Next Article in Journal
Metabolome and Transcriptome Integrated Analysis of Mulberry Leaves for Insight into the Formation of Bitter Taste
Previous Article in Journal
A Comprehensive Pan-Cancer Analysis of the Potential Biological Functions and Prognosis Values of RICTOR
Previous Article in Special Issue
Integration of Phosphoproteomics and Transcriptome Studies Reveals ABA Signaling Pathways Regulate UV-B Tolerance in Rhododendron chrysanthum Leaves
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multi-Omics Pipeline and Omics-Integration Approach to Decipher Plant’s Abiotic Stress Tolerance Responses

1
Department of Plant Pathology and Weed Research, Institute of Plant Protection, Agricultural Research Organization (ARO)—The Volcani Institute, Rishon Lezion 7505101, Israel
2
School of Bioscience, Seacom Skills University, Bolpur 731236, West Bengal, India
3
Dr. Vikram Sarabhai Institute of Cell and Molecular Biology, Faculty of Science, Maharaja Sayajirao University of Baroda, Vadodara 390002, Gujarat, India
4
Department of Biotechnology and Bioscience, Banasthali Vidyapith, Banasthali 304022, Rajasthan, India
5
Department of Plant Biochemistry and Biotechnology, Sri Krishnadevaraya College of Agricultural Sciences (SKCAS), Affiliated to Acharya N.G. Ranga Agricultural University (ANGRAU), Guntur 522034, Andhra Pradesh, India
6
Department of Genetics and Plant Breeding, Faculty of Agriculture, Bangabandhu Sheikh Mujibur Rahman Agricultural University, Gazipur 1706, Bangladesh
7
Department of Botany, Maharshi Vishwamitra (M.V.) College, Buxar 802102, Bihar, India
8
Department of Biotechnology, Mizoram University, Pachhunga University College Campus, Aizawl 796001, Mizoram, India
9
Plant Molecular Biology Laboratory, Department of Botany, Sri Krishnadevaraya University, Anantapur 515003, Andhra Pradesh, India
*
Author to whom correspondence should be addressed.
Genes 2023, 14(6), 1281; https://doi.org/10.3390/genes14061281
Submission received: 25 April 2023 / Revised: 3 June 2023 / Accepted: 14 June 2023 / Published: 16 June 2023
(This article belongs to the Special Issue Abiotic Stress in Plants: Present and Future)

Abstract

:
The present day’s ongoing global warming and climate change adversely affect plants through imposing environmental (abiotic) stresses and disease pressure. The major abiotic factors such as drought, heat, cold, salinity, etc., hamper a plant’s innate growth and development, resulting in reduced yield and quality, with the possibility of undesired traits. In the 21st century, the advent of high-throughput sequencing tools, state-of-the-art biotechnological techniques and bioinformatic analyzing pipelines led to the easy characterization of plant traits for abiotic stress response and tolerance mechanisms by applying the ‘omics’ toolbox. Panomics pipeline including genomics, transcriptomics, proteomics, metabolomics, epigenomics, proteogenomics, interactomics, ionomics, phenomics, etc., have become very handy nowadays. This is important to produce climate-smart future crops with a proper understanding of the molecular mechanisms of abiotic stress responses by the plant’s genes, transcripts, proteins, epigenome, cellular metabolic circuits and resultant phenotype. Instead of mono-omics, two or more (hence ‘multi-omics’) integrated-omics approaches can decipher the plant’s abiotic stress tolerance response very well. Multi-omics-characterized plants can be used as potent genetic resources to incorporate into the future breeding program. For the practical utility of crop improvement, multi-omics approaches for particular abiotic stress tolerance can be combined with genome-assisted breeding (GAB) by being pyramided with improved crop yield, food quality and associated agronomic traits and can open a new era of omics-assisted breeding. Thus, multi-omics pipelines together are able to decipher molecular processes, biomarkers, targets for genetic engineering, regulatory networks and precision agriculture solutions for a crop’s variable abiotic stress tolerance to ensure food security under changing environmental circumstances.

1. Introduction

Stress could be defined as potentially unfavorable changes in environmental and/or biological factors that negatively affect plant growth, development and productivity [1]. Stress is categorized as abiotic and biotic. Living organisms such as fungi, bacteria, viruses, parasites, insects, weeds and native plants can induce biotic stress [2]. Whereas abiotic stress is caused by the effect of non-living environmental factors such as drought, salinity, temperature, cold, waterlogging, heavy metal, high light intensity, etc. [3], the impact of stresses on crop plants, either abiotic or biotic, is multidimensional and causes significant yield loss [4,5]. For the adaptation to environmental stresses, plants evolve several sophisticated mechanisms. Plants’ responsiveness to stress conditions entails sensing the stress, which activates a signal transduction pathway, activating stress-responsive genes via secondary messengers of signal transduction cascades and finally activating several stress-responsive genes and their products, which respond transcriptionally and translationally to the concerned abiotic stress. Due to the complex polygenic nature of the stress responses, there is a need to unravel the possible mechanisms of stress tolerance [6,7,8]. Deciphering the complex molecular regulatory network for a plant’s stress response requires cutting-edge genomic and molecular biology techniques such as high-throughput analysis of expressed sequence tags (EST), large-scale parallel analysis of gene expression, targeted or random mutagenesis and loss/gain-of-function or mutant complementation. These techniques have the potential to significantly improve our understanding of plants’ response to abiotic stress tolerance [9,10,11]. This technological advancement generates enormous amounts of information with thousands of new algorithms, tools and software, improvement in storage, processing and sharing of large datasets. “Omics” emerged as a new field that boosted the interaction of different modern biological approaches [12,13,14,15]. Omics is a term that refers to a set of molecular, system and computational biology tools that are used to assess the roles and interconnections of biological information in various clusters of life [16,17]. Multi-omics approaches include genomics, transcriptomics, proteomics, metabolomics, epigenomics, bioinformatics, proteogenomics, lipidomics, ionomics, interactomics and phenomics that provide a magnitude of data for understanding the physiological processes in plants under stress conditions and the tricky way to combat the harsh effect of those stresses [16,17,18,19]. Nevertheless, utilizing only a mono-omics approach does not provide sufficient knowledge to understand the complexity of plant responses under stress conditions. Therefore it is required to apply or integrate the multi-omics approaches for some promising output.
Innovative approaches that integrate the data from multi-omics layers, i.e., panomics, such as genomics, transcriptomics, proteomics and metabolomics hold great promise for enhancing crop improvement strategies [20]. By combining information from these different layers, it is possible to gain a comprehensive understanding of complex biological processes and identify key molecular players involved in a plant’s abiotic stress tolerance response. Several potential tools and approaches have been created for omics-data integration and interpretation due to the availability of multi-omics-pipeline-generated data collected from a wide range of samples and the introduction of high-throughput screening procedures by using data repositories and visualization portals [21,22]. While integrating multi-omics data, panomics can provide a more holistic and in-depth analysis of abiotic stress tolerance through data integration, system biological analytics, functional annotation and pathway analysis, data mining and machine learning for precise genomic prediction of crop germplasm [22]. It overcomes the difficulties encountered when integrating data from several different sources. Such integration of multi-omics data enables researchers to unravel intricate regulatory networks, identify candidate genes, proteins and metabolites, and discover potential biomarkers or targets for crop enhancement. This knowledge can aid in the development of stress-tolerant crop varieties through targeted breeding or genetic engineering approaches [21,22].
To decipher crops’ abiotic stress tolerance response, both available marker-assisted breeding (MAB) and advanced multiomics-based analysis are complementary to each other, but omics-based tools provide a broader and more comprehensive understanding of plant biology of stress responses than MAB [23]. Omics technologies provide a comprehensive view of the entire plant system to study the complex interactions between genes, proteins and metabolites underlying molecular mechanisms of abiotic stress tolerance. It can identify novel genes and pathways associated with abiotic stress tolerance that may not have been previously recognized or targeted by marker-assisted breeding approaches [24]. In addition, omics pipelines enable high-throughput screening and selection of germplasm based on their genetic makeup, gene expression patterns, protein profiles or metabolite compositions which allows breeders to efficiently identify and select individuals with desirable stress tolerance properties [25]. Omics-based analysis can integrate data from various omics platforms, providing a systems-level understanding that enables the identification of key regulatory networks, biomarkers and candidate genes that can be targeted for breeding efforts. This holistic approach increases the chances of success in developing stress-tolerant crop varieties. The big data obtained from the multi-omics layers, combined with advanced bioinformatics and computational tools, can be used for predictive modelling and precision breeding by applying machine learning algorithms [26]. In the following sections, a detailed view on each omics approach and multi-omics integration has been discussed to bring a clear picture of abiotic stress tolerance mechanism in plants (Figure 1).

2. Genomics

A genome is an organism’s comprehensive collection of nucleic acids (DNA or RNA), which contains all of its genes. Genomic science or genomics is the study of the genome’s structure, function, evolution, mapping and modifications. Recent breakthroughs in molecular biological techniques have accelerated the pace of high-throughput genome sequencing, genomic characterization and gene expression analysis [27]. The technique of decoding the genome using high throughput next-generation sequencing (NGS) technology comprises the isolation of genomic DNA, the multiplication of DNA using polymerase chain reaction (PCR), the sequencing of the DNA and the assessment of the sequence’s integrity [28]. The sequencing and assembly of DNA, followed by the structural and functional annotation of the gene, permits large-scale investigations into the activities of genes and elucidates the interactions of gene products at the cellular and organismal levels [29]. The field of genomics has been discussed under three categories in the following sections (Figure 2):

2.1. Functional Genomics

Functional genomics analyzes the data generated by complete or partial genome sequencing to describe gene functions and interactions and employs two complementary approaches to the determination of individual genes, viz., forward and reverse genetics [30]. The forward genetic approach investigates a randomly obtained mutant of an interesting phenotype and identifies the responsible gene(s). On the other hand, the reverse genetic approach is the analysis of an organism’s phenotype by disruption of a known gene [31]. The technique of functional genomics helps in unravelling the gene interactions as well as the regulatory networks of genes and this technique employs the below-mentioned methodologies:

2.1.1. Sequencing-Based Approaches

Exploring the expressed gene catalogue has been possible by analyzing the ESTs, i.e., the gene sequence produced from the cDNA clones by the single-pass method [32]. Utilizing the ESTs is a cost-effective as well as rapid method and is thus considered mainly in functional genomics studies. Deokar et al. [33] conducted an EST-based investigation in which they found differentially expressed genes (DEGs) in drought-susceptible and tolerant plants using the suppression subtraction hybridization (SSH) approach to build the analyzed plant’s EST library. After obtaining the ESTs, it gets submitted to the National Center for Biotechnological Information (NCBI), which serves as the source for EST sequencing to reveal the genes that are differentially expressed. Another sequence-based approach is Serial Analysis of Gene Expression (SAGE) which helps in quantifying the abundance of several transcripts together. In the SAGE, sequence tags of small stretch are joined and thereafter sequenced to analyze the gene expression [34] and the identification by these short tags depends on the presence of the EST database for a given species of consideration. SAGE technique is not very applicable to plant systems and thus has been modified as either SuperSAGE or DeepSAGE [35]. Similar to SAGE, the Massively Parallel Signature Sequencing (MPSS) approach has been used to study the long sequences with tags that are affixed to microbeads and then sequenced in parallel, allowing for the analysis of millions of transcripts simultaneously [36]. MPSS technique has high throughput and thus enables identifications to be performed with more specificity.

2.1.2. Hybridization-Based Approaches

Another approach to studying the sequence is an array-based technique, where the hybridization of the DNA that has to be studied is carried out with the cDNA/oligonucleotide probes to assess the gene expression [37]. The limitation of this approach is that designing the probe requires the knowledge of the transcript either in the form of a sequence or a clone. Array-based data exist extensively for model plant species but there is a lack of data for the economically important crop plants and thus unravelling the stress responses utilizing these methods in crop plants becomes a difficult task.

2.1.3. Expansions to Functional Genomics Approaches

Genome-wide association studies (GWAS) are an experimental and statistical examination of a large number of genetic variations across the genome in different organisms (or individuals) to determine whether any variant is related to a trait of interest. GWAS examines the entire genome to identify DNA variations related to the trait of interest [38]. GWAS has been successfully used in deciphering abiotic tolerance in rice [39], soybean [40], wheat [41], maize [42], sesame [43], barley [44], chickpea [45], rapeseed [46], cotton [47,48,49] and sorghum [50,51]. The primary goal of GWAS is to identify genomic areas linked with agronomic or morphological features or any phenotypes that can be markers, genes or quantitative trait loci (QTL) for gene discovery, introgressive hybridization and MAB [52] (Table 1). Advances in genomics and phenomics have resulted in a more precise and comprehensive characterization of QTLs, often referred to as QTLomes [53]. Presently, the QTLome concept is being utilized in specific QTL alleles associated with traits. In addition, numerous statistical methods, such as meta-QTL analysis, have aided in the collection of QTL data from various studies on the same linkage to pinpoint the precise QTL region [41]. This meta-analysis has been applied to study important crop plants such as wheat, soybean, etc. Utilizing the meta-analysis study, Ha et al. [54] identified loci for salt tolerance in soybean on chromosome 3 and used simple sequence repeat (SSR) and single nucleotide polymorphism (SNP) markers to analyze the RIL population (PI 483463 Hutcheson). Sheoran et al. [55] identify the candidate genes of maize for abiotic stress tolerance and utilization in future breeding for crop improvement. Moreover, such meta-QTL analysis also helped to screen the genomic loci of rice for salinity and drought tolerance in different growth and developmental phases—seedling and flowering stages [56,57].
In the 21st century, gene editing has emerged very alarmingly for functional characterization and validation of newly identified genes or genetic regions associated with stress-responsive genes in plants [58]. Success in manipulating a specific gene with a respective function may be achieved by the use of the clustered regularly interspaced short palindromic repeat (CRISPR)—Cas (CRISPR-associated system), which is a more concise, less labor-intensive alternative to traditional methods such as meganucleases (MNs), zinc-finger nucleases (ZFNs), transcription-activator-like effector nucleases (TALENs) [59]. The CRISPR-Cas system of gene editing approach becomes very efficient to characterize the functionality of plant-responsive genes for drought [60,61], salinity [62,63], heat [64] and cold stress [65].

2.2. Structural Genomics

Functional genomics is concerned with the function of genes and their interactions. Structural genomics is concerned with determining the three-dimensional structure of genes to identify, locate and determine their order along the chromosome [60]. Functional and structural genomics studies corroborate the intricate links between sequence and structure, ultimately offering the complete genome, which can aid in the understanding of a wide range of biological issues [66].

2.2.1. Genomic Selection (GS)

Genomic selection (GS) permits the quick selection of better genotypes by utilizing high-density markers distributed across the genome [67]. GS is a novel strategy for optimizing quantitative characteristics; it utilizes marker and phenotypic data from observed populations to assess the impact of all loci [68,69]. Generally, the method of genomic selection relies on two types of datasets: a training set and a validation set [70,71]. The training data set is the reference population and is used for the estimation of marker effects, whereas the validation set possesses the selected candidates that have been genotyped [72].

2.2.2. Genome Sequencing and Mapping

DNA sequencing has provided many details on the sequence, including the whole genome. There are several platforms, such as Roche 454GS FLX Titanium or Illumina Solexa Genome Analyzer, that are said to be NGS platforms that have helped in reducing the sequencing cost as well as time in comparison to conventional sequencing methods such as the Sanger method [73]. The sequencing method helps in developing improved varieties of crops by sequencing and resequencing processes. To date, the genomic sequence for several crop plants such as rice, wheat, maize, sorghum, soybean and tomato has been published. Apart from these crop plants, the sequence of model plants such as Arabidopsis thaliana and Brachypodium distachyon has also been published [74]. Genome sequencing provides detailed data on the features of genomes (coding as well as non-coding genes), GC content, repetitive elements as well as regulatory sequences [75]. Although genome sequencing provides important details for improving crops using molecular breeding, its usefulness is limited to species that have a smaller genome. To facilitate a complex genome study, another technology that is chromosome-specific has helped in developing Bacterial Artificial Chromosome (BAC) libraries to help in studying the complex genome. Mapping of 1 Gb chromosome of wheat has been possible with the help of the chromosome-by-chromosome approach only [76]. Mapping compiles genetic mappings into physical contigs as well as providing a framework for the assembly of sequences into the whole genome, and in the absence of a reference genome sequence, this BAC-end shotgun sequence gives details of genome evolution as well as structure [77,78]. Another interesting method for developing a whole genome sequence has been carried out by detecting the QTL using a methodology named QTL-seq. A QTL is a polymorphic locus that differentially affects the trait. QTL mapping is the technique of utilizing DNA markers to generate linkage maps and identify genomic areas linked with certain characteristics. QTL mapping is used to characterize the organization and evolution of the chromosomes [79,80]. So far, several QTLs have been reported for tolerance to drought [81,82], salinity [39,83,84,85,86], heat [87,88], cold [46,89], etc. (Table 1).
Table 1. Important QTLs/markers identified for abiotic stress response in field crops.
Table 1. Important QTLs/markers identified for abiotic stress response in field crops.
PlantsQTLs/MarkersChr. LocationMethods UsedAbiotic StressesReferences
RiceOsHKT1;1Chr 1GWASSalinity[39]
qWUE.STI6Chr 6Linkage mappingDrought[80]
SaltolChr 1Linkage mappingSalinity[87]
qCTBB2
qCTBB3
Chr 2
Chr 3
Linkage mappingCold[89]
qSTS4Chr 4QTL-seqSalinity[90]
SoybeanAX-93897192Chr 19GWASPhosphorus efficiency[40]
qGI10-1Chr 10GWASDrought[79]
qSFT_3-38,
qSFT_7-3
Chr 3
Chr 7
Linkage mappingFlooding[91]
qST6
qST10
Chr 6
Chr 10
Genotype-based sequencing (GBS)Salinity[92]
WheatMQTL1D.4
MQTL2D.5
MQTL3A.1
Chr 1D
Chr 2D
Chr 3A
MetaQTLDrought stress
Heat, Salinity
Waterlogging
[41]
QNa.asl-2AChr 2AGenotype-based sequencing (GBS)Salinity[81]
YIELD_MQTL4B.2_DChr 4BMetaQTLHeat, Drought[85]
qWMs108_7-1Chr 7-3ALinkage mappingDrought[93]
QSpad3.ua-1D.5Chr 1DGWASWaterlogging[94]
QMrl3B(T2|T1)Chr 3BLinkage mappingSalinity[95]
MaizeZm00001eb013650Chr 1-10GWAS + RNAseqSalinity[42]
qPOD2bChr 2Genome-Wide Association Study (GWAS)Cold[96]
RapeseedSA07_23415428Chr SA07GWASFreezing[45]
qDSI_SL-11-3 qDSI_RL-11-1 qDSI_RL-11-4 qDSI_SL11-3Chr C01Linkage mappingDrought, Freezing[97]
qRRL.A3bChr A03Linkage mappingWaterlogging[98]
BarleyQcRWC.3H_2.1
QcWC.3H_1 3H
Chr 3HLinkage mappingDrought[76]
HORVU2Hr1G111780.3Chr 2HLinkage mappingSalinity[82]
qSLS-4Chr 4HLinkage mappingSalinity[88]
QBIO.2HChr 2HGWASWaterlogging[99]
CottonqtlCSI01Chr 3Composite interval mappingDrought[47]
qGR-Chr4-3, qFER-Chr12-3, qFER-Chr15-1Chr 4
Chr 12
Chr 15
Linkage mappingSalinity[48]
qEC_A02_ck
qFW_A06_150.1
Chr 2
Chr 6
Genotyping by Sequencing (GBS)Salinity[49]
qFSHa1Chr 15Composite interval mappingHeat[86]
SorghumqPH-6
qMC2-9
Chr 6
Chr 9
Genotype-based sequencing (GBS)Excess soil nitrogen[50]
qTB45_4.SChr 4Linkage mappingSalinity[51]

2.2.3. Molecular Marker Resources

DNA markers are short areas of DNA sequences that have the ability to identify variations in a population’s DNA or polymorphisms (base deletion, insertion and substitution), including base deletions, insertions and substitutions. DNA markers are also known as genetic markers [11]. Molecular markers aid in tagging genomic traits such as pathogen resistance, abiotic stress tolerance, quantitative analysis, etc. Recent advancement in this resource has provided a new horizon for the genetic improvement of traits for stresses such as drought, salt, etc. [11]. To date, several molecular markers have been reported that help in identifying polymorphism in plants, and these markers include random amplified polymorphic DNA (RAPD), restriction fragment length polymorphism (RFLP), amplified fragment length polymorphism (AFLP), SNP, SSR and sequence-tagged sites (STS) [11,100]. Restriction fragment length polymorphism (RFLP) is the most basic marker that helps in identifying the polymorphism arising due to mutation or deletion/insertion leading to either formation/deletion of endonuclease recognition sites in restriction fragment length [101]. Another marker, RAPD, which is generated via random primers, identifies complementary sites at a short distance within the genome, while AFLP combines the restriction digestion as well as the PCR amplification and thus helps in identifying the linkages [102]. The SSR or microsatellite markers are tandem repeats of short mono-, di-, tri- and tetra-nucleotides and help in measuring the genetic diversity among species and also differentiate alleles that are homozygotic and heterozygotic between the lines from the same origin [103]. The SNPs are used for the characterization of germplasm as well as gene mapping. Due to their high abundance, codominance and sequence tagging they help in understanding complex traits utilizing microarrays such as Affymetrix GeneChip. Marker-assisted selection (MAS) is a genomic approach to identifying and breeding associated allelic markers [104]. During the process of marker-assisted selection, a characteristic of interest is chosen on the basis of a marker that has been associated with a particular or multiple abiotic stress [105,106]. Previously, success in MAS for abiotic stress tolerance was lagging due to the limited availability of genomic data. Genome databases and datasets that are very valuable in the construction of SSRs and SNP markers have been produced thanks to recent improvements in high-throughput DNA sequencing and genotyping technology [107]. Such availability of various high-throughput molecular markers and genome sequencing technologies leads to genomics-assisted breeding [108] and SNP difference-based haplotype mapping [109] to improve crops with stress tolerance properties.

2.3. Comparative Genomics

Comparative genomics is the science of comparing entire genomes or parts of genomes to find out basic biological similarities and differences as well as investigating evolutionary relationships between organisms [110]. Comparative genomics compares biological sequences by aligning them and detecting conserved sequences. Thus, studies of comparative genomics have revealed considerable synteny in related species [111]. Moreover, as comparative genomics can detect small-scale changes within different genomes, comparative studies of protein-coding regions and their consequences on protein structure and function identify important regulatory elements within DNA [112]. Comparative genomics gave rise to the “genome zipper” concept that helps in determining the virtual gene order within the partially sequenced genome. Genome zipper links the annotated and fully sequenced genome of sorghum, Brachypodium and rice with the data of less-studied species to predict the gene order and organization of the gene [113,114].

