Next Article in Journal
The Potential Role of Cellular Senescence in Non-Alcoholic Fatty Liver Disease
Next Article in Special Issue
Multiple Abiotic Stresses Applied Simultaneously Elicit Distinct Responses in Two Contrasting Rice Cultivars
Previous Article in Journal
Heat Shock Protein 90 as Therapeutic Target for CVDs and Heart Ageing
Previous Article in Special Issue
The SlHB8 Acts as a Negative Regulator in Stem Development and Lignin Biosynthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Entailing the Next-Generation Sequencing and Metabolome for Sustainable Agriculture by Improving Plant Tolerance

1
State Key Laboratory of Subtropical Silviculture, Zhejiang A&F University, 666 Wusu Street, Lin’An, Hangzhou 311300, China
2
Colleges of Agriculture and Horticulture, South China Agricultural University, Guangzhou 510642, China
3
State Key Laboratory of Microbial Resources, Institute of Microbiology, Chinese Academy of Sciences, Beijing 100101, China
4
Maize Research Station, AARI, Faisalabad 38000, Pakistan
5
Guangdong Technology Research Center for Marine Algal Bioengineering, Guangdong Key Laboratory of Plant Epigenetics, College of Life Sciences and Oceanography, Shenzhen University, Shenzhen 518055, China
6
Department of Botany, Government College University, Faisalabad 38000, Pakistan
*
Author to whom correspondence should be addressed.
Authors have shared the first authorship.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(2), 651; https://doi.org/10.3390/ijms23020651
Submission received: 26 November 2021 / Revised: 23 December 2021 / Accepted: 29 December 2021 / Published: 7 January 2022
(This article belongs to the Collection Feature Papers in Molecular Plant Sciences)

Abstract

:
Crop production is a serious challenge to provide food for the 10 billion individuals forecasted to live across the globe in 2050. The scientists’ emphasize establishing an equilibrium among diversity and quality of crops by enhancing yield to fulfill the increasing demand for food supply sustainably. The exploitation of genetic resources using genomics and metabolomics strategies can help generate resilient plants against stressors in the future. The innovation of the next-generation sequencing (NGS) strategies laid the foundation to unveil various plants’ genetic potential and help us to understand the domestication process to unmask the genetic potential among wild-type plants to utilize for crop improvement. Nowadays, NGS is generating massive genomic resources using wild-type and domesticated plants grown under normal and harsh environments to explore the stress regulatory factors and determine the key metabolites. Improved food nutritional value is also the key to eradicating malnutrition problems around the globe, which could be attained by employing the knowledge gained through NGS and metabolomics to achieve suitability in crop yield. Advanced technologies can further enhance our understanding in defining the strategy to obtain a specific phenotype of a crop. Integration among bioinformatic tools and molecular techniques, such as marker-assisted, QTLs mapping, creation of reference genome, de novo genome assembly, pan- and/or super-pan-genomes, etc., will boost breeding programs. The current article provides sequential progress in NGS technologies, a broad application of NGS, enhancement of genetic manipulation resources, and understanding the crop response to stress by producing plant metabolites. The NGS and metabolomics utilization in generating stress-tolerant plants/crops without deteriorating a natural ecosystem is considered a sustainable way to improve agriculture production. This highlighted knowledge also provides useful research that explores the suitable resources for agriculture sustainability.

1. Introduction

The current global population is forecasted to cross ~9.8 billion in 2050 [1,2,3]. Several parts of the world are at risk of food insecurity [3]. After a consistent decline in crop production, the frequency of malnutrition in different world areas overturned the passage beginning in 2015 and has continued to climb. Malnutrition is predicted to increase up to 9.8% in 2030, presently soaring at ~9% worldwide, leading to a hunger crisis among ~850 million persons. Furthermore, agriculture production endures consuming a vast resource footmark, captivating ~38% of the surface area of the Earth and utilizing approximately 70% and 1.2% of fresh water and global energy resources, respectively, of the world [1,4]. Besides agriculture consumption, other challenges include the degradation of agricultural land, urbanization, increasing water shortage, and dependence on carbon-economy-based synthetic inputs [1,5]. Agriculture production should be increased more as compared to the current progress in an ecofriendly, sustainable, and safe way [6,7]. After that, the food supply can be maintained to deliver enough food worldwide and avoid food insecurity events. Different types of plants have been domesticated to use as a food source and confront the huger events across the globe. Still, environmental alterations and biotic stresses have been the off-putting reason for reaching the targeted, sustainable crop yield. For example, biotic (pests, microbes, etc.) and abiotic (temperature variations, incidents of drought, salinity, etc.) stressors adversely affect agriculture production [4,8,9,10]. Therefore, feeding such a huge population will be a serious test along with creating livelihood opportunities, limited resources, and various global challenges as aforementioned to gain sustainable agriculture production (Figure 1).
Therefore, humanity is under the threat of increasing world hunger, and the United Nations commission has set a goal, which is the Zero-Hunger Target by 2030. It is obligatory to achieve this goal by employing sustainable resources via safeguarding crop production in extreme environments while decreasing the resources indispensable to nourish a burgeoning global population. To entail a comprehensive system-centered technology that integrates innovative farming approaches, long-term sustainable agronomic practices, and value-added climate-resilient crops, genomic-based technologies offer, for this task, solid foundational tools and genetic tools insights for shaping the future agriculture [4]. The whole-genome sequence (WGS) of Arabidopsis thaliana was developed 21 years ago, and later on, rice was the first crop in 2002 with available WGS and so on [11,12,13,14,15] (Figure 2).
Access to the sequenced genome of various plants has exploited the potential genomic targets for improving the agronomic traits in crops. Genetic manipulations for desirable variations permit crop production, improving flexibility against harsh environments and pathogen stressors, and resulting in the generation of novel types of a plant [16,17,18]. Moreover, genomic is vital for advances in the crop sciences to fulfill the agriculture demands. Strategies related to genome sequencing have been improved to offer knowledge for crop enhancements during the last century [16]. Now, WGS data of the complex/polyploidy crops can be generated using NGS strategies such as long-read single-molecule sequencing strategy. For example, the wheat genome (hexaploid) was generated through NGS [19,20,21,22]. It is the fruit of the advancement in technologies, setting the stage to obtain elaborative information (info) by performing the genome-based interpretation of epigenomic data, consisting of the 3-D validation of the nucleus genome, the huge metabolome, transcriptome, and proteome [23,24,25]. Robot-based technologies also help in gaining agriculture sustainability. For example, geosatellite imaging can forecast heatwave, drought, etc., events, and high-throughput phenology technologies and the involvement of drone technology have been used for planning a better strategy to improve crop production. Recent developments in the computational approaches to obtain detailed results about an individual or a big dataset by involving artificial intelligence are further strengthening our understanding of sustainable crop production [26,27,28].
For now, CRISPR/Cas9 technology is a valuable editing system to generate modifications in genes and manipulate new genomes with precision to explore unknown mechanisms and is also aimed at the de novo domestication of an important crop to produce a high yield and short breeding cycle, etc. [29,30]. Progress in the genomic techniques provides new dimensions and inspires prospects for crop improvement by implying the genomic resources in the upcoming years. Thus, the assimilation among various streams of congruent and/or inconsistent data is key to cultivating an innovative approach into crop science toward the practical application in agriculture. Furthermore, these technologies have a broad spectrum application in plant biology and other fields of life sciences, such as the biomedical field, and will possibly affect the future of agriculture. Therefore, formulating the recent advances related to the metabolome and sequencing approaches for sustainability in production is crucial. It is also required to obtain a better understanding of a crop and to define a strategy to solve the global issues associated with current and past agriculture.

