapa CAX1a transporter. These mutants, along with the original strain (R-o-18), were cultivated in a saline environment, and various parameters were assessed, including biomass, photosynthesis efficiency, glucose-6-phosphate dehydrogenase (G6PDH), and soluble carbohydrates. It was revealed that the BraA.cax1a-7 mutation negatively impacted alkalinity tolerance, resulting in reduced plant biomass, increased oxidative stress, and partial inhibition of antioxidant responses and photosynthesis. Conversely, the BraA.cax1a-12 mutation enhanced plant biomass, increased calcium (Ca2+) accumulation, reduced oxidative stress, and improved both antioxidant responses and photosynthesis performance [33]. Thus, BraA.cax1a-12 emerges as a promising mutation for enhancing alkaline tolerance in plants. Also, a mutant of the tomato HSBP1 (heat shock binding protein 1) gene was identified in the TILLING population, resulting in a partial loss of protein function. This mutation led to improved resistance to high temperatures in young plants and increased resilience in mature plants under repeated heat stress [34]. A list of tools and databases available for plant gene inactivation or gene editing is given in Table 1.

2.1.2. Gain-of-Function Approach through Gene Overexpression

The technique of gene overexpression has become a pivotal tool in genomics, offering valuable insights into gene function and potential applications in crop improvement. In the past decades, numerous significant studies have underscored the efficacy of this approach, particularly in enhancing stress tolerance and overall crop performance. For instance, in rice, OsCYP19-4 was found to be significantly upregulated under cold stress. Overexpression of OsCYP19-4 in rice demonstrated enhanced tolerance to cold stress and improved grain yield [48]. Tomato plants overexpressing SlGRAS7 (GAI, RGA, and SCR-like protein 7) exhibited enhanced resistance to drought and salt stress compared to the wild type (WT) [49]. Elongation factor 1 A (EF1A), a crucial regulator for protein synthesis, has been found to participate in plant responses to abiotic stress and environmental adaptation. Overexpression of MtEF1A1 in Medicago truncatula resulted in increased salt stress resistance and reduced levels of reactive oxygen species (ROS) in leaves [50]. Additionally, the expression of abiotic stress-responsive genes (MtRD22A and MtCOR15A) and calcium-binding genes (MtCaM and MtCBL4) was upregulated in M. truncatula lines overexpressing MtEF1A1 [50]. These abiotic stress-related genes could be important for developing stress resilient crops. Some important genes could impart tolerance to multiple abiotic stresses in plants when overexpressed. For example, in white birch (Betula platyphylla Suk.), overexpression of Ethylene response factor 1.1 (ERF1.1) showed increased tolerance to cold, drought, and salt stress compared to WT. RNA-Seq analysis has shown 689 differentially expressed genes (DEGs) in transgenic birch compared to WT. This shows overexpression of single gene results in triggering of cascade of gene networks leading to stress tolerance [51]. Similarly, Arabidopsis ICE1 (inducer of CBF expression 1) was overexpressed in Indica rice to improve cold tolerance in cold-sensitive rice. AtICE1 lines showed lower accumulation of H2O2 and higher membrane stability, thus increased seed survival rate under cold stress. Also, AtICE1 lines had increased grain yield under cold, drought, and salt stresses [52]. In rice group-A PP2Cs, OsPP108 conferred tolerance to salt and drought stresses when overexpressed in Arabidopsis. Interestingly, this gene rendered plants highly insensitive to ABA and proposed to regulate abiotic stress tolerance possibly in an ABA-independent manner [53]. Similarly, tomato overexpression of the transgenic line of dwarf and delayed flowering 2 (SlDDF2) under stress-inducible RD29a promoter showed better growth performance and tolerance under abiotic stresses including salinity and drought [54]. Overall, these studies suggested that identification of the key gene inducible under multiple stresses, and its overexpression in plants, is an effective strategy for developing multi-stress resilient crops. Details of several other genes that regulate a plant’s abiotic stress response through overexpression or gene knockout/silencing is given in Table 2.

2.2. Structural Genomics

Functional genomics primarily addresses gene functionality, whereas structural genomics focuses on elucidating the physical structure of genomes. Understanding an individual genome’s structure is valuable for gene manipulation and DNA segment control [11]. Structural genomics encompasses the creation of high-resolution genetic and physical maps. In the context of crop improvement, genome sequencing and molecular mapping hold significant importance [74]. Molecular markers found on polymorphisms within DNA play a crucial role in assessing genetic diversity in germplasm [75]. These DNA markers find extensive utility in plant breeding, aiding in gene mapping, the identification of quantitative trait loci (QTL), germplasm evaluation, and marker-assisted breeding (MAS). Recent advances in genotyping techniques based on single nucleotide polymorphisms have accelerated MAS. The advent of next-generation sequencing (NGS) has further facilitated genome resequencing and the comparison of various genotypes, leading to the identification of thousands of SNPs [76]. A noteworthy development in molecular markers is insertion site-based polymorphisms (ISBPs), capitalizing on polymorphisms generated by insertion regions at repeat junctions [75].

Molecular Mapping and Marker-Assisted Breeding in Crops for Abiotic Stress Tolerance

Molecular mapping techniques play a pivotal role in traditional breeding, facilitating the development of elite crop varieties. These techniques have unveiled significant structural genomic variations associated with stress conditions, thus aiding in the identification of genotypes resilient to such stressors [11]. These variations serve as valuable indicators for pinpointing the target genes responsible for abiotic stress responses. Furthermore, these advancements have led to the localization of stress-related quantitative trait loci (QTLs). The evolution of next-generation sequencing (NGS) technologies and DNA polymorphism detection techniques, alongside map-based cloning, has further enhanced our ability to identify additional QTLs and develop the associated markers. These methodologies collectively expedite the breeding process, significantly boosting breeding efficiency [75]. For instance, in the context of salinity tolerance in temperate japonica rice accessions, the evaluation of 235 accessions marked with 30,000 SNP markers facilitated a genome-wide association study (GWAS). This study resulted in the identification of 27 QTLs, several of which were closely positioned to genes linked with calcium signaling and kinases. [77]. In a separate study employing bulk segregant analysis–next generation sequencing (BSA-Seq), four candidate regions were associated with thousand-grain weight (TGW) under alkali stress conditions. Notably, QTL-qATGW 2-2 was precisely mapped within a 116Kb range between molecular markers RM13592 and Indel3 on Chr.2, encompassing 18 predictive genes. This analysis highlighted Os02g39884 as the prime candidate gene in QTL-qATGW 2-2, representing an alkali-tolerant gene locus in rice [78,79] aiming to enhance stress tolerance in improved white ponni (IWP) rice by pyramiding QTLs against drought (qDTY 1.1; qDTY 2.1), salinity (Saltol), and submergence (Sub1) through marker-assisted selection (MAS). These QTLs offer promising avenues for developing triple-stress-tolerant crop varieties.
Meta analysis was conducted to refine the locations of 195 major QTLs related to drought, salinity, and waterlogging tolerance in barley from various mapping populations. Identified meta-QTLs (MQTLs) were used to search for candidate genes linked to these tolerances. These refined MQTLs and candidate genes are crucial for successful MAS in barley breeding [13]. A comprehensive meta-QTL analysis on maize revealed the presence of 32 meta-QTLs associated with different abiotic stresses, with a total of 1907 candidate genes identified [80]. Notably, the meta-QTLs designated as MQTL2.1, 5.1, 5.2, 5.6, 7.1, 9.1, and 9.2 were found to control various stress-related traits, contributing to combined abiotic stress tolerance. Recently, a total of 33 QTLs associated with drought tolerance were identified across eight chromosomes in sunflower. Notably, four genes located on chromosome 13 were found to be associated with drought stress in both the germination and seedling stages. These genes, namely LOC110898128 (aquaporin SIP1-2-like), LOC110898092 (cytochrome P450 94C1), LOC110898071 (GABA transporter 1-like), and LOC110898072 (GABA transporter 1-like isoform X2) have been annotated and will undergo further functional validation [81]. This study contributes insights into the molecular mechanisms underlying sunflower’s response to drought stress, providing a foundation for drought tolerance breeding and genetic improvement in sunflower. Thus, in this decade, many molecular markers and QTLs have been developed and this can help in finding crop accessions with desired traits and faster breeding for the development of new hybrid lines with multiple abiotic stress tolerance. A list of tools and databases related to QTLs and molecular markers in plants are given in Table 3.

2.3. Comparative Genomic

Comparative genomics involves the comparison of two or more genomes to uncover both their similarities and differences. In this context, gene annotations derived from model plants can be applied to newly sequenced crop species that are yet to undergo functional studies. Essential to this process is knowledge about orthologs, genes that have evolved from a common ancestor and serve similar functions among species descended from that ancestor [97]. Furthermore, comparative genomics finds utility in analyzing the expression profiles of less-studied plants under diverse stress conditions, enabling the identification of stress-related genes and facilitating inter-species expression profile comparisons. Both intra- and inter-specific sequence comparisons rely on a range of computational methods, including multiple sequence alignment, genome-wide comparisons, analyses of orthology and paralogy, as well as the construction of phylogenetic trees [98]. Various tools and databases are available for comparative genomic studies. Ensembl, an extensive database, provides access to numerous annotated genomes and comparative analysis tools, making it a cornerstone for studying genetic variation across species [99,100]. The UCSC Genome Browser offers visualization tools for a diverse range of genomes, simplifying cross-species comparisons and aiding in the exploration of genomic features [101,102]. Another critical resource is OrthoDB, a comprehensive catalog of orthologous genes spanning various species, facilitating the identification of conserved genes with potential roles in evolutionary processes [103]. These tools play a pivotal role in comparative genomics, enabling researchers to uncover evolutionary relationships, conserved genetic elements, and functional insights across genomes. In the context of crop improvement, these resources are invaluable for identifying candidate genes, regulatory elements, and conserved pathways that can inform breeding strategies, enhance crop resilience, and improve agricultural productivity.

Genome-Wide Identification of Genes Families and Promoter Elements Responsible for Abiotic Stress Response

Genome-wide identification and analysis of gene families in crop genomes typically rely on the sequence homology to known genes. Expression analysis further aids in identifying functional members and pseudogenes [98]. This approach has successfully identified numerous gene families in crops. In rice, numerous gene families have been identified through genome-wide approaches. Moreover, expression analysis helped in marking genes related to specific functions such as biotic stress, abiotic stress, plant development, etc. For instance, genome-wide analysis revealed 491 Pentatricopeptide-repeat proteins (PPRs), categorized into subclass P (246 genes) and subclass PLS (245 genes). Expression analysis showed induction of many PPR genes under both biotic and abiotic stress [104]. DUF221 domain-containing genes (DDP genes) play essential roles in plant development, hormone signaling, and stress responses. Comparative genomics in rice identified at least nine DDP gene members in both domesticated and wild rice, with various expression analyses showing their upregulation under salt stress [105]. Comparative genomic tools identified 81 Ca2+ transport element genes in rice, with their expression established during abiotic stresses and different developmental stages using microarray and qRT-PCR techniques [106]. Similarly, many gene family cell signaling components, such as protein phosphatases [107], phospholipases [108,109], Ca2+ dependent protein kinases (CDPKs) [110], and receptor-like cytoplasmic kinases (RLCKs) [111] were identified and analyzed in rice. Additionally, several other gene families, such as, MADS-box, Phytocyanin, BURP, Arabinogalactan, Nuclear-factor Y, ABA repressor, and various transcription factors, have been identified as abiotic stress-responsive genes in rice through comparative genomics approaches [112,113,114,115,116,117,118]. Besides rice, the genome-wide approach has been used to identify and analyze of a number of gene families in different crops like soybean [119,120]), cotton [121,122], Brassica Spp. [52], Chickpea [123,124,125,126] and Maize [127,128,129].
Identifying stress-inducible promoter regions is important for deploying transgenes with specific promoters for optimum expression under stress conditions. Studies have shown the effectiveness of promoters from stress-responsive genes, such as OsABA2, RAB16A, RD29A, and HP1 in driving strong expression of genes under abiotic stress conditions [130]. Transcription profiles of rice, soybean, and Arabidopsis revealed conserved sequences in cold and dehydration-inducible promoters, including the abscisic acid-responsive element (ABRE). Novel cold-inducible cis elements CGTACG and GTAGTA were identified in rice genes promoters [131]. The AL resistance transcription factor1 (ART1), a C2H2 type zinc finger transcription factor with specific cis-acting elements in the promoter, was found to be involved in detoxifying aluminum in rice [132]. Comparative studies in rice and Arabidopsis revealed the occurrence and arrangements of cis-regulatory elements ABRE and CE3 in gene promoters, with ABRE forming ABA-responsive complexes and exhibiting distinct combinations with CE3 in rice [133]. These findings significantly contribute to our understanding of the genetic mechanisms involved in plant response to environmental stresses, paving the way for future research and potential applications in crop improvement. In summary, comparative genomics approaches have successfully identified various gene families involved in abiotic stress responses. These findings provide valuable insights into the genetic mechanisms underlying stress tolerance and can facilitate the development of stress-tolerant crop varieties.