3. Transcriptomics

The whole collection of transcripts that are present in a cell or organism is referred to as the transcriptome, and the study of the transcriptome is referred to as transcriptomics [115]. It mainly helps in finding gene transcripts or RNA that are associated with a plant’s phenotypic expression under different environmental conditions [116]. A variety of methods, including DNA microarrays, SAGE or high-throughput technologies relying on NGS, may be used in the process of conducting a transcriptome study such as RNA sequencing (RNAseq) and digital gene expression (DGE) [117,118,119,120,121]. To date, transcriptomic analysis has identified many stress-responsive genes, and their mode of expression under abiotic stress conditions in many plants including wheat [122], maize [123], rice [124], barley [125], sorghum [126], cotton [127] and soybean [128]. Moreover, transcriptome analysis in tomatoes has revealed the discovery of regulators of SGA pathways such as GLYCOALKALOID METABOLISM (GAME) 9, also called JRE4, an AP2/ERF transcription factor in response to various abiotic stresses [129]. Meta-transcriptome analysis in rice revealed the expression of 6956 abiotic stress tolerance (ASTR) genes, some transcription factors (TFs) and a few functional modules such as cis-motifs. Out of the expressed ASTR genes, 1% were found to be colocalized within the trait-associated QTL and over 65% of the genes in the tolerant genotypes showed differential expression under saline, high temperature and drought stress environments [130]. Similar to this, Azzouz-Olden et al. with the help of transcriptome analysis and RNA sequencing data, reported two sorghum lines, viz., SC56 (drought-tolerant) and Tx7000 (drought-sensitive), the former showed overexpression of antioxidant genes such as SOD1, SOD2, VTC1, MDAR1, MSRB2, ABC1K1, regulatory factors such as CIPK1 and CRK7 and repressors of senescence, i.e., SAUL1 [131]. Initially, microarray experiments detected the co-expressed genes during abiotic stress conditions. However, the most significant disadvantage of using microarray analysis is that information about the transcripts cannot be obtained from the genome as a whole. As a consequence of this, the studies using microarrays are unable to fully decode the regulatory gene networks that are involved in the abiotic stress response of plants. Fortunately, recent developments in molecular methods have made it possible to construct high-throughput procedures that are based on sequence-based methodologies [132]. The most widely used method for analyzing transcriptomes is called RNAseq. This is because it provides comprehensive coverage of the genome and ubiquitous expression of transcripts [133]. RNAseq detects DEG and consequently deciphers the regulatory mechanism of plant abiotic stress tolerance. Tiwari et al. [134] carried out transcriptome analysis using Illumina NextSeq500, and the outputs exhibited DEGs in different parts of plants. For example, there was up-regulation of 761, 572 and 688 DEGs and down-regulation of 280, 292 and 230 DEGs in the shoot, root and stolon, respectively. Moreover, fewer DEGs such as Myb-like DNA-binding protein, WRKY transcription factor 16, glutaredoxin family protein, malate synthase, CLE7, 2-oxoglutarate-dependent dioxygenase FLOWERING LOCUS T BTB/POZ domain-containing protein, F-box family protein and aquaporin TIP1;3 responsive to N-deficiency were also found to be expressed [134]. A list of potent genes associated with plants’ abiotic stress response is presented in Table 2 and those are highly important for further transcriptome analysis under varied and combined stress environments.

4. Proteomics

The study and characterization of the whole sequence of proteins that are present in a cell, organ or species at a particular point in time are referred to as proteomics. Changes in plant proteomes are an extremely significant topic to research since proteins are the primary key regulators of the plant’s stress response. Various stages, such as development, cellular differentiation and the cell cycle, as well as distinct environmental factors, such as abiotic stressors, may cause the same genes to express themselves in a variety of different ways [155]. As a result, various sets of proteins are produced by cells depending on the environment they are in. Because of this, some proteins might be regarded to be unique biomarkers for certain environmental factors such as abiotic stressors [156]. Therefore, proteomics research may lead to the discovery of these putative protein markers and variations in their abundance may be correlated with quantitative shifts in specific physiological indicators related to a person’s capacity to withstand stress [157]. Under environmental stresses, proteomics allows for the detection and characterization of proteins, as well as their activity profiles, post-translational modifications (PTMs) and protein–protein interactions [158,159,160,161,162]. A fundamental framework for a comparative analysis of the drought stress proteome changes in cereal crops such as wheat, rice, maize, barley, sorghum and pearl millet has been provided by a comprehensive study of the existing proteomics data sets [161]. The two most common laboratory techniques are used extensively in proteomics studies—protein electrophoresis and protein identification using mass spectrometry. Conventional gel-based protein electrophoresis methods include techniques such as two-dimensional electrophoresis (2-DE) and Difference In-Gel Electrophoresis (DIGE) [160,162]. Gel-based methods’ primary benefits consist of their ease of use, repeatability, broad molecular mass coverage and sensitivity to the detection of post-translational changes [163]. However, the 2-DE method possesses some inherent limitations such as reproducibility, detection of less abundant and hydrophobic proteins, identification of basic proteins, co-migration of proteins and presence of exceedingly large or small proteins [164]. Multi-dimensional Protein Identification Technology (MudPIT), a non-gel method of both qualitative and quantitative proteomic analyses has been popularly used to overcome the limitations of 2-DE. The mass spectrometry (MS) approach includes techniques such as liquid chromatography-MS (LC-MS/MS), Ion Trap–MS (IT-MS) and matrix-assisted laser desorption/ionization–MS (MALDI-MS). Fluorophore-tagged protein immune-precipitation and label-free MS-based quantification approaches have been developed [165] to achieve a higher level of precision in the identification of low-abundance signalling and regulatory protein complexes. In addition to this, Laser-Capture Micro-dissection, commonly known as LCM, has been used for the identification of tissue- and cell-specific proteins that play a crucial role in the response of crops to environmental stresses [166]. In the process of regulating a plant’s response to environmental stress, post-translational protein modifications, such as phosphorylation, redox and glycosylation, perform an essential role. The technique of phosphoproteomics, which is based on mass spectrometry, is an extremely useful instrument for determining the in vivo kinase activity of proteins. Immobilized metal affinity chromatography (IMAC) and immunoprecipitation utilizing antibodies towards phosphorylated amino acids study are two more methods that have allowed the discovery of hundreds of novel in vivo phosphorylation sites [163]. Stable isotope labelling by/with amino acids in cell culture (SILAC) and isobaric tags for absolute and relative quantification (iTRAQ) seem to be more sophisticated techniques that may monitor changes in the specific kinase domain. So far, various proteomic studies revealed abiotic stress tolerance of the plant, such as in drought [167,168], heat [158,169], chilling [159,170], salinity [156,171] and waterlogging [172,173]. Comparative proteome analysis for Medicago sativa cv. Zhongmu-1 and Medicago truncatula cv. Jemalong A17 roots were experimented with by using 2D gel electrophoresis and mass spectrometry under salt stress and revealed the abundance of 93 and 30 proteins was affected, while tandem spectrometry revealed expression of 60 and 26 proteins, respective cultivars and these proteins have been supposed to play important role in salinity stress [174]. Quantitative proteome analysis by Zhu et al. [175], revealed the expression of 179 salt–alkali responsive proteins and it was suggested that these proteins have a role in the tri-carboxylic acid (TCA) cycle, oxidative phosphorylation, glycolysis, sucrose metabolism as well as being involved in reactive oxygen species (ROS) homeostasis. Another, study considering the comparative proteome analysis revealed the expression of 37 proteins under salinity stress and these proteins were identified with their role in the process of photosynthesis, stress response and phytohormone biosynthesis [176].

5. Bioinformatics

Bioinformatics is a wide multidisciplinary field that encompasses both theoretical and practical methods to comprehend, generate, analyze and disseminate biological information. Bioinformatics is the first link between biological data and the application of computational methods [177]. To analyze and alter resources from databases [178], computational tools and methodologies provided by bioinformatics may be used. This may result in the production of novel findings or hypotheses, even though these may need proper evaluation [179]. The Integrative Omics–Metabolic Analysis (IOMA) platform was developed to combine proteomic data with cellular metabolic information. Thus, with time, the field of bioinformatics has evolved and provides the platform for interactive “Omics” technologies [180]. Several studies have involved bioinformatics tools to understand the overall underlying mechanism under abiotic stress conditions [178,180,181]. Although bioinformatics provides the platform for integrating the omics approaches by evolving with more and more novel technologies to provide in-depth knowledge of the complex data set to understand the individual analyses as well as a comparative analysis [182]. For example, the modern genome editing tool CRISPR-Cas9 approach can be utilized for analyzing different abiotic stress conditions and improving crop plants but it requires appropriate development of genomics as well as bioinformatics pipelines to provide more detailed information on how this process can be done [62,183]. A number of software packages such as E-CRISP, TIDE, CHOPCHOP and CCTop have been developed to envisage and select CRISPR-Cas9 for genome editing [62,184]. Recently, for agricultural communities, the gene ontology (GO)-analysis toolkit earlier called AgriGO has been rereleased with an updated tool under the name AgriGO2.0. This toolkit provides additional bioinformatics analysis such as singular enrichment analysis (SEA), transfer IDs by BLAST (BLAST4ID), parametric analysis of gene set enrichment (PAGE), etc. [185].

6. Epigenetics-Aided Epigenomics

The term ‘epigenetics’ was originally coined by Waddington in the middle of the 20th century by combining genetics and epigenesis to explain the phenotypic features of the plant due to the genetic interactions and its products [186]. In general, epigenetics refers to the non-heritable/heritable changes, cell division (mitotic or meiotic), methylation pattern of cytosine (C) nucleotide in DNA and histone protein modification which is concerned with various epialleles in the genomic region [187]. Epialleles have been reported as a major factor for phenotypic diversity [188]. Post-translational modifications are another major player that affects gene expression and makes changes in phenotypes. The study of phenotypes in a species due to epigenetics modification refers to epigenomics [189]. Acetylation, sumoylation, phosphorylation, methylation, ubiquitination, glycosylation, carbonylation and ADP ribosylation are the major post-translational modifications to modify the histone tails [190]. Apart from post-translational modification, chromatin remodelling is another factor that contributes to epigenetic modification in crop species. In addition to the modification, it is related to some sort of trans-generational inheritance, which affects the accessibility of DNA transcription factors (TFs) [191]. Simply, the transmission of epigenetic modification traits from one generation to the next is known as “transgenerational memory” which is independent of DNA sequences [192]. Alternatively, epigenetics also refer to the changes in gene activity without any alteration in DNA sequence. Epigenetic mechanisms result in the modification of chromatin structure which regulates mRNA accumulation at the transcriptional level [193]. Recent research showed that epigenetic mechanisms play a critical role in the regulation of plant genes’ expression under various abiotic stress conditions [194,195]. Changes in environmental factors such as temperature, day length, ultraviolet (UV) radiation, availability of water and soil salinity in plants lead to modifications in the (de)methylation pattern of the coding regions in many stress-responsive genes which consequently regulate their expression [193,194,195,196]. Nevertheless, in addition to DNA methylation, recent evidence showed that histone modification, small regulatory RNA (sRNA) and long non-coding RNA (lncRNA) associated regulatory pathways are other adaptive mechanisms that regulate gene expression under stress conditions [197]. In A. thaliana, four DNA methylases have been characterized: ROS 1 (REPRESSOR OF SILENCING 1), DME (DEMETER), DML2 (DEMETER LIKE 2) and DML3 (DEMETER LIKE 3). The DNA demethylase (glycosylase) replaces the 5- methylated cytosine with unmethylated cytosine through a base excision repair mechanism [198]. AtDME has been reported as an epigenetic regulator that is essential for maternal allelic expression of the MEA (MEDEA) gene which encodes a repressive H3K27 methyltransferase, in the embryo and endosperm central cell [199]. DEMETER LIKE 2 and DEMETER LIKE 3 prevent hypermethylation at specific genomic regions. These two are expressed in vegetative and reproductive tissue [200]. The ROS1 DNA demethylase has been identified as targeting particularly TEs in a genomic region closer to protein-coding genes, revealing activation of nearby genes through the demethylation process [201]. It is now recognized that short RNAs may also guide DNA methylation, a process known as the RNA-directed DNA methylation pathway (RdDM) through the activity of two kinds of RNA polymerases—PolIV and PolV [202]. A number of studies have shown that, when plants are exposed to heat stress, DNA methyltransferase (MET1, DRM2 and CMT3—which are the largest subunits of PolIV and PolV) is upregulated. The upregulation of these methyltransferases leads to more methylation under heat stress conditions in Arabidopsis [203].
It is well documented that modifications in histones are reversible, as the overlapping process of methylation and acetylation are very much important for stress response in plants. The acetylation process is governed by Histone acetyltransferase (HATs) and Histone deacetylases (HDACs) that help the plant adapt to changing environments [204]. In Arabidopsis, a total of 12 HAT genes represent four HAT families (GENERAL CONTROL NONDEREPRESSIBLE5 [GCN5]-like [GCN5/HISTONEACETYLTRANSFERASE OF THE GNAT FAMILY {HAG}1, 2, 3}, p300/CBP [CREB binding protein]-like [HISTONE ACETYLTRANSFERASE OF THE CBP FAMILY {HAC}1, 2, 4, 5, 12], TAFII250-like [HISTONE ACETYLTRANSFERASE OF THE TAFII250 FAMILY {HAF}1, 2] and, MYST-like [HISTONE ACETYLTRANSFERASE OF THE MYST FAMILY (HAM) 1, 2] [202,205]. For example, in soybean, the high salinity stress leads to methylation and histone modifications required for the activation/repression of stress-responsive TFs [206]. Salinity stress has a significant impact on genome-wide histone modifications and methylation for providing tolerance against such osmotic stress [204]. GCN5 was first characterized in maize root tissue under salt stress response. Up-regulation of cell-wall-related genes such as ZmXYLOGLUCAN ENDOTRANSGLUCOSYLASE/HYDROLASE1 and ZmEXPANSIN B2 are linked with acetylation of H3K9, which is a histone protein (H3) with lysine (K) on the 9th position, in both the promoter and coding region. Mutant gcn5 exhibits salt sensitive response in maize because of the cell wall integrity [207]. In another study, Arabidopsis GCN5 plays an important role in heat tolerance through H3K9/k14 acetylation process in the promoter sequence of the ULTRAVIOLET HYPERSENSITIVE6 and, HEAT SHOCK TRANSCRIPTION FACTOR A3 genes [208]. Under cold stress, the C-REPEAT BINDING FACTOR (CBF)-COLD RESPONSIVE (COR) pathway augments to plant for survival. Induced expression of TFs, such as CBF family proteins under cold stress, bind over the COR gene promoter to facilitate its COR expression [209]. Heat shock proteins (HSPs) play a wide role in heat stress tolerance in plants. Accumulation of H3K4me3 and H3K9Ac has also been reported on HSP70, HSP22, HSP18 and APX2 [210]. Under different abiotic stresses, drought is well documented as a major abiotic stress factor that makes histone alterations [210,211]. NCED3 (NINE CISEPOXYCAROTENOID DIOXYGENASE 3) is a well-reported gene that responds to ABA synthesis under water scarcity conditions. Accumulation of H3K4me3 in NCED3 gene regions helps with drought resistance in plants accompanied by NCED3 gene expression [193]. Overall histone modification is a complicated process that plays an important role in epigenetic regulation. For example, histone H2A.Z is essential for the repression of unwanted transcription of drought-inducible genes’ expression in Arabidopsis [212]. On the other hand, H2A.Z is contributing to the grain yield of the Brachypodium under heat stress [213]. These contemporary data for H2A.Z suggest the diverse role of epigenetic-mediated gene regulation through structural divergence or PTMs of histone proteins that define a crucial topic to understand [213].

7. Metabolomics

Metabolites are concerned with the quantitative, qualitative and dynamic study of all endogenous, low molecular weight compounds (less than 1000–1500 dalton) within the organismal cells, tissue or organs and perform essential activities in a spatio-temporal manner. The plant kingdom contains approximately 0.2–1.0 million distinct metabolites whose concentrations vary from one species to another. These compounds vary from each other by their classes, physiochemical properties, chemical structure and polarity level [214]. A quantitative and qualitative study of plant metabolites responding to several environmental and biotic stress is not only a descriptive feature of plants but also reflects the genetic and biochemical background in stress-responding plants which brings the difference among plant species according to their level of tolerance and adaptation in particular stress [215,216]. Metabolomics has its advantage over the field of genomics, transcriptomics and proteomics. Metabolites are the downstream product of gene and protein activity that define the effect on living phenotype and other physiological activity [217,218,219]. Metabolites can be classified into two classes—primary and secondary metabolites. Primary metabolites are essential for plant growth and play a wide role in physiological activity [220], while secondary metabolites are essential for defence response under a wide range of abiotic stresses [221]. Plants have been studied for their ability to adapt to a variety of environmental conditions by looking at how they modify metabolites such as osmoprotectants (proline, glycine betaine, trehalose, etc.) and antioxidant enzymes (superoxide dismutase, peroxidase, ascorbate peroxidase, catalase, glutathione reductase, guaiacol peroxidase, etc.) (Figure 3) [222,223,224,225]. Primary metabolites such as sugars, amino acids and TCA (Krebs) cycle intermediates (citric acid, α-ketoglutarate) are directly involved in plants’ normal growth and development; whereas the secondary metabolites are genera-specific and condition-specific. Thus, the total metabolite profile of a given plant species indicates how many regulatory systems, such as gene expression and gene–protein interaction, have been integrated. Under any adverse environmental conditions, plants exhibit an array of responses that lead the particular stress tolerance, all of which are associated with metabolic modifications. Therefore, the study of stress-associated changes in metabolites is given particular attention in the 21st century [226,227]. Bioactive chemicals including antioxidants, signalling compounds, biosynthesis intermediates for cellular structures and storage compounds are produced when a metabolic pathway is activated. The production of these compounds, in turn, regulates or activates other compounds or intermediates that can feedback activate or inactivate different metabolic steps [228]. Polyamines are one of the well-reported secondary metabolites that contribute to plant growth and provide resistance under abiotic stress conditions in angiosperm plants. Spermidine (SPD), spermine (SPM) and putrescine (Put) are the common well-known polyamines found in almost all land plants. In cotton, the Put gene is expressed under the regulation of the Arginine decarboxylase2 (ADC2) gene promoter in a salinity environment [229]. It has been reported that drought favors the accumulation of several secondary metabolites such as alkaloids, terpenes and complex phenols. For example, drought stress induces the phenolic content in Hypericum polyanthemum (hypericum), Salvia officinalis (garden sage), rice and barley. Likewise, the monoterpenes or terpenoids amount also increases in Barley and Salvia officinalis under drought stress [230,231,232]. Similarly, trehalose, which is a non-reducing disaccharide, plays a beneficial role to maintain membrane integrity and stabilization of macromolecules under drought conditions. The rate of photosynthesis is also increased under overexpression of trehalose and PSII is protected against photo-oxidation through trehaloses [233]. Variation of the polyamines profile has been observed in salt-tolerant rice and tomato species under diverse stress conditions [234]. Heat stress also induces the overproduction of flavonoids, phenylpropanoids and phenolic metabolites through the upregulated overexpression of the associated genes [235]. In leaves of tomatoes, heat stress factor: HsfB1 suppression or overexpression increases thermo-tolerance capacity in the plant. The overexpression of HsfB1 leads to the accumulation of phenylpropanoid products and the pathway of flavonoid in addition to various isomers of caffeoyl quinic acid [236]. In an aspect of salinity stress, omeprazole (proton pump inhibitor) helps to enhance the salt stress tolerance in tomato and makes several hormonal changes in leaves such as abscisic acid increment while decrementing auxin, cytokinin and gibberellic acid levels. Additionally, alkaloids and sesquiterpenes are conjugated with polyamines under the response of Omeprazole [237]. Under oxidative stress, ROS are overproduced in the plants, causing oxidative deterioration of cellular macromolecular structures such as DNA, RNA, protein and lipids [238]. To overcome this ROS-mediated oxidative stress, a powerful ROS scavenger, glutathione, is an essential antioxidative metabolite [239]. Various other metabolites such as proline, polyphenols, ascorbic acid, carotenoids and tocopherols act as nonenzymatic antioxidant molecules [240]. Under salinity stress, the cellular antioxidant level is increased [241]. The ascorbate–glutathione cycle is an important biochemical pathway as well as the potent non-enzymatic antioxidant system that is used to detoxify several toxic compounds and ROS in living cells generated under abiotic stresses. The ascorbate–glutathione pathway detoxifies the methylglyoxal (MG), which is a highly reactive cytotoxic compound. MG is accumulated under adverse abiotic stress conditions [242]. Several other primary metabolites such as organic acids have their importance for different abiotic stress. For example, malic acid provides drought resistance in different plant species such as tropical grasses, cotton and spare grasses [243]. Overexpression of galacturonic acid reductase in potato genotypes helps to increase the content of ascorbic acid and water stress tolerance [244].
Alteration in many metabolic pathways in plant cells and organs contributes to the balance of the metabolite profile in the organism. Presently, several detections and analytical separation techniques are used in combination for the visualization of an organism’s metabolomic profile. With the advancement in MS, nuclear magnetic resonance (NMR) and chromatographic techniques, a large number of metabolite analyses and studies have become quite handy for scientists [245,246]. Ghatak et al. [246] provide detailed information about plant metabolomic methods, libraries used in the analysis, data mining and processing, chemical identification and the limits of metabolomics. LC-MS/MS and gas chromatography-MS (GC-MS) are popular techniques for metabolomics study because of their unparalleled level of sensitivity and extensive coverage of huge metabolites [136]. Development of new analytical techniques such as GC, LC coupled to MS, NMR, Fourier Transform Infrared spectroscopy (FTIR) or capillary electrophoresis (CE) provides a more accurate description of metabolite interactions in a given plant species [247]. Characteristics of several stress-responsive metabolites can be detected and quantified concurrently using mass-spectrometry-based metabolomics methods [248]. However, the molecular heterogeneity and broad-spectrum metabolome significantly hamper compound identification and meaningful measurement. Ion suppression, fragmentation and the existence of isomers may further complicate the simultaneous measurement of multiple metabolites within complex phytochemical mixtures. To facilitate high-quality reporting of data derived from LC and GC-MS-based metabolomics, Alseekh et al. [248] recommended encompassing the preparation of samples, reproduction and randomization, quantization, restoration and recombination, ion suppression and identifying incorrect peaks. Moreover, a few more techniques such as targeted analysis, metabolic fingerprinting and metabolite profiling are utilized for speed, improved comprehensiveness, better resolution, the throughput of analytical assays and miniaturization equipment [249]. Meanwhile the combination of several techniques such as LC-MS/MS, CE-MS/MS, ultra-high performance liquid chromatography-quadrupole time-of-flight mass spectrometry (UHPLC-TOF-MS) or nuclear magnetic resonance [such as LC-NMR or LC with PhotoDiode Array Detection-Solid phase extraction-NMR-MS/MS (LC- DAD-SPE-NMR-MS/MS)] joint with bioinformatics tool is of great assistance for the study of the natural products of plants and clears the vision of comprehensive profiles of metabolites [165,250].