2. Progress in Sequencing

DNA sequences store a bulk quantity of genetic information of life. The strategies that decode this genetic information can make a paradigm revolution in the multiscientific disciplines. Frederick Sanger in 1977 elaborated on the DNA sequencing technology (DNA-seq). DNA-seq was also called Sanger’s sequencing, which was established on the basis of a chain termination system [31]. More improvements were introduced by Maxam and Gilbert [32] using chemical amendments of DNA and following cleavage at a particular nucleotide base(s).

2.1. First-Generation Sequencing

Sanger’s sequencing displayed high productiveness and less radioactivity. Later, Sanger’s sequencing was termed first-generation sequencing (FGS). FGS was utilized at the commercial level in many fields [33]. Sanger’s method has been used to generate small and large sets of FGS data of organisms, such as bacteria and human, respectively [34,35,36]. Previously, sequencing was challenging and required radioactive reaction reagents. Other constraints were limited output data using a single reaction, laborious work required to sequence diploid/haploid DNA by performing subcloning. Subcloning produces a specific template of DNA for sequencing. Despite many efforts to improve FGS, the sequencing strategy had touched its ceiling due to time consumption and cost [37,38]. For example, in the past, 10 million US dollars was invested to create an additional genome of human [39]. The above-described limitations eventuated the exploration of new approaches for sequencing.
In the past, the sequencing field had relied on the induction of the first automatic machine in 1987 by Applied Biosystems. This sequencing machine named AB370 contains capillary electrophoresis (CE). CE enhanced the speed of sequencing and accuracy of the AB370. The reported detection of nucleotides at one time and a day were 96 bases and 500 kilo bases, respectively, by AB370. Overall, AB370 can generate a read-length up to 600 bases. An upgraded model (AB373xl) can produce 900 base read-lengths by detecting 2.88 megabases per day since 1995. At that time, powerful sequencing machines were developed to save time and the cost of consumption [38]. FGS had been utilized to determine the expressed-sequence tags (ESTs), and genomic region related to the single-nucleotide-polymorphism (SNP) exploitation, and simple-sequence repeat (SSR) markers. Molecular markers related to the agronomic traits had been demonstrated among different plant species [40,41,42]. Further, the integration of automatic-sequencing machine, Sanger’s sequencing method, and linked data analyzing software had laid the foundation for improvements in sequencing strategies [37,38]. Since then, sequencing technology has revolutionized consistently from the cottage business to the big production enterprise, which demands a sophisticated and dedicated research setup comprising robots with upgrading artificial intelligence and a strong integration with bioinformatics, database setup, up-to-date chemicals, and instruments [43,44].

2.2. Next-Generation Sequencing

In 2007, DNA sequencing covered a marvelous milestone to achieve a great step forward for understanding the genomic composition of an organism after the invention of new sequencing strategies. These strategies are designated as next-generation sequencing (NGS) using the high-throughput sequencing methods [43]. In this way, researchers can obtain billions of sequenced DNA nucleotides simultaneously under millions of self-directed chemical reactions by decoding a specific target with high quality and more detailed coverage of short/long sequenced reads of plant species with time-saving and low expense costs [43]. NGS has also been designated as deep sequencing, high-throughput sequencing, or massively parallel sequencing [45,46].
Thus, NGS might require one or two devices or machine deployments to obtain the sequence data during an experiment. It is flexible due to the lack of demand for the precloned DNA region and highly competitive for conducting the genomic data interpretation in contrast to the microarray strategy that depends on tailored arrays of a subject [47]. NGS forums can create genome sequences using the libraries that were constructed by fragmented as well as adapter-attached RNA/DNA/amplicon. It is also better than the conventional vector-constructed approach of cloning and resulting in avoiding or minimizing impurities that appeared due to cloned-DNA sequences under genome sequencing projects [48,49]. Particularly, sequencing strategies follow an ordinary workflow irrespective of a sequencing research forum: such as, (i) library construction using the nucleic acid, (ii) running sequencing machine and aggregate sequenced data, and (iii) finally making data interpretation by bioinformatics or software. Library construction during NGS is a vital step to define sequencing technologies on the basis of the chemical composition of (i) synthesis reaction system, (ii) single-molecule long read, and (iii) ligation chemistry [22,50]. Herein, we briefly described the short-read sequencing and long-read sequencing, which are also known as the second-generation sequencing (SGS) and third-generation sequencing (TGS), respectively.

2.2.1. Second-Generation Sequencing

SGS-short-read forums depend on the construction type of the nucleic acid libraries generated by integrating the DNA strings with the help of adaptors and/or linkers under a ligation reaction. Hence, these DNA regions are not inserted or cloned into a vector or host cells before obtaining decoding sequence data [51,52]. Especially in plant species, the commonly utilized SGS-short-read-associated NGS technologies/models are; (i) Roche 454 (pyrosequencing), (ii) Illumina (Solexa) such as HiSeq and MiSeq methods of sequencing, (iii) oligonucleotide-ligation and detection (SOLiD) sequencing, (iv) BGI Retrovolocity strategy for DNA-nanoball sequencing, and Ion-torrent sequencing. All these formal NGS forums have advantages and disadvantages [53,54,55,56]. So far, several plant species have been sequenced using SGS research forums. For example, the genetic information relating to the several model plants such as Arabidopsis thaliana, rice (Oryza sativa), maize (Zea mays), and papaya (Carica papaya) was generated using NGS [40,50]. The genome data can also be accessed by online websites, e.g., https://plabipd.de/index.ep, http://planttfdb.gao-lab.org/, https://phytozome-next.jgi.doe.gov/, http://www.bamboogdb.org/#/, https://www.ncbi.nlm.nih.gov/, etc., which are the edible, medicinal, ornamental, and so on.
The trademark of NGS is more turnout, with countless reactions as compared to the Sanger’s sequencing (FGS), as well as the clonal sequencing. Sample multiplexing in SGS forums can remarkably decrease the cost per sample. NGS also has the potential to overcome the problem of sequencing the haploid fragments, which was a serious problem during Sanger’s sequencing. Until the SGS-short-read has a wide range of applications and dominates the present sequencing market. Many bioinformatic tools are programmed according to the SGS-short-read data analysis and considered more accurate as compared to the TGS-long-read sequencing. Major constraints of the SGS-short-read are; (i) long running times, (ii) generation of de novo assembly is difficult, (iii) structural variations, (iv) the determination of a true isoform of a transcript which is also difficult, (v) haplotype phasing, and (vi) being unable to sequence long fragments of DNA.