3. Transcriptomics Techniques

Transcriptomics encompasses the comprehensive analysis of all RNAs transcribed by a cell or tissue, including both coding and non-coding RNA at a specific functional state. It involves the study of the type, structure, function, and regulation of gene transcription [134]. Transcriptomics provides valuable insights by quantitatively analyzing changes in plant gene expression. It enables the exploration of regulatory networks and whole-genome expression patterns, which help in revealing novel stress tolerance-associated genes in crop plants [135].
Advancements in transcript sequencing and analysis technologies have provide a significant upthrust to the field of transcriptomics. Traditional methods like northern blotting and RT-PCR are limited to analyzing single transcripts or small groups of transcripts at a time [136]. However, the introduction of microarrays in the mid-1990s revolutionized transcript profiling, enabling simultaneous analysis of thousands of genes [137]. Subsequently, real-time RT-PCR or RT-qPCR emerged as a sensitive technique for detecting low-abundance transcripts and became popular for both absolute and relative quantification of gene expression [134]. A major breakthrough happened with the advent of next-generation sequencing (NGS), which profoundly impacted gene expression profiling. NGS-based RNA sequencing (RNA-Seq) has widened our horizon of understanding gene regulatory networks and epigenetics. This powerful technology enables the detection and quantification of known, novel, and less abundant transcripts, encompassing both coding and non-coding RNA [136]. The transcriptomics approaches involved in crop improvement against abiotic stress are summarized in Figure 1.
ESTs (expressed sequence tags) are random individual transcripts obtained from cDNA libraries and sequenced using one-time, low-throughput Sanger sequencing. Initially, ESTs were considered an efficient means of determining the gene content of an organism without the need for whole-genome sequencing, making them a valuable tool in early transcriptomics research [134]. Another sequencing-based gene expression analysis technique emerged in 1995, known as serial analysis of gene expression (SAGE). This method involved Sanger sequencing of concatenated random transcript fragments, and quantification is achieved by matching the transcripts with known genes. A variant of SAGE utilizing high-throughput sequencing techniques, called digital gene expression analysis, was also briefly employed [136]. However, these techniques have been overtaken by contemporary approaches such as microarrays and RNA-Seq, which offer more advanced capabilities and broader applications in transcriptomics research. These modern transcriptomics technologies have significantly advanced our understanding of gene expression regulation and the complexity of cellular processes. They have become essential tools for researchers studying various biological phenomena, including stress responses, development, and disease mechanisms [138]. The continuous development of transcriptomics methods promises to unveil further layers of gene regulation and functional insights in the future.

3.1. ESTs and Microarray for Identification of Abiotic Stress Responsive Genes

Over the years, numerous candidate genes related to abiotic stress in different plant species have been identified using ESTs and microarray techniques. For example, transcript profiling using microarrays and ESTs revealed that Glutathione S-transferases (GSTs) exhibit similar and specific functions during various stages of rice development, while also mediating cross-talk between different stress and hormone response pathways [139]. Microarray expression analysis revealed significant differential expression under abiotic stresses such as drought, salinity, and cold and during reproductive developmental stages in important gene families like protein phosphatases [107], phospholipase A and D [108,140], phospholipase C [110], Calcium transport elements [106], the MADS box family [112], and the C2H2 zinc finger TF family [141] in rice. The expression patterns of most of these differentially expressed genes were validated using qRT-PCR. Heat shock proteins (HSPs) are another important group involved in plant response to heat stress and regulated by heat shock factors (HSFs). In rice, HSFs are categorized into three classes: A, B, and C [142]. Expression profiling through ESTs, microarrays, and qRT-PCR showed that eight OsHSFs are upregulated during seed development and six HSFs during abiotic stress in both roots and shoots. OsHSFA2a and OsHSFA3 are upregulated in response to cold and drought stress, while OsHSFB4a showed little or no change in expression [143]. Affymetrix microarray technology was used to analyze the expression of the AP2/EREBP gene family in Chaling wild rice and cultivated cold-sensitive rice cultivar Pei’ai64S. Microarray revealed that 36 AP2/EREBP genes had a much higher expression level in Chaling wild rice than in Pei’ai64S [144]. Rice is a C3 plant and has become a model organism for studying the genetic engineering of the C4 pathway. Several C4 gene families were identified in the rice genome through sequence homology using maize C4 gene sequences as queries [145]. Expression analysis using EST and FL-cDNA databases indicated the presence of at least one EST or FL-cDNA for all the identified genes. In addition, abiotic stress-related genes were also identified [145].
Using EST and microarray analysis, researchers identified genes associated with drought stress tolerance in cotton. For instance, genes encoding dehydration-responsive element binding (DREB) transcription factors were found to be upregulated under drought conditions, suggesting their role in regulating drought-responsive genes [146]. Additionally, using EST and microarray analysis in cotton, many abiotic stress responsive genes were identified, which included MYB-related, C2H2, FAR1, bHLH, bZIP, MADS [147] and RTNLB5, and PRA1 [148]. Similarly, in barley, a pre-mRNA processing (Prp1) gene was found to be crucial during seed development and abiotic stress [149]. Moreover, in barley, EST databases helped in revealing the roles of HKT (high-affinity K+ transporter) proteins in regulating potassium (K+) uptake under saline conditions [150]. Also, SbDREB2A, encoding a DREB transcription factor, was found to be upregulated under drought conditions using ESTs and microarray analysis. Its overexpression enhanced drought tolerance in transgenic sorghum plants [151]. To facilitate such research, there are two large EST repositories, edbEST, and UniGene, both hosted by NCBI, which include EST data from a variety of organisms [152]. These repositories serve as valuable resources for scientists to access and analyze transcriptomic data for various research purposes.

3.2. RNA-Seq for Identification of Genes Involved in Abiotic Stress Responses

RNA-Seq is a powerful technique that combines high-throughput sequencing with computational methods to quantify and analyze transcripts within an RNA pool. The sequencing process generates nucleotide sequences, typically around 100 base pairs in length, although the actual read length can vary depending on the specific sequencing method used [153].
The fundamental principle of RNA-Seq involves the alignment of these generated sequences either to a reference genome or to each other. By mapping the RNA transcripts, one can identify the specific genes present within a genome and also determine which genes are actively expressed at a particular point in time [153]. One of the major advantages of RNA-Seq is its ability to accurately measure gene expression levels. Unlike microarrays, which rely on hybridization to predefine probes, RNA-Seq provides a more comprehensive and quantitative view of gene expression. It can detect both known and novel transcripts, including non-coding RNA, thereby offering a more complete understanding of the transcriptome. Additionally, RNA-Seq is not limited by the predefined nature of microarrays, allowing researchers to explore previously unannotated regions of the genome and discover novel transcripts and isoforms. This capability has been instrumental in identifying alternative splicing events, alternative transcription start sites, and other post-transcriptional regulatory processes [138]. RNA-Seq analysis between imbibed seeds and dry seeds of rice showed that genes related to the cell wall, abiotic stress, and antioxidants were associated with stress response during imbibition and germination [154]. Genes like receptor kinase (e.g., OsCRK2), pectinesterase (e.g., OsPME3), polygalacturonase (e.g., OsPGIP1), cupin-domain protein (e.g., OsCP1), methyltransferases (e.g., OsTRMT1), SPX domain (e.g., OsPHR2), GSTs (e.g., OsGSTL3), and peroxidase (e.g., OsAPX2) were significantly expressed. GSTs were particularly implicated in preventing H2O2 accumulation during the initial imbibition stage, contributing to successful seed germination [154]. RNA-Seq analysis of Aus, a drought and heat-tolerant cultivar of rice, identified 56 differentially expressed genes in developing seeds under combined drought and heat stresses [155]. Among them, B12288 (RAB21), a dehydrin family LEA protein, was significantly induced. Although sequence differences were not large, functional effects were observed, highlighting the role of dehydrins in stress regulation responses. RNA-Seq analysis in tomatoes showed that C2H2-type zinc finger protein genes C2H2-ZFP3, -5, and -8 were involved in cold, salt, and drought resistance [156]. RNA-Seq analysis of rice treated with Cd and As revealed genes associated with redox control, stress response, transcriptional regulation, transmembrane transport, signal transduction, biosynthesis, and metabolism of macromolecules and sulfur compounds [157].
In wheat (Triticum aestivum), RNA-Seq analysis revealed that R2R3-MYB family genes are abiotic stress-responsive [158]. In maize (Zea mays), RNA-Seq analysis under abiotic stress showed involvement of various transcription factors like ERF, NAC, ARF and HD-ZIP to initiate abiotic stress response [159]. Soybean (Glycine max) experiencing water deficit stress displayed significant changes in gene expression in RNA-Seq analysis, with upregulation of genes involved in ABA signaling, including GmPYLs and GmPP2Cs, both part of the ABA receptor complex [160]. Recently, RNA-seq analysis of two contrasting cultivars of chickpea, i.e., K+ deficiency-sensitive (Pusa362) and -tolerant (Pusa372), identified hundreds of differentially expressed genes in both the cultivars [161]. These genes belonged to different functionally categories and pathways. This analysis provided a significant insight into the K+ deficiency-tolerant mechanism in the important legume crop chickpea.
The wealth of information gained from these RNA-Seq studies contributes to advancing our knowledge of plant stress responses and has laid a foundation for targeted strategies to improve crop resilience and ensure global food security. Various databases that encompass mRNA sequences obtained from crops have been developed over the years. A list of databases related to plant expression is given in Table 4.

3.3. Third Generation Sequencing for Identification of Abiotic Stress Related Genes

The recent surge in the third-generation sequencing (TGS) technologies has significantly impacted the field of transcriptomics, as TGS offer several advantages over traditional first- and second-generation sequencing [175]. Third-generation sequencing methods, such as Pacific Biosciences (PacBio) and Oxford Nanopore Technologies (ONT) provide longer reads and can directly sequence RNA molecules. This feature significantly facilitates the transcriptional profiling as they enable the identification of full-length transcripts without the need for assembly or the use of sophisticated bioinformatics tools [175]. Importantly, TGS help in identifying novel genes involved in abiotic stress responses in plants. Recently, PacBio sequencing was employed to sequence the transcriptomes of ten rice cultivars belonging to three distinct subspecies under normal and abiotic stress conditions. It was able to reconstruct high-quality, plant-specific isoforms, ranging from 37,500 to 54,600 isoforms per cultivar. To reduce redundancy in the sequences, the isoforms were consolidated and assessed for protein completeness. Approximately 40% of the identified transcripts represented novel isoforms not present in the reference transcriptome of Nipponbare rice. Moreover, for the drought- and heat-tolerant aus cultivar N22, 56 differentially expressed genes were identified in developing seeds under the combined stress of heat and drought [155]. Similarly in poplar PacBio sequencing and RNA-Seq was combinedly used to identify the role of alternative splicing (AS) in cold stress tolerance. It was found that 1261 AS events in Populus trichocarpa and 2101 in P. ussuriensis among which intron retention, with a frequency of more than 30% was the most prominent type under cold stress [176]. Similarly in cotton, long read transcripts were sequenced using PacBio to identify novel transcripts involved in salt stress. A significant number of DEGs involved in various ion homeostasis, hormone signaling, cell wall modification and transcription factors were found [136]. The Arachis glabarata transcriptome was sequenced using PacBio sequencing and several DEGs related to abiotic stress at various organs/tissues were obtained. This study identified 30 polyphenol oxidase (PPO) encoding genes, and most of them were proposed to be involved in biotic or abiotic stresses responses [177].
The cost effectiveness and high-throughput nature of TGS has made it useful for whole genome sequencing of many under-explored crops. Rehmannia glutinosa, a medicinal crop was recently sequenced using ONT sequencing. The assembly genome is 2.49 Gb long with a scaffold N50 length of 70 Mb and high heterozygosity (2%). The newly generated reference genome sequence of R. glutinosa increases the genomic resources in the Lamiales order [178]. Cunninghamia lanceolata (China Fir) belongs to Gymnospermae, which are fast-growing and have desirable wood properties. However, in this species genes involved in stress regulation are little known. Direct RNA sequencing using nanopore technologies has revealed a total of 51 AP2/ERF, 29 NAC, and 37 WRKY transcription factors in C. lanceolata. The expression of most of the NAC and WRKY TFs increased under cold stress. These provided preliminary clues about genes involved in stress regulation in Cunninghamia [179]. This evidence suggests that TGS has not only simplified transcriptomics studies but also provide valuable and novel insight into stress responses in plants.

4. Comparative Analysis of Tools Employed in Genomics and Transcriptomics Studies

As discussed earlier, genomics and transcriptomics encompasses several tools and techniques, each with its own advantages and disadvantages. For instance, TILLING is used for precise and systematic identification of specific mutations in a target gene, making it a valuable tool for functional analysis. However, TILLING can be labor-intensive and may not be as high-throughput as desired, particularly for large-scale screening [180]. QTL mapping is a genetic analysis approach focused on identifying genomic regions associated with specific traits or quantitative character. Thus, it is instrumental in understanding the genetic basis of complex traits. However, QTL mapping has limited resolution and often relies on the presence of genetic markers, which can restrict its application in cases where markers are not readily available [181]. Comparative genomics, while invaluable for understanding gene function and evolution, may have limitations in comparing vastly different species or distant evolutionary relatives. The interpretation of observed similarities and differences can be complex and may not always provide straightforward answers about gene function or regulatory elements (Tam et al., 2019). Gene inactivation methods, such as CRISPR-Cas9, ZFNs, and TALENs, are highly precise in gene editing and inactivation, but they may require complex design and validation procedures. Also, there are potential off-target effects, where unintended and untargeted genetic modifications or disruptions can occur, necessitating careful scrutiny and validation of the edited genes [182].
Microarray analysis enables simultaneous analysis of gene expression patterns, making it useful for studying gene regulation on a large scale. However, it is limited to known probes, and there is a risk of cross-hybridization between closely related sequences [183]. ESTs are valuable for gene expression analysis, identifying, and cataloging genes. Nevertheless, they provide only partial gene information, which can be limited in comprehensive studies [184]. RNA-Seq, on the other hand, offers high-throughput transcriptome analysis and is exceptionally valuable for understanding gene expression patterns. However, it requires complex data analysis and can be costly [185]. RT-qPCR is highly sensitive in gene expression analysis, but it necessitates prior knowledge of the target sequence, making it most suitable for studying known genes [186]. When selecting a specific tool for research, one should take into account the scale of their project, including both the cost implications and the number of genes they intend to target. A comprehensive list of tools and techniques involved in genomics and transcriptomics studies, with their year of origin, in chronological order, along with their advantages and disadvantages are given in Table 5.