8. Proteogenomics, Lipidomics, Ionomics and Interactomics

Proteogenomics is a new and integrative approach that combines the technological advancement of genomics, transcriptomics and proteomics together [251]. The goal of typical proteogenomics research is to catalogue the proteins that are already being expressed in the cell by combining high-throughput NGS data with MS-based proteomics applications [252]. Particularly, proteogenomics is useful for the identification of proteins by integrating genomic, transcriptomic and MS data of the same crop species and/or sample. Thus, the integrative proteogenomics approach has identified novel proteins and provided a hardcore knowledge of the genes’ regulatory expression and cell signalling for abiotic stress tolerance [253]. Proteogenomics has shed light on the mechanisms of plants’ response to abiotic stress and adaptation to changing the environment. Phosphorylation of protein molecules is to be considered in the proteogenomics phenomenon [254]. Under salinity stress, the primary and secondary transporters depend upon phosphorylation for being active and to regulate the sodium ion (Na+) but, still, the information on responsive kinases is not explored much for most economically important crops [255].
Lipidome describes the whole profile of lipids in the cellular, tissue or organ level of an organism. Lipidomics is an emerging field of science that studies the structure and function of the lipidome as well as their interactions with other lipids, proteins and metabolites [256]. Lipids play a dual role in plants’ abiotic stress response. Besides being important signalling mediators, lipid molecules play significant roles in the alleviation of stress [257]. Signalling lipids such as fatty acids, sphingolipids, diacylglycerols, lysophospholipids, phosphatidic acid, inositol phosphate, oxylipins and N-acylethanolamine are quickly synthesized under stress conditions. Simultaneously, lipids are also involved in the remodelling of cell membranes under abiotic stress and mitigate cell damage [258]. Improvement of analytical methods, particularly liquid chromatography and mass spectrometry, enables systems-level analysis of lipids and their interacting partners [259]. The tools of lipidomics are categorized into two broad categories—MS prediction tools and structure-drawing tools [260]. The MS-based methods have been shown to be highly efficient for the characterization and quantification of molecular lipids. The different MS techniques are categorized into three groups—global lipidomic analysis (GLA), targeted lipidomics analysis (TLA) and novel lipid discovery (NLD). The GLA is a high-throughput method of identification and quantification of cellular lipid species. It is particularly useful to analyze and decipher pathways and networks associated with lipid metabolism, trafficking and homeostasis [261]. TLA employed LC-MS and LC-MS/MS based on the identification of lipid molecules, whereas NLD uses LC coupled with MS and is involved in the finding of novel lipid species. Methods such as MALDI-TOF MS coupled with thin-layer chromatography (TLC) are presently being used for imaging lipids from tissue slides [262]. A particular lipid profile of a given crop species under certain abiotic stress conditions can act as a lipid biomarker or lipidotype. Sun et al. [263] identified the leaf lipid profile in Begonia and its alterations under heat stress, which helps to understand the stress adaptive mechanism in plants. The lipidomics process may be paired with MS-based methods and robust GWAS (lipidomics-aided GWAS or LiA-GWAS) to find membrane lipid remodelling-associated genes and likely relationships that may be exploited to generate stress-tolerant plants [256]. However, because of its complexity and specificity, lipidomics studies are complicated and quite challenging. Moreover, due to the great diversity in lipid classes, the structural identification of lipids is a complicated process. Hopefully, in the near future, the development of comprehensive lipidomics technologies developed will expand the sphere of plant lipidomics and shed more light on the involvement of lipid molecules during abiotic stresses.
In ionomics, elements are profiled in a high-throughput manner and deal with the studies of inorganic components, mineral nutrients and trace element composition of a living being [264]. Genomic data aided with ionomics, particularly in combinations with forward and reverse genetic approaches, can detect cellular changes during abiotic stress conditions [265]. Recent reports showed that ionomics studies revealed the mechanisms of ion uptake under abiotic stresses in plants [266,267,268]. Moreover, transport, compartmentalization and exclusion of ions during adverse environmental conations were also monitored by ionomics. Although ionomics is a new field and there are only a few reports of ionomics studies under abiotic stress available, the trends are growing for ionomics studies [266,267,268]. In a general sense, ionomics are related to the ion content of an organism (here in plants) that is required for its growth and developmental processes under different environments and growth stages [269]. In plants, ions are classified under two major categories: essential nutrient ions (macro and micronutrients) and non-essential nutrient ions [270]. Apart from this, some of the non-essential ions harm normal physiological conditions in the plant. For example, the sodium ion (Na+) is a well-reported causal factor of salinity stress in many glycophytes, including rice, in which the productivity of grain yield is drastically affected [271]. The role of ion transport regarding how Na+ is regulated by primary and secondary transporter under salt stress in both tolerant and sensitive rice lines has been identified [151]. There is also a positive role of an essential ion such as Ca2+ that antagonizes the salt stress (Na+) effect in rice salt-tolerant landraces Nona Bokra and sensitive cultivar IR-64 [272]. In terms of essential ions, phosphorus (P) is one of the most important ions that help in major biological activities in the plant system [273]. The importance of phosphorus nutrients has been revealed in EMS-induced mutants compared to Nagina N-22 rice under low and normal P soil conditions and could have accounted for improved physiological and biochemical activity under low P field conditions [274].
Interactomics is the study of the interactions and the consequences of those interactions among the biomolecules in a cell [275]. Complex physical, biochemical and functional interactions between DNA, RNA, proteins, lipids and tiny metabolites mediate cellular processes. The term interactome most usually refers to a network of protein–protein interactions (PPIN) [276]. However, another important interactome is the protein–DNA interactome which is also known as the gene-regulatory network. Thus, the plant interactome constitutes TFs factors and chromatin regulatory proteins with the genes of their target site [277]. Since protein–protein and protein–DNA interactions are central to all cellular processes, understanding these interactions in both normal and stress conditions facilitate the identification of underneath regulatory mechanism of stress tolerance. In recent years, many different technologies have been developed for interactomics study. All these technological approaches are broadly categorized into three heads: in silico, in vivo and in vitro. The in silico methods are carried out by computer simulation and consist of text mining and computational analyses. The in vivo methods are performed on intact living individuals. The yeast two-hybrid (Y2H), protein-fragment complementation assay (PFCA) and protein–protein interaction trap (MAPPIT) are common in vivo approaches to interactomics. The experiments of in vitro methods are performed outside a living organism and under controlled conditions. The in vitro approaches of interactomics include techniques such as tandem affinity purification-MS (TAP-MS), protein microarray and the luminescence-based mammalian interactome (LUMIER) tools [278]. Though there are few instances of involvement of interactomics in plants for abiotic stress tolerance, its proper high-throughput applications are still underway [279,280].

9. Phenomics

Phenomics deals with omics of phene (phenotypes), a product of genes and utilizes high-throughput analysis of organismal phenotype by evaluating the morphological, physiological, and biochemical traits [281]. In the case of plants, phenomics correlates growth, performance and composition with genetic, epigenetic and environmental factors. Therefore, phenomics integrated with other omics unveils cellular biochemical or bio-physical networks that result in the final desirable phenotype [282]. As most phenotypic traits are determined by the interactions between genes and the environment (G × E), collections of large numbers of phenotypic data across multiple environmental conditions revealed the relationships between phenotypic traits and prevailing abiotic stresses [283,284]. Both forward and reverse phenomics strategies were employed for the analysis of various traits. Forward phenomics uses high-throughput and fully automated phenotyping tools for the rapid identification of interesting, unique or desired traits [285], whereas the reverse phenomics method investigates the selected traits in detail and subsequently discovers the underneath mechanism [286]. In fact, phenomics can be used at the cellular and tissue level and also used on a bigger scale, i.e., plant organ, whole plant, plant community in the field, the vegetation of the particular area and ecosystem basis (Figure 4).
In recent years, for large-scale precise, accurate and rapid trait phenotyping, high-throughput non-invasive imaging technologies became quite popular [287,288,289]. High-throughput phenotyping (HTP) estimates the quantification of chlorophyll fluorescence, the water content in the leaf and other associated plant parts and geographical parameters [290]. HTP encompasses image-based techniques such as visible light imaging, fluorescence imaging, hyperspectral imaging, infrared (IR) imaging and X-ray computed tomography, and these techniques are controlled by a robust software system [291]. Image-based automated HTP integrates advanced software that is feasible to access for plant biology research [292]. Visible light imaging techniques are based upon a two-dimensional (2D) digital imaging system to measure leaf morphology, canopy coverage, above-ground dry matter, seed and panicle morphology, the architecture of root, shoot tip extension and yield-associated traits [293,294]. Apart from the 2D imaging system, 3D imaging techniques have also been reported to measure different characteristics in plants, such as leaf morphology, plant height, above-ground dry matter, crop structure and stature [295]. Plant eye is a 3D-based imaging technique that has been reported to observe the area of the leaf with the wet (fresh) and dry matter in wheat under salinity stress [296]. A study of a photosynthetic function under abiotic stresses by Chlorophyll Fluorescence Analysis (CFA) can discriminate between susceptible and tolerant genotypes [297]. Fluorescence Imaging is also widely used for detecting the stress in plants system at the primary level [298], which becomes helpful to resolve the heterogeneity in photosynthetic performance based on chlorophyll fluorescence in the leaf [299]. Likewise, the Chl F transient (Chlorophyll fluorescence) technique is used to discriminate the cold-sensitive and tolerant A. thaliana species [300]. Pulse Amplitude Modulated (PAM) or fluorometry can measure fluorescence parameters in plants [301] and is successfully used for screening Arabidopsis, tobacco and cotton (Gossypium ssp.) [302,303,304].
Recently, leaf spectroscopy, hyperspectral reflectance spectroscopy and imaging sensors related to chlorophyll fluorescence have also been successfully used for the study of phenomics under abiotic-stress conditions [305]. Digital imaging is a popular method of in situ plant phenotyping. In the last decade, numerous techniques and methodologies have been developed for automated phenotyping [306,307], which will provide valuable information about the abiotic stress tolerance of plants. PlantDIP (plant digital image processing) related to Scanalyzer HTS has been demonstrated to estimate the high ascorbic acid (vitamin C) content for osmotic stress response in Arabidopsis model [308]. Red, green, and blue color-based phenotyping (RGB phenotyping) are accounted for the measurement in various crops under abiotic stress conditions with the parallel use of computational software such as WIWAM (https://www.wiwam.be/ accessed on 6 March 2023) and PHENOPSIS [309]. Similarly, Lemna Tech is reported to use in barley and maize for drought stress [310,311,312] and in rice and wheat for salinity [313,314]. Li-Cor 6400 is a modern noninvasive HTP tool that has been considered for leaf gas exchange parameter study as reported in grapevine under drought stress [315]. It is widely used to study physiological parameters such as intercellular CO2, transpiration rate, stomatal conductance and photosynthetic rate under different abiotic stress environmental conditions [316]. The application of spectroscopy imaging is widely used for field phenotyping through aerial platforms. Hyperspectral high-spatial resolution satellite data are very effective in analyzing the physical and empirical analysis of water content in the canopy [317]. Unmanned aerial vehicle (UAV) is a thermal-based remote sensing noninvasive method that has accounted for drought stress response in poplar plants [318]. Apart from the drought stress, UAV-based HTP also accounted for salinity stress response in wild-type tomatoes and date palms [319,320]. In addition, the heat stress effect on crops and lodging should be accounted for. Common indicators for agricultural plants include the vegetation indices (VIs)—normalized difference vegetation index (NDVI), excessive green (ExG) and green leaf index (GLI), all of which represent canopy features and crop phenology that are significantly influenced by stress situations [321,322]. NDVI is a graphical form of obtained remote sensing data that are widely used in crop phenotyping [323]. Its calculation is based on the reflectance spectrum under red and near-infrared (NIR) regions of light in the plant [28]. It is one of the phenotypic tools that can be used for the estimation of canopy temperature and height and has high applicability to analyze heat, drought, water and salinity stress [324,325,326,327]. The ExG index is one of the parameters (based on RGB images) that has been reported for crop water stress index (CWSI) as well as water potential of leaf analysis in maize canopy [328]. Forster resonance energy transfer (FRET) is another advanced noninvasive method that has been applied to track the zinc and calcium dynamics in the root tissue during the transport of sugar [329]. Positron emission tomography (PET) is also reported to measure the stress effect on photosynthetic performance [292]. Infrared thermography has been reported to study stomatal activity under drought and salinity stresses through differences in plant canopy temperature and structure [330]. Infrared-image-based techniques provide high-quality measurements with high spatial resolution images under a broad range of climatic conditions [331]. Temperature differences of the canopy among different plant species can be applied for drought stress tolerance under dry (arid and semi-arid) environmental situations. The thermal-based infrared imaging system is widely performed in laboratory and field conditions to characterize drought, salinity and heat stress based on Na+ exclusion and osmotic imbalance [332]. With the application of infrared imaging, it is feasible to detect the significant differences among leaf, canopy and environmental temperature under high temperature and drought stress which is well reported in fruits and vegetables [292,327,328,329,330].

10. Integration of Multi-Omics Data and Interpretation for Abiotic Stress Response in Plants

Integrating multi-omics methods is basically to connect genotype to phenotype for a proper understanding of the biological processes such as abiotic stress response in plants. Multi-omics data integration is a powerful approach that combines information from multiple high-throughput omics technologies, such as genomics, transcriptomics, proteomics, metabolomics, epigenomics, etc., to gain a comprehensive understanding of complex biological systems (Figure 5). This approach has been widely used in diverse crops to investigate the molecular mechanisms underlying abiotic stress tolerance responses which are needed to make tolerant crop varieties [333]. Integrating multi-omics data can provide a more holistic view of how plants respond to abiotic stresses at the molecular level. Here a general overview of the steps involved in multi-omics data integration is represented for studying crop abiotic stress tolerance responses.
‘Experimental design’ is the first step to planning and designing experiments that expose crops to specific abiotic stress conditions while considering appropriate control sets for a quick compare. It needs to ensure the collection of samples for multiple omics platforms, including DNA, RNA, proteins and metabolites. ‘Data generation’ utilizes high-throughput omics technologies such as whole genome sequencing (WGS) for genomics, RNAseq for transcriptomics, MS for proteomics and metabolomics profiling for metabolomics, to generate large-scale datasets for each omics layer [334]. ‘Data pre-processing’ performs quality control and pre-processing steps specific to each omics platform. This may involve read-trimming and alignment for genomics and transcriptomics data, quality control (QC), data normalization and missing value imputation for proteomics and metabolomics data and removing batch effects if multiple experiments are involved. ‘Data integration’ is a very vital step which can apply computational methods to integrate the multi-omics datasets. Such big-data-driven analysis requires high statistical significance to integrate different omics layers. For easy visualization and analysis, these interconnections are analyzed using functional and statistical networks to validate results obtained by multi-omics layers. Different strategies can be employed, including correlation-based approaches, network-based approaches and machine learning (ML) algorithms. PaintOmics 4 is a new web-based server to integrate multi-omics datasets using biological pathway maps [335]. It is crucial in data integration to combine data from several sources in order to build a model that can be used to predict complicated features and increase prediction accuracy. In order to predict phenotypes, an increasing variety of statistical models, including both linear and nonlinear models, have been created and are currently in use [336]. Several linear models, such as Genomic Best Linear Unbiased Prediction (GBLUP), Linear mixed models (LMMs), Bayesian sparse linear mixed model (BSLMM) and Penalized linear mixed model with generalized method of moments estimator (MpLMMGMM) model, are widely used to model multi-omics data with higher phenotypic prediction [336]. On the other hand, ML is one of the nonlinear methods that use both supervised and unsupervised learning programming paradigms with statistical inference from big complex data. Two main objectives to be predicted in supervised learning are categorization and regression. Unsupervised learning is frequently employed to look for data interpretations including grouping, association and dimensionality reduction (DR) which is significant in high spatial biology since it minimizes the number of random variables to take into account [337]. These methods aim to identify relationships and interactions between molecules across different omics layers. ‘Functional analysis’ can interpret the integrated multi-omics data to gain insights into the molecular mechanisms underlying abiotic stress tolerance responses and may involve GO analysis, pathway enrichment analysis and functional annotation of key genes, proteins and metabolites [23]. ‘Network analysis’ is to construct biological networks that capture the interactions between different molecules identified in the integrated data. Network analysis techniques, such as co-expression networks or protein–protein interaction networks, can help identify key hub genes or proteins involved in stress response [338]. In the end, ‘experimental verification and validation’ involve the selection of candidate genes, proteins or metabolites identified from the integrated analysis for experimental validation. Techniques such as qPCR, Western blotting or targeted metabolomics can be used to validate the findings and confirm their roles in abiotic stress tolerance. This approach is also necessary to validate data and reveal post-transcriptional and post-translational mechanisms of gene expression regulation [339].
Omics-integration is much more positive when it can apply to early plant life, i.e., at the seedling stage. It has validated the possibility of applying a non-targeted integration approach to non-model plant Quercus ilex for early response to drought [340]. Transcriptomics, proteomics and metabolomics data use two integrative approaches, Principal Component Analysis (PCA) and Data Integration Analysis for Biomarker discovery using Latent variable approaches for Omics studies (DIABLO), which permits interconnections between the different omics-layers to be inferred and enables the discovery of key processes such as transcriptional control and to identify the key function TFs [340]. Multi-omics integration was also evident in oil palm for drought and salinity response by applying transcriptomic, proteomics and metabolomics [341,342]. Differential enzymes and metabolites identified from the analysis highly correlate (r ≥ 90) with cysteine and methionine metabolism pathways affected by the osmotic stress [341,342]. Integration of root multi-omics (transcriptomics, proteomics and metabolomics) reveals drought stress tolerance response in chickpeas. Integration of transcriptomics and proteomics data was able to identify enriched proteins hubs and integration of root-omics data also revealed some key candidate genes underlying drought-responsive ‘QTL-hotspot’ [343]. By integrating multi-omics data, a deeper understanding of the regulatory networks and molecular mechanisms governing crop responses to abiotic stress is possible. This knowledge can be leveraged for the development of stress-tolerant crop varieties through targeted breeding and genetic engineering (transgenic and genome editing) strategies, as well as for the identification of potential biomarkers or targets for future crop development.

11. Conclusions and Perspectives

Different existing omics approaches are overlapping and are interconnected with each other; and allow the identification of integrated cellular activities leading to stress responses and tolerance levels of a plant. To conclude and fully understand the primary cell response cascades that may vary between tolerant and sensitive plants under certain abiotic stress conditions, it is essential to integrate multi-omics data gathered through various omics pipelines. After revealing the crop’s response through multi-omics-aided non-DNA markers such as transcripts, proteins, metabolites, etc., those crops can be used as important genetic resources and incorporated into the breeding and genetic engineering strategies for making stress-tolerant plants. Here, we have broadly reviewed diverse multi-omics approaches for studying stress response and adaptive mechanisms of a plant under abiotic stress conditions. For the practical utility in breeding, we may consider marker assisted breeding (MAB) or its advanced version—genome assisted breeding (GAB) for plant’s abiotic stress tolerance, but these are only dealing with genomics. However, multi-omics and omics integration facilitate to open a new avenue of ‘Omics-assisted breeding’ which can also utilize GAB to enhance crop yield, quality attributes and other associated agronomic parameters along with the particular abiotic stress tolerance. Multi-omics-based analysis can integrate data from various omics platforms and provide a comprehensive systems-level understanding of abiotic stress tolerance in crops, offering the identification of key regulatory networks, biomarkers and candidate genes that can be targeted for breeding efforts, enabling precision agriculture strategies. Such multi-omics integration output can be a reliable strategy for linking genotype by the phenotype of a plant. This holistic approach increases the chances of success in developing stress-tolerant crop varieties. The big data obtained from the multi-omics layers, combined with advanced bioinformatics and computational tools, can be used for predictive modelling and precision breeding by applying machine learning algorithms. These achievements contribute to the development of stress-tolerant crop varieties and sustainable agricultural practices, ensuring food security in the face of changing environmental conditions.