2.2.2. Third-Generation Sequencing

TGS-long-read forums have the potential to generate 5 kb (kilobases) to 30 kb read lengths. Previously the longest read-length using the TGS forum is 2 Gb (gigabase pairs) [21]. Thus, the TGS-long-read technology can sequence the single molecule to produce a considerable overlapping read-length for sequence assembly by avoiding the amplification bias [57]. Scientists encountered a persistent problem in dealing with polyploidy genomes of crop plants due to the extensive DNA sequence repetition, a huge genome size to create an assembly of a long chromosome through short-DNA regions. The resulting sequenced DNA data are unable to be mapped according to their genomic or/and chromosome positions [58,59,60]. The above-described reasons created more curiosity among researchers and paved the way for the invention of TGS technologies of NGS. After that, it was common practice to obtain the sequence of a single molecule via TGS-long-read technology that can create the long-DNA sequences or reads and/or scaffolds to encompass the whole chromosome or even genome of an organism. TGS-long-read forum linked methods are (i) DNA dilution constructed, (ii) optical mapping, and (iii) chromosomal-conformation arrest technologies. TGS-long-read research forums are (i) the single-molecule real-time (SMART) sequenced data generated by Pacific-Bioscience, (ii) nanopore sequenced data via Oxford-Nanopore platform (such as MinION and PromethION), (iii) Helico-sequenced data by a genetic analysis system (GAS), and (iv) electron microscopy to generate TGS-long-read data [61,62,63,64]. Currently, TGS-long-read technologies are rapidly taking the place of the SGS-short-read technologies due to the more efficiency of the sequenced data and very low cost of consumption in contrast to the past DNA sequencing technologies. The detailed description of each method, model, and cost of consumption per sample have been reviewed elsewhere [21,65,66,67,68].

2.2.3. Challenges and Limitations of SGS and TGS Forums

The plant itself may cause hurdles for producing a continuous good quality assembly of the genome due to the intrinsic factor(s) (IF). IF can be high heterozygosity, whole-genome duplication (WGD), and polyploidy episodes in organisms under changing climate [69]. The projects relating to the SGS and TGS were executed, numerous polymorphic molecular markers consisting of SNPs were determined, while swiftly creating de-novo genome maps, genotyping-by-sequencing, and transcriptomes, which help to assess the genetic diversity and investigate the traits in plants that can be domesticated [15,49,62,70]. Several isoforms of the transcripts may result from alternative splicing that leads to compositional and functional modifications in protein [71,72]. For example, SGS-short-reads forums possess inherent read-length restrictions that can cause positional genetic information loss and may also undervalue the diversity of an isoform of the transcript [68,73]. Transcripts-isoforms, which possess a different starting transcriptional site and respective RNA processing forms, may require extensive bioinformatic work to precisely process SGS-short-reads into the complete whole-length transcripts [74]. The SGS-short-read technology may be unable to generate the whole-transcriptome annotation due to the WGD event that produces highly similar isoforms of transcripts [74]. While the quantification and identification of transcripts can be possible using SGS-short-read (RNA sequencing) by transcriptome mapping, the novel isoforms cannot be discovered because of the spanning fragments of a transcript [75]. Another challenge of the SGS-short-read RNA sequencing during workflow is the RNA conversion to cDNA that can possibly introduce several library constructions linked biases, i.e., (i) reverse transcription and (ii) amplification and sequence target bias (GC contents) [76]. All the above limitations can be managed using the TGS-long-read technology that can create high-quality whole-length transcripts using a single-RNA molecule by lower coverage-depth. The whole-transcriptome annotation may require full-length sequences, which can be generated by TGS-long-read technology [74,77]. Extrinsic factors (EF) that can affect TGS-long-read are (i) poor quality sample preparations, (ii) sequencing technology (read-length, sequencing coverage, and depth), and (iii) assembly avenues. It also has the potential to resolve the problem of isoform determinations from the long reads or the whole transcripts.

3. NGS and Its Promising Aspects

Sequencing technologies have always been the foundation of genomics, and during the last 20 years, whole genome or draft of several plant species has been characterized after sequencing or refining by resequencing through various NGS forums/technologies (Figure 3).

Significance of the NGS

Progress in the NGS technologies has delivered a wide range of research forums with enriched genomic information of the plant species by decoding more complex genomes, e.g., maize, barley, pea, cotton (allotetraploid), wheat (hexaploid), and sugarcane (octoploid), etc., producing long-reads rather than short-reads with limited decoding facts [78]. This progress has also saved time and the cost of expenses and upgraded genome assemblies, leading to normally managing WGS (whole-genome-wide sequencing) tasks using NGS technologies [79].
Currently, it is feasible to generate the high-quality reference genome sequences (RGS) of a plant [65,79]. Another unique advantage of sequencing is to explore the biological niche, local abundance of a species, orphan, or minor crops, which are crucial for national or international ecosystems and participate in the food system. Moreover, the relative of the major crops such as past and current genetic diversity, including wild types, have a special status and genetic information (Table 1). Now, plant scientists can access this genetic information with more detail to find the possibilities or solutions of the current ongoing problems among domesticated plants [66,67]. This acquired knowledge could be more useful in developing climate-resilient crops in the future.
Additionally, NGS has the potential to generate massive data to dissect the novel genes or fragments. Expression profiles among various parts or organs of a plant are determined using NGS to facilitate more specific improvement in plants. For example, the identified genomic maps or regions can be used in developing the potential marker under marker-assisted breeding. A determined expression pattern also helps in uncovering the molecular regulatory processes in a plant under certain stress or normal environments [80]. Thus, NGS provides a research forum along with improving analyzing tools/software or methods [81,82,83] to understand the evolutionary aspects and functions of plants that are taking place under normal or stressed environments by conducting a detailed characterization of a desirable gene or fragment to reveal the complex regulatory mechanism.

4. NGS Can Promote Sustainable Crop Production

NGS technologies have now been applied at a large scale in searching for the genetic resources that can pave the path and promise to eradicate food hunger across the world by helping produce or improve crop yield to gain self-sufficiency in food resources. These technologies are creating the genome recourses, improving specificity and efficiency in predicting and designing targeted ESTs, SSR, and SNP markers, genome editing or gene engineering strategies in plants to attain sustainability in production by accurately determining the cause of an appeared trait and/or phenotype under harsh environments [70,84].