5. Conclusions and Future Prospects

In conclusion, the synergy between genomics and transcriptomics has ushered in a new era of understanding and enhancing crop responses to abiotic stress conditions. Genomics, encompassing functional, structural, and comparative genomics, empowers researchers with a suite of powerful tools. Functional genomics techniques, including RNAi, VIGS, and genome editing with CRISPR-Cas9, enable precise gene manipulation and identification of pivotal stress-responsive genes. Structural genomics contributes to the development of molecular markers, QTLs, and marker-assisted breeding, while comparative genomics unveils conserved genes and pathways across species. These insights facilitate the selection of candidate genes for crop improvement, thereby expediting the breeding of stress-tolerant varieties crucial for sustainable agriculture in the face of climate change. Transcriptomics techniques, on the other hand, have revolutionized our comprehension of gene expression regulation under abiotic stress. Traditional methods like northern blotting and microarray paved the way for advanced approaches like RNA-Seq and third generation sequencing allowing simultaneous analysis of thousands of genes and the detection of both known and novel transcripts. These modern techniques have uncovered a plethora of abiotic stress-responsive genes, shedding light on the intricate mechanisms underlying stress tolerance. The wealth of information generated through transcriptomics provides invaluable insights into potential targets for crop improvement. However, there has been substantial progress, but certain crucial findings are still lacking. While numerous stress-responsive genes have been identified, the critical functional characterization of these genes remains a significant challenge, hindering our ability to precisely target crop improvement efforts. Most studies have focused on identifying major genes or pathways involved in stress responses. In-depth investigation is needed to understand the quantitative nature of stress tolerance traits and how multiple genes with small effects contribute to overall tolerance. Additionally, understanding the role of epigenetic modifications and non-coding RNAs in stress responses is an emerging area, warranting further investigation. Addressing these research gaps, along with considerations of environmental variability and the validation of findings in real-world field trials, is crucial for realizing the full potential of genomics and transcriptomics in enhancing crop resilience and food security.

Author Contributions

Conceptualization, A.S.; methodology, S.K., formal analysis, S.K. and A.S.; resources, A.S.; data curation, S.K.; writing—S.K.; writing—review and editing, A.S. and S.K.; supervision, A.S.; project administration, A.S.; funding acquisition, A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No data is associated with this article.