Author Contributions

Conceptualization, R.R. and S.P.D.; equally contributed to writing the original draft, R.R., S.P.D., A.G. and P.P.; writing—review, editing, valuable feedback, R.R., K.C., U.S., A.K. and C.S.; manuscript corrections and draft formatting, D.P.R.; supervision, funding acquisition, visualization, R.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding. The APC was jointly funded by R.R. and K.C.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are thankful to Agricultural Research Organization (ARO-Volcani Center), Israel. The authors are also thankful to the editors and anonymous reviewers for their critical comments to improvise the manuscript to its current form. We are apologizing to those researchers whose relevant research and publications are not cited in this manuscript due to the space limitation in the present form.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abobatta, W.F. The influence of climate change on interactions between environmental stresses and plants. In Plant Stress Mitigators; Ghorbanpour, M., Shahid, M.A., Eds.; Academic Press: Cambridge, MA, USA, 2023; pp. 425–434. [Google Scholar] [CrossRef]
  2. Li, Y.; Roychowdhury, R.; Govta, L.; Jaiwar, S.; Wei, Z.Z.; Shams, I.; Fahima, T. Intracellular reactive oxygen species (intraROS)-aided localized cell death contributing to immune responses against wheat powdery mildew pathogen. Phytopathology 2023. [Google Scholar] [CrossRef]
  3. Del Buono, D.; Regni, L.; Proietti, P. Abiotic stresses, biostimulants and plant activity. Agriculture 2023, 13, 191. [Google Scholar] [CrossRef]
  4. Roychowdhury, R. Crop Improvement in the Era of Climate Change; IK International Publisher: New Delhi, India, 2014; p. 496. ISBN 978-939-045-590-4. [Google Scholar]
  5. Roychowdhury, R.; Choudhury, S.; Hasanuzzaman, M.; Srivastava, S. Sustainable Agriculture in the Era of Climate Change; Springer-Nature: Basel, Switzerland, 2020; p. 690. ISBN 978-303-045-669-6. [Google Scholar]
  6. Hasanuzzaman, M.; Roychowdhury, R.; Karmakar, J.; Dey, N.; Nahar, K.; Fujita, M. Recent advances in biotechnology and genomic approaches for abiotic stress tolerance in crop plants. In Genomics and Proteomics: Concepts, Technologies and Applications, 1st ed.; Thangadurai, D., Sangeetha, J., Eds.; Apple Academic Press: Burlington, ON, Canada, 2015; pp. 333–366. ISBN 978-177-463-537-7. [Google Scholar]
  7. Rai, K.K.; Kumar, A.; Rai, A.; Rai, V.P.; Rai, A.C. Conventional breeding approaches for abiotic stress management in horticultural crops. In Stress Tolerance in Horticultural Crops; Rai, A.C., Rai, A., Rai, K.K., Rai, V.P., Kumar, A., Eds.; Woodhead Publishing: Cambridge, UK; Elsevier: Alpharetta, GA, USA, 2021; pp. 21–32. [Google Scholar] [CrossRef]
  8. Rai, A.C.; Rai, A.; Rai, K.K.; Rai, V.P.; Kumar, A. Stress Tolerance in Horticultural Crops; Woodhead Publishing: Cambridge, UK; Elsevier: Alpharetta, GA, USA, 2021. [Google Scholar] [CrossRef]
  9. Roychowdhury, R.; Tah, J. Mutagenesis—A potential approach for crop improvement. In Crop Improvement—New Approaches and Modern Techniques, 1st ed.; Hakeem, K.R., Ahmad, P., Ozturk, M., Eds.; Springer: Boston, MA, USA, 2013; pp. 149–187. ISBN 978-148-997-357-3. [Google Scholar]
  10. Roychowdhury, R.; Karmakar, J.; Adak, M.K.; Dey, N. Physio-biochemical and microsatellite-based profiling of lowland rice (Oryza sativa L.) landraces for osmotic stress tolerance. Am. J. Plant Sci. 2013, 4, 52–63. [Google Scholar] [CrossRef] [Green Version]
  11. Roychowdhury, R.; Taoutaou, A.; Hakeem, K.R.; Gawwad, M.R.A.; Tah, J. Molecular marker-assisted technologies for crop improvement. In Crop Improvement in the Era of Climate Change, 1st ed.; Roychowdhury, R., Ed.; IK International Publishing House: New Delhi, India, 2014; pp. 241–258. ISBN 978-939-045-590-4. [Google Scholar]
  12. Deshmukh, R.; Sonah, H.; Patil, G.; Chen, W.; Prince, S.; Mutava, R.; Vuong, T.; Valliyodan, B.; Nguyen, H.T. Integrating omic approaches for abiotic stress tolerance in soybean. Front. Plant Sci. 2014, 5, 244. [Google Scholar] [CrossRef] [PubMed]
  13. Henry, V.J.; Bandrowski, A.E.; Pepin, A.S.; Gonzalez, B.J.; Desfeux, A. OMICtools: An informative directory for multi-omic data analysis. Database 2014, 2014, bau069. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Sharma, T.; Gupta, A. Integrated OMICS approaches to ameliorate the abiotic stress in Brassica napus. In Plant Stress: Challenges and Management in the New Decade; Roy, S., Mathur, P., Chakraborty, A.P., Saha, S.P., Eds.; Springer-Nature: Basel, Switzerland, 2022; pp. 361–373. ISBN 978-303-095-365-2. [Google Scholar]
  15. Khandagale, K.; Krishna, R.; Roylawar, P.; Ade, A.B.; Benke, A.; Shinde, B.; Singh, M.; Gawande, S.J.; Rai, A. Omics approaches in Allium research: Progress and way ahead. PeerJ 2020, 8, e9824. [Google Scholar] [CrossRef]
  16. Muthuramalingam, P.; Jeyasri, R.; Rakkammal, K.; Satish, L.; Shamili, S.; Karthikeyan, A.; Valliammai, A.; Priya, A.; Selvaraj, A.; Gowri, P.; et al. Multi-Omics and integrative approach towards understanding salinity tolerance in rice: A review. Biology 2022, 11, 1022. [Google Scholar] [CrossRef]
  17. Raza, A. Plant biotechnological tools: Solutions for raising climate-resilient crop plants. Mod. Phytomorphol. 2022, 15, 132–133. [Google Scholar]
  18. Ashraf, U.; Mahmood, S.; Shahid, N.; Imran, M.; Siddique, M.; Abrar, M. Multi-omics approaches for strategic improvements of crops under changing climatic conditions. In Principles and Practices of Omics and Genome Editing for Crop Improvement; Prakash, C.S., Fiaz, S., Fahad, S., Eds.; Springer-Nature: Basel, Switzerland, 2022; pp. 57–92. ISBN 978-303-096-925-7. [Google Scholar]
  19. Zhou, R.; Jiang, F.; Niu, L.; Song, X.; Yu, L.; Yang, Y.; Wu, Z. Increase crop resilience to heat stress using omic strategies. Front. Plant. Sci. 2022, 13, 891861. [Google Scholar] [CrossRef]
  20. Weckwerth, W.; Ghatak, A.; Bellaire, A.; Chaturvedi, P.; Varshney, R.K. PANOMICS meets germplasm. Plant Biotechnol. J. 2020, 18, 1507–1525. [Google Scholar] [CrossRef] [Green Version]
  21. Subramanian, I.; Verma, S.; Kumar, S.; Jere, A.; Anamika, K. Multi-omics data integration, interpretation, and its application. Bioinform. Biol. Insights 2020, 14, 1177932219899051. [Google Scholar] [CrossRef] [Green Version]
  22. Derbyshire, M.C.; Batley, J.; Edwards, D. Use of multiple ‘omics techniques to accelerate the breeding of abiotic stress tolerant crops. Curr. Plant Biol. 2022, 32, 100262. [Google Scholar] [CrossRef]
  23. Khan, M.K.R.; Ditta, A.; Wang, B.; Fang, L.; Anwar, Z.; Ijaz, A.; Ahmed, S.R.; Khan, S.M. The intervention of multi-omics approaches for developing abiotic stress resistance in cotton crop under climate change. In Sustainable Agriculture in the Era of the OMICs Revolution; Prakash, C.S., Fiaz, S., Nadeem, M.A., Baloch, F.S., Qayyum, A., Eds.; Springer: Cham, Switzerland, 2023; pp. 37–82. [Google Scholar] [CrossRef]
  24. Vennapusa, A.R.; Nimmakayala, P.; Zaman-Allah, M.A.; Ratnakumar, P. Physiological, molecular, and genetic perspectives of environmental stress response in plants. Front. Plant Sci. 2023, 14, 1213762. [Google Scholar] [CrossRef]
  25. Parida, S.K.; Mondal, N.; Yadav, R.; Vishwakarma, H.; Rana, J.C. Mining legume germplasm for genetic gains: An Indian perspective. Front. Genet. 2023, 14, 996828. [Google Scholar] [CrossRef]
  26. Parray, J.A.; Yaseen Mir, M.; Shameem, N.; Parray, J.A.; Yaseen Mir, M.; Shameem, N. Advancement in sustainable agriculture: Computational and bioinformatics tools. In Sustainable Agriculture: Biotechniques in Plant Biology; Parray, J.A., Yaseen Mir, M., Shameem, N., Eds.; Springer: Singapore, 2019; pp. 465–547. [Google Scholar] [CrossRef]
  27. Thakkar, S.; Banerjee, A.; Goel, S.; Roy, S.; Bansal, K.C. Genomics-based approaches to improve abiotic stress tolerance in plants: Present status and future prospects. In Plant Perspectives to Global Climate Changes—Developing Climate-Resilient Plants; Aftab, T., Roychoudhury, A., Eds.; Academic Press: Cambridge, MA, USA, 2022; pp. 195–219. [Google Scholar] [CrossRef]
  28. Xu, W.; Zhang, H.; Zhang, Y.; Shen, P.; Li, X.; Li, R.; Yang, L. A paired-end whole-genome sequencing approach enables comprehensive characterization of transgene integration in rice. Commun. Biol. 2022, 5, 667. [Google Scholar] [CrossRef]
  29. Zanini, S.F.; Bayer, P.E.; Wells, R.; Snowdon, R.J.; Batley, J.; Varshney, R.K.; Nguyen, H.T.; Edwards, D.; Golicz, A.A. Pangenomics in crop improvement-from coding structural variations to finding regulatory variants with pangenome graphs. Plant Genome 2022, 15, e20177. [Google Scholar] [CrossRef]
  30. Sharma, N.; Siddappa, S.; Malhotra, N.; Thakur, K.; Salaria, N.; Sood, S.; Bhardwaj, V. Advances in potato functional genomics: Implications for crop improvement. Plant Cell Tissue Organ Cult. 2022, 148, 447–464. [Google Scholar] [CrossRef]
  31. Singh, R.; Kumar, K.; Bharadwaj, C.; Verma, P.K. Broadening the horizon of crop research: A decade of advancements in plant molecular genetics to divulge phenotype governing genes. Planta 2022, 255, 46. [Google Scholar] [CrossRef]
  32. Jain, D.; Ashraf, N.; Khurana, J.P.; Kameshwari, S. The ‘omics’ approach for crop improvement against drought stress. In Genetic Enhancement of Crops for Tolerance to Abiotic Stress: Mechanisms and Approaches; Rajpal, V.R., Sehgal, D., Kumar, A., Raina, S.N., Eds.; Springer-Nature: Basel, Switzerland, 2019; Volume I, pp. 183–204. ISBN 978-331-991-956-0. [Google Scholar]
  33. Deokar, A.A.; Kondawar, V.; Jain, P.K.; Karuppayil, S.M.; Raju, N.L.; Vadez, V.; Varshney, R.K.; Srinivasan, R. Comparative analysis of expressed sequence tags (ESTs) between drought-tolerant and -susceptible genotypes of chickpea under terminal drought stress. BMC Plant Biol. 2011, 11, 70. [Google Scholar] [CrossRef] [Green Version]
  34. Lin, S.; Scholtens, D.; Datta, S. Bioinformatics Methods: From Omics to Next Generation Sequencing; Chapman and Hall/CRC Press: London, UK, 2022; p. 336. ISBN 978-149-876-515-2. [Google Scholar]
  35. Girma, G.; Natsume, S.; Carluccio, A.V.; Takagi, H.; Matsumura, H.; Uemura, A.; Muranaka, S.; Takagi, H.; Stavolone, L.; Gedil, M.; et al. Identification of candidate flowering and sex genes in white Guinea yam (D. rotundata Poir.) by SuperSAGE transcriptome profiling. PLoS ONE 2019, 14, e0216912. [Google Scholar] [CrossRef] [Green Version]
  36. El-Sappah, A.H.; Rather, S.A. Genomics approaches to study abiotic stress tolerance in plants. In Plant Abiotic Stress Physiology; Aftab, T., Hakeem, R., Eds.; Apple Academic Press: Palm Bay, FL, USA, 2022; pp. 25–46. ISBN 978-100-318-057-9. [Google Scholar]
  37. Kulwal, P.L.; Mir, R.R.; Varshney, R.K. Efficient breeding of crop plants. In Fundamentals of Field Crop Breeding; Yadava, D.K., Dikshit, H.K., Mishra, G.P., Tripathi, S., Eds.; Springer-Nature: Singapore, 2022; pp. 745–777. ISBN 978-981-169-257-4. [Google Scholar]
  38. Zhu, F.; Ahchige, M.W.; Brotman, Y.; Alseekh, S.; Zsögön, A.; Fernie, A.R. Bringing more players into play: Leveraging stress in genome wide association studies. J. Plant Physiol. 2022, 271, 153657. [Google Scholar] [CrossRef] [PubMed]
  39. Lv, Y.; Ma, J.; Wei, H.; Xiao, F.; Wang, Y.; Jahan, N.; Hazman, M.; Qian, Q.; Shang, L.; Guo, L. Combining GWAS, genome-wide domestication and a transcriptomic analysis reveals the loci and natural alleles of salt tolerance in rice (Oryza sativa L.). Front. Plant Sci. 2022, 13, 912637. [Google Scholar] [CrossRef] [PubMed]
  40. Wang, Q.; Ning, L.; Yu, W.; Zhao, W.; Huang, F.; Yu, D.; Wang, H.; Cheng, H. Detection of candidate loci and genes related to phosphorus efficiency at maturity through a genome-wide association study in Soybean. Agronomy 2022, 12, 2031. [Google Scholar] [CrossRef]
  41. Tanin, M.J.; Saini, D.K.; Sandhu, K.S.; Pal, N.; Gudi, S.; Chaudhary, J.; Sharma, A. Consensus genomic regions associated with multiple abiotic stress tolerance in wheat and implications for wheat breeding. Sci. Rep. 2022, 12, 13680. [Google Scholar] [CrossRef]
  42. Liu, P.; Zhu, Y.; Liu, H.; Liang, Z.; Zhang, M.; Zou, C.; Yuan, G.; Gao, S.; Pan, G.; Shen, Y.; et al. A Combination of a genome-wide association study and a transcriptome analysis reveals circRNAs as new regulators involved in the response to salt stress in maize. Int. J. Mol. Sci. 2022, 23, 9755. [Google Scholar] [CrossRef]
  43. Zhang, X.; You, J.; Miao, H.; Zhang, H. Genomic designing for sesame resistance to abiotic stresses. In Genomic Designing for Abiotic Stress Resistant Oilseed Crops; Kole, C., Ed.; Springer-Nature: Basel, Switzerland, 2022; pp. 219–234. ISBN 978-303-090-044-1. [Google Scholar]
  44. Fatemi, F.; Kianersi, F.; Pour-Aboughadareh, A.; Poczai, P.; Jadidi, O. Overview of identified genomic regions associated with various agronomic and physiological traits in barley under abiotic stresses. Appl. Sci. 2022, 12, 5189. [Google Scholar] [CrossRef]
  45. Samineni, S.; Mahendrakar, M.D.; Hotti, A.; Chand, U.; Rathore, A.; Gaur, P.M. Impact of heat and drought stresses on grain nutrient content in chickpea: Genome-wide marker-trait associations for protein, Fe and Zn. Environ. Exp. Bot. 2022, 194, 104688. [Google Scholar] [CrossRef]
  46. Chao, W.S.; Li, X.; Horvath, D.P.; Anderson, J.V. Genetic loci associated with freezing tolerance in a European rapeseed (Brassica napus L.) diversity panel identified by genome-wide association mapping. Plant Direct 2022, 6, e405. [Google Scholar] [CrossRef]
  47. Shukla, R.P.; Tiwari, G.J.; Joshi, B.; Song-Beng, K.; Tamta, S.; Boopathi, N.M.; Jena, S.N. GBS-SNP and SSR based genetic mapping and QTL analysis for drought tolerance in upland cotton. Physiol. Mol. Biol. Plants 2021, 27, 1731–1745. [Google Scholar] [CrossRef]
  48. Guo, A.; Su, Y.; Nie, H.; Li, B.; Ma, X.; Hua, J. Identification of candidate genes involved in salt stress response at germination and seedling stages by QTL mapping in upland cotton. G3 2022, 12, jkac099. [Google Scholar] [CrossRef]
  49. Diouf, L.; Pan, Z.; He, S.P.; Gong, W.F.; Jia, Y.H.; Magwanga, R.O.; Romy, K.R.E.; Or Rashid, H.; Kirungu, J.N.; Du, X. High-density linkage map construction and mapping of salt-tolerant QTLs at seedling stage in upland cotton using genotyping by sequencing (GBS). Int. J. Mol. Sci. 2017, 18, 2622. [Google Scholar] [CrossRef] [Green Version]
  50. Gelli, M.; Konda, A.R.; Liu, K.; Zhang, C.; Clemente, T.E.; Holding, D.R.; Dweikat, I.M. Validation of QTL mapping and transcriptome profiling for identification of candidate genes associated with nitrogen stress tolerance in sorghum. BMC Plant Biol. 2017, 17, 123. [Google Scholar] [CrossRef] [Green Version]
  51. Hostetler, A.N.; Govindarajulu, R.; Hawkins, J.S. QTL mapping in an interspecific sorghum population uncovers candidate regulators of salinity tolerance. Plant Stress 2021, 2, 100024. [Google Scholar] [CrossRef]
  52. Parihar, A. Molecular breeding and marker-assisted selection for crop improvement. In Plant Genomics for Sustainable Agriculture; Singh, R.L., Mondal, S., Parihar, A., Singh, P.K., Eds.; Springer-Nature: Singapore, 2022; pp. 129–164. ISBN 978-981-166-974-3. [Google Scholar]
  53. Salvi, S.; Tuberosa, R. The crop QTLome comes of age. Curr. Opin. Biotechnol. 2015, 32, 179–185. [Google Scholar] [CrossRef]
  54. Ha, B.K.; Vuong, T.D.; Velusamy, V.; Nguyen, H.T.; Shannon, J.G.; Lee, J.D. Genetic mapping of quantitative trait loci conditioning salt tolerance in wild soybean (Glycine soja) PI 483463. Euphytica 2013, 193, 79–88. [Google Scholar] [CrossRef]
  55. 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]
  56. Selamat, N.; Nadarajah, K.K. Meta-analysis of quantitative traits loci (QTL) identified in drought response in rice (Oryza sativa L.). Plants 2021, 10, 716. [Google Scholar] [CrossRef]
  57. Prakash, N.R.; Lokeshkumar, B.M.; Rathor, S.; Warraich, A.S.; Yadav, S.; Vinaykumar, N.M.; Dushynthkumar, B.M.; Krishnamurthy, S.L.; Sharma, P.C. Meta-analysis and validation of genomic loci governing seedling and reproductive stage salinity tolerance in rice. Physiol. Plant. 2022, 174, e13629. [Google Scholar] [CrossRef]
  58. Khan, I.; Zhang, Y.; Akbar, F.; Khan, J. Abiotic stress tolerance in cereals through genome editing. In Omics Approach to Manage Abiotic Stress in Cereals; Roychoudhury, A., Aftab, T., Acharya, K., Eds.; Springer-Nature: Singapore, 2022; pp. 295–319. ISBN 978-981-190-140-9. [Google Scholar]
  59. Singh, A.; Roychowdhury, R.; Singh, T.; Wang, W.; Yadav, D.; Kumar, A.; Modi, A.; Rai, A.C.; Ghughe, S.; Kumar, A.; et al. Improvement of crop’s stress tolerance by gene editing CRISPR/CAS9 system. In Sustainable Agriculture in the Era of Climate Change; Roychowdhury, R., Choudhury, S., Hasanuzzaman, M., Srivastava, S., Eds.; Springer-Nature: Basel, Switzerland, 2020; pp. 557–587. ISBN 978-303-045-669-6. [Google Scholar]
  60. Lou, D.; Wang, H.; Liang, G.; Yu, D. OsSAPK2 confers abscisic acid sensitivity and tolerance to drought stress in rice. Front. Plant Sci. 2017, 8, 993. [Google Scholar] [CrossRef] [Green Version]
  61. Jain, A.; Bhar, A.; Das, S. Improving biotic and abiotic stress tolerance in plants: A CRISPR-Cas approach. In Genome Engineering for Crop Improvement; Sarmah, B.K., Borah, B.K., Eds.; Springer-Nature: Basel, Switzerland, 2021; pp. 217–237. ISBN 978-303-063-372-1. [Google Scholar]
  62. 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] [Green Version]
  63. Bouzroud, S.; Gasparini, K.; Hu, G.; Barbosa, M.A.M.; Rosa, B.L.; Fahr, M.; Bendaou, N.; Bouzayen, M.; Zsögön, A.; Smouni, A.; et al. Down regulation and loss of auxin response factor 4 function using CRISPR/Cas9 alters plant growth, stomatal function and improves tomato tolerance to salinity and osmotic stress. Genes 2020, 11, 272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Debbarma, J.; Sarki, Y.N.; Saikia, B.; Boruah, H.P.D.; Singha, D.L.; Chikkaputtaiah, C. Ethylene Response Factor (ERF) family proteins in abiotic stresses and CRISPR-Cas9 genome editing of ERFs for multiple abiotic stress tolerance in crop plants: A review. Mol. Biotechnol. 2019, 61, 153–172. [Google Scholar] [CrossRef] [PubMed]
  65. Shen, C.; Que, Z.; Xia, Y.; Tang, N.; Li, D.; He, R.; Cao, M. Knock out of the annexin gene OsAnn3 via CRISPR/Cas9-mediated genome editing decreased cold tolerance in rice. J. Plant Biol. 2017, 60, 539–547. [Google Scholar] [CrossRef]
  66. Zhou, J. Sequence-based modeling of three-dimensional genome architecture from kilobase to chromosome scale. Nat. Genet. 2022, 54, 725–734. [Google Scholar] [CrossRef] [PubMed]
  67. Mérot, C.; Oomen, R.A.; Tigano, A.; Wellenreuther, M. A roadmap for understanding the evolutionary significance of structural genomic variation. Trends Ecol. Evol. 2020, 35, 561–572. [Google Scholar] [CrossRef] [PubMed]
  68. DoVale, J.C.; Carvalho, H.F.; Sabadin, F.; Fritsche-Neto, R. Reduction of genotyping marker density for genomic selection is not an affordable approach to long-term breeding in cross-pollinated crops. bioRxiv 2021. [Google Scholar] [CrossRef]
  69. Merrick, L.F.; Herr, A.W.; Sandhu, K.S.; Lozada, D.N.; Carter, A.H. Optimizing plant breeding programs for genomic selection. Agronomy 2022, 12, 714. [Google Scholar] [CrossRef]
  70. Rio, S.; Gallego-Sánchez, L.; Montilla-Bascón, G.; Canales, F.J.; Sánchez, J.I.Y.; Prats, E. Genomic prediction and training set optimization in a structured Mediterranean oat population. Theor. Appl. Genet. 2021, 134, 3595–3609. [Google Scholar] [CrossRef]
  71. Rogers, A.R.; Holland, J.B. Environment-specific genomic prediction ability in maize using environmental covariates depends on environmental similarity to training data. G3 2022, 12, jkab440. [Google Scholar] [CrossRef]
  72. Rio, S.; Akdemir, D.; Carvalho, T.; Sánchez, J.I.Y. Assessment of genomic prediction reliability and optimization of experimental designs in multi-environment trials. Theor. Appl. Genet. 2022, 135, 405–419. [Google Scholar] [CrossRef]
  73. Beyene, Y.; Gowda, M.; Pérez-Rodríguez, P.; Olsen, M.; Robbins, K.R.; Burgueño, J.; Prasanna, B.M.; Crossa, J. Application of genomic selection at the early stage of breeding pipeline in tropical maize. Front. Plant Sci. 2021, 12, 685488. [Google Scholar] [CrossRef]
  74. Udriște, A.A.; Iordachescu, M.; Ciceoi, R.; Bădulescu, L. Next-generation sequencing of local Romanian tomato varieties and bioinformatics analysis of the Ve locus. Int. J. Mol. Sci. 2022, 23, 9750. [Google Scholar] [CrossRef]
  75. Kress, W.J.; Soltis, D.E.; Kersey, P.J.; Wegrzyn, J.L.; Leebens-Mack, J.H.; Gostel, M.R.; Liu, X.; Soltis, P.S. Green plant genomes: What we know in an era of rapidly expanding opportunities. Proc. Natl. Acad. Sci. USA 2022, 119, e2115640118. [Google Scholar] [CrossRef]
  76. Giacopuzzi, E.; Popitsch, N.; Taylor, J.C. GREEN-DB: A framework for the annotation and prioritization of non-coding regulatory variants from whole-genome sequencing data. Nucleic Acids Res. 2022, 50, 2522–2535. [Google Scholar] [CrossRef]
  77. Pérez-Wohlfeil, E.; Diaz-Del-Pino, S.; Trelles, O. Ultra-fast genome comparison for large-scale genomic experiments. Sci. Rep. 2019, 9, 10274. [Google Scholar] [CrossRef] [Green Version]
  78. Ramkumar, T.R.; Arya, S.S.; Kumari, D.D.; Lenka, S.K. Brassica juncea genome sequencing: Structural and functional insights. In The Brassica juncea Genome; Kole, C., Mohapatra, T., Eds.; Springer-Nature: Basel, Switzerland, 2022; pp. 221–240. ISBN 978-303-091-507-0. [Google Scholar]
  79. Song, J.M.; Zhang, Y.; Zhou, Z.W.; Lu, S.; Ma, W.; Lu, C.; Chen, L.L.; Guo, L. Oil plant genomes: Current state of the science. J. Exp. Bot. 2022, 73, 2859–2874. [Google Scholar] [CrossRef]
  80. Song, J.; Xu, D.; Dong, Y.; Li, F.; Bian, Y.; Li, L.; Luo, X.; Fei, S.; Li, L.; Zhao, C.; et al. Fine mapping and characterization of a major QTL for grain weight on wheat chromosome arm 5DL. Theor. Appl. Genet. 2022, 135, 3237–3246. [Google Scholar] [CrossRef]
  81. Satrio, R.D.; Fendiyanto, M.H.; Supena, E.D.J.; Suharsono, S.; Miftahudin, M. Genome-wide SNP discovery, linkage mapping, and analysis of QTL for morpho-physiological traits in rice during vegetative stage under drought stress. Physiol. Mol. Biol. Plants 2021, 27, 2635–2650. [Google Scholar] [CrossRef]
  82. Sun, M.; Li, Y.; Zheng, J.; Wu, D.; Li, C.; Li, Z.; Zang, Z.; Zhang, Y.; Fang, Q.; Li, W.; et al. A nuclear factor Y-B transcription factor, GmNFYB17, regulates resistance to drought stress in soybean. Int. J. Mol. Sci. 2022, 23, 7242. [Google Scholar] [CrossRef]
  83. Mwando, E.; Han, Y.; Angessa, T.; Zhang, X.Q.; Li, C. Fine-mapping and characterisation of genes on barley (Hordeum vulgare) chromosome 2H for salinity stress tolerance during germination. Crop. J. 2022, 10, 754–766. [Google Scholar] [CrossRef]
  84. Makhtoum, S.; Sabouri, H.; Gholizadeh, A.; Ahangar, L.; Katouzi, M. QTLs controlling physiological and morphological traits of barley (Hordeum vulgare L.) seedlings under salinity, drought, and normal conditions. BioTech 2022, 11, 26. [Google Scholar] [CrossRef] [PubMed]
  85. Singh, R.K.; Gregorio, G.B.; Jain, R.K. QTL mapping for salinity tolerance in rice. Physiol. Mol. Biol. Plant. 2007, 13, 87–99. [Google Scholar]
  86. Asif, M.A.; Garcia, M.; Tilbrook, J.; Brien, C.; Dowling, K.; Berger, B.; Schilling, R.K.; Short, L.; Trittermann, C.; Gilliham, M.; et al. Identification of salt tolerance QTL in a wheat RIL mapping population using destructive and non-destructive phenotyping. Funct. Plant Biol. 2022, 49, 672. [Google Scholar] [CrossRef] [PubMed]