4.1. Exploiting the Molecular Markers, Genetic Maps, and Phylogenetic Relationships Using NGS Technologies

The genome of an organism retains altering nucleic acid sequences, which could control the phenome or a specific necessary feature as behaving molecular markers, e.g., SNPs are naturally existing several nucleotide or point mutations in the genome of a plant species and resulting to enhance genetic manipulation possibilities to fine-tune a desirable character [38]. Before the creation of molecular markers, it was very difficult to construct a library, perform many cloning works, and finally to sequence [85]. Since then, ESTs have been utilized to discover the SNPs and/or SSRs (microsatellite markers) in the genome, but technologies related to the ESTs also need more funds to generate sequenced data with low genetic coverage [86]. Molecular markers are important in exploring variations among various genomes and determining the quality trait loci (QTLs) in plants to reveal plant traits [87]. For example, QTLs responsible for the grain weight and numbers, yield, sugar accretion, flowering induction or timing, contents of the proline, and other stress-related proteins to increase plant resistance against harsh environments are the key agronomic aspects for sustainable crop yield [88].
Previously, many types of research were carried out to create high-density genetic linkage maps to determine the important QTLs associated with agronomic characters, pinpointing, and isolating the key candidate genes, as well as map-associated gene engineering. Breeders faced several problems as genetic maps with very little information related to the QTLs/molecular markers and low-density by laborious molecular work and higher cost of consumption [89,90,91]. Specifying a molecular marker and respective candidate genes for improving the agronomic traits was not easy in crops before the application of SGS technologies. That said, the accumulating data related to the molecular markers (e.g., in rice QTLs related to the tiller number, etc.) are useful for drawing genetic maps in plant species [91]. SGS technologies have been utilized to generate transcriptome data in various crops [92,93]. Now, researchers can define a particular fragment and/or molecular markers (SNPs and SSRs) by selecting the region(s) of a candidate gene or genome to improve agronomic features of the crop [94]. NGS technologies have transformed the identifications and established the genetic maps by interpreting data associated with the SNPs, SSRs, and QTLs by conducting genome-wide integrated investigations in plant species [95].
Importantly, NGS technologies provide a delightful research forum for scientists to reveal many markers by improving genetic maps with more information about agronomic traits. In this way, these makers participate in genomic selection (GS) for stress-related traits in crops. The genetic tools that have been employed by several plant breeders remain vital for finding the solution of the causes that are drastically influencing plant health, e.g., biotic and abiotic stressors, climate change, etc. [20,96,97,98,99]. Furthermore, these genetic tools can also assist scientists to boost plant yield by determining the desirable traits in crops for improvement. Now, researchers have access to the unique treasure of genome information that can display an important method of genetic manipulation to enhance crop production, after the invention of NGS. For example, genome-assisted breeding (GAB) is forecasted to permit precise and efficient plant breeding to create superior quality cultivars, promoting crop sustainability [80]. Most recently, NGS technologies are fast, less expensive, and have the capacity to sequence many samples within a short time [55]. These salient features promote genotyping technologies such as GAB, marker-assisted selection (MAS; to associate a marker with an appeared trait in a plant), and breeding-assisted genomics (BAG) in an ultrahigh-throughput manner by NGS to determine several molecular markers [17,24,28,86,87]. For example, transcriptome data were generated using NGS technology in lentils and recruited 376 out of 50,960 SNPs, which represent potential targets to control plant traits, e.g., resistance against ascochyta blight, flowering time and color, pigmentation of the stem, seed coat and size, etc. [100,101]. These study outcomes were utilized to construct high-density molecular maps and other markers such as SSRs and ISSR. By analyzing NGS-based transcriptome data, the average space and total coverage distance were improved among two molecular markers, such as 1.11 cM [101,102].
NGS constructed markers permitted a comprehensive phylogenetic genetic analysis of intraspecies and/or interspecies and estimated the divergence time of plants following the wild types and cultivated landraces among various geographic backgrounds. For example, the speciation divergence gaps among L. culinaris and M. truncatula, as well as L. ervoides and L. culinaris, were 38 million years ago (MYA) and 0.0677 MYA, respectively [103,104]. Genotyping based on the NGS data revealed geographic distribution and gene pool associated with a specific trait that helps understand the postdomestication pattern and assortment of the improved current morphology of the plant cultivars or species [70,105]. Such outcomes of the past studies laid the foundation for determining the diverse and suitable plant types to perform hybridization during breeding projects to improve genetic resources [106]. Precise divergence analysis can be improved more by generating genome data through NGS technologies in the future. The development in markers (SNPs/SSRs) using NGS could allow the tremendous advancement in designing a plant fingerprinting and forensic science, GS, evolutionary studies, phylogenetic networks, gene flow, genetic maps, etc. [107,108,109,110,111,112,113,114,115,116]. Furthermore, the generation of WGS information of a plant species can lead to more development associated the maker breeding among no-reference available plants.
The noticeable limits to crop breeding evolvement are very slow genetic progress using the crossing, multifaceted characters, and ignored minor or clash crops, which were affected due to the lack of reference genome and/or genetic information until the commencement of NGS technologies. Developing genome data or databases and integrating with the developing analytic tools participate in technology advancement to augment the understanding of genetic resources to respond against several environments by adjusting multi-agronomic traits. Nowadays, formulating quick and precise genotyping strategies to link genome information with phenomes is the pivotal aspect of desirable genetic improvements with the normal affordable expenditure of high-throughput accurate phenotyping. Hence, NGS technologies fuel up the phenotyping strategies with more accuracy to explore genetic or heritable variations in plants under varying environments and decrease the cost of traits’ determinations. More efficiency in NGS technologies is achieved through artificial-intelligence-based robotic ways, standardizing protocols for screening, and launching phenotyping centers for biotic (hotspots for the disease spreading pathogens, insects, pests, etc.) and abiotic (heat waves, drought, land erosion, salinity, and nutrient uptake efficiency) stressors [49,105,116,117,118,119], which are really important factors in formulating precise phenome system to dissect the genomics of quantifiable traits. In the future, portable devices integrated with advanced technologies (ATech) can promote sustainable crop production.