Acknowledgments

Authors acknowledge financial support from the NIPGR core research grant and from the DBT-BUILDER project (DBT-BUILDER BT/INF/22/SP45382/2022) of the School of Life Sciences, JNU. Research in the AS lab is supported by project funds from the Science & Engineering Research Board (SERB), India (Grant No.: CRG/2021/000694 and EEQ/2022/000062). SK acknowledges a research fellowship from the Council of Scientific and Industrial Research (CSIR), India.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Food and Agriculture Organization. World Food and Agriculture—Statistical Yearbook 2022; Food & Agriculture Organization: Rome, Italy, 2022. [Google Scholar]
  2. Rivero, R.M.; Mittler, R.; Blumwald, E.; Zandalinas, S.I. Developing climate-resilient crops: Improving plant tolerance to stress combination. Plant J. 2021, 109, 373–389. [Google Scholar] [CrossRef] [PubMed]
  3. Dietz, K.-J.; Vogelsang, L. A general concept of quantitative abiotic stress sensing. Trends Plant Sci. 2023, in press, corrected proof. [Google Scholar] [CrossRef] [PubMed]
  4. Cole, M.B.; Augustin, M.A.; Robertson, M.J.; Manners, J.M. The science of food security. Npj Sci. Food 2018, 2, 14. [Google Scholar] [CrossRef] [PubMed]
  5. Gao, C. Genome engineering for crop improvement and future agriculture. Cell 2021, 184, 1621–1635. [Google Scholar] [CrossRef] [PubMed]
  6. Borlaug, N.E. The Green Revolution Revisited and the Road Ahead; Nobelprize.org: Stockholm, Sweden, 2002. [Google Scholar]
  7. Yang, Y.; Saand, M.A.; Huang, L.; Abdelaal, W.B.; Zhang, J.; Wu, Y.; Li, J.; Sirohi, M.H.; Wang, F. Applications of multi-omics technologies for crop improvement. Front. Plant Sci. 2021, 12, 563953. [Google Scholar] [CrossRef]
  8. Bustamante, C.D.; De La Vega, F.M.; Burchard, E.G. Genomics for the world. Nature 2011, 475, 163–165. [Google Scholar] [CrossRef]
  9. Akpınar, B.A.; Lucas, S.J.; Budak, H. Genomics approaches for crop improvement against abiotic stress. Sci. World J. 2013, 2013, 361921. [Google Scholar] [CrossRef]
  10. Joshi, R.; Gupta, B.K.; Pareek, A.; Singh, M.B.; Singla-Pareek, S.L. Functional genomics approach towards dissecting out abiotic stress tolerance trait in plants. In Sustainable Development and Biodiversity; Springer International Publishing: Berlin/Heidelberg, Germany, 2019; pp. 1–24. [Google Scholar] [CrossRef]
  11. Rashid, B.; Husnain, T.; Riazuddin, S. Genomic approaches and abiotic stress tolerance in plants. In Emerging Technologies and Management of Crop Stress Tolerance; Elsevier: Amsterdam, The Netherlands, 2014; pp. 1–37. [Google Scholar] [CrossRef]
  12. Xu, H.-R.; Liu, Y.; Yu, T.-F.; Hou, Z.-H.; Zheng, J.-C.; Chen, J.; Zhou, Y.-B.; Chen, M.; Fu, J.-D.; Ma, Y.-Z.; et al. Comprehensive profiling of tubby-like proteins in soybean and roles of the gmtlp8 gene in abiotic stress responses. Front. Plant Sci. 2022, 13, 844545. [Google Scholar] [CrossRef]
  13. Zhang, D.; Zhang, Z.; Unver, T.; Zhang, B. CRISPR/Cas: A powerful tool for gene function study and crop improvement. J. Adv. Res. 2021, 29, 207–221. [Google Scholar] [CrossRef]
  14. Gao, Y.; Han, D.; Jia, W.; Ma, X.; Yang, Y.; Xu, Z. Molecular characterization and systematic analysis of NtAP2/ERF in tobacco and functional determination of NtRAV-4 under drought stress. Plant Physiol. Biochem. 2020, 156, 420–435. [Google Scholar] [CrossRef]
  15. Raffeiner, M.; Üstün, S.; Guerra, T.; Spinti, D.; Fitzner, M.; Sonnewald, S.; Baldermann, S.; Börnke, F. The Xanthomonas type-III effector XopS stabilizes CaWRKY40a to regulate defense responses and stomatal immunity in pepper (Capsicum annuum). Plant Cell 2022, 34, 1684–1708. [Google Scholar] [CrossRef]
  16. Hen-Avivi, S.; Savin, O.; Racovita, R.C.; Lee, W.-S.; Adamski, N.M.; Malitsky, S.; Almekias-Siegl, E.; Levy, M.; Vautrin, S.; Bergès, H.; et al. A Metabolic Gene Cluster in the Wheat W1 and the Barley Cer-cqu Loci Determines β-Diketone Biosynthesis and Glaucousness. Plant Cell 2016, 28, 1440–1460. [Google Scholar] [CrossRef] [PubMed]
  17. Zhou, J.; Deng, K.; Cheng, Y.; Zhong, Z.; Tian, L.; Tang, X.; Tang, A.; Zheng, X.; Zhang, T.; Qi, Y.; et al. CRISPR-Cas9 based genome editing reveals new insights into microrna function and regulation in rice. Front. Plant Sci. 2017, 8, 1598. [Google Scholar] [CrossRef] [PubMed]
  18. Senthil-Kumar, M.; Mysore, K.S. Caveat of rnai in plants: The off-target effect. In Methods in Molecular Biology; Humana Press: Totowa, NJ, USA, 2011; pp. 13–25. [Google Scholar] [CrossRef]
  19. Krasnodębski, C.; Sawuła, A.; Kaźmierczak, U.; Żuk, M. Oligo—Not only for silencing: Overlooked potential for multidirectional action in plants. Int. J. Mol. Sci. 2023, 24, 4466. [Google Scholar] [CrossRef] [PubMed]
  20. Liu, R.; Long, Q.; Zou, X.; Wang, Y.; Pei, Y. DNA methylation occurring in Cre-expressing cells inhibits loxP recombination and silences loxP-sandwiched genes. New Phytol. 2021, 231, 210–224. [Google Scholar] [CrossRef] [PubMed]
  21. Anand, A.; Wu, E.; Li, Z.; TeRonde, S.; Arling, M.; Lenderts, B.; Mutti, J.S.; Gordon-Kamm, W.; Jones, T.J.; Chilcoat, N.D. High efficiency Agrobacterium-mediated site-specific gene integration in maize utilizing the FLP-FRT recombination system. Plant Biotechnol. J. 2019, 17, 1636–1645. [Google Scholar] [CrossRef] [PubMed]
  22. Ran, Y.; Patron, N.; Kay, P.; Wong, D.; Buchanan, M.; Cao, Y.; Sawbridge, T.; Davies, J.P.; Mason, J.; Webb, S.R.; et al. Zinc finger nuclease-mediated precision genome editing of an endogenous gene in hexaploid bread wheat (Triticum aestivum) using a DNA repair template. Plant Biotechnol. J. 2018, 16, 2088–2101. [Google Scholar] [CrossRef]
  23. Zhou, S.; Jiang, W.; Long, F.; Cheng, S.; Yang, W.; Zhao, Y.; Zhou, D.-X. Rice homeodomain protein WOX11 recruits a histone acetyltransferase complex to establish programs of cell proliferation of crown root meristem. Plant Cell 2017, 29, 1088–1104. [Google Scholar] [CrossRef]
  24. Zhang, A.; Liu, Y.; Wang, F.; Li, T.; Chen, Z.; Kong, D.; Bi, J.; Zhang, F.; Luo, X.; Wang, J.; et al. Enhanced rice salinity tolerance via CRISPR/Cas9-targeted mutagenesis of the OsRR22 gene. Mol. Breed. 2019, 39, 47. [Google Scholar] [CrossRef]
  25. Santosh Kumar, V.V.; Verma, R.K.; Yadav, S.K.; Yadav, P.; Watts, A.; Rao, M.V.; Chinnusamy, V. CRISPR-Cas9 mediated genome editing of drought and salt tolerance (OsDST) gene in indica mega rice cultivar MTU1010. Physiol. Mol. Biol. Plants 2020, 26, 1099–1110. [Google Scholar] [CrossRef]
  26. Tran, M.T.; Son, G.H.; Song, Y.J.; Nguyen, N.T.; Park, S.; Thach, T.V.; Kim, J.; Sung, Y.W.; Das, S.; Pramanik, D.; et al. CRISPR-Cas9-based precise engineering of SlHyPRP1 protein towards multi-stress tolerance in tomato. Front. Plant Sci. 2023, 14, 1186932. [Google Scholar] [CrossRef]
  27. Jain, M.; Garg, R. Enhancers as potential targets for engineering salinity stress tolerance in crop plants. Physiol. Plant. 2021, 173, 1382–1391. [Google Scholar] [CrossRef] [PubMed]
  28. Li, R.; Liu, C.; Zhao, R.; Wang, L.; Chen, L.; Yu, W.; Zhang, S.; Sheng, J.; Shen, L. CRISPR/Cas9-Mediated SlNPR1 mutagenesis reduces tomato plant drought tolerance. BMC Plant Biol. 2019, 19, 38. [Google Scholar] [CrossRef] [PubMed]
  29. Casella, L.; Greco, R.; Bruschi, G.; Wozniak, B.; Dreni, L.; Kater, M.; Cavigiolo, S.; Lupotto, E.; Piffanelli, P. TILLING in European rice: Hunting mutations for crop improvement. Crop Sci. 2013, 53, 2550–2562. [Google Scholar] [CrossRef]
  30. Negrão, S.; Almadanim, C.; Pires, I.; McNally, K.L.; Oliveira, M.M. Use of EcoTILLING to identify natural allelic variants of rice candidate genes involved in salinity tolerance. Plant Genet. Resour. 2011, 9, 300–304. [Google Scholar] [CrossRef]
  31. Hwang, J.E.; Jang, D.-S.; Lee, K.J.; Ahn, J.-W.; Kim, S.H.; Kang, S.-Y.; Kim, D.S.; Kim, J.-B. Identification of gamma ray irradiation-induced mutations in membrane transport genes in a rice population by TILLING. Genes Genet. Syst. 2016, 91, 245–256. [Google Scholar] [CrossRef]
  32. Cho, H.Y.; Park, S.J.; Kim, D.S.; Jang, C.S. A TILLING Rice Population Induced by Gamma-ray Irradiation and its Genetic Diversity. Korean J. Breed. Sci. 2010, 42, 365–373. [Google Scholar]
  33. Navarro-León, E.; Grazioso, A.; Atero-Calvo, S.; Rios, J.J.; Esposito, S.; Blasco, B. Evaluation of the alkalinity stress tolerance of three Brassica rapa CAX1 TILLING mutants. Plant Physiol. Biochem. 2023, 198, 107712. [Google Scholar] [CrossRef]
  34. Marko, D.; El-shershaby, A.; Carriero, F.; Summerer, S.; Petrozza, A.; Iannacone, R.; Schleiff, E.; Fragkostefanakis, S. Identification and characterization of a thermotolerant TILLING allele of heat shock binding protein 1 in tomato. Genes 2019, 10, 516. [Google Scholar] [CrossRef]
  35. Minkenberg, B.; Zhang, J.; Xie, K.; Yang, Y. CRISPR-PLANT v2: An online resource for highly specific guide RNA spacers based on improved off-target analysis. Plant Biotechnol. J. 2018, 17, 5–8. [Google Scholar] [CrossRef]
  36. Park, J.; Bae, S.; Kim, J.S. Cas-Designer: A web-based tool for choice of CRISPR-Cas9 target sites. Bioinformatics 2015, 31, 4014–4016. [Google Scholar] [CrossRef]
  37. Hough, S.H.; Ajetunmobi, A.; Brody, L.; Humphryes-Kirilov, N.; Perello, E. Desktop genetics. Pers. Med. 2016, 13, 517–521. [Google Scholar] [CrossRef]
  38. Heigwer, F.; Kerr, G.; Boutros, M. E-CRISP: Fast CRISPR target site identification. Nat. Methods 2014, 11, 122–123. [Google Scholar] [CrossRef]
  39. Perez, A.R.; Pritykin, Y.; Vidigal, J.A.; Chhangawala, S.; Zamparo, L.; Leslie, C.S.; Ventura, A. GuideScan software for improved single and paired CRISPR guide RNA design. Nat. Biotechnol. 2017, 35, 347–349. [Google Scholar] [CrossRef]
  40. Sander, J.D.; Zaback, P.; Joung, J.K.; Voytas, D.F.; Dobbs, D. Zinc Finger Targeter (ZiFiT): An engineered zinc finger/target site design tool. Nucleic Acids Res. 2007, 35, W599–W605. [Google Scholar] [CrossRef]
  41. Labun, K.; Montague, T.G.; Krause, M.; Torres Cleuren, Y.N.; Tjeldnes, H.; Valen, E. CHOPCHOP v3: Expanding the CRISPR web toolbox beyond genome editing. Nucleic Acids Res. 2019, 47, W171–W174. [Google Scholar] [CrossRef]
  42. Concordet, J.-P.; Haeussler, M. CRISPOR: Intuitive guide selection for CRISPR/Cas9 genome editing experiments and screens. Nucleic Acids Res. 2018, 46, W242–W245. [Google Scholar] [CrossRef] [PubMed]
  43. Liu, H.; Wei, Z.; Dominguez, A.; Li, Y.; Wang, X.; Qi, L.S. CRISPR-ERA: A comprehensive design tool for CRISPR-mediated gene editing, repression and activation. Bioinformatics 2015, 31, 3676–3678. [Google Scholar] [CrossRef] [PubMed]
  44. Liu, H.; Ding, Y.; Zhou, Y.; Jin, W.; Xie, K.; Chen, L.-L. CRISPR-P 2.0: An Improved CRISPR-Cas9 Tool for Genome Editing in Plants. Mol. Plant 2017, 10, 530–532. [Google Scholar] [CrossRef] [PubMed]
  45. Bae, S.; Park, J.; Kim, J.-S. Cas-OFFinder: A fast and versatile algorithm that searches for potential off-target sites of Cas9 RNA-guided endonucleases. Bioinformatics 2014, 30, 1473–1475. [Google Scholar] [CrossRef] [PubMed]
  46. Upadhyay, S.K.; Sharma, S. SSFinder: High throughput crispr-cas target sites prediction tool. BioMed Res. Int. 2014, 2014, 742482. [Google Scholar] [CrossRef] [PubMed]
  47. Castro-Mondragon, J.A.; Riudavets-Puig, R.; Rauluseviciute, I.; Berhanu Lemma, R.; Turchi, L.; Blanc-Mathieu, R.; Lucas, J.; Boddie, P.; Khan, A.; Manosalva Pérez, N.; et al. JASPAR 2022: The 9th release of the open-access database of transcription factor binding profiles. Nucleic Acids Res. 2021, 50, D165–D173. [Google Scholar] [CrossRef] [PubMed]
  48. Yoon, D.H.; Lee, S.S.; Park, H.J.; Lyu, J.I.; Chong, W.S.; Liu, J.R.; Kim, B.-G.; Ahn, J.C.; Cho, H.S. Overexpression of OsCYP19-4increases tolerance to cold stress and enhances grain yield in rice (Oryza sativa). J. Exp. Bot. 2016, 67, 69–82. [Google Scholar] [CrossRef] [PubMed]
  49. Habib, S.; Waseem, M.; Li, N.; Yang, L.; Li, Z. Overexpression of slgras7 affects multiple behaviors leading to confer abiotic stresses tolerance and impacts gibberellin and auxin signaling in tomato. Int. J. Genom. 2019, 2019, 4051981. [Google Scholar] [CrossRef] [PubMed]
  50. Xu, L.; Zhang, L.; Liu, Y.; Sod, B.; Li, M.; Yang, T.; Gao, T.; Yang, Q.; Long, R. Overexpression of the elongation factor MtEF1A1 promotes salt stress tolerance in Arabidopsis thaliana and Medicago truncatula. BMC Plant Biol. 2023, 23, 138. [Google Scholar] [CrossRef] [PubMed]
  51. Zhang, X.; Lin, X.; Chen, S.; Chen, S. Overexpression of BpERF1.1 in Betula Platyphylla enhanced tolerance to multiple abiotic stresses. Physiol. Mol. Biol. Plants 2022, 28, 1159–1172. [Google Scholar] [CrossRef] [PubMed]
  52. Verma, R.K.; Kumar, V.V.S.; Yadav, S.K.; Kumar, T.S.; Rao, M.V.; Chinnusamy, V. Overexpression of Arabidopsis ICE1 enhances yield and multiple abiotic stress tolerance in indica rice. Plant Signal. Behav. 2020, 15, 1814547. [Google Scholar] [CrossRef]