  87. Touzy, G.; Lafarge, S.; Redondo, E.; Lievin, V.; Decoopman, X.; Le Gouis, J.; Praud, S. Identification of QTLs affecting post-anthesis heat stress responses in European bread wheat. Theor. Appl. Genet. 2022, 135, 947–964. [Google Scholar] [CrossRef]
  88. Rani, S.; Baber, M.; Naqqash, T.; Malik, S.A. Identification and genetic mapping of potential QTLs conferring heat tolerance in cotton (Gossypium hirsutum L.) by using micro satellite marker’s approach. Agronomy 2022, 12, 1381. [Google Scholar] [CrossRef]
  89. Yang, L.; Lei, L.; Li, P.; Wang, J.; Wang, C.; Yang, F.; Chen, J.; Liu, H.; Zheng, H.; Xin, W.; et al. Identification of candidate genes conferring cold tolerance to rice (Oryza sativa L.) at the bud-bursting stage using bulk segregant analysis sequencing and linkage mapping. Front. Plant Sci. 2021, 12, 647239. [Google Scholar] [CrossRef]
  90. Lei, L.; Cao, L.; Ding, G.; Zhou, J.; Luo, Y.; Bai, L.; Xia, T.; Chen, L.; Wang, J.; Liu, K.; et al. OsBBX11 on qSTS4 links to salt tolerance at the seeding stage in Oryza sativa L. ssp. Japonica. Front. Plant Sci. 2023, 14, 1139961. [Google Scholar] [CrossRef]
  91. Dhungana, S.K.; Kim, H.S.; Kang, B.K.; Seo, J.H.; Kim, H.T.; Shin, S.O.; Oh, J.H.; Baek, I.Y. Identification of QTL for tolerance to flooding stress at seedling stage of soybean (Glycine max L. Merr.). Agronomy 2021, 11, 908. [Google Scholar] [CrossRef]
  92. Cho, K.H.; Kim, M.Y.; Kwon, H.; Yang, X.; Lee, S.H. Novel QTL identification and candidate gene analysis for enhancing salt tolerance in soybean (Glycine max (L.) Merr.). Plant Sci. 2021, 313, 111085. [Google Scholar] [CrossRef]
  93. Khaled, K.A.M.; Habiba, R.M.M.; Bashasha, J.A.; El-Aziz, M.H.A. Identification and mapping of QTL associated with some traits related for drought tolerance in wheat using SSR markers. Beni-Suef Univ. J. Basic Appl. Sci. 2022, 11, 38. [Google Scholar] [CrossRef]
  94. Xu, L.; Zhao, C.; Pang, J.; Niu, Y.; Liu, H.; Zhang, W.; Zhou, M. Genome-wide association study reveals quantitative trait loci for waterlogging-triggered adventitious roots and aerenchyma formation in common wheat. Front. Plant Sci. 2022, 13, 1066752. [Google Scholar] [CrossRef]
  95. Guo, X.; Wu, C.; Wang, D.; Wang, G.; Jin, K.; Zhao, Y.; Tian, J.; Deng, Z. Conditional QTL mapping for seed germination and seedling traits under salt stress and candidate gene prediction in wheat. Sci. Rep. 2022, 12, 21010. [Google Scholar] [CrossRef]
  96. Jin, Y.; Zhang, Z.; Xi, Y.; Yang, Z.; Xiao, Z.; Guan, S.; Qu, J.; Wang, P.; Zhao, R. Identification and functional verification of cold tolerance genes in spring maize seedlings based on a genome-wide association study and quantitative trait locus mapping. Front. Plant Sci. 2021, 12, 776972. [Google Scholar] [CrossRef]
  97. Gad, M.; Chao, H.; Li, H.; Zhao, W.; Lu, G.; Li, M. QTL Mapping for seed germination response to drought stress in Brassica napus. Front. Plant Sci. 2020, 11, 629970. [Google Scholar] [CrossRef]
  98. Ding, X.Y.; Xu, J.S.; Huang, H.; Xing, Q.I.A.O.; Shen, M.Z.; Cheng, Y.; Zhang, X. Unraveling waterlogging tolerance-related traits with QTL analysis in reciprocal intervarietal introgression lines using genotyping by sequencing in rapeseed (Brassica napus L.). J. Integr. Agric. 2020, 19, 1974–1983. [Google Scholar] [CrossRef]
  99. Borrego-Benjumea, A.; Carter, A.; Zhu, M.; Tucker, J.R.; Zhou, M.; Badea, A. Genome-wide association study of waterlogging tolerance in barley (Hordeum vulgare L.) under controlled field conditions. Front. Plant Sci. 2021, 12, 711654. [Google Scholar] [CrossRef]
  100. Sarkar, B.; Varalaxmi, Y.; Vanaja, M.; Kumar, N.R.; Prabhakar, M.; Jyothilakshmi, N.; Yadav, S.K.; Maheswari, M.; Singh, V.K. Genome-wide SNP discovery, identification of QTLs and candidate genes associated with morpho-physiological and yield related traits for drought tolerance in maize. ResearchSquare 2022. [Google Scholar] [CrossRef]
  101. Wijerathna-Yapa, A.; Ramtekey, V.; Ranawaka, B.; Basnet, B.R. Applications of in vitro tissue culture technologies in breeding and genetic improvement of wheat. Plants 2022, 11, 2273. [Google Scholar] [CrossRef]
  102. Maldonado-Alconada, A.M.; Castillejo, M.Á.; Rey, M.D.; Labella-Ortega, M.; Tienda-Parrilla, M.; Hernández-Lao, T.; Honrubia-Gómez, I.; Ramírez-García, J.; Guerrero-Sanchez, V.M.; López-Hidalgo, C.; et al. Multiomics molecular research into the recalcitrant and orphan Quercus ilex tree species: Why, what for, and how. Int. J. Mol. Sci. 2022, 23, 9980. [Google Scholar] [CrossRef]
  103. Brake, M.; Al-Qadumii, L.; Hamasha, H.; Migdadi, H.; Awad, A.; Haddad, N.; Sadder, M.T. Development of SSR markers linked to stress responsive genes along tomato chromosome 3 (Solanum lycopersicum L.). BioTech 2022, 11, 34. [Google Scholar] [CrossRef]
  104. Shamim, M.; Kumar, M.; Srivastava, D. Molecular markers mediated heat stress tolerance in crop plants. In Thermotolerance in Crop Plants; Kumar, R.R., Praveen, S., Rai, G.K., Eds.; Springer-Nature: Singapore, 2022; pp. 23–44. ISBN 978-981-193-800-9. [Google Scholar]
  105. Kumar, P.; Patni, B.; Singh, M. Wheat genome sequence opens new opportunities to understand the genetic basis of frost tolerance (FT) and marker-assisted breeding in wheat (Triticum aestivum L.). J. Stress Physiol. Biochem. 2022, 18, 17–27. [Google Scholar]
  106. Chugh, V.; Kaur, D.; Purwar, S.; Kaushik, P.; Sharma, V.; Kumar, H.; Rai, A.; Singh, C.M.; Kamaluddin; Dubey, R.B. Applications of molecular markers for developing abiotic-stress-resilient oilseed crops. Life 2023, 13, 88. [Google Scholar] [CrossRef] [PubMed]
  107. Bhatia, D.; Bajwa, G.S. Molecular marker techniques and recent advancements. In Genotyping by Sequencing for Crop Improvement; Sonah, H., Goyal, V., Shivaraj, S.M., Deshmukh, R.K., Eds.; John Wiley & Sons Ltd.: Chichester, UK, 2022; pp. 1–21. ISBN 978-111-974-568-6. [Google Scholar]
  108. Das, A.; Singh, S.; Islam, Z.; Munshi, A.D.; Behera, T.K.; Dutta, S.; Weng, Y.; Dey, S.S. Current progress in genetic and genomics-aided breeding for stress resistance in cucumber (Cucumis sativus L.). Sci. Hortic. 2022, 300, 111059. [Google Scholar] [CrossRef]
  109. Khan, N.; You, F.M.; Cloutier, S. Designing genomic solutions to enhance abiotic Stress resistance in flax. In Genomic Designing for Abiotic Stress Resistant Oilseed Crops; Kole, C., Ed.; Springer-Nature: Basel, Switzerland, 2022; pp. 251–283. ISBN 978-303-090-044-1. [Google Scholar]
  110. Avni, R.; Lux, T.; Minz-Dub, A.; Millet, E.; Sela, H.; Distelfeld, A.; Deek, J.; Yu, G.; Steuernagel, B.; Pozniak, C.; et al. Genome sequences of three Aegilops species of the section Sitopsis reveal phylogenetic relationships and provide resources for wheat improvement. Plant J. 2022, 110, 179–192. [Google Scholar] [CrossRef] [PubMed]
  111. Kalyana Babu, B.; Sood, S.; Gaur, V.S.; Kumar, A. Comparative genomics of finger millet. In The Finger Millet Genome; Kumar, A., Sood, S., Kalyana Babu, B., Gupta, S.M., Dayakar Rao, B., Eds.; Springer-Nature: Basel, Switzerland, 2022; pp. 113–121. ISBN 978-303-100-868-9. [Google Scholar]
  112. Fang, Y.; Qin, X.; Liao, Q.; Du, R.; Luo, X.; Zhou, Q.; Li, Z.; Chen, H.; Jin, W.; Yuan, Y.; et al. The genome of homosporous maidenhair fern sheds light on the euphyllophyte evolution and defences. Nat. Plants 2021, 8, 1024–1037. [Google Scholar] [CrossRef] [PubMed]
  113. Hill, M.J.; Penning, B.W.; McCann, M.C.; Carpita, N.C. COMPILE: A GWAS computational pipeline for gene discovery in complex genomes. BMC Plant Biol. 2022, 22, 315. [Google Scholar] [CrossRef]
  114. Kaur, B.; Sandhu, K.S.; Kamal, R.; Kaur, K.; Singh, J.; Röder, M.S.; Muqaddasi, Q.H. Omics for the improvement of abiotic, biotic, and agronomic traits in major cereal crops: Applications, challenges, and prospects. Plants 2021, 10, 1989. [Google Scholar] [CrossRef]
  115. McGettigan, P.A. Transcriptomics in the RNA-seq era. Curr. Opin. Chem. Biol. 2013, 17, 4–11. [Google Scholar] [CrossRef]
  116. Kwasniewski, M.; Daszkowska-Golec, A.; Janiak, A.; Chwialkowska, K.; Nowakowska, U.; Sablok, G.; Szarejko, I. Transcriptome analysis reveals the role of the root hairs as environmental sensors to maintain plant functions under water-deficiency conditions. J. Exp. Bot. 2016, 67, 1079–1094. [Google Scholar] [CrossRef]
  117. Lowe, R.; Shirley, N.; Bleackley, M.; Dolan, S.; Shafee, T. Transcriptomics technologies. PLoS Comput. Biol. 2017, 13, e1005457. [Google Scholar] [CrossRef] [Green Version]
  118. Karkute, S.G.; Gujjar, R.S.; Rai, A.; Akhtar, M.; Singh, M.; Singh, B. Genome wide expression analysis of WRKY genes in tomato (Solanum lycopersicum) under drought stress. Plant Gene 2018, 13, 8–17. [Google Scholar] [CrossRef]
  119. Rai, A.C.; Rai, A.; Shah, K.; Singh, M. Engineered BcZAT12 gene mitigates salt stress in tomato seedlings. Physiol. Mol. Biol. Plants 2021, 27, 535–541. [Google Scholar] [CrossRef]
  120. Mishra, P.; Singh, P.; Rai, A.; Abhishek, K.; Shanmugam, V.; Aamir, M.; Kumar, A.; Malik, M.Z.; Singh, S.K. Abiotic stress-mediated transcription regulation, chromatin dynamics, and gene expression in plants: Arabidopsis as a role model. In Mitigation of Plant Abiotic Stress by Microorganisms; Santoyo, G., Kumar, A., Aamir, M., Uthandi, S., Eds.; Academic Press: Cambridge, MA, USA, 2022; pp. 321–345. [Google Scholar] [CrossRef]
  121. Tyagi, S.; Kabade, P.G.; Gnanapragasam, N.; Singh, U.M.; Gurjar, A.K.S.; Rai, A.; Sinha, P.; Kumar, A.; Singh, V.K. Codon usage provide insights into the adaptation of rice genes under stress condition. Int. J. Mol. Sci. 2023, 24, 1098. [Google Scholar] [CrossRef]
  122. Shah, T.; Xu, J.; Zou, X.; Cheng, Y.; Nasir, M.; Zhang, X. Omics approaches for engineering wheat production under abiotic stresses. Int. J. Mol. Sci. 2018, 19, 2390. [Google Scholar] [CrossRef] [Green Version]
  123. Zhang, X.; Liu, J.; Huang, Y.; Wu, H.; Hu, X.; Cheng, B.; Ma, Q.; Zhao, Y. Comparative transcriptomics reveals the molecular mechanism of the parental lines of Maize hybrid An’nong876 in response to salt stress. Int. J. Mol. Sci. 2022, 23, 5231. [Google Scholar] [CrossRef]
  124. Ramkumar, M.K.; Mulani, E.; Jadon, V.; Sureshkumar, V.; Krishnan, S.G.; Senthil Kumar, S.; Raveendran, M.; Singh, A.K.; Solanke, A.U.; Singh, N.K.; et al. Identification of major candidate genes for multiple abiotic stress tolerance at seedling stage by network analysis and their validation by expression profiling in rice (Oryza sativa L.). 3 Biotech 2022, 12, 127. [Google Scholar] [CrossRef]
  125. Zinati, Z.; Sazegari, S. Identification of important genes involved in priming induced drought tolerance in barley through transcriptomic data mining. Crop Pasture Sci. 2022, 73, 1011–1025. [Google Scholar] [CrossRef]
  126. Chen, C.; Shang, X.; Sun, M.; Tang, S.; Khan, A.; Zhang, D.; Yan, H.; Jiang, Y.; Yu, F.; Wu, Y.; et al. Comparative transcriptome analysis of two sweet Sorghum genotypes with different salt tolerance abilities to reveal the mechanism of salt tolerance. Int. J. Mol. Sci. 2022, 23, 2272. [Google Scholar] [CrossRef]
  127. Han, B.; Wang, F.; Liu, Z.; Chen, L.; Yue, D.; Sun, W.; Lin, Z.; Zhang, X.; Zhou, X.; Yang, X. Transcriptome and metabolome profiling of interspecific CSSLs reveals general and specific mechanisms of drought resistance in cotton. Theor. Appl. Genet. 2022, 135, 3375–3391. [Google Scholar] [CrossRef]
  128. Wang, X.; Song, S.; Wang, X.; Liu, J.; Dong, S. Transcriptomic and metabolomic analysis of seedling-stage soybean responses to PEG-simulated drought stress. Int. J. Mol. Sci. 2022, 23, 6869. [Google Scholar] [CrossRef]
  129. Guo, H.; Mao, M.; Deng, Y.; Sun, L.; Chen, R.; Cao, P.; Lai, J.; Zhang, Y.; Wang, C.; Li, C.; et al. Multi-omics analysis reveals that SlERF.D6 synergistically regulates SGAs and fruit development. Front. Plant Sci. 2022, 13, 860577. [Google Scholar] [CrossRef] [PubMed]
  130. Smita, S.; Katiyar, A.; Lenka, S.K.; Dalal, M.; Kumar, A.; Mahtha, S.K.; Yadav, G.; Chinnusamy, V.; Pandey, D.M.; Bansal, K.C. Gene network modules associated with abiotic stress response in tolerant rice genotypes identified by transcriptome meta-analysis. Funct. Integr. Genom. 2020, 20, 29–49. [Google Scholar] [CrossRef] [PubMed]
  131. Azzouz-Olden, F.; Hunt, A.G.; Dinkins, R. Transcriptome analysis of drought-tolerant Sorghum genotype SC56 in response to water stress reveals an oxidative stress defense strategy. Mol. Biol. Rep. 2020, 47, 3291–3303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Zhang, X.; Liu, X.; Zhang, D.; Tang, H.; Sun, B.; Li, C.; Hao, L.; Liu, C.; Li, Y.; Shi, Y.; et al. Genome-wide identification of gene expression in contrasting maize inbred lines under field drought conditions reveals the significance of transcription factors in drought tolerance. PLoS ONE 2017, 12, e0179477. [Google Scholar] [CrossRef] [Green Version]
  133. Satya, P.; Bhattacharjee, S.; Sarkar, D.; Roy, S.; Sharma, L.; Mandal, N.A. Transcriptomics in plant. In Plant Genomics for Sustainable Agriculture; Singh, R.L., Mondal, S., Parihar, A., Singh, P.K., Eds.; Springer-Nature: Singapore, 2022; pp. 99–127. ISBN 978-981-166-974-3. [Google Scholar]
  134. Tiwari, J.K.; Buckseth, T.; Zinta, R.; Saraswati, A.; Singh, R.K.; Rawat, S.; Dua, V.K.; Chakrabarti, S.K. Transcriptome analysis of potato shoots, roots and stolons under nitrogen stress. Sci. Rep. 2020, 10, 1152. [Google Scholar] [CrossRef] [Green Version]
  135. Song, Y.; Lv, J.; Ma, Z.; Dong, W. The mechanism of alfalfa (Medicago sativa L.) response to abiotic stress. Plant Growth Regul. 2019, 89, 239–249. [Google Scholar] [CrossRef]
  136. Samtani, H.; Sharma, A.; Khurana, P. Overexpression of HVA1 enhances drought and heat stress tolerance in Triticum aestivum doubled haploid plants. Cells 2022, 11, 912. [Google Scholar] [CrossRef]
  137. El-Esawi, M.A.; Al-Ghamdi, A.A.; Ali, H.M.; Ahmad, M. Overexpression of AtWRKY30 transcription factor enhances heat and drought stress tolerance in wheat (Triticum aestivum L.). Genes 2019, 10, 163. [Google Scholar] [CrossRef] [Green Version]
  138. Sun, J.Q.; Jiang, H.L.; Xu, Y.X.; Li, H.M.; Wu, X.Y.; Xie, Q.; Li, C.Y. The CCCH-type zinc finger proteins AtSZF1 and AtSZF2 regulate salt stress responses in Arabidopsis. Plant Cell Physiol. 2007, 48, 1148–1158. [Google Scholar] [CrossRef] [Green Version]
  139. Yun, S.D.; Kim, M.H.; Oh, S.A.; Soh, M.S.; Park, S.K. Overexpression of C-Repeat Binding Factor1 (CBF1) gene enhances heat stress tolerance in Arabidopsis. J. Plant Biol. 2022, 65, 253–260. [Google Scholar] [CrossRef]
  140. Yang, R.; Hong, Y.; Ren, Z.; Tang, K.; Zhang, H.; Zhu, J.K.; Zhao, C. A role for PICKLE in the regulation of cold and salt stress tolerance in Arabidopsis. Front. Plant Sci. 2019, 10, 900. [Google Scholar] [CrossRef] [Green Version]
  141. Xu, Z.; Wang, F.; Ma, Y.; Dang, H.; Hu, X. Transcription factor SlAREB1 is involved in the antioxidant regulation under saline–alkaline stress in tomato. Antioxidants 2022, 11, 1673. [Google Scholar] [CrossRef]
  142. Klay, I.; Gouia, S.; Liu, M.; Mila, I.; Khoudi, H.; Bernadac, A.; Bouzayen, M.; Pirrello, J. Ethylene Response Factors (ERF) are differentially regulated by different abiotic stress types in tomato plants. Plant Sci. 2018, 274, 137–145. [Google Scholar] [CrossRef]
  143. Liu, X.; Zhang, Q.; Yang, G.; Zhang, C.; Dong, H.; Liu, Y.; Yin, R.; Lin, L. Pivotal roles of tomato photoreceptor SlUVR8 in seedling development and UV-B stress tolerance. Biochem. Biophys. Res. Commun. 2020, 522, 177–183. [Google Scholar] [CrossRef]
  144. Munir, S.; Liu, H.; Xing, Y.; Hussain, S.; Ouyang, B.; Zhang, Y.; Li, H.; Ye, Z. Overexpression of calmodulin-like (ShCML44) stress-responsive gene from Solanum habrochaites enhances tolerance to multiple abiotic stresses. Sci. Rep. 2016, 6, 31772. [Google Scholar] [CrossRef]
  145. Lee, S.S.; Jung, W.Y.; Park, H.J.; Lee, A.; Kwon, S.Y.; Kim, H.S.; Cho, H.S. genome-wide analysis of alternative splicing in an inbred cabbage (Brassica oleracea L.) line ’HO’ in response to heat stress. Curr. Genom. 2018, 19, 12–20. [Google Scholar] [CrossRef]
  146. Meena, R.P.; Ghosh, G.; Vishwakarma, H.; Padaria, J.C. Expression of a Pennisetum glaucum gene DREB2A confers enhanced heat, drought and salinity tolerance in transgenic Arabidopsis. Mol. Biol. Rep. 2022, 49, 7347–7358. [Google Scholar] [CrossRef]
  147. Hu, J.; Zhou, J.; Peng, X.; Xu, H.; Liu, C.; Du, B.; Yuan, H.; Zhu, L.; He, G. The Bphi008a gene interacts with the ethylene pathway and transcriptionally regulates MAPK genes in the response of rice to brown planthopper feeding. Plant Physiol. 2011, 156, 856–872. [Google Scholar] [CrossRef] [Green Version]
  148. Yan, L.; Baoxiang, W.; Jingfang, L.; Zhiguang, S.; Ming, C.; Yungao, X.; Bo, X.; Bo, Y.; Jian, L.; Jinbo, L.; et al. A novel SAPK10-WRKY87-ABF1 biological pathway synergistically enhance abiotic stress tolerance in transgenic rice (Oryza sativa). Plant Physiol. Biochem. 2021, 168, 252–262. [Google Scholar] [CrossRef]
  149. Gupta, A.; Shaw, B.P. Biochemical and molecular characterisations of salt tolerance components in rice varieties tolerant and sensitive to NaCl: The relevance of Na+ exclusion in salt tolerance in the species. Funct. Plant Biol. 2020, 48, 72–87. [Google Scholar] [CrossRef]
  150. Zhao, H.; Li, Z.; Wang, Y.; Wang, J.; Xiao, M.; Liu, H.; Quan, R.; Zhang, H.; Huang, R.; Zhu, L.; et al. Cellulose synthase-like protein OsCSLD4 plays an important role in the response of rice to salt stress by mediating abscisic acid biosynthesis to regulate osmotic stress tolerance. Plant Biotechnol. J. 2022, 20, 468–484. [Google Scholar] [CrossRef] [PubMed]
  151. Zhong, R.; Wang, Y.; Gai, R.; Xi, D.; Mao, C.; Ming, F. Rice SnRK protein kinase OsSAPK8 acts as a positive regulator in abiotic stress responses. Plant Sci. 2020, 292, 110373. [Google Scholar] [CrossRef] [PubMed]
  152. Li, S.; Han, X.; Lu, Z.; Qiu, W.; Yu, M.; Li, H.; He, Z.; Zhuo, R. MAPK cascades and transcriptional factors: Regulation of heavy metal tolerance in plants. Int. J. Mol. Sci. 2022, 23, 4463. [Google Scholar] [CrossRef] [PubMed]
  153. Kokkanti, R.R.; Vemuri, H.; Gaddameedi, A.; Rayalacheruvu, U. Variability in drought stress-induced physiological, biochemical responses and expression of DREB2A, NAC4 and HSP70 genes in groundnut (Arachis hypogaea L.). S. Afr. J. Bot. 2022, 144, 448–457. [Google Scholar] [CrossRef]
  154. Chen, Y.; Li, C.; Zhang, B.; Yi, J.; Yang, Y.; Kong, C.; Lei, C.; Gong, M. The role of the late embryogenesis-abundant (LEA) protein family in development and the abiotic stress response: A comprehensive expression analysis of potato (Solanum Tuberosum). Genes 2019, 10, 148. [Google Scholar] [CrossRef] [Green Version]
  155. Conde, D.; Kirst, M. Decoding exceptional plant traits by comparative single-cell genomics. Trends Plant Sci. 2022, 27, 1095–1098. [Google Scholar] [CrossRef]
  156. Barkla, B.J.; Vera-Estrella, R.; Raymond, C. Single-cell-type quantitative proteomic and ionomic analysis of epidermal bladder cells from the halophyte model plant Mesembryanthemum crystallinum to identify salt-responsive proteins. BMC Plant Biol. 2016, 16, 110. [Google Scholar] [CrossRef] [Green Version]
  157. Ahmad, P.; Abdel Latef, A.A.; Rasool, S.; Akram, N.A.; Ashraf, M.; Gucel, S. Role of proteomics in crop stress tolerance. Front. Plant Sci. 2016, 7, 1336. [Google Scholar] [CrossRef] [Green Version]
  158. Jan, N.; Rather, A.M.; John, R.; Chaturvedi, P.; Ghatak, A.; Weckwerth, W.; Zargar, S.M.; Mir, R.A.; Khan, M.A.; Mir, R.R. Proteomics for abiotic stresses in legumes: Present status and future directions. Crit. Rev. Biotechnol. 2022, 43, 171–190. [Google Scholar] [CrossRef]
  159. Li, J.; Sohail, H.; Nawaz, M.A.; Liu, C.; Yang, P. Physiological and proteomic analyses reveals that Brassinosteroids application improves the chilling stress tolerance of pepper seedlings. Plant Growth Regul. 2022, 96, 315–329. [Google Scholar] [CrossRef]
  160. Lohani, N.; Singh, M.B.; Bhalla, P.L. Biological parts for engineering abiotic stress tolerance in plants. BioDesign Res. 2022, 2022, 9819314. [Google Scholar] [CrossRef]
  161. Ghatak, A.; Chaturvedi, P.; Weckwerth, W. Cereal crop proteomics: Systemic analysis of crop drought stress responses towards marker-assisted selection breeding. Front. Plant Sci. 2017, 8, 757. [Google Scholar] [CrossRef] [Green Version]
  162. Kausar, R.; Wang, X.; Komatsu, S. Crop proteomics under abiotic stress: From data to insights. Plants 2022, 11, 2877. [Google Scholar] [CrossRef]
  163. Yan, S.; Bhawal, R.; Yin, Z.; Thannhauser, T.W.; Zhang, S. Recent advances in proteomics and metabolomics in plants. Mol. Hortic. 2022, 2, 17. [Google Scholar] [CrossRef]
  164. Ahmad, J.; Ali, A.A.; Iqbal, M.; Ahmad, A.; Qureshi, M.I. Proteomics of mercury-induced responses and resilience in plants: A review. Environ. Chem. Lett. 2022, 20, 3335–3355. [Google Scholar] [CrossRef]
  165. Blaženović, I.; Kind, T.; Ji, J.; Fiehn, O. Software tools and approaches for compound identification of LC-MS/MS data in metabolomics. Metabolites 2018, 8, 31. [Google Scholar] [CrossRef] [Green Version]
  166. Vats, S.; Bansal, R.; Rana, N.; Kumawat, S.; Bhatt, V.; Jadhav, P.; Kale, V.; Sathe, A.; Sonah, H.; Jugdaohsingh, R.; et al. Unexplored nutritive potential of tomato to combat global malnutrition. Crit. Rev. Food Sci. Nutr. 2022, 62, 1003–1034. [Google Scholar] [CrossRef]
  167. Singh, P.K.; Indoliya, Y.; Agrawal, L.; Awasthi, S.; Deeba, F.; Dwivedi, S.; Chakrabarty, D.; Shirke, P.A.; Pandey, V.; Singh, N.; et al. Genomic and proteomic responses to drought stress and biotechnological interventions for enhanced drought tolerance in plants. Curr. Plant Biol. 2022, 29, 100239. [Google Scholar] [CrossRef]
  168. Mancini, I.; Domingo, G.; Bracale, M.; Loreto, F.; Pollastri, S. Isoprene emission influences the proteomic profile of Arabidopsis plants under well-watered and drought-stress conditions. Int. J. Mol. Sci. 2022, 23, 3836. [Google Scholar] [CrossRef]
  169. Camp, E.F.; Kahlke, T.; Signal, B.; Oakley, C.A.; Lutz, A.; Davy, S.K.; Suggett, D.J.; Leggat, W.P. Proteome metabolome and transcriptome data for three Symbiodiniaceae under ambient and heat stress conditions. Sci. Data 2022, 9, 153. [Google Scholar] [CrossRef]
  170. Zhang, X.; Feng, Y.; Khan, A.; Ullah, N.; Li, Z.; Zaheer, S.; Zhou, R.; Zhang, Z. Quantitative proteomics-based analysis reveals molecular mechanisms of chilling tolerance in grafted cotton seedlings. Agronomy 2022, 12, 1152. [Google Scholar] [CrossRef]
  171. Zhao, Y.; Zhang, F.; Mickan, B.; Wang, D.; Wang, W. Physiological, proteomic, and metabolomic analysis provide insights into Bacillus sp.-mediated salt tolerance in wheat. Plant Cell Rep. 2022, 41, 95–118. [Google Scholar] [CrossRef] [PubMed]
  172. Khan, M.N.; Ahmed, I.; Ud Din, I.; Noureldeen, A.; Darwish, H.; Khan, M. Proteomic insight into soybean response to flooding stress reveals changes in energy metabolism and cell wall modifications. PLoS ONE 2022, 17, e0264453. [Google Scholar] [CrossRef] [PubMed]
  173. Komatsu, S.; Yamaguchi, H.; Hitachi, K.; Tsuchida, K.; Rehman, S.U.; Ohno, T. Morphological, biochemical, and proteomic analyses to understand the promotive effects of plant-derived smoke solution on wheat growth under flooding stress. Plants 2022, 11, 1508. [Google Scholar] [CrossRef] [PubMed]
  174. Long, R.; Li, M.; Zhang, T.; Kang, J.; Sun, Y.; Cong, L.; Gao, Y.; Liu, F.; Yang, Q. Comparative proteomic analysis reveals differential root proteins in Medicago sativa and Medicago truncatula in response to salt stress. Front. Plant Sci. 2016, 7, 424. [Google Scholar] [CrossRef] [Green Version]