4.2. Creating the Pan- or Super-Pan-Genome Based on NGS Technology

Precise genetic manipulation requires more effects to analyze the genomic variations of a population to choose a suitable novel gene to improve agronomic traits and help dissect evolutionary relationships between the species. Researchers cannot only rely on the reference genome of a plant to explore the genetic variations and determine the most suitable population for cultivation. Additionally, sequence coverage or quality of the assembled and annotated reference genomes can also allow the comprehensive understanding of a genome. Generally, the pan-genome describes the genome of a plant/species acquired using the comparative analysis of the huge resequenced genomes, usually the genus [15,18,19,29,44,60,115,116,120]. Such genes can be categorized among various fundamental and rudimentary genes. The fundamental genes are conserved crosswise between various plants and designated as the housekeeping that participate in the key cellular functions, which can be ubiquitin (UBQ), β-actin (ACT), α-tubulin (TUA), ribosomal RNA (subunits such as 18S or 26S) elongation factors (EF), and glyceraldehyde-3-phosphate-dehydrogenase (GAPDH), etc., in different plants. These housekeeping genes can serve as an internal control during validating the expression profile of the targeted genes in the plants under an imposed or naturally occurring stress [121,122,123,124,125]. The rudimentary genes determined using the pan-genome can be conserved within the genome of a particular species or just in some members, however not in the whole genus. By contrast, intercontrol acting essential genes are conserved and integrated with several characters that can be adjustable under stressors in plants, e.g., tolerance to stressors, the activity of the antioxidants and receptors, signal transduction, gene regulation, etc. [126,127,128]. These genes actively shape the diversity of a species and aid in the fast evolution among plants to cope with stressors [129]. Now, targeted resequencing of particular tissues of plants such as mitochondria, plastids (whole plastomes), or even a few regions permits one to establish more precise phylogenetic relationships at the species or subspecies level to reveal genetic diversity and understand the domestication in crops and find ways to boost yield [130,131].
In some plants, high homozygosity and heterozygosity among the candidate genes creates more genetic variations by influencing molecular markers. Therefore, the idea of pan-genome relies on catching variations within or among genomes of different or same species by knowing the genomic structure of the genes [38,116,128]. The NGS has made possible the sequencing or resequencing of many accessions that indicate a plant or species to display structural modifications or variations, varying copy numbers, and including many alterations such as interchromosomal and/or intrachromosomal re-arrangements, transversion, and inversions [43,124,130]. Progress in pan-genome analysis can show the conserved or nonconserved regions between many accessions of a planting species, whereas genomic variations of the accessions belonging to a specific species can be examined by analyzing the super pan-genome [132]. Importantly, super-pan-genome analysis of wild-type plants can present the novel treasure of information relating to the genomic structural variation that can be deployed to mend the agronomic traits in crops by exploring the whole genome of a genus [92]. Both pan- and super-pan-genomes can allow scientists to enhance crop production by creating climate-tolerant plants.

4.3. Sustainability by Exploiting the World Genetic Resources

Globally, genetic resources or gene banks are important assets with broad nutritional prospects and safeguarding food security by preserving the vital information that can be utilized to tackle an outbreak of the disease in plants and/or determining the keys genetic factors, which can permit us to create harsh-environment-resilient crops. These genetic resources consist of primitive landraces and plant species, new, extinct, and vanishing plant types, germplasm/lines under breeding programs, wild species of crops and weeds, etc. [16,129], and the articulated evidence about germplasm provision for research purposes by gene stock and other national or international organizations across the world is very limited. Thus, the world’s huge population is deprived of the benefits of the latest innovations relating to the genomic fields due to slow development in developing countries and lack of advanced research bodies.
Although agriculture future relies on crop production within the limited land resources to produce sufficient food to feed the ever-increasing population, it requires more uniformity among genetic materials to cultivate the large-scale area with the same type of crop or species, which can be planted to tackle the increasing events of crop damage by stressors (biotic and abiotic) [8,123]. Even preserved genetic variations can enlighten us to find the cause of an outbreak of disease by exploring the genomic data of the wild or parents’ plants. Many vulnerable, endangered plant species worldwide can show new ways to improve the agriculture future by benefiting from the recent advances in DNA sequencing technologies, especially NGS. Consequently, NGS can also allow scientists to reclassify the bulk of misleading accessions or duplicated by improving identification methods that permit breeders to enhance gene bank management and overcome the common challenges to differentiate the mislabeled genotypes [60,133,134,135,136,137]. The precise assessment of genetic diversity through NGS can lead the researchers to mine the desirable genetic pool, defining the grouping of genetic material, also designating the main or minor target germplasm for research, which could offer the novel cultivars with high resistance to environmental alteration and assist in boosting the sustainability in crop yield.
The generation of sequencing data from the accessions (gene banks) using NGS technologies and the improvements in the bioinformatic tools is the key to determining or choosing the suitable or climate-resilient plants to combat the devastating events forecasted by several researchers [62,83,133]. This sequenced data strengthen more understanding in designing the molecular markers and enhance the accuracy in shaping the allelic variations or detailed genotyping–phenotyping that can improve trait-specific breeding programs. However, gene pools are the treasure that needs to be exploited by NGS around the world, which would develop a clear understanding of inter- or intra-species evolution and reveal the association between wild types and current cultivars that can expand germplasm knowledge to attain sustainable production.

4.4. Mining the Novel Genes and Regularity Pathways Using NGS to Generate Transcriptome

Rapid progress in the genomic field has laid the foundation in mining the desirable candidate genes that can help in developing climate-resilient plants, and the utilization of the NGS technologies can lead researchers to adjust or modify the key agronomic traits in crops [62,109,114,135]. The latest developments in technologies have also been employed to obtain sequenced data at a particular stage or time frame of a crop. The acquired data have been submitted or annotated with the determined regularity function to display the information relating to the candidate genes or metabolic pathways integrated with the phenome of a plant. Now, it is relatively easy to find the up-/down-regulated genes or pathways under stress by analyzing the transcriptome generated using the NGS. Studies have revealed several influenced genes or metabolites under abiotic stress at the most critical stage of a crop, e.g., plant reproduction, grain filling, assimilate storage, etc. [138]. Multiple genes or proteins have also been recognized by performing transcriptome analysis such as glycerol-3-phosphate acyltransferase-2 (GPAT2), O-acyltransferase, phosphatidylinositol, or/and phosphatidylcholine transfer protein (SFH13), phosphatidylcholine-diacylglycerol-choline-phosphotransferase (PDCP), plasmodesmata-callose-binding-protein 3 (PDCB), etc. [93]. Under stress, the genes mentioned above were activated, and a few of them encode pyruvate-di-phosphate that participates in shikimate pathways, and this pathway is responsible for producing secondary metabolites in plants under stress [139].
However, such studies reveal more detail about plants’ response to external stimuli by deploying the transcripts to influence the molecular functions and biological and cellular processes that improve the plant tolerance against harsh environments [140]. Synthesized transcriptome data also allow the finding of variations or novel genes among the cultivated cultivars or reference genome of a plant to know altering genetic factors and metabolite under stress. Transcriptome generated by NGS has been employed to unveil the role of genetic material (susceptible and resistant) to resist disease by proving the candidate genes, e.g., NPR1, NPR2, calcium-transporting ATPase (CT-ATPase), glutamate-receptor 3.2, etc., and many genes related to the phytohormones (jasmonic acid, abscisic acid, and gibberellic acid) that participate in augmenting the plant immunity against pathogen attack [141,142]. Thus, such investigations can be utilized for the genetic manipulation of the desirable crops or plants to create climate-resilient plants that could withstand adverse growth conditions to meet future demands of more food.