  53. Singh, A.; Jha, S.K.; Bagri, J.; Pandey, G.K. ABA inducible rice protein phosphatase 2C confers ABA insensitivity and abiotic stress tolerance in Arabidopsis. PLoS ONE 2015, 10, e0125168. [Google Scholar] [CrossRef]
  54. Al-Deeb, T.; Abo Gamar, M.; El-Assi, N.; Al-Debei, H.; Al-Sayaydeh, R.; Al-Abdallat, A.M. Stress-Inducible overexpression of slddf2 gene improves tolerance against multiple abiotic stresses in tomato plant. Horticulturae 2022, 8, 230. [Google Scholar] [CrossRef]
  55. Yang, X.; Kim, M.Y.; Ha, J.; Lee, S.-H. Overexpression of the soybean NAC gene gmnac109 increases lateral root formation and abiotic stress tolerance in transgenic Arabidopsis plants. Front. Plant Sci. 2019, 10, 1036. [Google Scholar] [CrossRef]
  56. Patel, P.; Yadav, K.; Srivastava, A.K.; Suprasanna, P.; Ganapathi, T.R. Overexpression of native Musa-miR397 enhances plant biomass without compromising abiotic stress tolerance in banana. Sci. Rep. 2019, 9, 16434. [Google Scholar] [CrossRef]
  57. Wytynck, P.; Lambin, J.; Chen, S.; Demirel Asci, S.; Verbeke, I.; De Zaeytijd, J.; Subramanyam, K.; Van Damme, E.J.M. Effect of RIP overexpression on abiotic stress tolerance and development of rice. Int. J. Mol. Sci. 2021, 22, 1434. [Google Scholar] [CrossRef]
  58. Lu, T.; Fan, Z.; He, X.; Yu, H.; Liang, R.; Huang, X.; Zhang, Y.; Zhu, J.; Wang, J.; Luo, C. Overexpression of mango MiMFT inhibits seed germination and enhances abiotic stress tolerance in transgenic Arabidopsis. Sci. Hortic. 2023, 307, 111495. [Google Scholar] [CrossRef]
  59. Chang, Y.; Song, X.; Li, M.; Zhang, Q.; Zhang, P.; Lei, X.; Pei, D. Characterization of walnut JrWOX11 and its overexpression provide insights into adventitious root formation and development and abiotic stress tolerance. Front. Plant Sci. 2022, 13, 951737. [Google Scholar] [CrossRef]
  60. Wan, J.; Zhang, J.; Zan, X.; Zhu, J.; Chen, H.; Li, X.; Zhou, Z.; Gao, X.; Chen, R.; Huang, Z.; et al. Overexpression of rice histone H1 gene reduces tolerance to cold and heat stress. Plants 2023, 12, 2408. [Google Scholar] [CrossRef]
  61. Liu, Y.; Li, A.; Liang, M.; Zhang, Q.; Wu, J. Overexpression of the maize genes ZmSKL1 and ZmSKL2 positively regulates drought stress tolerance in transgenic Arabidopsis. Plant Cell Rep. 2022, 42, 521–533. [Google Scholar] [CrossRef]
  62. Chen, X.; Jiang, X.; Niu, F.; Sun, X.; Hu, Z.; Gao, F.; Zhang, H.; Jiang, Q. Overexpression of lncRNA77580 Regulates Drought and Salinity Stress Responses in Soybean. Plants 2023, 12, 181. [Google Scholar] [CrossRef] [PubMed]
  63. Kilwake, J.W.; Umer, M.J.; Wei, Y.; Mehari, T.G.; Magwanga, R.O.; Xu, Y.; Hou, Y.; Wang, Y.; Shiraku, M.L.; Kirungu, J.N.; et al. Genome-Wide characterization of the SAMS gene family in cotton unveils the putative role of ghsams2 in enhancing abiotic stress tolerance. Agronomy 2023, 13, 612. [Google Scholar] [CrossRef]
  64. He, Y.; Zhang, X.; Tan, Y.; Si, D.; Zhao, T.; Xu, X.; Jiang, J.; Yang, H.; Li, J. Virus-Induced gene silencing of slwrky79 attenuates salt tolerance in tomato plants. Agronomy 2021, 11, 1519. [Google Scholar] [CrossRef]
  65. Liu, T.; Chen, T.; Kan, J.; Yao, Y.; Guo, D.; Yang, Y.; Ling, X.; Wang, J.; Zhang, B. The GhMYB36 transcription factor confers resistance to biotic and abiotic stress by enhancing PR1 gene expression in plants. Plant Biotechnol. J. 2021, 20, 722–735. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, T.; Xun, H.; Wang, W.; Ding, X.; Tian, H.; Hussain, S.; Dong, Q.; Li, Y.; Cheng, Y.; Wang, C.; et al. Mutation of gmaitr genes by crispr/cas9 genome editing results in enhanced salinity stress tolerance in soybean. Front. Plant Sci. 2021, 12, 779598. [Google Scholar] [CrossRef] [PubMed]
  67. Wei, W.; Wang, H.; Liu, X.; Kou, W.; Liu, Z.; Wang, H.; Yang, Y.; Zhao, L.; Zhang, H.; Liu, B.; et al. Transcriptome Profiling of Stem-Differentiating Xylem in Response to Abiotic Stresses Based on Hybrid Sequencing in Cunninghamia lanceolata. Int. J. Mol. Sci. 2022, 23, 13986. [Google Scholar] [CrossRef]
  68. Yue, E.; Cao, H.; Liu, B. OsmiR535, a Potential Genetic Editing Target for Drought and Salinity Stress Tolerance in Oryza sativa. Plants 2020, 9, 1337. [Google Scholar] [CrossRef] [PubMed]
  69. Kim, M.-S.; Ko, S.-R.; Jung, Y.J.; Kang, K.-K.; Lee, Y.-J.; Cho, Y.-G. Knockout mutants of ospub7 generated using crispr/cas9 revealed abiotic stress tolerance in rice. Int. J. Mol. Sci. 2023, 24, 5338. [Google Scholar] [CrossRef]
  70. Yan, X.; Guan, Y.; Liu, X.; Yu, J.; Lei, B.; Wang, Z.; Zhang, H.; Cui, H. NtCycB2 gene knockout enhances resistance to high salinity stress in Nicotiana tabacum. Ind. Crops Prod. 2021, 171, 113886. [Google Scholar] [CrossRef]
  71. Habib, S.; Lwin, Y.Y.; Li, N. Down-Regulation of slgras10 in tomato confers abiotic stress tolerance. Genes 2021, 12, 623. [Google Scholar] [CrossRef] [PubMed]
  72. Zhang, A.; Yang, X.; Lu, J.; Song, F.; Sun, J.; Wang, C.; Lian, J.; Zhao, L.; Zhao, B. OsIAA20, an Aux/IAA protein, mediates abiotic stress tolerance in rice through an ABA pathway. Plant Sci. 2021, 308, 110903. [Google Scholar] [CrossRef]
  73. Lv, A.; Su, L.; Wen, W.; Fan, N.; Zhou, P.; An, Y. Analysis of the function of the alfalfa mslea-d34 gene in abiotic stress responses and flowering time. Plant Cell Physiol. 2020, 62, 28–42. [Google Scholar] [CrossRef]
  74. Choudhary, M.; Wani, S.H.; Kumar, P.; Bagaria, P.K.; Rakshit, S.; Roorkiwal, M.; Varshney, R.K. QTLian breeding for climate resilience in cereals: Progress and prospects. Funct. Integr. Genom. 2019, 19, 685–701. [Google Scholar] [CrossRef]
  75. Nogoy, F.M.; Song, J.-Y.; Ouk, S.; Rahimi, S.; Kwon, S.W.; Kang, K.-K.; Cho, Y.-G. Current Applicable DNA Markers for Marker Assisted Breeding in Abiotic and Biotic Stress Tolerance in Rice (Oryza sativa L.). Plant Breed. Biotechnol. 2016, 4, 271–284. [Google Scholar] [CrossRef]
  76. Lei, L.; Zheng, H.; Bi, Y.; Yang, L.; Liu, H.; Wang, J.; Sun, J.; Zhao, H.; Li, X.; Li, J.; et al. Identification of a Major QTL and Candidate Gene Analysis of Salt Tolerance at the Bud Burst Stage in Rice (Oryza sativa L.) Using QTL-Seq and RNA-Seq. Rice 2020, 13, 55. [Google Scholar] [CrossRef]
  77. Frouin, J.; Languillaume, A.; Mas, J.; Mieulet, D.; Boisnard, A.; Labeyrie, A.; Bettembourg, M.; Bureau, C.; Lorenzini, E.; Portefaix, M.; et al. Tolerance to mild salinity stress in japonica rice: A genome-wide association mapping study highlights calcium signaling and metabolism genes. PLoS ONE 2018, 13, e0190964. [Google Scholar] [CrossRef]
  78. Sun, J.; Wang, J.; Guo, W.; Yin, T.; Zhang, S.; Wang, L.; Xie, D.; Zou, D. Identification of alkali-tolerant candidate genes using the NGS-assisted BSA strategy in rice. Mol. Breed. 2021, 41, 44. [Google Scholar] [CrossRef]
  79. Muthu, V.; Abbai, R.; Nallathambi, J.; Rahman, H.; Ramasamy, S.; Kambale, R.; Thulasinathan, T.; Ayyenar, B.; Muthurajan, R. Pyramiding QTLs controlling tolerance against drought, salinity, and submergence in rice through marker assisted breeding. PLoS ONE 2020, 15, e0227421. [Google Scholar] [CrossRef]
  80. Sheoran, S.; Gupta, M.; Kumari, S.; Kumar, S.; Rakshit, S. Meta-QTL analysis and candidate genes identification for various abiotic stresses in maize (Zea mays L.) and their implications in breeding programs. Mol. Breed. 2022, 42, 26. [Google Scholar] [CrossRef]
  81. Shi, H.; Wu, Y.; Yi, L.; Hu, H.; Su, F.; Wang, Y.; Li, D.; Hou, J. Analysis of QTL mapping for germination and seedling response to drought stress in sunflower (Helianthus annuus L.). Plant Biol. 2023, 11, e15275. [Google Scholar] [CrossRef]
  82. Tello-Ruiz, M.K.; Naithani, S.; Stein, J.C.; Gupta, P.; Campbell, M.; Olson, A.; Wei, S.; Preece, J.; Geniza, M.J.; Jiao, Y.; et al. Gramene 2018: Unifying comparative genomics and pathway resources for plant research. Nucleic Acids Res. 2017, 46, D1181–D1189. [Google Scholar] [CrossRef] [PubMed]
  83. Grant, D.; Nelson, R.T.; Cannon, S.B.; Shoemaker, R.C. SoyBase, the USDA-ARS soybean genetics and genomics database. Nucleic Acids Res. 2009, 38, D843–D846. [Google Scholar] [CrossRef]
  84. Yu, J.; Jung, S.; Cheng, C.-H.; Lee, T.; Zheng, P.; Buble, K.; Crabb, J.; Humann, J.; Hough, H.; Jones, D.; et al. CottonGen: The community database for cotton genomics, genetics, and breeding research. Plants 2021, 10, 2805. [Google Scholar] [CrossRef] [PubMed]
  85. Fernandez-Pozo, N.; Menda, N.; Edwards, J.D.; Saha, S.; Tecle, I.Y.; Strickler, S.R.; Bombarely, A.; Fisher-York, T.; Pujar, A.; Foerster, H.; et al. The Sol Genomics Network (SGN)—From genotype to phenotype to breeding. Nucleic Acids Res. 2014, 43, D1036–D1041. [Google Scholar] [CrossRef] [PubMed]
  86. Woodhouse, M.R.; Cannon, E.K.; Portwood, J.L., II; Harper, L.C.; Gardiner, J.M.; Schaeffer, M.L.; Andorf, C.M. A pan-genomic approach to genome databases using maize as a model system. BMC Plant Biol. 2021, 21, 385. [Google Scholar] [CrossRef] [PubMed]
  87. Jahnke, G.; Smidla, J.; Poczai, P. MolMarker: A simple tool for DNA fingerprinting studies and polymorphic information content calculation. Diversity 2022, 14, 497. [Google Scholar] [CrossRef]
  88. Blake, V.C.; Birkett, C.; Matthews, D.E.; Hane, D.L.; Bradbury, P.; Jannink, J. The triticeae toolbox: Combining phenotype and genotype data to advance small-grains breeding. Plant Genome 2016, 9, plantgenome2014.12.0099. [Google Scholar] [CrossRef] [PubMed]
  89. Zheng, Y.; Wu, S.; Bai, Y.; Sun, H.; Jiao, C.; Guo, S.; Zhao, K.; Blanca, J.; Zhang, Z.; Huang, S.; et al. Cucurbit Genomics Database (CuGenDB): A central portal for comparative and functional genomics of cucurbit crops. Nucleic Acids Res. 2018, 47, D1128–D1136. [Google Scholar] [CrossRef] [PubMed]
  90. Meng, L.; Li, H.; Zhang, L.; Wang, J. QTL IciMapping: Integrated software for genetic linkage map construction and quantitative trait locus mapping in biparental populations. Crop J. 2015, 3, 269–283. [Google Scholar] [CrossRef]
  91. Broman, K.W.; Wu, H.; Sen, Ś.; Churchill, G.A. R/qtl: QTL mapping in experimental crosses. Bioinformatics 2003, 19, 889–890. [Google Scholar] [CrossRef]
  92. Zhu, T.; Liang, C.; Meng, Z.; Sun, G.; Meng, Z.; Guo, S.; Zhang, R. CottonFGD: An integrated functional genomics database for cotton. BMC Plant Biol. 2017, 17, 101. [Google Scholar] [CrossRef]
  93. Zeng, H.; Luo, L.; Zhang, W.; Zhou, J.; Li, Z.; Liu, H.; Zhu, T.; Feng, X.; Zhong, Y. PlantQTL-GE: A database system for identifying candidate genes in rice and Arabidopsis by gene expression and QTL information. Nucleic Acids Res. 2007, 35, D879–D882. [Google Scholar] [CrossRef]
  94. Dash, S.; Cannon, E.K.S.; Kalberer, S.R.; Farmer, A.D.; Cannon, S.B. PeanutBase and other bioinformatic resources for peanut. In Peanuts; Elsevier: Amsterdam, The Netherlands, 2016; pp. 241–252. [Google Scholar] [CrossRef]
  95. Steinbach, D.; Alaux, M.; Amselem, J.; Choisne, N.; Durand, S.; Flores, R.; Keliet, A.-O.; Kimmel, E.; Lapalu, N.; Luyten, I.; et al. GnpIS: An information system to integrate genetic and genomic data from plants and fungi. Database 2013, 2013, bat058. [Google Scholar] [CrossRef]
  96. Alaux, M.; Rogers, J.; Letellier, T.; Flores, R.; Alfama, F.; Pommier, C.; Mohellibi, N.; Durand, S.; Kimmel, E.; Michotey, C.; et al. Linking the International Wheat Genome Sequencing Consortium bread wheat reference genome sequence to wheat genetic and phenomic data. Genome Biol. 2018, 19, 111. [Google Scholar] [CrossRef]
  97. Wei, L.; Liu, Y.; Dubchak, I.; Shon, J.; Park, J. Comparative genomics approaches to study organism similarities and differences. J. Biomed. Inform. 2002, 35, 142–150. [Google Scholar] [CrossRef]
  98. Haubold, B.; Wiehe, T. Comparative genomics: Methods and applications. Naturwissenschaften 2004, 91, 405–421. [Google Scholar] [CrossRef] [PubMed]
  99. Howe, K.L.; Achuthan, P.; Allen, J.; Allen, J.; Alvarez-Jarreta, J.; Amode, M.R.; Armean, I.M.; Azov, A.G.; Bennett, R.; Bhai, J.; et al. Ensembl 2021. Nucleic Acids Res. 2020, 49, D884–D891. [Google Scholar] [CrossRef] [PubMed]
  100. Flicek, P.; Amode, M.R.; Barrell, D.; Beal, K.; Billis, K.; Brent, S.; Carvalho-Silva, D.; Clapham, P.; Coates, G.; Fitzgerald, S.; et al. Ensembl 2014. Nucleic Acids Res. 2014, 42, D749–D755. [Google Scholar] [CrossRef] [PubMed]
  101. Haeussler, M.; Zweig, A.S.; Tyner, C.; Speir, M.L.; Rosenbloom, K.R.; Raney, B.J.; Lee, C.M.; Lee, B.T.; Hinrichs, A.S.; Gonzalez, J.N.; et al. The UCSC Genome Browser database: 2019 update. Nucleic Acids Res. 2018, 47, D853–D858. [Google Scholar] [CrossRef] [PubMed]
  102. Kent, W.J.; Sugnet, C.W.; Furey, T.S.; Roskin, K.M.; Pringle, T.H.; Zahler, A.M.; Haussler, D. The human genome browser at UCSC. Genome Res. 2002, 12, 996–1006. [Google Scholar] [CrossRef] [PubMed]