  175. Zhu, Y.; Jia, X.; Wu, Y.; Hu, Y.; Cheng, L.; Zhao, T.; Huang, Z.; Wang, Y. Quantitative proteomic analysis of Malus halliana exposed to salt-alkali mixed stress reveals alterations in energy metabolism and stress regulation. Plant Growth Regul. 2020, 90, 205–222. [Google Scholar] [CrossRef]
  176. Liu, Y.L.; Shen, Z.J.; Simon, M.; Li, H.; Ma, D.N.; Zhu, X.Y.; Zheng, H.L. Comparative proteomic analysis reveals the regulatory effects of H2S on salt tolerance of mangrove plant Kandelia obovata. Int. J. Mol. Sci. 2019, 21, 118. [Google Scholar] [CrossRef] [Green Version]
  177. Laha, A.; Chakraborty, P.; Banerjee, C.; Panja, A.S.; Bandopadhyay, R. Application of bioinformatics for crop stress response and mitigation. In Sustainable Agriculture in the Era of Climate Change; Roychowdhury, R., Choudhury, S., Hasanuzzaman, M., Srivastava, S., Eds.; Springer-Nature: Basel, Switzerland, 2020; pp. 589–614. ISBN 978-303-045-669-6. [Google Scholar] [CrossRef]
  178. Ambrosino, L.; Colantuono, C.; Diretto, G.; Fiore, A.; Chiusano, M.L. Bioinformatics resources for plant abiotic stress responses: State of the art and opportunities in the fast evolving—Omics era. Plants 2020, 9, 591. [Google Scholar] [CrossRef]
  179. Orozco, A.; Morera, J.; Jiménez, S.; Boza, R. A review of bioinformatics training applied to research in molecular medicine, agriculture and biodiversity in Costa Rica and Central America. Brief Bioinform. 2013, 14, 661–670. [Google Scholar] [CrossRef] [Green Version]
  180. Noor, W.; Sadia, B.; Baloch, I.A.; Ahmad, A.J.; Safia, B.; Farida. Identification and characterization of abiotic stress responsive genes in Ricinus communis L. using bioinformatics tools. Int. J. Biosci. 2020, 16, 23–34. [Google Scholar]
  181. Raza, A.; Tabassum, J.; Fakhar, A.Z.; Sharif, R.; Chen, H.; Zhang, C.; Ju, L.; Fotopoulos, V.; Siddique, K.H.M.; Singh, R.K.; et al. Smart reprograming of plants against salinity stress using modern biotechnological tools. Crit. Rev. Biotechnol. 2022, 1–28. [Google Scholar] [CrossRef]
  182. Larmande, P.; Todorov, K. Revealing genotype-phenotype interactions: The AgroLD experience and challenges. In Integrative Bioinformatics—History and Future; Chen, M., Hofestädt, R., Eds.; Springer-Nature: Singapore, 2022; pp. 321–342. ISBN 978-981-166-795-4. [Google Scholar]
  183. Hussain, B.; Akpınar, B.A.; Alaux, M.; Algharib, A.M.; Sehgal, D.; Ali, Z.; Aradottir, G.I.; Batley, J.; Bellec, A.; Bentley, A.R.; et al. Capturing wheat phenotypes at the genome level. Front. Plant Sci. 2022, 13, 851079. [Google Scholar] [CrossRef]
  184. Li, C.; Chu, W.; Gill, R.A.; Sang, S.; Shi, Y.; Hu, X.; Yang, Y.; Zaman, Q.U.; Zhang, B. Computational tools and resources for CRISPR/Cas genome editing. Genom. Proteom. Bioinform. 2022, in press. [CrossRef]
  185. Bhati, J.; Avashthi, H.; Kumar, A.; Majumdar, S.G.; Budhlakoti, N.; Mishra, D.C. Protocol for identification and annotation of differentially expressed genes using reference-based transcriptomic approach. In Genomics of Cereal Crops; Wani, S.H., Anuj Kumar, A., Eds.; Springer-Nature: Humana, NY, USA, 2022; pp. 175–193. ISBN 978-107-162-533-0. [Google Scholar]
  186. Waddington, C.H. The epigenotype. 1942. Int. J. Epidemiol. 2012, 41, 10–13. [Google Scholar] [CrossRef] [Green Version]
  187. Manning, K.; Tör, M.; Poole, M.; Hong, Y.; Thompson, A.J.; King, G.J.; Giovannoni, J.J.; Seymour, G.B. A naturally occurring epigenetic mutation in a gene encoding an SBP-box transcription factor inhibits tomato fruit ripening. Nat. Genet. 2006, 38, 948–952. [Google Scholar] [CrossRef]
  188. Schmitz, R.J.; Schultz, M.D.; Lewsey, M.G.; O’Malley, R.C.; Urich, M.A.; Libiger, O.; Schork, N.J.; Ecker, J.R. Transgenerational epigenetic instability is a source of novel methylation variants. Science 2011, 334, 369–373. [Google Scholar] [CrossRef] [Green Version]
  189. Lloyd, J.P.B.; Lister, R. Epigenome plasticity in plants. Nat. Rev. Genet. 2022, 23, 55–68. [Google Scholar] [CrossRef]
  190. Zentner, G.E.; Henikoff, S. Regulation of nucleosome dynamics by histone modifications. Nat. Struct. Mol. Biol. 2013, 20, 259–266. [Google Scholar] [CrossRef]
  191. Fitz-James, M.H.; Cavalli, G. Molecular mechanisms of transgenerational epigenetic inheritance. Nat. Rev. Genet. 2022, 23, 325–341. [Google Scholar] [CrossRef]
  192. Cong, W.; Miao, Y.; Xu, L.; Zhang, Y.; Yuan, C.; Wang, J.; Zhuang, T.; Lin, X.; Jiang, L.; Wang, N.; et al. Transgenerational memory of gene expression changes induced by heavy metal stress in rice (Oryza sativa L.). BMC Plant Biol. 2019, 19, 282. [Google Scholar] [CrossRef] [Green Version]
  193. Yung, W.S.; Huang, C.; Li, M.W.; Lam, H.M. Changes in epigenetic features in legumes under abiotic stresses. Plant Genome 2022, e20237. [Google Scholar] [CrossRef] [PubMed]
  194. Dar, F.A.; Mushtaq, N.U.; Saleem, S.; Rehman, R.U.; Dar, T.U.H.; Hakeem, K.R. Role of epigenetics in modulating phenotypic plasticity against abiotic stresses in plants. Int. J. Genom. 2022, 2022, 1092894. [Google Scholar] [CrossRef] [PubMed]
  195. Tiwari, A.; Pandey-Rai, S.; Rai, K.K.; Tiwari, A.; Pandey, N. Molecular and epigenetic basis of heat stress responses and acclimatization in plants. Nucleus 2023, 66, 69–79. [Google Scholar] [CrossRef]
  196. Ali, S.; Khan, N.; Tang, Y. Epigenetic marks for mitigating abiotic stresses in plants. J. Plant Physiol. 2022, 275, 153740. [Google Scholar] [CrossRef] [PubMed]
  197. De Kort, H.; Toivainen, T.; Van Nieuwerburgh, F.; Andrés, J.; Hytönen, T.P.; Honnay, O. Signatures of polygenic adaptation align with genome-wide methylation patterns in wild strawberry plants. New Phytol. 2022, 235, 1501–1514. [Google Scholar] [CrossRef] [PubMed]
  198. Ma, X.; Zhao, F.; Zhou, B. The Characters of Non-Coding RNAs and Their Biological Roles in Plant Development and Abiotic Stress Response. Int. J. Mol. Sci. 2022, 23, 4124. [Google Scholar] [CrossRef]
  199. Lu, X.; Hyun, T.K. The role of epigenetic modifications in plant responses to stress. Bot. Serbica 2021, 45, 3–12. [Google Scholar] [CrossRef]
  200. Choi, Y.; Gehring, M.; Johnson, L.; Hannon, M.; Harada, J.J.; Goldberg, R.B.; Jacobsen, S.E.; Fischer, R.L. DEMETER, a DNA glycosylase domain protein, is required for endosperm gene imprinting and seed viability in Arabidopsis. Cell 2002, 110, 33–42. [Google Scholar] [CrossRef] [Green Version]
  201. Van Oosten, M.J.; Bressan, R.A.; Zhu, J.K.; Bohnert, H.J.; Chinnusamy, V. The role of the epigenome in gene expression control and the epimark changes in response to the environment. Critic. Rev. Plant Sci. 2014, 33, 64–87. [Google Scholar] [CrossRef]
  202. Luo, M.; Cheng, K.; Xu, Y.; Yang, S.; Wu, K. Plant responses to abiotic stress regulated by histone deacetylases. Front. Plant Sci. 2017, 8, 2147. [Google Scholar] [CrossRef] [Green Version]
  203. Tang, K.; Lang, Z.; Zhang, H.; Zhu, J.K. The DNA demethylase ROS1 targets genomic regions with distinct chromatin modifications. Nat. Plants 2016, 2, 16169. [Google Scholar] [CrossRef] [Green Version]
  204. He, L.; Huang, H.; Bradai, M.; Zhao, C.; You, Y.; Ma, J.; Zhao, L.; Lozano-Durán, R.; Zhu, J.K. DNA methylation-free Arabidopsis reveals crucial roles of DNA methylation in regulating gene expression and development. Nat. Commun. 2022, 13, 1335. [Google Scholar] [CrossRef]
  205. Pandey, R.; Müller, A.; Napoli, C.A.; Selinger, D.A.; Pikaard, C.S.; Richards, E.J.; Bender, J.; Mount, D.W.; Jorgensen, R.A. Analysis of histone acetyltransferase and histone deacetylase families of Arabidopsis thaliana suggests functional diversification of chromatin modification among multicellular eukaryotes. Nucleic Acids Res. 2002, 30, 5036–5055. [Google Scholar] [CrossRef] [Green Version]
  206. Earley, K.W.; Shook, M.S.; Brower-Toland, B.; Hicks, L.; Pikaard, C.S. In vitro specificities of Arabidopsis co-activator histone acetyltransferases: Implications for histone hyperacetylation in gene activation. Plant J. 2007, 52, 615–626. [Google Scholar] [CrossRef]
  207. Song, Y.; Ji, D.; Li, S.; Wang, P.; Li, Q.; Xiang, F. The dynamic changes of DNA methylation and histone modifications of salt responsive transcription factor genes in soybean. PLoS ONE 2012, 7, e41274. [Google Scholar] [CrossRef]
  208. Zheng, M.; Liu, X.; Lin, J.; Liu, X.; Wang, Z.; Xin, M.; Yao, Y.; Peng, H.; Zhou, D.X.; Ni, Z.; et al. Histone acetyltransferase GCN5 contributes to cell wall integrity and salt stress tolerance by altering the expression of cellulose synthesis genes. Plant J. 2019, 97, 587–602. [Google Scholar] [CrossRef]
  209. Hu, Z.; Song, N.; Zheng, M.; Liu, X.; Liu, Z.; Xing, J.; Ma, J.; Guo, W.; Yao, Y.; Peng, H.; et al. Histone acetyltransferase GCN5 is essential for heat stress-responsive gene activation and thermotolerance in Arabidopsis. Plant J. 2015, 84, 1178–1191. [Google Scholar] [CrossRef] [Green Version]
  210. Hwarari, D.; Guan, Y.; Ahmad, B.; Movahedi, A.; Min, T.; Hao, Z.; Lu, Y.; Chen, J.; Yang, L. ICE-CBF-COR signaling cascade and its regulation in plants responding to cold stress. Int. J. Mol. Sci. 2022, 23, 1549. [Google Scholar] [CrossRef]
  211. Singh, R.K.; Prasad, M. Delineating the epigenetic regulation of heat and drought response in plants. Crit. Rev. Biotechnol. 2022, 42, 548–561. [Google Scholar] [CrossRef]
  212. Sharma, M.; Kumar, P.; Verma, V.; Sharma, R.; Bhargava, B.; Irfan, M. Understanding plant stress memory response for abiotic stress resilience: Molecular insights and prospects. Plant Physiol. Biochem. 2022, 179, 10–24. [Google Scholar] [CrossRef]
  213. Boden, S.A.; Kavanová, M.; Finnegan, E.J.; Wigge, P.A. Thermal stress effects on grain yield in Brachypodium distachyon occur via H2A.Z-nucleosomes. Genome Biol. 2013, 14, R65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Obata, T.; Fernie, A.R. The use of metabolomics to dissect plant responses to abiotic stresses. Cell Mol. Life Sci. 2012, 69, 3225–3243. [Google Scholar] [CrossRef] [Green Version]
  215. Hong, J.; Yang, L.; Zhang, D.; Shi, J. Plant metabolomics: An indispensable system biology tool for plant science. Int. J. Mol. Sci. 2016, 17, 767. [Google Scholar] [CrossRef] [PubMed]
  216. Chugh, V.; Kaur, N.; Gupta, A.K.; Rai, A. The seed biochemical signature as a potent marker for water logging tolerance in maize. Plant Stress 2022, 4, 100085. [Google Scholar] [CrossRef]
  217. Rinschen, M.M.; Ivanisevic, J.; Giera, M.; Siuzdak, G. Identification of bioactive metabolites using activity metabolomics. Nat. Rev. Mol. Cell Biol. 2019, 20, 353–367. [Google Scholar] [CrossRef]
  218. Fiehn, O.; Kopka, J.; Dörmann, P.; Altmann, T.; Trethewey, R.N.; Willmitzer, L. Metabolite profiling for plant functional genomics. Nat. Biotechnol. 2000, 18, 1157–1161. [Google Scholar] [CrossRef]
  219. Fiehn, O. Metabolomics—The link between genotypes and phenotypes. Plant Mol. Biol. 2002, 48, 155–171. [Google Scholar] [CrossRef]
  220. Verslues, P.E.; Bailey-Serres, J.; Brodersen, C.; Buckley, T.N.; Conti, L.; Christmann, A.; Dinneny, J.R.; Grill, E.; Hayes, S.; Heckman, R.W.; et al. Burning questions for a warming and changing world: 15 unknowns in plant abiotic stress. Plant Cell 2023, 35, 67–105. [Google Scholar] [CrossRef]
  221. Mashabela, M.D.; Masamba, P.; Kappo, A.P. Metabolomics and chemoinformatics in agricultural biotechnology research: Complementary probes in unravelling new metabolites for crop improvement. Biology 2022, 11, 1156. [Google Scholar] [CrossRef]
  222. Hasanuzzaman, M.; Nahar, K.; Alam, M.M.; Roychowdhury, R.; Fujita, M. Physiological, biochemical, and molecular mechanisms of heat stress tolerance in plants. Int. J. Mol. Sci. 2013, 14, 9643–9684. [Google Scholar] [CrossRef]
  223. Roychowdhury, R.; Khan, M.H.; Choudhury, S. Arsenic in rice: An overview on stress implications, tolerance and mitigation strategies. In Plants under Metal and Metalloid Stress: Responses, Tolerance and Remediation; Hasanuzzaman, M., Nahar, K., Fujita, M., Eds.; Springer-Nature: Singapore, 2018; pp. 401–416. ISBN 978-981-132-242-6. [Google Scholar] [CrossRef]
  224. Roychowdhury, R.; Khan, M.H.; Choudhury, S. Physiological and molecular responses for metalloid stress in rice—A Comprehensive Overview. In Advances in Rice Research for Abiotic Stress Tolerance; Hasanuzzaman, M., Fujita, M., Nahar, K., Biswas, J., Eds.; Woodhead Publishing: Cambridge, UK; Elsevier: Amsterdam, The Netherlands, 2019; pp. 341–369. ISBN 978-012-814-333-9. [Google Scholar] [CrossRef]
  225. Gujjar, R.S.; Karkute, S.G.; Rai, A.; Singh, M.; Singh, B. Proline-rich proteins may regulate free cellular proline levels during drought stress in tomato. Curr. Sci. 2018, 114, 915–920. [Google Scholar] [CrossRef]
  226. Patel, J.; Khandwal, D.; Choudhary, B.; Ardeshana, D.; Jha, R.K.; Tanna, B.; Yadav, S.; Mishra, A.; Varshney, R.K.; Siddique, K.H.M. Differential physio-biochemical and metabolic responses of peanut (Arachis hypogaea L.) under multiple abiotic stress conditions. Int. J. Mol. Sci. 2022, 23, 660. [Google Scholar] [CrossRef]
  227. Xu, Y.; Fu, X. Reprogramming of Plant Central Metabolism in Response to Abiotic Stresses: A metabolomics view. Int. J. Mol. Sci. 2022, 23, 5716. [Google Scholar] [CrossRef]
  228. Huchzermeyer, B.; Menghani, E.; Khardia, P.; Shilu, A. Metabolic pathway of natural antioxidants, antioxidant enzymes and ROS providence. Antioxidants 2022, 11, 761. [Google Scholar] [CrossRef]
  229. Ren, G.; Mo, H.; Xu, R. Arginine decarboxylase gene ADC2 regulates fiber elongation in cotton. Genes 2022, 13, 784. [Google Scholar] [CrossRef]
  230. Quan, N.T.; Anh, L.H.; Khang, D.T.; Tuyen, P.T.; Toan, N.P.; Minh, T.N.; Minh, L.T.; Bach, D.T.; Ha, P.T.T.; Elzaawely, A.A.; et al. Involvement of secondary metabolites in response to drought stress of rice (Oryza sativa L.). Agriculture 2016, 6, 23. [Google Scholar] [CrossRef] [Green Version]
  231. Piasecka, A.; Sawikowska, A.; Kuczyńska, A.; Ogrodowicz, P.; Mikołajczak, K.; Krystkowiak, K.; Gudyś, K.; Guzy-Wróbelska, J.; Krajewski, P.; Kachlicki, P. Drought-related secondary metabolites of barley (Hordeum vulgare L.) leaves and their metabolomic quantitative trait loci. Plant J. 2017, 89, 898–913. [Google Scholar] [CrossRef] [Green Version]
  232. Radwan, A.; Kleinwächter, M.; Selmar, D. Impact of drought stress on specialised metabolism: Biosynthesis and the expression of monoterpene synthases in sage (Salvia officinalis). Phytochemistry 2017, 141, 20–26. [Google Scholar] [CrossRef]
  233. Nawaz, M.; Hassan, M.U.; Chattha, M.U.; Mahmood, A.; Shah, A.N.; Hashem, M.; Alamri, S.; Batool, M.; Rasheed, A.; Thabit, M.A.; et al. Trehalose: A promising osmo-protectant against salinity stress-physiological and molecular mechanisms and future prospective. Mol. Biol. Rep. 2022, 49, 11255–11271. [Google Scholar] [CrossRef]
  234. Biondi, S.; Antognoni, F.; Marincich, L.; Lianza, M.; Tejos, R.; Ruiz, K.B. The polyamine “multiverse” and stress mitigation in crops: A case study with seed priming in quinoa. Sci. Hort. 2022, 304, 111292. [Google Scholar] [CrossRef]
  235. Masouleh, S.S.S.; Sassine, Y.N. Molecular and biochemical responses of horticultural plants and crops to heat stress. Ornam. Hort. 2020, 26, 148–158. [Google Scholar] [CrossRef]
  236. Paupière, M.J.; Tikunov, Y.; Schleiff, E.; Bovy, A.; Fragkostefanakis, S. Reprogramming of tomato leaf metabolome by the activity of heat stress transcription factor HsfB1. Front. Plant Sci. 2020, 11, 610599. [Google Scholar] [CrossRef] [PubMed]
  237. Rouphael, Y.; Raimondi, G.; Lucini, L.; Carillo, P.; Kyriacou, M.C.; Colla, G.; Cirillo, V.; Pannico, A.; El-Nakhel, C.; De Pascale, S. Physiological and metabolic responses triggered by omeprazole improve tomato plant tolerance to NaCl stress. Front. Plant Sci. 2018, 9, 249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Sabreena; Hassan, S. Plant life under changing environment: An exertion of environmental factors in oxidative stress modulation. In Antioxidant Defense in Plants-Molecular Basis of Regulation; Aftab, T., Hakeem, K.R., Eds.; Springer-Nature: Singapore, 2022; pp. 421–433. ISBN 978-981-167-981-0. [Google Scholar]
  239. Rai, G.K.; Mushtaq, M.; Bhat, B.A.; Kumar, R.R.; Singh, M.; Rai, P.K. Reactive Oxygen Species: Friend or Foe. In Thermotolerance in Crop Plants; Kumar, R.R., Praveen, S., Rai, G.K., Eds.; Springer-Nature: Singapore, 2022; pp. 129–162. ISBN 978-981-193-800-9. [Google Scholar]
  240. Parida, A.K.; Kumari, A.; Rangani, J.; Patel, M. Halophytes: Potential resources of coastal ecosystems and their economic, ecological and bioprospecting significance. In Halophytes and Climate Change: Adaptive Mechanisms and Potential Uses; Hasanuzzaman, M., Shabala, S., Fujita, M., Eds.; CABI: Oxfordshire, UK, 2019; pp. 287–323. ISBN 978-178-639-433-0. [Google Scholar]
  241. Sudhakar, C.; Veeranagamallaiah, G.; Nareshkumar, A.; Sudhakarbabu, O.; Sivakumar, M.; Pandurangaiah, M.; Kiranmai, K.; Lokesh, U. Polyamine metabolism influences antioxidant defense mechanism in foxtail millet (Setaria italica L.) cultivars with different salinity tolerance. Plant Cell Rep. 2015, 34, 141–156. [Google Scholar] [CrossRef] [PubMed]
  242. Hossain, M.A.; Hoque, T.S.; Zaid, A.; Wani, S.H.; Mostofa, M.G.; Henry, R. Targeting the ascorbate-glutathione pathway and the glyoxalase pathway for genetic engineering of abiotic stress-tolerance in rice. In Molecular Breeding for Rice Abiotic Stress Tolerance and Nutritional Quality; Hossain, M.A., Hassan, L., Ifterkharuddaula, K.M., Kumar, A., Henry, R., Eds.; John Wiley & Sons, Ltd.: Chichester, UK, 2021; pp. 398–427. ISBN 978-111-963-317-4. [Google Scholar]
  243. Sharma, P.; Dubey, R.S. Protein synthesis by plants under stressful conditions. In Handbook of Plant and Crop Stress, 4th ed.; Pessarakli, M., Ed.; CRC Press: Boca Raton, UK, 2019; pp. 405–449. ISBN 978-135-110-460-9. [Google Scholar]
  244. Chaturvedi, S.; Khan, S.; Bhunia, R.K.; Kaur, K.; Tiwari, S. Metabolic engineering in food crops to enhance ascorbic acid production: Crop biofortification perspectives for human health. Physiol. Mol. Biol. Plants 2022, 28, 871–884. [Google Scholar] [CrossRef]
  245. Putri, S.P.; Yamamoto, S.; Tsugawa, H.; Fukusaki, E. Current metabolomics: Technological advances. J. Biosci. Bioeng. 2013, 116, 9–16. [Google Scholar] [CrossRef]
  246. Ghatak, A.; Chaturvedi, P.; Weckwerth, W. Metabolomics in plant stress physiology. In Plant Genetics and Molecular Biology (Advances in Biochemical Engineering/Biotechnology, Vol 164); Varshney, R., Pandey, M., Chitikineni, A., Eds.; Springer: Cham, Switzerland, 2018; pp. 187–236. [Google Scholar] [CrossRef]
  247. Jurowski, K.; Kochan, K.; Walczak, J.; Barańska, M.; Piekoszewski, W.; Buszewski, B. Analytical techniques in lipidomics: State of the art. Crit. Rev. Anal. Chem. 2017, 47, 418–437. [Google Scholar] [CrossRef]
  248. Alseekh, S.; Aharoni, A.; Brotman, Y.; Contrepois, K.; D’auria, J.; Ewald, J.; Ewald, J.C.; Fraser, P.D.; Giavalisco, P.; Hall, R.D.; et al. Mass spectrometry-based metabolomics: A guide for annotation, quantification and best reporting practices. Nat. Methods 2021, 18, 747–756. [Google Scholar] [CrossRef]
  249. Pua, A.; Goh, R.M.V.; Huang, Y.; Tang, V.C.Y.; Ee, K.H.; Cornuz, M.; Liu, S.Q.; Lassabliere, B.; Yu, B. Recent advances in analytical strategies for coffee volatile studies: Opportunities and challenges. Food Chem. 2022, 388, 132971. [Google Scholar] [CrossRef]
  250. Brunetti, A.E.; Carnevale Neto, F.; Vera, M.C.; Taboada, C.; Pavarini, D.P.; Bauermeister, A.; Lopes, N.P. An integrative omics perspective for the analysis of chemical signals in ecological interactions. Chem. Soc. Rev. 2018, 47, 1574–1591. [Google Scholar] [CrossRef]
  251. Zhu, F.Y.; Chen, M.X.; Ye, N.H.; Shi, L.; Ma, K.L.; Yang, J.F.; Cao, Y.Y.; Zhang, Y.; Yoshida, T.; Fernie, A.R.; et al. Proteogenomic analysis reveals alternative splicing and translation as part of the abscisic acid response in Arabidopsis seedlings. Plant J. 2017, 91, 518–533. [Google Scholar] [CrossRef] [Green Version]
  252. Ruggles, K.V.; Krug, K.; Wang, X.; Clauser, K.R.; Wang, J.; Payne, S.H.; Fenyö, D.; Zhang, B.; Mani, D.R. Methods, tools and current perspectives in proteogenomics. Mol. Cell. Proteom. 2017, 16, 959–981. [Google Scholar] [CrossRef] [Green Version]
  253. Ullah, M.A.; Abdullah-zawawi, M.R.; Zainal-abidin, R.A.; Sukiran, N.L.; Uddin, M.I.; Zainal, Z. A review of integrative omic approaches for understanding rice salt response mechanisms. Plants 2022, 11, 1430. [Google Scholar] [CrossRef]
  254. Low, T.Y.; Mohtar, M.A.; Ang, M.Y.; Jamal, R. Connecting proteomics to next-generation sequencing: Proteogenomics and its current applications in biology. Proteomics 2019, 19, e1800235. [Google Scholar] [CrossRef]
  255. Gupta, A.; Shaw, B.P.; Sahu, B.B. Post-translational regulation of the membrane transporters contributing to salt tolerance in plants. Funct. Plant Biol. 2021, 48, 1199–1212. [Google Scholar] [CrossRef]
  256. Pranneshraj, V.; Sangha, M.K.; Djalovic, I.; Miladinovic, J.; Djanaguiraman, M. Lipidomics-assisted GWAS (LGWAS) approach for improving high-temperature stress tolerance of crops. Int. J. Mol. Sci. 2022, 23, 9389. [Google Scholar] [CrossRef]
  257. Buffagni, V.; Zhang, L.; Senizza, B.; Rocchetti, G.; Ferrarini, A.; Miras-Moreno, B.; Lucini, L. Metabolomics and lipidomics insight into the effect of different polyamines on tomato plants under non-stress and salinity conditions. Plant Sci. 2022, 322, 111346. [Google Scholar] [CrossRef]
  258. Zhang, D.; Li, J.; Li, M.; Cheng, Z.; Xu, Q.; Song, X.; Shang, X.; Guo, W. Overexpression of a cotton nonspecific lipid transfer protein gene, GhLTP4, enhances drought tolerance by remodeling lipid profiles, regulating abscisic acid homeostasis and improving tricarboxylic acid cycle in cotton. Environ. Exp. Bot. 2022, 201, 104991. [Google Scholar] [CrossRef]
  259. Moradi, P.; Mahdavi, A.; Khoshkam, M.; Iriti, M. Lipidomics unravels the role of leaf lipids in thyme plant response to drought stress. Int. J. Mol. Sci. 2017, 18, 2067. [Google Scholar] [CrossRef] [Green Version]
  260. Tarazona, P.; Feussner, K.; Feussner, I. An enhanced plant lipidomics method based on multiplexed liquid chromatography-mass spectrometry reveals additional insights into cold- and drought-induced membrane remodeling. Plant J. 2015, 84, 621–633. [Google Scholar] [CrossRef]
  261. Liu, B.; Wang, X.; Li, K.; Cai, Z. Spatially resolved metabolomics and lipidomics reveal salinity and drought-tolerant mechanisms of cottonseeds. J. Agric. Food Chem. 2021, 69, 8028–8037. [Google Scholar] [CrossRef] [PubMed]
  262. Engel, K.M.; Prabutzki, P.; Leopold, J.; Nimptsch, A.; Lemmnitzer, K.; Vos, D.R.N.; Hopf, C.; Schiller, J. A new update of MALDI-TOF mass spectrometry in lipid research. Prog. Lipid Res. 2022, 86, 101145. [Google Scholar] [CrossRef] [PubMed]
  263. Sun, A.Z.; Chen, L.S.; Tang, M.; Chen, J.H.; Li, H.; Jin, X.Q.; Yi, Y.; Guo, F.Q. Lipidomic remodeling in Begonia grandis under heat stress. Front. Plant Sci. 2022, 13, 843942. [Google Scholar] [CrossRef] [PubMed]
  264. Baxter, I. Ionomics: Studying the social network of mineral nutrients. Curr. Opin. Plant Biol. 2009, 12, 381–386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Salt, D.E.; Baxter, I.; Lahner, B. Ionomics and the study of the plant ionome. Annu. Rev. Plant Biol. 2008, 59, 709–733. [Google Scholar] [CrossRef] [Green Version]