5. Role of Metabolomics in the Sustainable Crop Production

Metabolome refers to the extensive study of plant-secondary metabolites that regulate various cellular functions in the alive system. This field designates the wide-range set of plant metabolites produced through metabolic pathways in the plant [143,144,145]. Metabolome studies have been largely implicated in the genomic field, which allows us to exploit various phases of the ongoing biomolecular and physiological alteration triggered by climate-change or genetic unrest [146]. Metabolic modifications are directly suggested as the postgenomic changes among plants and facilitate determining the plant phenome or phenotype. Thus, the metabolites are becoming the emerging and the most reliable tools in the plants to unravel stress tolerance or resistance mechanism [147]. For example, respiratory amino acids such as glycine and serine, branched-chain amino-acids (BCAAs), and the few intermediates of the tricarboxylic acid cycle, are found to be assembled in various plants such as barley (Hordeum vulgare), (Oryza sativa), and Arabidopsis thaliana in response to harsh conditions [148,149,150]. Similarly, alterations among levels or numbers of proteins or metabolites such as proline, glycine-butane, tryptophan, phenolic, organic acids, sulpher-responsible metabolites including methionine, cysteine, and glutathione as well as phytohormones (Table 2) have also been affected in plants under stress [151,152,153,154].
Generation of knowledge about specific metabolites related to the critical stage or varying time points to examine the stress response can permit precise genetic manipulation by targeting the particular transcripts in a plant. However, NGS technologies have appeared to be promising breeding tools to describe the regulatory mechanisms and/or cellular reactions against the environmental stimuli that can be the biotic and abiotic stress [108,153]. Moreover, the association between the NGS and metabolomics has enhanced the forecasting abilities of the researchers to find the preliminary metabolic co-networks using the sequenced data of an organism. In this way, the fabricated information generated using NGS technologies and quantification (Figure 3) of the targeted metabolites helps create the most suitable strategy with more precision to increase crop yield. These technologies can speed up the genetic manipulation programs or projects to create climate-resilient smart crops that can be a source of nutritious food as well as the key to eradicating hunger across the world by safeguarding food security [143]. There is no doubt that phenotypic or phenome-directed genetic manipulation has been witnessed to improve the performance of a crop by performing metabolic engineering in the same manner as the genomic fields have demonstrated the substantial participation in accomplishing more advances in genomic research gains [144,145,182].

6. Diagnosis and Monitoring of the Disease-Causing Pathogens Using NGS and Metabolites

Multiple pathogens are casual disease agents including the vast diversity of bacteria, protozoa, mollicutes, fungi, virods, and viruses [183,184]. Researchers have employed several molecular techniques to diagnose and monitor the aforementioned pathogens to design control strategies and improve crop production. Previous techniques were not efficient enough to determine the minute residues of an organism, especially viruses and evolving pathogens. Then, the advent of genomic science assisted scientists to reveal the genetic composition of crops and the disease-causing pathogens and provided the key clues to monitor plant and pathogen interaction for crop improvements by employing the NGS and metabolomics to determine the host genetic factors that the disease-causing agent engineers during attack or infection and the ability to gain defense through reprogramming the genetic makeup [185]. For example, the sugarcane mosaic virus (SCMV) is a major threat to the maize-cultivating farming community in China’s maize-cultivating farming community; it was revealed by analyzing transcriptomic data that SCMV actively participates in downregulating the photosynthesis-responsible genes, leading to the phenotype of chlorotic lesions [186].
Recent NGS-based deep-sequencing technologies generated reliable sequenced genomes to carry out genomic analysis using the developed bioinformatic resources to improve disease control approaches by accurately diagnosing plant pathogens [112,155,156,157]. These ATech can be employed to produce metagenomics and estimate the infecting or developing microbial population in a crop [187]. Moreover, exploring the small RNA (sRNA) families such as interfering RNAs (siRNAs) might be utilized to perform identification as well as reconstruction of multiple virus genome (DNA or/and RNA) and belonging microvariants using the latest bioinformatics approaches. NGS technologies are also applicable to find harboring pathogens by insect vectors, crop certification, and quarantine programs by diagnosis techniques. Nowadays, it is possible to detect plant metabolites (Table 2), e.g., flavonoids, cyanogenic glycosides, benzoxazinoids, saponins, terpenes, and terpenoids, which are produced under a pathogen attack in various crops such as rice, maize, rye, barley, oat, millet, sorghum, etc. [146,151,153,154]. Therefore, a combined strategy based on the genomic and metabolomics analysis can stabilize the declining plant yield across the world.

7. Integration between the Transcriptome and Metabolome to Achieve Crop Sustainability

Integrated networks have been established by analyzing the transcriptome and metabolome data to pave the path for a more accurate engineering of the metabolites through genomics in plants [188,189]. This aspect enlightens us to reveal the importance of metabolites and transcripts in plants to acquire stress tolerance and provide more evidence to develop a comprehensive strategy to increase crop yield. Both metabolome and transcriptome also have the potential to show the key metabolites and/or cellular processes that can affect the plant architecture and biomass production and participate in plant adjustments by regulating the physiological state of an organism [146]. Furthermore, recent advances in technologies have displayed novel metabolic networks and have pinpointed the key regulatory genes by dissecting the genetic of transgene lines or mutants [139,190,191].
In addition, these techniques display the role of a gene to influence the metabolic pathways and unmask the underlying sophisticated mechanisms and coordination that are established among various pathways, which are hardly obtainable through conventional techniques such as microarray [192]. Successfully, these techniques have been employed to conduct metabolic engineering in various crops, e.g., anthocyanin production altered in tomato, and the nutritional reserved of rice’s endosperm has also improved by increasing the accumulation of β-carotene [193]. The availability of advanced genetic manipulation technology such as CRISPR/Cas9 may improve the metabolome and transcriptome related studies in plants. CRISPR/Cas9 has been used to create the germplasm with superior quality and traits in plants [194,195,196,197]. Moreover, development in bioinformatics fields can participate in a better assembly of the WGS of a plant. In this way, the strategies to reveal the genome-wide genetic alterations and the genotyping approaches will be improved in a cost-effective manner. Then, researchers will be able to precisely build integration among various technologies to bring a revolution in plant breeding and/or engineering. These efforts may facilitate attaining sustainability objectives without degrading our environment to feed the expanding population of the world.