  103. Kriventseva, E.V.; Kuznetsov, D.; Tegenfeldt, F.; Manni, M.; Dias, R.; Simão, F.A.; Zdobnov, E.M. OrthoDB v10: Sampling the diversity of animal, plant, fungal, protist, bacterial and viral genomes for evolutionary and functional annotations of orthologs. Nucleic Acids Res. 2018, 47, D807–D811. [Google Scholar] [CrossRef]
  104. Chen, G.; Zou, Y.; Hu, J.; Ding, Y. Genome-wide analysis of the rice PPR gene family and their expression profiles under different stress treatments. BMC Genom. 2018, 19, 720. [Google Scholar] [CrossRef]
  105. Ganie, S.A.; Pani, D.R.; Mondal, T.K. Genome-wide analysis of DUF221 domain-containing gene family in Oryza species and identification of its salinity stress-responsive members in rice. PLoS ONE 2017, 12, e0182469. [Google Scholar] [CrossRef]
  106. Singh, A.; Kanwar, P.; Yadav, A.K.; Mishra, M.; Jha, S.K.; Baranwal, V.; Pandey, A.; Kapoor, S.; Tyagi, A.K.; Pandey, G.K. Genome-wide expressional and functional analysis of calcium transport elements during abiotic stress and development in rice. FEBS J. 2014, 281, 894–915. [Google Scholar] [CrossRef]
  107. Singh, A.; Giri, J.; Kapoor, S.; Tyagi, A.K.; Pandey, G.K. Protein phosphatase complement in rice: Genome-wide identification and transcriptional analysis under abiotic stress conditions and reproductive development. BMC Genom. 2010, 11, 435. [Google Scholar] [CrossRef]
  108. Singh, A.; Pandey, G.K. Protein phosphatases: A genomic outlook to understand their function in plants. J. Plant Biochem. Biotechnol. 2012, 21, 100–107. [Google Scholar] [CrossRef]
  109. Singh, A.; Kanwar, P.; Pandey, A.; Tyagi, A.K.; Sopory, S.K.; Kapoor, S.; Pandey, G.K. Comprehensive genomic analysis and expression profiling of phospholipase C gene family during abiotic stresses and development in rice. PLoS ONE 2013, 8, e62494. [Google Scholar] [CrossRef]
  110. Ray, S.; Agarwal, P.; Arora, R.; Kapoor, S.; Tyagi, A.K. Expression analysis of calcium-dependent protein kinase gene family during reproductive development and abiotic stress conditions in rice (Oryza sativa L. ssp. indica). Mol. Genet. Genom. 2007, 278, 493–505. [Google Scholar] [CrossRef] [PubMed]
  111. Vij, S.; Giri, J.; Dansana, P.K.; Kapoor, S.; Tyagi, A.K. The receptor-like cytoplasmic kinase (osrlck) gene family in rice: Organization, phylogenetic relationship, and expression during development and stress. Mol. Plant 2008, 1, 732–750. [Google Scholar] [CrossRef] [PubMed]
  112. Arora, R.; Agarwal, P.; Ray, S.; Singh, A.K.; Singh, V.P.; Tyagi, A.K.; Kapoor, S. MADS-box gene family in rice: Genome-wide identification, organization and expression profiling during reproductive development and stress. BMC Genom. 2007, 8, 242. [Google Scholar] [CrossRef]
  113. Ma, H.; Zhao, H.; Liu, Z.; Zhao, J. The Phytocyanin Gene Family in Rice (Oryza sativa L.): Genome-Wide Identification, Classification and Transcriptional Analysis. PLoS ONE 2011, 6, e25184. [Google Scholar] [CrossRef]
  114. Ding, X.; Hou, X.; Xie, K.; Xiong, L. Genome-wide identification of BURP domain-containing genes in rice reveals a gene family with diverse structures and responses to abiotic stresses. Planta 2009, 230, 149–163. [Google Scholar] [CrossRef]
  115. Ma, H.; Zhao, J. Genome-wide identification, classification, and expression analysis of the arabinogalactan protein gene family in rice (Oryza sativa L.). J. Exp. Bot. 2010, 61, 2647–2668. [Google Scholar] [CrossRef]
  116. Yang, W.; Lu, Z.; Xiong, Y.; Yao, J. Genome-wide identification and co-expression network analysis of the OsNF-Y gene family in rice. Crop J. 2017, 5, 21–31. [Google Scholar] [CrossRef]
  117. Mishra, M.; Kanwar, P.; Singh, A.; Pandey, A.; Kapoor, S.; Pandey, G.K. Plant omics: Genome-Wide analysis of ABA repressor1 (ABR1) related genes in rice during abiotic stress and development. OMICS J. Integr. Biol. 2013, 17, 439–450. [Google Scholar] [CrossRef]
  118. Kim, S.-G.; Lee, S.; Seo, P.J.; Kim, S.-K.; Kim, J.-K.; Park, C.-M. Genome-scale screening and molecular characterization of membrane-bound transcription factors in Arabidopsis and rice. Genomics 2010, 95, 56–65. [Google Scholar] [CrossRef]
  119. Feng, C.; He, C.; Wang, Y.; Xu, H.; Xu, K.; Zhao, Y.; Yao, B.; Zhang, Y.; Zhao, Y.; Idrice Carther, K.F.; et al. Genome-wide identification of soybean Shaker K+ channel gene family and functional characterization of GmAKT1 in transgenic Arabidopsis thaliana under salt and drought stress. J. Plant Physiol. 2021, 266, 153529. [Google Scholar] [CrossRef]
  120. Almeida-Silva, F.; Venancio, T.M. Pathogenesis-related protein 1 (PR-1) genes in soybean: Genome-wide identification, structural analysis and expression profiling under multiple biotic and abiotic stresses. Gene 2022, 809, 146013. [Google Scholar] [CrossRef] [PubMed]
  121. Li, Z.; Liu, Z.; Wei, Y.; Liu, Y.; Xing, L.; Liu, M.; Li, P.; Lu, Q.; Peng, R. Genome-wide identification of the MIOX gene family and their expression profile in cotton development and response to abiotic stress. PLoS ONE 2021, 16, e0254111. [Google Scholar] [CrossRef]
  122. Sun, R.; Qin, T.; Wall, S.B.; Wang, Y.; Guo, X.; Sun, J.; Liu, Y.; Wang, Q.; Zhang, B. Genome-wide identification of KNOX transcription factors in cotton and the role of GhKNOX4-A and GhKNOX22-D in response to salt and drought stress. Int. J. Biol. Macromol. 2023, 226, 1248–1260. [Google Scholar] [CrossRef] [PubMed]
  123. Sagar, S.; Biswas, D.K.; Singh, A. Genomic and expression analysis indicate the involvement of phospholipase C family in abiotic stress signaling in chickpea (Cicer arietinum). Gene 2020, 753, 144797. [Google Scholar] [CrossRef] [PubMed]
  124. Sagar, S.; Deepika; Biswas, D.K.; Chandrasekar, R.; Singh, A. Genome-wide identification, structure analysis and expression profiling of phospholipases D under hormone and abiotic stress treatment in chickpea (Cicer arietinum). Int. J. Biol. Macromol. 2021, 169, 264–273. [Google Scholar] [CrossRef]
  125. Deepika, D.; Ankit; Jonwal, S.; Mali, K.V.; Sinha, A.K.; Singh, A. Molecular analysis indicates the involvement of Jasmonic acid biosynthesis pathway in low-potassium (K+) stress response and development in chickpea (Cicer arietinum). Environ. Exp. Bot. 2022, 194, 104753. [Google Scholar] [CrossRef]
  126. Deepika, D.; Poddar, N.; Kumar, S.; Singh, A. Molecular Characterization Reveals the Involvement of Calcium Dependent Protein Kinases in Abiotic Stress Signaling and Development in Chickpea (Cicer arietinum). Front. Plant Sci. 2022, 13, 831265. [Google Scholar] [CrossRef]
  127. He, X.; Xie, S.; Xie, P.; Yao, M.; Liu, W.; Qin, L.; Liu, Z.; Zheng, M.; Liu, H.; Guan, M.; et al. Genome-wide identification of stress-associated proteins (SAP) with A20/AN1 zinc finger domains associated with abiotic stresses responses in Brassica napus. Environ. Exp. Bot. 2019, 165, 108–119. [Google Scholar] [CrossRef]
  128. Song, W.; Zhao, H.; Zhang, X.; Lei, L.; Lai, J. Genome-Wide Identification of VQ Motif-Containing Proteins and their Expression Profiles Under Abiotic Stresses in Maize. Front. Plant Sci. 2016, 6, 1177. [Google Scholar] [CrossRef]
  129. Hu, W.; Ren, Q.; Chen, Y.; Xu, G.; Qian, Y. Genome-wide identification and analysis of WRKY gene family in maize provide insights into regulatory network in response to abiotic stresses. BMC Plant Biol. 2021, 21, 427. [Google Scholar] [CrossRef]
  130. Rai, M.; He, C.; Wu, R. Comparative functional analysis of three abiotic stress-inducible promoters in transgenic rice. Transgenic Res. 2009, 18, 787–799. [Google Scholar] [CrossRef]
  131. Maruyama, K.; Todaka, D.; Mizoi, J.; Yoshida, T.; Kidokoro, S.; Matsukura, S.; Takasaki, H.; Sakurai, T.; Yamamoto, Y.Y.; Yoshiwara, K.; et al. Identification of cis-acting promoter elements in cold- and dehydration-induced transcriptional pathways in arabidopsis, rice, and soybean. DNA Res. 2012, 19, 37–49. [Google Scholar] [CrossRef] [PubMed]
  132. Tsutsui, T.; Yamaji, N.; Feng Ma, J. Identification of a cis-acting element of ART1, a c2h2-type zinc-finger transcription factor for aluminum tolerance in rice. Plant Physiol. 2011, 156, 925–931. [Google Scholar] [CrossRef] [PubMed]
  133. Gómez-Porras, J.L.; Riaño-Pachón, D.M.; Dreyer, I.; Mayer, J.E.; Mueller-Roeber, B. Genome-wide analysis of ABA-responsive elements ABRE and CE3 reveals divergent patterns in Arabidopsis and rice. BMC Genom. 2007, 8, 260. [Google Scholar] [CrossRef] [PubMed]
  134. Nejat, N.; Ramalingam, A.; Mantri, N. Advances in transcriptomics of plants. In Plant Genetics and Molecular Biology; Springer International Publishing: Berlin/Heidelberg, Germany, 2018; pp. 161–185. [Google Scholar] [CrossRef]
  135. Wang, D.; Lu, X.; Chen, X.; Wang, S.; Wang, J.; Guo, L.; Yin, Z.; Chen, Q.; Ye, W. Temporal salt stress-induced transcriptome alterations and regulatory mechanisms revealed by PacBio long-reads RNA sequencing in Gossypium hirsutum. BMC Genom. 2020, 21, 838. [Google Scholar] [CrossRef] [PubMed]
  136. Wang, X.; Li, N.; Li, W.; Gao, X.; Cha, M.; Qin, L.; Liu, L. Advances in transcriptomics in the response to stress in plants. Glob. Med. Genet. 2020, 7, 30–34. [Google Scholar] [CrossRef] [PubMed]
  137. Lowe, R.; Shirley, N.; Bleackley, M.; Dolan, S.; Shafee, T. Transcriptomics technologies. PLOS Comput. Biol. 2017, 13, e1005457. [Google Scholar] [CrossRef] [PubMed]
  138. Wang, Z.; Gerstein, M.; Snyder, M. RNA-Seq: A revolutionary tool for transcriptomics. Nat. Rev. Genet. 2009, 10, 57–63. [Google Scholar] [CrossRef]
  139. Jain, M.; Ghanashyam, C.; Bhattacharjee, A. Comprehensive expression analysis suggests overlapping and specific roles of rice glutathione S-transferase genes during development and stress responses. BMC Genom. 2010, 11, 73. [Google Scholar] [CrossRef]
  140. Singh, A.; Pandey, A.; Baranwal, V.; Kapoor, S.; Pandey, G.K. Comprehensive expression analysis of rice phospholipase D gene family during abiotic stresses and development. Plant Signal. Behav. 2012, 7, 847–855. [Google Scholar] [CrossRef] [PubMed]
  141. Agarwal, P.; Arora, R.; Ray, S.; Singh, A.K.; Singh, V.P.; Takatsuji, H.; Kapoor, S.; Tyagi, A.K. Genome-wide identification of C2H2 zinc-finger gene family in rice and their phylogeny and expression analysis. Plant Mol. Biol. 2007, 65, 467–485. [Google Scholar] [CrossRef]
  142. Andrási, N.; Pettkó-Szandtner, A.; Szabados, L. Diversity of plant heat shock factors: Regulation, interactions, and functions. J. Exp. Bot. 2021, 72, 1558–1575. [Google Scholar] [CrossRef]
  143. Chauhan, H.; Khurana, N.; Agarwal, P.; Khurana, P. Heat shock factors in rice (Oryza sativa L.): Genome-wide expression analysis during reproductive development and abiotic stress. Mol. Genet. Genom. 2011, 286, 171–187. [Google Scholar] [CrossRef]
  144. Yang, S.; Zhou, J.; Li, Y.; Wu, J.; Ma, C.; Chen, Y.; Sun, X.; Wu, L.; Liang, X.; Fu, Q.; et al. AP2/EREBP Pathway Plays an Important Role in Chaling Wild Rice Tolerance to Cold Stress. Int. J. Mol. Sci. 2023, 24, 14441. [Google Scholar] [CrossRef] [PubMed]
  145. Muthusamy, S.K.; Lenka, S.K.; Katiyar, A.; Chinnusamy, V.; Singh, A.K.; Bansal, K.C. Genome-Wide identification and analysis of biotic and abiotic stress regulation of C4 photosynthetic pathway genes in rice. Appl. Biochem. Biotechnol. 2018, 187, 221–238. [Google Scholar] [CrossRef] [PubMed]
  146. Zhu, Y.-N.; Shi, D.-Q.; Ruan, M.-B.; Zhang, L.-L.; Meng, Z.-H.; Liu, J.; Yang, W.-C. Transcriptome Analysis Reveals Crosstalk of Responsive Genes to Multiple Abiotic Stresses in Cotton (Gossypium hirsutum L.). PLoS ONE 2013, 8, e80218. [Google Scholar] [CrossRef]
  147. Zhou, B.; Zhang, L.; Ullah, A.; Jin, X.; Yang, X.; Zhang, X. Identification of Multiple Stress Responsive Genes by Sequencing a Normalized cDNA Library from Sea-Land Cotton (Gossypium barbadense L.). PLoS ONE 2016, 11, e0152927. [Google Scholar] [CrossRef] [PubMed]
  148. Tahmasebi, A.; Ashrafi-Dehkordi, E.; Shahriari, A.G.; Mazloomi, S.M.; Ebrahimie, E. Integrative meta-analysis of transcriptomic responses to abiotic stress in cotton. Prog. Biophys. Mol. Biol. 2019, 146, 112–122. [Google Scholar] [CrossRef] [PubMed]
  149. Jiang, Q.-T.; Liu, T.; Ma, J.; Wei, Y.-M.; Lu, Z.-X.; Lan, X.-J.; Dai, S.-F.; Zheng, Y.-L. Characterization of barley Prp1 gene and its expression during seed development and under abiotic stress. Genetica 2011, 139, 1283–1292. [Google Scholar] [CrossRef] [PubMed]
  150. Mian, A.; Oomen, R.J.F.J.; Isayenkov, S.; Sentenac, H.; Maathuis, F.J.M.; Véry, A.-A. Over-expression of an Na+- and K+-permeable HKT transporter in barley improves salt tolerance. Plant J. 2011, 68, 468–479. [Google Scholar] [CrossRef]
  151. Akbudak, M.A.; Filiz, E.; Kontbay, K. DREB2 (dehydration-responsive element-binding protein 2) type transcription factor in sorghum (Sorghum bicolor): Genome-wide identification, characterization and expression profiles under cadmium and salt stresses. 3 Biotech 2018, 8, 426. [Google Scholar] [CrossRef] [PubMed]