  266. Cao, Y.; Du, P.; Ji, J.; He, X.; Zhang, J.; Shang, Y.; Liu, H.; Xu, J.; Liang, B. Ionomic combined with transcriptomic and metabolomic analyses to explore the mechanism underlying the effect of melatonin in relieving nutrient stress in apple. Int. J. Mol. Sci. 2022, 23, 9855. [Google Scholar] [CrossRef]
  267. Guo, J.; Lu, X.; Tao, Y.; Guo, H.; Min, W. Comparative ionomics and metabolic responses and adaptive strategies of cotton to salt and alkali stress. Front. Plant Sci. 2022, 13, 871387. [Google Scholar] [CrossRef]
  268. Hua, Y.P.; Wang, Y.; Zhou, T.; Huang, J.Y.; Yue, C.P. Combined morpho-physiological, ionomic and transcriptomic analyses reveal adaptive responses of allohexaploid wheat (Triticum aestivum L.) to iron deficiency. BMC Plant Biol. 2022, 22, 234. [Google Scholar] [CrossRef]
  269. Muszyńska, E.; Labudda, M. Dual role of metallic trace elements in stress biology—From negative to beneficial impact on plants. Int. J. Mol. Sci. 2019, 20, 3117. [Google Scholar] [CrossRef] [Green Version]
  270. Kirkby, E. Introduction, definition and classification of nutrients. In Marschner’s Mineral Nutrition of Higher Plants; Marschner, P., Ed.; Academic Press: Cambridge, MA, USA, 2012; pp. 3–5. ISBN 978-012-384-905-2. [Google Scholar] [CrossRef]
  271. Zhang, R.; Wang, Y.; Hussain, S.; Yang, S.; Li, R.; Liu, S.; Chen, Y.; Wei, H.; Dai, Q.; Hou, H. Study on the effect of salt stress on yield and grain quality among different rice varieties. Front. Plant Sci. 2022, 13, 918460. [Google Scholar] [CrossRef]
  272. Gupta, A.; Shaw, B.P. Augmenting salt tolerance in rice by regulating uptake and tissue specific accumulation of Na+- through Ca2+-induced alteration of biochemical events. Plant Biol. 2021, 23, 122–130. [Google Scholar] [CrossRef]
  273. Balemi, T.; Negisho, K. Management of soil phosphorus and plant adaptation mechanisms to phosphorus stress for sustainable crop production: A review. J. Soil Sci. Plant Nutr. 2012, 12, 547–561. [Google Scholar] [CrossRef] [Green Version]
  274. Yugandhar, P.; Veronica, N.; Subrahmanyam, D.; Brajendra, P.; Nagalakshmi, S.; Srivastava, A.; Voleti, S.R.; Sarla, N.; Sundaram, R.M.; Sevanthi, A.M.; et al. Revealing the effect of seed phosphorus concentration on seedling vigour and growth of rice using mutagenesis approach. Sci. Rep. 2022, 12, 1203. [Google Scholar] [CrossRef]
  275. Shafi, A.; Zahoor, I.; Habib, H. Omics technologies to unravel plant-microbe interactions. In Plant-Microbe Dynamics: Recent Advances for Sustainable Agriculture; Pirzadah, T.B., Malik, B., Hakeem, K.R., Eds.; CRC Press: Boca Raton, UK, 2021; pp. 201–220. ISBN 978-036-761-838-4. [Google Scholar]
  276. Gupta, O.P.; Deshmukh, R.; Kumar, A.; Singh, S.K.; Sharma, P.; Ram, S.; Singh, G.P. From gene to biomolecular networks: A review of evidences for understanding complex biological function in plants. Curr. Opin. Biotechnol. 2022, 74, 66–74. [Google Scholar] [CrossRef]
  277. Hajheidari, M.; Huang, S.C. Elucidating the biology of transcription factor–DNA interaction for accurate identification of Cis-regulatory elements. Curr. Opin. Plant Biol. 2022, 68, 102232. [Google Scholar] [CrossRef]
  278. Di Silvestre, D.; Bergamaschi, A.; Bellini, E.; Mauri, P.L. Large scale proteomic data and network-based systems biology approaches to explore the plant world. Proteomes 2018, 6, 27. [Google Scholar] [CrossRef] [Green Version]
  279. Kapoor, B.; Kumar, P.; Gill, N.S.; Sharma, R.; Thakur, N.; Irfan, M. Molecular mechanisms underpinning the silicon-selenium (Si-Se) interactome and cross-talk in stress-induced plant responses. Plant Soil 2023, 486, 45–68. [Google Scholar] [CrossRef]
  280. Yadav, R.; Santal, A.R.; Singh, N.P. Root protein interactomics of salt stress-induced proteins of wheat genotypes KH-65 (salt-tolerant) and PBW-373 (salt-susceptible). J. Appl. Biotechnol. Rep. 2022, 9, 632–639. [Google Scholar] [CrossRef]
  281. Perez-Sanz, F.; Navarro, P.J.; Egea-Cortines, M. Plant phenomics: An overview of image acquisition technologies and image data analysis algorithms. Gigascience 2017, 6, gix092. [Google Scholar] [CrossRef] [Green Version]
  282. Kumar, J.; Kumar, S.; Pratap, A. Phenomics in Crop Plants: Trends, Options and Limitations; Springer: New Delhi, India, 2015. [Google Scholar] [CrossRef]
  283. Plomin, R.; DeFries, J.C.; Loehlin, J.C. Genotype-environment interaction and correlation in the analysis of human behavior. Psychol. Bull. 1977, 84, 309–322. [Google Scholar] [CrossRef]
  284. Roychowdhury, R.; Zilberman, O.; Chandrasekhar, K.; Curzon, A.Y.; Nashef, K.; Abbo, S.; Slafer, G.A.; Bonfil, D.J.; Ben-David, R. Pre-anthesis spike growth dynamics and its association to yield components among elite bread wheat cultivars (Triticum aestivum L. spp.) under Mediterranean climate. Field Crops Res. 2023, 298, 108948. [Google Scholar] [CrossRef]
  285. Van Bezouw, R.F.H.M.; Keurentjes, J.J.B.; Harbinson, J.; Aarts, M.G.M. Converging phenomics and genomics to study natural variation in plant photosynthetic efficiency. Plant J. 2019, 97, 112–133. [Google Scholar] [CrossRef] [PubMed]
  286. Pasala, R.; Pandey, B.B. Plant phenomics: High-throughput technology for accelerating genomics. J. Biosci. 2020, 45, 111. [Google Scholar] [CrossRef]
  287. Furbank, R.T.; Tester, M. Phenomics—Technologies to relieve the phenotyping bottleneck. Trends Plant Sci. 2011, 16, 635–644. [Google Scholar] [CrossRef]
  288. Arya, S.; Sandhu, K.S.; Singh, J.; Kumar, S. Deep learning: As the new frontier in high-throughput plant phenotyping. Euphytica 2022, 218, 47. [Google Scholar] [CrossRef]
  289. Ninomiya, S. High-throughput field crop phenotyping: Current status and challenges. Breed. Sci. 2022, 72, 3–18. [Google Scholar] [CrossRef]
  290. Laha, S.D.; Naskar, A.J.; Sarkar, T.; Guha, S.; Mondal, H.A.; Das, M. Field phenotyping for salt tolerance and imaging techniques for crop stress biology. In Intelligent Image Analysis for Plant Phenotyping; Samal, A., Choudhury, S.D., Eds.; CRC Press: Boca Raton, UK, 2020; pp. 287–304. ISBN 978-131-517-730-4. [Google Scholar]
  291. Tayade, R.; Yoon, J.; Lay, L.; Khan, A.L.; Yoon, Y.; Kim, Y. Utilization of spectral indices for high-throughput phenotyping. Plants 2022, 11, 1712. [Google Scholar] [CrossRef]
  292. Al-Tamimi, N.; Langan, P.; Bernád, V.; Walsh, J.; Mangina, E.; Negrão, S. Capturing crop adaptation to abiotic stress using image-based technologies. Open Biol. 2022, 12, 210353. [Google Scholar] [CrossRef]
  293. Langan, P.; Bernád, V.; Walsh, J.; Henchy, J.; Khodaeiaminjan, M.; Mangina, E.; Negrão, S. Phenotyping for waterlogging tolerance in crops: Current trends and future prospects. J. Exp. Bot. 2022, 73, 5149–5169. [Google Scholar] [CrossRef]
  294. Vines, P.L.; Zhang, J. High-throughput plant phenotyping for improved turfgrass breeding applications. Grass Res. 2022, 2, 1. [Google Scholar] [CrossRef]
  295. Burchard-Levine, V.; Arribas, D.R.; del Hoyo, L.V.; Isabel, M.D.P.M.; Corral, J.B. A review of in-situ sampling protocols to support the remote sensing of vegetation. GeoFocus—Int. Rev. Geo. Info. Sci. Technol. 2022, 29, 59–87. [Google Scholar]
  296. Maphosa, L.; Emily Thoday-Kennedy, E.; Vakani, J.; Phelan, A.; Badenhorst, P.; Slater, A.; Spangenberg, G.; Kant, S. Phenotyping wheat under salt stress conditions using a 3D laser scanner. Israel J. Plant Sci. 2017, 64, 55–62. [Google Scholar] [CrossRef]
  297. Solovchenko, A.; Lukyanov, A.; Vasilieva, S.; Lobakova, E. Chlorophyll fluorescence as a valuable multitool for microalgal biotechnology. Biophys. Rev. 2022, 14, 973–983. [Google Scholar] [CrossRef]
  298. Berger, K.; Machwitz, M.; Kycko, M.; Kefauver, S.C.; Van Wittenberghe, S.; Gerhards, M.; Verrelst, J.; Atzberger, C.; van der Tol, C.; Damm, A.; et al. Multi-sensor spectral synergies for crop stress detection and monitoring in the optical domain: A review. Remote Sens. Environ. 2022, 280, 113198. [Google Scholar] [CrossRef]
  299. Xiao, Q.; Bai, X.; Zhang, C.; He, Y. Advanced high-throughput plant phenotyping techniques for genome-wide association studies: A review. J. Adv. Res. 2022, 35, 215–230. [Google Scholar] [CrossRef]
  300. Mishra, A.; Mishra, K.B.; Höermiller, I.I.; Heyer, A.G.; Nedbal, L. Chlorophyll fluorescence emission as a reporter on cold tolerance in Arabidopsis thaliana accessions. Plant Signal. Behav. 2011, 6, 301–310. [Google Scholar] [CrossRef] [Green Version]
  301. Zuo, G.; Aiken, R.M.; Feng, N.; Zheng, D.; Zhao, H.; Avenson, T.J.; Lin, X. Fresh perspectives on an established technique: Pulsed amplitude modulation chlorophyll a fluorescence. Plant-Environ. Interact 2022, 3, 41–59. [Google Scholar] [CrossRef]
  302. Calzadilla, P.I.; Carvalho, F.E.L.; Gomez, R.; Lima Neto, M.C.; Signorelli, S. Assessing photosynthesis in plant systems: A cornerstone to aid in the selection of resistant and productive crops. Environ. Exp. Bot. 2022, 201, 104950. [Google Scholar] [CrossRef]
  303. Fu, P.; Montes, C.M.; Siebers, M.H.; Gomez-Casanovas, N.; McGrath, J.M.; Ainsworth, E.A.; Bernacchi, C.J. Advances in field-based high-throughput photosynthetic phenotyping. J. Exp. Bot. 2022, 73, 3157–3172. [Google Scholar] [CrossRef]
  304. Osmolovskaya, N.; Shumilina, J.; Kim, A.; Didio, A.; Grishina, T.; Bilova, T.; Keltsieva, O.A.; Zhukov, V.; Tikhonovich, I.; Tarakhovskaya, E.; et al. Methodology of drought stress research: Experimental setup and physiological characterization. Int. J. Mol. Sci. 2018, 19, 4089. [Google Scholar] [CrossRef] [Green Version]
  305. Singh, B.; Mishra, S.; Bohra, A.; Joshi, R.; Siddique, K.H. Crop phenomics for abiotic stress tolerance in crop plants. In Biochemical, Physiological and Molecular Avenues for Combating Abiotic Stress Tolerance in Plants; Wani, S.H., Ed.; Academic Press: Cambridge, MA, USA, 2018; pp. 277–296. ISBN 978-012-813-066-7. [Google Scholar]
  306. Jansen, M.; Gilmer, F.; Biskup, B.; Nagel, K.A.; Rascher, U.; Fischbach, A.; Briem, S.; Dreissen, G.; Tittmann, S.; Braun, S.; et al. Simultaneous phenotyping of leaf growth and chlorophyll fluorescence via growscreen fluoro allows detection of stress tolerance in Arabidopsis thaliana and other rosette plants. Funct. Plant Biol. 2009, 36, 902–914. [Google Scholar] [CrossRef] [PubMed]
  307. Dornbusch, T.; Lorrain, Ś.; Kuznetsov, D.; Fortier, A.; Liechti, R.; Xenarios, I.; Fankhauser, C. Measuring the diurnal pattern of leaf hyponasty and growth in Arabidopsis a novel phenotyping approach using laser scanning. Funct. Plant Biol. 2012, 39, 860–869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Acosta-Gamboa, L.M.; Suxing, L.; Jarrod, W.C.; Zachary, C.C.; Raquel, T.; Jessica, P.Y.-C.; Argelia, L. Characterization of the response to abiotic stresses of high ascorbate Arabidopsis lines using phenomic approaches. Plant Physiol. Biochem. 2020, 151, 500–515. [Google Scholar] [CrossRef] [PubMed]
  309. Kuromori, T.; Fujita, M.; Takahashi, F.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Inter-tissue and inter-organ signaling in drought stress response and phenotyping of drought tolerance. Plant J. 2022, 109, 342–358. [Google Scholar] [CrossRef]
  310. Honsdorf, N.; March, T.J.; Berger, B.; Tester, M.; Pillen, K. High-throughput phenotyping to detect drought tolerance QTL in wild barley introgression lines. PLoS ONE 2014, 9, e97047. [Google Scholar] [CrossRef] [Green Version]
  311. Ge, Y.; Bai, G.; Stoerger, V.; Schnable, J.C. Temporal dynamics of maize plant growth, water use, and leaf water content using automated high throughput RGB and hyperspectral imaging. Comput. Electron. Agric. 2016, 127, 625–632. [Google Scholar] [CrossRef] [Green Version]
  312. Meng, R.; Saade, S.; Kurtek, S.; Berger, B.; Brien, C.; Pillen, K.; Tester, M.; Sun, Y. Growth curve registration for evaluating salinity tolerance in barley. Plant Methods 2017, 13, 18. [Google Scholar] [CrossRef] [Green Version]
  313. Hairmansis, A.; Berger, B.; Tester, M.; Roy, S.J. Image-based phenotyping for non-destructive screening of different salinity tolerance traits in rice. Rice 2014, 7, 16. [Google Scholar] [CrossRef] [Green Version]
  314. Humplík, J.F.; Lazár, D.; Husičková, A.; Spíchal, L. Automated phenotyping of plant shoots using imaging methods for analysis of plant stress responses—A review. Plant Methods 2015, 11, 29. [Google Scholar] [CrossRef] [Green Version]
  315. Briglia, N.; Nuzzo, V.; Petrozza, A.; Summerer, S.; Cellini, F.; Montanaro, G. Preliminary high-throughput phenotyping analysis in grapevines under drought. BIO Web Conf. 2019, 13, 02003. [Google Scholar] [CrossRef]
  316. Mutava, R.N.; Prince, S.J.K.; Syed, N.H.; Song, L.; Valliyodan, B.; Chen, W.; Nguyen, H.T. Understanding abiotic stress tolerance mechanisms in soybean: A comparative evaluation of soybean response to drought and flooding stress. Plant Physiol. Biochem. 2015, 86, 109–120. [Google Scholar] [CrossRef]
  317. Yu, H.; Kong, B.; Hou, Y.; Xu, X.; Chen, T.; Liu, X. A critical review on applications of hyperspectral remote sensing in crop monitoring. Exp. Agric. 2022, 58, e26. [Google Scholar] [CrossRef]
  318. Ludovisi, R.; Tauro, F.; Salvati, R.; Khoury, S.; Mugnozza, G.S.; Harfouche, A. UAV-based thermal imaging for high-throughput field phenotyping of black poplar response to drought. Front. Plant Sci. 2017, 8, 1681. [Google Scholar] [CrossRef]
  319. Al-Rahbi, S.; Al-Mulla, Y.A.; Jayasuriya, H.; Choudri, B. Analysis of true-color images from unmanned aerial vehicle to assess salinity stress on date palm. J. Appl. Remote Sens. 2019, 13, 34514. [Google Scholar] [CrossRef]
  320. Johansen, K.; Morton, M.J.L.; Malbeteau, Y.M.; Aragon, B.; Al-Mashharawi, S.K.; Ziliani, M.G.; Angel, Y.; Fiene, G.M.; Negrão, S.S.C.; Mousa, M.A.A.; et al. Unmanned aerial vehicle-based phenotyping using morphometric and spectral analysis can quantify responses of wild tomato plants to salinity stress. Front. Plant Sci. 2019, 10, 370. [Google Scholar] [CrossRef]
  321. Aharon, S.; Peleg, Z.; Argaman, E.; Ben-David, R.; Lati, R.N. Image-based high-throughput phenotyping of cereals early vigor and weed-competitiveness traits. Remote Sens. 2020, 12, 3877. [Google Scholar] [CrossRef]
  322. Aharon, S.; Fadida-Myers, A.; Nashef, K.; Ben-David, R.; Lati, R.N.; Peleg, Z. Genetic improvement of wheat early vigor promote weed-competitiveness under Mediterranean climate. Plant Sci. 2021, 303, 110785. [Google Scholar] [CrossRef]
  323. Wang, D.; Cao, W.; Zhang, F.; Li, Z.; Xu, S.; Wu, X. A review of deep learning in multiscale agricultural sensing. Remote Sens. 2022, 14, 559. [Google Scholar] [CrossRef]
  324. Lopes, M.S.; Reynolds, M.P. Stay-green in spring wheat can be determined by spectral reflectance measurements (normalized difference vegetation index) independently from phenology. J. Exp. Bot. 2012, 63, 3789–3798. [Google Scholar] [CrossRef] [Green Version]
  325. Galiano, S.G. Assessment of vegetation indexes from remote sensing: Theoretical basis. Options Méditerranéennes 2012, 67, 65–75. [Google Scholar]
  326. Sangwan, S.; Ram, K.; Rani, P.; Munjal, R. Effect of terminal high temperature on chlorophyll content and normalized difference vegetation index in recombinant inbred lines of bread wheat. Int. J. Curr. Microbiol. Appl. Sci. 2018, 7, 1174–1183. [Google Scholar] [CrossRef]
  327. Beisel, N.S.; Callaham, J.B.; Sng, N.J.; Taylor, D.J.; Paul, A.L.; Ferl, R.J. Utilization of single-image normalized difference vegetation index (SI-NDVI) for early plant stress detection. Appl. Plant Sci. 2018, 6, e01186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  328. De Castro, A.I.; Shi, Y.; Maja, J.M.; Peña, J.M. UAVs for vegetation monitoring: Overview and recent scientific contributions. Remote Sens. 2021, 13, 2139. [Google Scholar] [CrossRef]
  329. Jones, H.G.; Serraj, R.; Loveys, B.R.; Xiong, L.; Wheaton, A.; Price, A.H. Thermal infrared imaging of crop canopies for the remote diagnosis and quantification of plant responses to water stress in the field. Funct. Plant Biol. 2009, 36, 978–989. [Google Scholar] [CrossRef] [Green Version]
  330. Romero, P.; Navarro, J.M.; Ordaz, P.B. Towards a sustainable viticulture: The combination of deficit irrigation strategies and agroecological practices in Mediterranean vineyards. a review and update. Agric. Water Manag. 2022, 259, 107216. [Google Scholar] [CrossRef]
  331. Sweet, D.D.; Tirado, S.B.; Springer, N.M.; Hirsch, C.N.; Hirsch, C.D. Opportunities and challenges in phenotyping row crops using drone-based RGB imaging. Plant Phenome J. 2022, 5, e20044. [Google Scholar] [CrossRef]
  332. Tripodi, P.; Nicastro, N.; Pane, C.; Cammarano, D. Digital applications and artificial intelligence in agriculture toward next-generation plant phenotyping. Crop. Pasture Sci. 2022. [Google Scholar] [CrossRef]
  333. Duruflé, H.; Déjean, S. Multi-omics data integration in the context of plant abiotic stress signaling. In Plant Abiotic Stress Signaling; Methods in Molecular Biology; Couée, I., Ed.; Humana Press: New York, NY, USA, 2023; Volume 2642, pp. 295–318. [Google Scholar] [CrossRef]
  334. Gupta, S.; Kaur, R.; Sharma, T.; Bhardwaj, A.; Sharma, S.; Sohal, J.S.; Singh, S.V. Multi-omics approaches for understanding stressor-induced physiological changes in plants: An updated overview. Physiol. Mol. Plant Pathol. 2023, 126, 102047. [Google Scholar] [CrossRef]
  335. Liu, T.; Salguero, P.; Petek, M.; Martinez-Mira, C.; Balzano-Nogueira, L.; Ramšak, Ž.; McIntyre, L.; Gruden, K.; Tarazona, S.; Conesa, A. PaintOmics 4: New tools for the integrative analysis of multi-omics datasets supported by multiple pathway databases. Nucleic Acids Res. 2022, 50, W551–W559. [Google Scholar] [CrossRef]
  336. Mahmood, U.; Li, X.; Fan, Y.; Chang, W.; Niu, Y.; Li, J.; Qu, C.; Lu, K. Multi-omics revolution to promote plant breeding efficiency. Front. Plant Sci. 2022, 13, 1062952. [Google Scholar] [CrossRef]
  337. Yoosefzadeh Najafabadi, M.; Hesami, M.; Eskandari, M. Machine learning-assisted approaches in modernized plant breeding programs. Genes 2023, 14, 777. [Google Scholar] [CrossRef]
  338. Reimer, J.J.; Shaaban, B.; Drummen, N.; Sanjeev Ambady, S.; Genzel, F.; Poschet, G.; Wiese-Klinkenberg, A.; Usadel, B.; Wormit, A. Capsicum leaves under stress: Using multi-omics analysis to detect abiotic stress network of secondary metabolism in two species. Antioxidants 2022, 11, 671. [Google Scholar] [CrossRef]
  339. Gouesbet, G. Deciphering macromolecular interactions involved in abiotic stress signaling: A review of bioinformatics analysis. In Plant Abiotic Stress Signaling; Methods in Molecular Biology; Couée, I., Ed.; Humana Press: New York, NY, USA, 2023; Volume 2642, pp. 257–294. [Google Scholar] [CrossRef]
  340. Guerrero-Sánchez, V.M.; López-Hidalgo, C.; Rey, M.D.; Castillejo, M.Á.; Jorrín-Novo, J.V.; Escandón, M. Multiomic Data integration in the analysis of drought-responsive mechanisms in Quercus ilex seedlings. Plants 2022, 11, 3067. [Google Scholar] [CrossRef]
  341. Bittencourt, C.B.; Carvalho da Silva, T.L.; Rodrigues Neto, J.C.; Vieira, L.R.; Leão, A.P.; de Aquino Ribeiro, J.A.; Abdelnur, P.V.; de Sousa, C.A.F.; Souza, M.T. Insights from a multi-omics integration (MOI) study in oil palm (Elaeis guineensis Jacq.) response to abiotic stresses: Part one-salinity. Plants 2022, 11, 1755. [Google Scholar] [CrossRef]
  342. Leão, A.P.; Bittencourt, C.B.; Carvalho da Silva, T.L.; Rodrigues Neto, J.C.; Braga, Í.O.; Vieira, L.R.; de Aquino Ribeiro, J.A.; Abdelnur, P.V.; de Sousa, C.A.F.; Souza Júnior, M.T. Insights from a multi-omics integration (MOI) study in oil palm (Elaeis guineensis Jacq.) response to abiotic stresses: Part two-drought. Plants 2022, 11, 2786. [Google Scholar] [CrossRef]
  343. Kudapa, H.; Ghatak, A.; Barmukh, R.; Chaturvedi, P.; Khan, A.; Kale, S.; Fragner, L.; Chitikineni, A.; Weckwerth, W.; Varshney, R.K. Integrated multi-omics analysis reveals drought stress response mechanism in chickpea (Cicer arietinum L.). Plant Genome 2023, e20337. [Google Scholar] [CrossRef]
Figure 1. Integrative multi-omics approaches to confer abiotic stress tolerance in plants. The diagram was created using BioRender (https://biorender.com/) premium version.
Figure 1. Integrative multi-omics approaches to confer abiotic stress tolerance in plants. The diagram was created using BioRender (https://biorender.com/) premium version.
Genes 14 01281 g001
Figure 2. Different cohorts of genomics for crop assessment and improvement in relation to abiotic-stress tolerance response. The diagram was created using BioRender (https://biorender.com/) premium version.
Figure 2. Different cohorts of genomics for crop assessment and improvement in relation to abiotic-stress tolerance response. The diagram was created using BioRender (https://biorender.com/) premium version.
Genes 14 01281 g002
Figure 3. Involvement of different phytohormones, metabolites and other bioactive chemical components for abiotic stress response in plants.
Figure 3. Involvement of different phytohormones, metabolites and other bioactive chemical components for abiotic stress response in plants.
Genes 14 01281 g003
Figure 4. Phenomics platforms are represented schematically to assess agricultural productivity for abiotic stress-responsive future breeding.
Figure 4. Phenomics platforms are represented schematically to assess agricultural productivity for abiotic stress-responsive future breeding.
Genes 14 01281 g004
Figure 5. Integrated multi-omics pipeline for abiotic stress tolerance response in plants.
Figure 5. Integrated multi-omics pipeline for abiotic stress tolerance response in plants.
Genes 14 01281 g005
Table 2. Important genes involved in plants’ responses to abiotic stresses.
Table 2. Important genes involved in plants’ responses to abiotic stresses.
PlantsGenesFunctionAbiotic StressesNote/RemarksReferences
Barleyrhl1.aLimit root hairDroughtLoss of function mutation[114]
MaizeZm00001d010956Induce TFsSalinity [123]
TomatoSlERF.D6Steroidal glycoalkaloids (SGAs) biosynthesisDrought, Salinity [129]
AlfalfaHAMKSignaling pathwayHeat [135]
Bread wheatHVA1Signaling pathwayDrought, HeatHordeum vulgare aleurone 1[136]
WheatAtWRKY30TFsHeat, DroughtAntioxidant, osmolytes biosynthesis[137]
ArabidopsisAtSZF2Regulate salt-responsive genesSalinityLoss of function mutation[138]
ArabidopsisCBF1TFsCold, HeatC-Repeat Binding Factor1[139]
ArabidopsisPICKLE (PKL)CHD3-type chromatin remodelerColdTrimethylation of histone H3 lysine 27 (H3K27me3)[140]
TomatoSlAREB1TFsSalinity [141]
TomatoEthylene Response Factors (ERF)TFsCold, Heat, Salinity, Drought, Submergence [142]
TomatoSlUVR8UV-B photoreceptorUV-B [143]
Wild tomato
(Solanum habrochaites)
Calmodulin-like (ShCML44)Calmodulin-like (CML) proteins Ca2+ sensorsCold, Drought, SalinityEnhances Antioxidants Capacity[144]
CabbageHSP70Heat Shock ProteinHeat [145]
Pearl milletPgDREB2ATranscription FactorHeat, Drought, SalinityTransgenic overexpression in tobacco[146]
RiceOsMAPK5Signaling pathwayDrought, Salinity, Cold [147]
RiceOsWRKY87Increase DNA-binding abilityDrought, Salinity [148]
RicePM H+ ATPaseProton pumpSalinityPrimary transporter[149]
RiceCellulose synthase-like protein OsCSLD4Cell wall polysaccharide synthesisSalinityABA-induced osmotic response[150]
RiceSucrose non-fermenting 1-related protein kinases (OsSAPK8)Serine/threonine (Ser/Thr) protein kinaseCold, Drought, Salinity [151]
SoybeanMAPKSignaling pathwayExcess light [152]
Ground nutNAC4TFsDrought [153]
Ground nutHSP70ChaperonHeatHeat shock protein[153]
PotatoLEACellular protection during stressDrought, Salinity, HeatLate Embryogenesis-Abundant protein[154]
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