8. Understanding the Bamboos’ Tolerance Using NGS and Metabolome

Climate variations are badly affecting food security in all continents. According to a prediction, more than 50% of agriculture losses occurred due to stressors that are considered a big future challenge to fulfill the demands of the ever-increasing population by producing enough resources [198,199]. Forests have significant potential to aid the mitigation of anthropogenic activities of climate disturbance and offer several co-benefits by building a healthy society [200]. Unfortunately, prevalent climate-persuaded forest die-off has also been estimated among forests worldwide. It forms an unsafe carbon-cycle feedback, discharging a huge quantity of stored carbon from the forest ecosystem to the atmosphere, as well as declining the volume of carbon sink for future forests. Plant mortality occasions have been witnessed across the world during the last decades due to climate change [118,201]. The direct impacts of climate on plants, such as drought, salt, high-temperature incidences, including other affecting agents such as wildfire and pathogen/insect epidemics, are vulnerable to climate alterations and have a major influence on forests [117,202,203]. Therefore, forests of bamboos have a significant role in sustainable agriculture production. Especially, Phyllostachys edulis belongs to a novel forest resource with an extraordinarily fast growth, monopodial-rhizome network, and immense significance in the ecosystem, restoring degraded lands, being a source of income for ~2.5 billion people, a raw material for industrial use, and a fighting weapon against climate alterations in several American, African, and Asian countries [204,205]. Internationally, bamboo trade accounts for at least $2.5 billion per year, and a consistent increase in trade has been observed [204]. Forests of moso bamboo cover approximately 73.8% of the total bamboo forest area in China and have a great cultural impact [206]. Bamboo is utilized in decorating recreational places, manufacturing furniture, musical instruments, houses, and food dishes, with the shoots being especially very popular as nutrient-enriched tasty food (bamboo shoots) in Southern China [207,208].
Recently, high-quality genome data of moso bamboo (Phyllostachys edulis) were reported; they can be the key to identifying the various TFs/genes having more similarity with the reported Arabidopsis’s genes. These identified targets in bamboo may perform similar or different functions upon functional characterization in homologous or heterologous systems. More duplication events are also found in bamboo’s genome to display more copies of a gene than Arabidopsis. For example, NAC TFs related to the fiber development in rice, Arabidopsis, and moso bamboo are different in numbers such as 6, 8, and 16, respectively. Similarly, the homologous genes of OsNAP and OsNAC10 are six and three in moso bamboo [209,210,211]. It is also noted that moso bamboo presented more duplication events during genome evolution than rice; many genes display a specific function such as floral organ development-related genes in bamboo, and their respective high homology genes in rice, lignin, jasmonic acid, and stress-responsive genes [212] have been reported in bamboo (Figure 4). Bamboo growth requires indispensable energy resources and the integrity of the cell wall to perform normal physiochemical processes for maintaining bamboo’s fast growth (Figure 4) [211]. Furthermore, abiotic stresses create a serious impact on bamboo growth and development [213]. Access to the moso bamboo genome [210,211] provides a chance to many researchers for genome-wide classifications of TFs such as aquaporin, AAAP, UBP, IQD, HD-Zip, Hexokinase, Aux/IAA and ARF, NAC, PeUGE, HSF, and CONSTANS-like in moso bamboo [188,214,215,216,217,218,219,220,221,222,223].
Genome-wide classifications of TFs’ families in moso bamboo have been carried out and have demonstrated a limited molecular characterization in model plants by exogenous gene transfer in rice and Arabidopsis and the expression profile against stresses [224,225,226]. The expression of PeLAC was high in stem and its promoter sequence contains ABRE cisregulatory element that responds to ABA and GA treatments by upregulating and downregulating the transcript of PeLAC, respectively [227]. Stress tolerance in bamboo (Phyllostachys edulis) was investigated by characterizing the tonoplast-intrinsic proteins (TIPs). Results suggested that PeTIPs may improve abiotic stress in bamboo [228]. PheWRKY86 coordinates with NCED1 by binding the W-box within the promoter region and improve stress tolerance in transgenic rice and Arabidopsis [229]. Osmotic adjustment participates in plant stress tolerance and development mechanism; e.g., Ca2+ translocation in and out from the vacuole, cell wall, and intercellular compartments regulate the development of phloem ganglion in P. edulis during the active early growth phase. Later, at the maturation stage of phloem, more accumulation of vacuolar Ca2+ was observed as compared to mature cells of cytoplast. Results suggested the important role of Ca2+ in generating cells and the osmatic actions of phloem ganglion in P. edulis [230]. PeNAC-1 has been suggested to regulate Na+ across the cellular membrane and may affect the Na+/K+ homeostasis [206] in the heterologous systems as several other phytohormones and ion transporters perform the function in plants under normal and stressed condition plants (Figure 5) [231,232]. Many TFs/genes and transporters have been identified in bamboo, e.g., genes responsive to multiple hormones (cytokinins, gibberellic acid, jasmonic acid, abscisic acid, ethylene, and auxin), and auxin biosynthesis-related transporters (PhIAA, PhPIN, PhPILS, PhAFB, PhLAX, etc.) have been identified in bamboo [233].
The accumulation of metabolites also helps plants to encounter stress conditions. An increase in anthocyanin contents was examined in Ma bamboo in the overexpressing leaf-color-related gene, which improved stress resilience against cold and drought conditions by enhancing stress-related antioxidant activity [225]. Very limited research is available on the heterologous system investigations of genes retrieved using transcriptome data of bamboo [228,234,235]. For example, the high expression of PeUGE was validated in shoots [228]. Later on, PeUGE investigations in overexpressing Arabidopsis plants suggested it regulates cell wall biosynthesis and improves stress tolerance. A significant expression alteration of Phehdz1 was induced by drought, salinity, and ABA treatments, but the high expression was in roots. Overexpressing transgene rice (OE-Phehdz1) displayed improved DS tolerance and altered secondary metabolism [225]. The lignification of tissues in bamboo is a unique character that contains monolignol glucosides (i.e., syringin, coniferin, p-glucocoumaryl alcohol, and guaiacyl) that have been examined as the key storage components of lignin precursors, which are transported to the outer cells [236,237]. ATP-binding cassette transporters have been suggested to regulate translocation of lignifying agents using vascular cation transporters by establishing a gradient across and within cellular membrane and cytoplasm [237] that needs to be exploited in detail.
Furthermore, investigations related to cutin and suberin biosynthesis in bamboo can promote a clear understanding of the stress tolerance mechanism in bamboo [238,239,240,241,242,243]. Despite advances in technology, the enhancement of agricultural resources requires consistent improvements in genome editing and metabolome technologies. During the last decade, several scientists used the CRISPR/Cas9 genome editing tool to obtain desirable traits in plant species [244]. However, very limited research revealed successful genome editing by CRISPR/Cas9 after selecting a desirable gene in ma bamboo [245]. Reports have been demonstrated that in ma bamboo, plantlets can be regenerated by callus induction in germinating embryos [246]. This literature review supports further development in gaining a better understanding through NGS and metabolome knowledge to manipulate bamboo’s genome to meet the future demands of sustainable forest resources across the globe.