  152. Singh, R.B.; Singh, B.; Singh, R.K. Development of potential dbEST-derived microsatellite markers for genetic evaluation of sugarcane and related cereal grasses. Ind. Crops Prod. 2019, 128, 38–47. [Google Scholar] [CrossRef]
  153. Hrdlickova, R.; Toloue, M.; Tian, B. RNA-Seq methods for transcriptome analysis. WIREs RNA 2017, 8, e1364. [Google Scholar] [CrossRef]
  154. Zhao, J.; He, Y.; Li, X.; Weng, X.; Feng, D.; Ying, J.; Wang, Z. An integrated RNA-Seq and physiological study reveals gene responses involving in the initial imbibition of seed germination in rice. Plant Growth Regul. 2020, 90, 249–263. [Google Scholar] [CrossRef]
  155. Schaarschmidt, S.; Fischer, A.; Lawas, L.M.F.; Alam, R.; Septiningsih, E.M.; Bailey-Serres, J.; Jagadish, S.V.K.; Huettel, B.; Hincha, D.K.; Zuther, E. Utilizing PacBio Iso-Seq for Novel Transcript and Gene Discovery of Abiotic Stress Responses in Oryza sativa L. Int. J. Mol. Sci. 2020, 21, 8148. [Google Scholar] [CrossRef]
  156. Zhao, T.; Wu, T.; Zhang, J.; Wang, Z.; Pei, T.; Yang, H.; Li, J.; Xu, X. Genome-Wide Analyses of the Genetic Screening of C2H2-Type Zinc Finger Transcription Factors and Abiotic and Biotic Stress Responses in Tomato (Solanum lycopersicum) Based on RNA-Seq Data. Front. Genet. 2020, 11, 540. [Google Scholar] [CrossRef]
  157. Huang, Y.; Chen, H.; Reinfelder, J.R.; Liang, X.; Sun, C.; Liu, C.; Li, F.; Yi, J. A transcriptomic (RNA-seq) analysis of genes responsive to both cadmium and arsenic stress in rice root. Sci. Total Environ. 2019, 666, 445–460. [Google Scholar] [CrossRef]
  158. 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]
  159. Li, P.; Cao, W.; Fang, H.; Xu, S.; Yin, S.; Zhang, Y.; Lin, D.; Wang, J.; Chen, Y.; Xu, C.; et al. Transcriptomic Profiling of the Maize (Zea mays L.) Leaf Response to Abiotic Stresses at the Seedling Stage. Front. Plant Sci. 2017, 8, 290. [Google Scholar] [CrossRef] [PubMed]
  160. Zhang, Z.; Ali, S.; Zhang, T.; Wang, W.; Xie, L. Identification, evolutionary and expression analysis of pyl-pp2c-snrk2s gene families in soybean. Plants 2020, 9, 1356. [Google Scholar] [CrossRef] [PubMed]
  161. Ankit, A.; Singh, A.; Kumar, S.; Singh, A. Morphophysiological and transcriptome analysis reveal that reprogramming of metabolism, phytohormones and root development pathways governs the potassium (K+) deficiency response in two contrasting chickpea cultivars. Front. Plant Sci. 2023, 13, 1054821. [Google Scholar] [CrossRef] [PubMed]
  162. Edgar, R.; Micheal, D.; Alex, L. Gene Expression Omnibus: NCBI gene expression and hybridization array data repository. Nucleic Acids Res. 2002, 30, 207–210. [Google Scholar] [CrossRef]
  163. Kolesnikov, N.; Hastings, E.; Keays, M.; Melnichuk, O.; Tang, Y.A.; Williams, E.; Dylag, M.; Kurbatova, N.; Brandizi, M.; Burdett, T.; et al. ArrayExpress update—Simplifying data submissions. Nucleic Acids Res. 2015, 43, D1113–D1116. [Google Scholar] [CrossRef] [PubMed]
  164. Dash, S.; Van Hemert, J.; Hong, L.; Wise, R.P.; Dickerson, J.A. PLEXdb: Gene expression resources for plants and plant pathogens. Nucleic Acids Res. 2012, 40, D1194–D1201. [Google Scholar] [CrossRef]
  165. Hruz, T.; Laule, O.; Szabo, G.; Wessendorp, F.; Bleuler, S.; Oertle, L.; Widmayer, P.; Gruissem, W.; Zimmermann, P. Genevestigator V3: A reference expression database for the meta-analysis of transcriptomes. Adv. Bioinform. 2008, 2008, 420747. [Google Scholar] [CrossRef]
  166. Sato, Y.; Takehisa, H.; Kamatsuki, K.; Minami, H.; Namiki, N.; Ikawa, H.; Ohyanagi, H.; Sugimoto, K.; Antonio, B.A.; Nagamura, Y. RiceXPro Version 3.0: Expanding the informatics resource for rice transcriptome. Nucleic Acids Res. 2012, 41, D1206–D1213. [Google Scholar] [CrossRef]
  167. Portwood, J.L., II; Woodhouse, M.R.; Cannon, E.K.; Gardiner, J.M.; Harper, L.C.; Schaeffer, M.L.; Walsh, J.R.; Sen, T.Z.; Cho, K.T.; Schott, D.A.; et al. MaizeGDB 2018: The maize multi-genome genetics and genomics database. Nucleic Acids Res. 2018, 47, D1146–D1154. [Google Scholar] [CrossRef]
  168. Ramírez-González, R.H.; Borrill, P.; Lang, D.; Harrington, S.A.; Brinton, J.; Venturini, L.; Davey, M.; Jacobs, J.; van Ex, F.; Pasha, A.; et al. The transcriptional landscape of hexaploid wheat across tissues and cultivar. Science 2018, 361, 3083–3109. [Google Scholar] [CrossRef]
  169. Fernandez-Pozo, N.; Zheng, Y.; Snyder, S.I.; Nicolas, P.; Shinozaki, Y.; Fei, Z.; Catala, C.; Giovannoni, J.J.; Rose, J.K.C.; Mueller, L.A. The tomato expression atlas. Bioinformatics 2017, 33, 2397–2398. [Google Scholar] [CrossRef]
  170. Tian, T.; You, Q.; Zhang, L.; Yi, X.; Yan, H.; Xu, W.; Su, Z. SorghumFDB: Sorghum functional genomics database with multidimensional network analysis. Database 2016, 2016, baw099. [Google Scholar] [CrossRef]
  171. Shen, L.; Gong, J.; Caldo, R.A.; Nettleton, D.; Cook, D.; Cook, R.P.; Dickerson, J. BarleyBase--an expression profiling database for plant genomics. Nucleic Acids Res. 2005, 33, D614–D618. [Google Scholar] [CrossRef] [PubMed]
  172. Dai, X.; Zhuang, Z.; Boschiero, C.; Dong, Y.; Zhao, P.X. LegumeIP V3: From models to crops—An integrative gene discovery platform for translational genomics in legumes. Nucleic Acids Res. 2021, 49, D1472–D1479. [Google Scholar] [CrossRef]
  173. Hamada, K.; Hongo, K.; Suwabe, K.; Shimizu, A.; Nagayama, T.; Abe, R.; Kikuchi, S.; Yamamoto, N.; Fujii, T.; Yokoyama, K.; et al. OryzaExpress: An integrated database of gene expression networks and omics annotations in rice. Plant Cell Physiol. 2011, 52, 220–229. [Google Scholar] [CrossRef] [PubMed]
  174. García-Ruiz, S.; Gil-Martínez, A.L.; Cisterna, A.; Jurado-Ruiz, F.; Reynolds, R.H.; Cookson, M.R.; Hardy, J.; Ryten, M.; Botía, J.A. CoExp: A web tool for the exploitation of co-expression networks. Front. Genet. 2021, 12, 630187. [Google Scholar] [CrossRef]
  175. Athanasopoulou, K.; Boti, M.A.; Adamopoulos, P.G.; Skourou, P.C.; Scorilas, A. Third-Generation sequencing: The spearhead towards the radical transformation of modern genomics. Life 2021, 12, 30. [Google Scholar] [CrossRef] [PubMed]
  176. Yang, J.; Lv, W.; Zeng, M.; Fu, Y.; Li, C. PacBio and Illumina RNA sequencing identify alternative splicing events in response to cold stress in two poplar species. Authorea, 2020; preprint. [Google Scholar] [CrossRef]
  177. Zhao, C.; He, L.; Xia, H.; Zhou, X.; Geng, Y.; Hou, L.; Li, P.; Li, G.; Zhao, S.; Ma, C.; et al. De novo full length transcriptome analysis of Arachis glabrata provides insights into gene expression dynamics in response to biotic and abiotic stresses. Genomics 2021, 113, 1579–1588. [Google Scholar] [CrossRef]
  178. Ma, L.; Dong, C.; Song, C.; Wang, X.; Zheng, X.; Niu, Y.; Chen, S.; Feng, W. De novo genome assembly of the potent medicinal plant Rehmannia glutinosa using nanopore technology. Comput. Struct. Biotechnol. J. 2021, 19, 3954–3963. [Google Scholar] [CrossRef]
  179. Wei, H.; Xu, H.; Su, C.; Wang, X.; Wang, L. Rice CIRCADIAN CLOCK ASSOCIATED 1 transcriptionally regulates ABA signaling to confer multiple abiotic stress tolerance. Plant Physiol. 2022, 190, 1057–1073. [Google Scholar] [CrossRef]
  180. Chen, L.; Hao, L.; Parry, M.A.J.; Phillips, A.L.; Hu, Y. Progress in TILLING as a tool for functional genomics and improvement of crops. J. Integr. Plant Biol. 2014, 56, 425–443. [Google Scholar] [CrossRef] [PubMed]
  181. Boopathi, N.M. Genetic Mapping and Marker Assisted Selection: Basics, Practice and Benefits; Springer Nature: Berlin/Heidelberg, Germany, 2020. [Google Scholar]
  182. Khalil, A.M. The genome editing revolution: Review. J. Genet. Eng. Biotechnol. 2020, 18, 68. [Google Scholar] [CrossRef]
  183. Hira, Z.M.; Gillies, D.F. A review of feature selection and feature extraction methods applied on microarray data. Adv. Bioinform. 2015, 2015, 198363. [Google Scholar] [CrossRef] [PubMed]
  184. Nagaraj, S.H.; Gasser, R.B.; Ranganathan, S. A hitchhiker’s guide to expressed sequence tag (EST) analysis. Brief. Bioinform. 2006, 8, 6–21. [Google Scholar] [CrossRef]
  185. Van Verk, M.C.; Hickman, R.; Pieterse, C.M.J.; Van Wees, S.C.M. RNA-Seq: Revelation of the messengers. Trends Plant Sci. 2013, 18, 175–179. [Google Scholar] [CrossRef] [PubMed]
  186. Paterson, A.H.; Lander, E.S.; Hewitt, J.D.; Peterson, S.; Lincoln, S.E.; Tanksley, S.D. Resolution of quantitative traits into Mendelian factors by using a complete linkage map of restriction fragment length polymorphisms. Nature 1988, 335, 721–726. [Google Scholar] [CrossRef]
  187. Pardue, M.L.; Gall, J.G. Molecular Hybridization of Radioactive DNA to the DNA of Cytological Preparations. Proc. Natl. Acad. Sci. USA 1969, 64, 600–604. [Google Scholar] [CrossRef] [PubMed]
  188. Kleppe, K.; Ohtsuka, E.; Kleppe, R.; Molineux, I.; Khorana, H.G. Studies on polynucleotides. J. Mol. Biol. 1971, 56, 341–361. [Google Scholar] [CrossRef]
  189. Smith, H.O.; Birnstiel, M.L. A simple method for DNA restriction site mapping. Nucleic Acids Res. 1976, 3, 2387–2398. [Google Scholar] [CrossRef]
  190. Sanger, F.; Nicklen, S.; Coulson, A.R. DNA sequencing with chain-terminating inhibitors. Proc. Natl. Acad. Sci. USA 1977, 74, 5463–5467. [Google Scholar] [CrossRef]
  191. Adams, M.D.; Kelley, J.M.; Gocayne, J.D.; Dubnick, M.; Polymeropoulos, M.H.; Xiao, H.; Merril, C.R.; Wu, A.; Olde, B.; Moreno, R.F.; et al. Complementary DNA sequencing: Expressed sequence tags and human genome project. Science 1991, 252, 1651–1656. [Google Scholar] [CrossRef]
  192. Higuchi, R.; Dollinger, G.; Walsh, P.S.; Griffith, R. Simultaneous amplification and detection of specific DNA sequences. Bio/Technol. 1992, 10, 413–417. [Google Scholar] [CrossRef]
  193. Blank, C.E.; Kessler, P.S.; Leigh, J.A. Genetics in methanogens: Transposon insertion mutagenesis of a Methanococcus maripaludis nifH gene. J. Bacteriol. 1995, 177, 5773–5777. [Google Scholar] [CrossRef]
  194. Velculescu, V.E.; Zhang, L.; Vogelstein, B.; Kinzler, K.W. Serial analysis of gene expression. Science 1995, 270, 484–487. [Google Scholar] [CrossRef]
  195. De Mesmaeker, A.; Haener, R.; Martin, P.; Moser, H.E. Antisense oligonucleotides. Acc. Chem. Res. 1995, 28, 366–374. [Google Scholar] [CrossRef]
  196. Schena, M.; Shalon, D.; Davis, R.W.; Brown, P.O. Quantitative monitoring of gene expression patterns with a complementary DNA microarray. Science 1995, 270, 467–470. [Google Scholar] [CrossRef] [PubMed]
  197. Kumagai, M.H.; Donson, J.; della-Cioppa, G.; Harvey, D.; Hanley, K.; Grill, L.K. Cytoplasmic inhibition of carotenoid biosynthesis with virus-derived RNA. Proc. Natl. Acad. Sci. USA 1995, 92, 1679–1683. [Google Scholar] [CrossRef]
  198. Fire, A.; Xu, S.; Montgomery, M.K.; Kostas, S.A.; Driver, S.E.; Mello, C.C. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature 1998, 391, 806–811. [Google Scholar] [CrossRef] [PubMed]
  199. Mardis, E.R. The impact of next-generation sequencing technology on genetics. Trends Genet. 2008, 24, 133–141. [Google Scholar] [CrossRef] [PubMed]
  200. Bozas, A.; Beumer, K.J.; Trautman, J.K.; Carroll, D. Genetic analysis of zinc-finger nuclease-induced gene targeting in drosophila. Genetics 2009, 182, 641–651. [Google Scholar] [CrossRef] [PubMed]
  201. Tang, F.; Barbacioru, C.; Wang, Y.; Nordman, E.; Lee, C.; Xu, N.; Wang, X.; Bodeau, J.; Tuch, B.B.; Siddiqui, A.; et al. mRNA-Seq whole-transcriptome analysis of a single cell. Nat. Methods 2009, 6, 377–382. [Google Scholar] [CrossRef] [PubMed]
  202. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A programmable dual-rna–guided DNA endonuclease in adaptive bacterial immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef]
  203. Sun, N.; Zhao, H. Transcription activator-like effector nucleases (TALENs): A highly efficient and versatile tool for genome editing. Biotechnol. Bioeng. 2013, 110, 1811–1821. [Google Scholar] [CrossRef]
  204. Wenger, A.M.; Peluso, P.; Rowell, W.J.; Chang, P.-C.; Hall, R.J.; Concepcion, G.T.; Ebler, J.; Fungtammasan, A.; Kolesnikov, A.; Olson, N.D.; et al. Accurate circular consensus long-read sequencing improves variant detection and assembly of a human genome. Nat. Biotechnol. 2019, 37, 1155–1162. [Google Scholar] [CrossRef]
Figure 1. Genomics and Transcriptomics Approaches for Enhancing Abiotic Stress Tolerance in Plants: an overview of the genomics and transcriptomics methodologies in the study of abiotic stress tolerance. Genomics plays a crucial role in the identification and functional analysis of stress-responsive genes, while transcriptomics enables the comprehensive analysis of gene expression patterns. Together, these approaches contribute significantly to crop improvement efforts.