Roychowdhury, R.; Das, S.P.; Gupta, A.; Parihar, P.; Chandrasekhar, K.; Sarker, U.; Kumar, A.; Ramrao, D.P.; Sudhakar, C. Multi-Omics Pipeline and Omics-Integration Approach to Decipher Plant’s Abiotic Stress Tolerance Responses. Genes 2023, 14, 1281. https://doi.org/10.3390/genes14061281

AMA Style

Roychowdhury R, Das SP, Gupta A, Parihar P, Chandrasekhar K, Sarker U, Kumar A, Ramrao DP, Sudhakar C. Multi-Omics Pipeline and Omics-Integration Approach to Decipher Plant’s Abiotic Stress Tolerance Responses. Genes. 2023; 14(6):1281. https://doi.org/10.3390/genes14061281

Chicago/Turabian Style

Roychowdhury, Rajib, Soumya Prakash Das, Amber Gupta, Parul Parihar, Kottakota Chandrasekhar, Umakanta Sarker, Ajay Kumar, Devade Pandurang Ramrao, and Chinta Sudhakar. 2023. "Multi-Omics Pipeline and Omics-Integration Approach to Decipher Plant’s Abiotic Stress Tolerance Responses" Genes 14, no. 6: 1281. https://doi.org/10.3390/genes14061281

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