Figure 1. Genomics and Transcriptomics Approaches for Enhancing Abiotic Stress Tolerance in Plants: an overview of the genomics and transcriptomics methodologies in the study of abiotic stress tolerance. Genomics plays a crucial role in the identification and functional analysis of stress-responsive genes, while transcriptomics enables the comprehensive analysis of gene expression patterns. Together, these approaches contribute significantly to crop improvement efforts.
Agronomy 13 02903 g001
Table 1. List of tools and database available for plant gene inactivation or gene editing.
Table 1. List of tools and database available for plant gene inactivation or gene editing.
Tool/DatabaseURLFunctionCitation
CRISPR-PLANThttp://omap.org/crispr/ (accessed on 3 October 2023)Selects suitable CRISPR target sites for gene editing in various plant species.[35]
Cas-Designerwww.rgenome.net/cas-designer/ (accessed on 3 October 2023)Chooses CRISPR-Cas9 target sites for various organisms, including plants.[36]
DESKGEN Cloudwww.deskgen.com/ (accessed on 3 October 2023)Online platform for gRNA design and CRISPR experiment planning, supporting various CRISPR applications.[37]
E-CRISPwww.e-crisp.org/E-CRISP/ (accessed on 3 October 2023)Fast CRISPR target site identification tool.[38]
GuideScan2www.guidescan.com/ (accessed on 3 October 2023)Software for designing CRISPR guide RNA (gRNA) libraries.[39]
ZiFiThttps://mybiosoftware.com/zifit-4-2-zinc-finger-engineering-tool.html (accessed on 3 October 2023)Software tool for designing custom zinc finger proteins for gene editing.[40]
CHOPCHOPchopchop.cbu.uib.no/ (accessed on 3 October 2023)Web tool for precise design of CRISPR/Cas9 targets.[41]
CRISPORhttp://crispor.gi.ucsc.edu/ (accessed on 3 October 2023)Web tool offering efficiency and specificity scores for CRISPR-Cas9 genome editing.[42]
CRISPR-ERAcrispr-era.stanford.edu/ (accessed on 3 October 2023)Tool for evaluating and annotating CRISPR experiment results and designing gRNAs for gene editing.[43]
CRISPR-Pcrispr.hzau.edu.cn/CRISPR2/ (accessed on 3 October 2023)Web tool for identifying potential CRISPR target sites in input DNA sequences.[44]
Cas-OFFinderwww.rgenome.net/cas-offinder/ (accessed on 3 October 2023)Web-based tool for searching potential off-target sites in the genome for CRISPR/Cas-derived RNA-guided endonucleases.[45]
SSFinderhttps://code.google.com/p/ssfinder/ (accessed on 3 October 2023)Web-based tool for detecting sequences suitable for base editing using CRISPR/Cas-derived base editors.[46]
JASPARjaspar.genereg.net/ (accessed on 3 October 2023)Database of transcription factor binding profiles, including those for Cre and FLP recombinases.[47]
Table 2. Genes involved in regulating abiotic stress response through gene overexpression and gene silencing approaches.
Table 2. Genes involved in regulating abiotic stress response through gene overexpression and gene silencing approaches.
GenePlantAbiotic Stress ToleranceFunctional Analysis MethodCitation
GmNAC10SoybeanCold, salt and
dehydration
Overexpression[55]
AtICE1RiceCold and droughtOverexpression[52]
MP-mi397BananaCopper and iron
deficiencies, salt and drought
Overexpression[56]
BPERF1.1White BirchCold, salt and droughtOverexpression[51]
OsRIP1RiceOsmotic stressOverexpression[57]
MiMFTsMangoSalt and osmotic stressOverexpression[58]
JrWOX11WalnutSalt and osmotic stressOverexpression[59]
SlDDF2TomatoDrought, salinity and coldOverexpression.[54]
OsHis1.1RiceHeat and coldOverexpression[60]
ZmSLK1 and ZmSLK2ArabidopsisDroughtOverexpression[61]
GmlncRNA77580SoybeanDroughtOverexpression[62]
GhSAMSCottonDrought and salinityVGIS[63]
SlWRKY79TomatoSaltVGIS[64]
GhMYB36CottonDroughtVGIS[65]
GmAITRSoybeanSalinityCRISPR[66]
OsCCA1RiceSalinity, osmotic and droughtCRISPR[67]
OsmiR535RiceSalt and osmoticCRISPR[68]
OsPUB7RiceDrought and salinityCRISPR[69]
NtCycB2TobaccoSalinityCRISPR[70]
SlGRAS10TomatoOsmotic stressRNAi[71]
OsIAA20RiceDrought and salt stressRNAi[72]
Mslea-D34AlfalfaDrought and saltRNAi[73]
Table 3. Database of molecular markers and QTL-based tools in plants.
Table 3. Database of molecular markers and QTL-based tools in plants.
Tool/Database NameURLDescriptionCitation
Gramenewww.gramene.org (accessed on 3 October 2023)A resource for comparative analysis of grass genomes, includes a vast array of molecular markers.[82]
SoyBasewww.soybase.org (accessed on 3 October 2023)A comprehensive database of soybean genetic and genomic information, including molecular markers.[83]
CottonGenwww.cottongen.org (accessed on 3 October 2023)A database of genetic and genomic information for cotton, includes molecular markers.[84]
SOL Genomics Network (SGN)solgenomics.net (accessed on 3 October 2023)A database with genomic, genetic, and taxonomic information for the Solanaceae family and more distantly related species. Includes molecular markers.[85]
MaizeGDBhttps://www.maizegdb.org/ (accessed on 3 October 2023)The Maize Genetics and Genomics Database, a resource for maize sequence, stock, phenotype, and polymorphism data, including molecular markers.[86]
MolMarkerhttps://sourceforge.net/projects/molmarker/ (accessed on 3 October 2023)This software evaluates plant molecular marker information as well as the related QTL information.[87]
The Triticeae Toolbox (T3)https://triticeaetoolbox.org/ (accessed on 3 October 2023)A database for Triticeae (wheat and barley) genetic and genomic data, including molecular markers and QTL information.[88]
Cucurbit Genomics Database (CuGenDB)http://cucurbitgenomics.org/ (accessed on 3 October 2023)A centralized platform for cucurbit genomics and genetic data, including molecular markers and QTLs.[89]
QTL IciMappinghttps://isbreedingen.caas.cn/software/qtllcimapping/294607.htm (accessed on 3 October 2023)A software used for the genetic mapping of QTLs.[90]
R/QTLhttps://rqtl.org/ (accessed on 3 October 2023)R/QTL is an interactive QTL mapping software, implemented as an R package.[91]
CottonFGDhttps://cottonfgd.org/ (accessed on 3 October 2023)A functional genomics database for cotton, including molecular markers and QTLs.[92]
PlantQTL-GEhttp://www.scbit.org/qtl2gene/new/ (accessed on 3 October 2023)A database that stores QTLs and genetically mapped genes in plant species and provides a platform to perform comparative studies on the genetic architecture of complex traits.[93]
PeanutBasehttps://peanutbase.org/ (accessed on 3 October 2023)A peanut community resource providing genetic, genomic, gene function, and germplasm data, including molecular markers and QTLs.[94]
GnpIS-Ephesishttps://urgi.versailles.inra.fr/gnpis (accessed on 3 October 2023)An information system that allows querying and visualizing genotyping data and phenotypic scores for plant species. It includes QTL data.[95]
Wheat@URGIhttps://wheat-urgi.versailles.inra.fr/Seq-Repository/Annotations (accessed on 3 October 2023)A database providing a complete view of genetic, physical, and functional wheat sequence resources, including molecular markers and QTLs.[96]
Table 4. RNA-Seq-based and gene expression databases.
Table 4. RNA-Seq-based and gene expression databases.
Database NameURLShort ParticularsCitation
NCBI Gene Expression Omnibus (GEO)https://www.ncbi.nlm.nih.gov/geo/ (accessed on 3 October 2023)A repository for gene expression data[162]
ArrayExpresshttps://www.ebi.ac.uk/arrayexpress/ (accessed on 3 October 2023)A public repository for gene expression data[163]
Plant Expression Database (PLEXdb)https://www.plexdb.org/ (accessed on 3 October 2023)A resource for plant gene expression data[164]
Genevestigatorhttps://genevestigator.com/ (accessed on 3 October 2023)A gene expression database and analysis platform[165]
Rice Expression Database (RiceXPro)http://ricexpro.dna.affrc.go.jp/ ( accessed on 3 October 2023)A repository for rice gene expression data[166]
SoyBasehttps://www.soybase.org/ (accessed on 3 October 2023)A repository for soybean genomics data[83]
MaizeGDBhttps://www.maizegdb.org/ (accessed on 3 October 2023)A database for maize genetics and genomics[167]
Wheat Expression Browserhttps://wheat.pw.usda.gov/ (accessed on 3 October 2023)A platform for wheat gene expression data[168]
Tomato Expression Atlashttp://tea.solgenomics.net/ (accessed on 3 October 2023)A resource for tomato gene expression data[169]
CottonFGDhttps://cottonfgd.net/ (accessed on 3 October 2023)A functional genomics database for cotton[83]
SorghumFDBhttp://structuralbiology.cau.edu.cn/sorghum/index.html (accessed on 3 October 2023)A repository for sorghum genomics data[170]
BarleyBasehttps://www.plexdb.org/ (accessed on 3 October 2023)A database for barley genomics data[171]
Legum IP V3https://plantgrn.noble.org/LegumeIP (accessed on 3 October 2023)A database for legume RNA-Seq data[172]
Oryza Expresshttps://rice.plantbiology.msu.edu/ (accessed on 3 October 2023)A repository for rice gene expression data[173]
Coexpression Browser (BAR)https://bar.utoronto.ca/ (accessed on 3 October 2023)A database for gene coexpression in plants[174]
PLEXdbhttps://www.plexdb.org/ (accessed on 3 October 2023)A repository for plant gene expression data[164]
Table 5. A list of tools involved in genomics and transcriptomics with their advantages and disadvantages.
Table 5. A list of tools involved in genomics and transcriptomics with their advantages and disadvantages.
TechniqueYear of OriginApplicationAdvantagesDisadvantagesCitation
FISH (Fluorescence In Situ Hybridization)1969Gene mappingVisualizes specific DNA sequences on chromosomesLimited to fixed cells, labor-intensive[187]
PCR (Polymerase Chain Reaction)1971In vitro DNA amplificationAmplifies DNA quicklySusceptible to contamination, limited to short sequences[188]
Restriction Enzyme Mapping1976Enzymatic gene mappingMaps DNA fragmentsLimited resolution, labor-intensive[189]
Sanger Sequencing (DNA Sequencing)1977Chain-termination sequencingReads DNA base by baseSlow, expensive for whole genomes[190]
QTL Mapping (Quantitative Trait Loci Mapping)1988Genetic linkage analysisIdentifies genomic regions associated with traitsLimited resolution, requires genetic markers[186]
ESTs (Expressed Sequence Tags)1991Gene expression analysisIdentifies and catalogs genesProvides only partial gene information[191]
RT-PCR (Reverse Transcription PCR)1992Gene expression analysisHighly sensitiveRequires prior knowledge of the target sequence[192]
Transposon-Mediated Insertional Mutagenesis1995Gene disruptionRandom gene disruptionLack of precise control, potential for multiple insertions[193]
SAGE (Serial Analysis of Gene Expression)1995Quantification of mRNA tagsQuantifies gene expressionRequires a significant amount of starting material[194]
Antisense Oligonucleotides1995Inhibition of gene expressionSpecific gene silencingTransient effect, variable efficiency[195]
Microarray Analysis1995Hybridization-based gene expression analysisSimultaneous analysis of gene expressionLimited to known probes, cross-hybridization risk[196]
VIGS (Virus-Induced Gene Silencing)1995Gene silencingSilences gene expression in plantsLimited to plants, viral interactions[197]
RNAi Technology (RNA interference)1998Gene silencingSpecific inhibition of gene expressionOff-target effects, variability in silencing efficiency[198]
RNA-Seq2008Next-Generation Sequencing (NGS)High-throughput transcriptome analysisData analysis complexity, cost[138]
NGS (Next-Generation Sequencing)2008High-throughput sequencing (e.g., Illumina)High-throughput genome sequencingData storage and analysis demands[199]
ZFNs (Zinc Finger Nucleases)2009Genome editingPrecise gene targeting, reduced off-target effectsDesign complexity, higher cost[200]
Single-Cell RNA sequencing2009Single-cell analysisReveals cellular heterogeneityComplex data analysis, limited to single cells[201]
CRISPR-Cas9 Genome Editing2012Genome editingPrecise gene editing and inactivationOff-target effects, ethical concerns[202]
TALENs (Transcription Activator-Like Effector Nucleases)2013Genome editingPrecise gene targeting, reduced off-target effectsDesign complexity, higher cost[203]
Long-Read Sequencing (PacBio)2019Long-read sequencingSequences longer DNA fragmentsHigher error rate, cost[204]
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

Kamali, S.; Singh, A. Genomic and Transcriptomic Approaches to Developing Abiotic Stress-Resilient Crops. Agronomy 2023, 13, 2903. https://doi.org/10.3390/agronomy13122903

AMA Style

Kamali S, Singh A. Genomic and Transcriptomic Approaches to Developing Abiotic Stress-Resilient Crops. Agronomy. 2023; 13(12):2903. https://doi.org/10.3390/agronomy13122903

Chicago/Turabian Style

Kamali, Saravanappriyan, and Amarjeet Singh. 2023. "Genomic and Transcriptomic Approaches to Developing Abiotic Stress-Resilient Crops" Agronomy 13, no. 12: 2903. https://doi.org/10.3390/agronomy13122903

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