9. Concluding Remarks and Promising Future Perspectives

Currently, crop production is unable to feed the growing world population due to deteriorating natural resources or mismanagement of the available genetic resources rather than other agriculture challenges. Under a perfect scenario, the pure wild-type progenitor individuals are necessary for the marking out crop domestication information and the clear identification of crop or a plant desirable sweep fragments or genes among the genome. Therefore, the improvements of genetic makeup are heralded as the prominent aspect to boost crop yield, generate climate-resilient plants/crops, and enhance nutritional value. The application of advanced bioinformatics and genetic tools grasp the considerable assurance for more agriculture output, increased livelihoods, and multiple prospects for food security by exploring the potential of the cash and orphan crops.
Improved genome information of plants will allow one to carry out effective and accurate genetic engineering in crops by revealing the mechanism of the most important agronomic traits. However, NGS technologies are evolving rapidly and claim to generate long-length strings of genomic reads, within less time and lower cost per sample or unit. Now, scientists can produce whole-genome assembly, de-novo assembly metagenomics, transcriptome, methylome sequencing, etc. Hence, NGS application is very important in many fields related to agriculture to find targets for genetic manipulation, evolution studies, exploring the bionetworks, and understanding the fundamental principles of functional genomics. Improved molecular markers will speed up breeding programs. Furthermore, metabolome studies are vital for producing knowledge about physiological mechanisms and/or metabolic pathways regulated in a crop under a stress condition to allow plant adaptation against harsh environments, to build integrated networks among genomic and metabolomics and to help the researcher to accomplish enriched information to estimate phenotype and manipulate the agronomic trait by genome and/or metabolite engineering in plants. However, a broad-spectrum application of the super- and pan-genome remains to be exploited to attain more precision in determining the phenotypes in plants related to the agronomic traits.
Particularly, developing countries are lagging to obtain benefits from the advanced technologies (ATech) due to very few investments to establish the state-of-the-art sequencing genomic center and large-scale or even small-scale robotic laboratories and provide services to the farming communities for the pathogen identification to design a better control strategy. Lack of funds for establishing infrastructure, less awareness about advanced technologies, training workshops, or increasing corruption may be the causes of decline in agriculture production. Now, the government and private sectors are actively participating in transfer-technology collaborative programs to disseminate the ATech among farming communities and open Hi-Technology centers to provide services and utilize bioinformatics and biotechnological tools to enhance crop production. Universities are also training students and producing the researchers that can construct NGS libraries and generate sequence data to interpret a better understanding of stress regulation mechanism in a crop. People can visit the online-accessible portal to acquire more information related to crop diseases or pathogens using artificial-intelligence-equipped devices or smart phones due to high-speed internet availability.
Additionally, it is necessary to refine genome assemblies by resequencing reference genomes with advanced NGS technologies and determining metabolites among cultivated and wild plants. Integration among various fields such as transcriptomes, epigenomes, proteomes, phenomes, genomes, and metabolomes, also demands advanced programming language and database creation to build a network that can assist in understanding the molecular mechanism of an appeared phenotype in plants. Bulk data generation requires an automated pipeline or operating systems that can perform routine tasks such as the generation of sequence data and interpretation of the results using the developed build-in bioinformatic tools. Consequently, comprehensive knowledge in ATech will facilitate comparative genomics to explore unmined genomes of plants and improve or find new ways of sustainable crop production.

Author Contributions

All authors contributed to the article and approved the submitted version. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was supported by National Key R & D Program of China (2021YFD2200503), Natural Science Foundation of Zhejiang Province (LZ20C160002), National Natural Science Foundation of China (31971735,32001377), and State Key Laboratory of Subtropical Silviculture (ZY20180203).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Foley, J.A.; Ramankutty, N.; Brauman, K.A.; Cassidy, E.S.; Gerber, J.S.; Johnston, M.; Mueller, N.D.; O’Connell, C.; Ray, D.K.; West, P.C.; et al. Solutions for a cultivated planet. Nature 2011, 478, 337–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Godfray, H.C.J.; Beddington, J.R.; Crute, I.R.; Haddad, L.; Lawrence, D.; Muir, J.F.; Pretty, J.; Robinson, S.; Thomas, S.M.; Toulmin, C. Food Security: The Challenge of Feeding 9 Billion People. Science 2010, 327, 812–818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. FAO. The State of Food Security and Nutrition in the World 2020; FAO: Rome, Italy, 2019. [Google Scholar]
  4. Purugganan, M.D.; Jackson, S.A. Advancing crop genomics from lab to field. Nat. Genet. 2021, 53, 595–601. [Google Scholar] [CrossRef] [PubMed]
  5. Abubakar, M.S.; Attanda, M.L. The Concept of Sustainable Agriculture: Challenges and Prospects. IOP Conf. Series Mater. Sci. Eng. 2013, 53, 012001. [Google Scholar] [CrossRef] [Green Version]
  6. FAO. The State of Food and Agriculture; FAO: Rome, Italy, 2017. [Google Scholar]
  7. Kah, M.; Tufenkji, N.; White, J.C. Nano-enabled strategies to enhance crop nutrition and protection. Nat. Nanotechnol. 2019, 14, 532–540. [Google Scholar] [CrossRef]
  8. Ma, X.; Su, Z.; Ma, H. Molecular genetic analyses of abiotic stress responses during plant reproductive development. J. Exp. Bot. 2020, 71, 2870–2885. [Google Scholar] [CrossRef]
  9. Ashraf, M.F.; Peng, G.; Liu, Z.; Noman, A.; Alamri, S.; Hashem, M.; Qari, S.H.; Al Zoubi, O.M. Molecular Control and Application of Male Fertility for Two-Line Hybrid Rice Breeding. Int. J. Mol. Sci. 2020, 21, 7868. [Google Scholar] [CrossRef]
  10. Khakimov, B.; Rasmussen, M.A.; Kannangara, R.M.; Jespersen, B.M.; Munck, L.; Engelsen, S.B. From metabolome to phenotype: GC-MS metabolomics of developing mutant barley seeds reveals effects of growth, temperature and genotype. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef] [Green Version]
  11. Kaul, S.; Koo, H.L.; Jenkins, J.; Rizzo, M.; Rooney, T.; Tallon, L.J.; Feldblyum, T.; Nierman, W.; Benito, M.-I.; Lin, X.J.N. Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 2000, 408, 796–815. [Google Scholar]
  12. Yu, J.; Hu, S.; Wang, J.; Wong, G.K.S.; Li, S.; Liu, B.; Deng, Y.; Dai, L.; Zhou, Y.; Zhang, X.; et al. A Draft Sequence of the Rice Genome (Oryza sativa L. ssp. indica). Science 2002, 296, 79–92. [Google Scholar] [CrossRef]
  13. Goff, S.A.; Ricke, D.; Lan, T.-H.; Presting, G.; Wang, R.; Dunn, M.; Glazebrook, J.; Sessions, A.; Oeller, P.; Varma, H.; et al. A Draft Sequence of the Rice Genome (Oryza sativa L. ssp. japonica). Science 2002, 296, 92–100. [Google Scholar] [CrossRef] [Green Version]
  14. Hamilton, J.; Buell, C.R. Advances in plant genome sequencing. Plant J. 2012, 70, 177–190. [Google Scholar] [CrossRef] [PubMed]
  15. Li, C.; Lin, F.; An, D.; Wang, W.; Huang, R. Genome Sequencing and Assembly by Long Reads in Plants. Genes 2017, 9, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Bevan, M.W.; Uauy, C.; Wulff, B.; Zhou, J.; Krasileva, K.; Clark, M. Genomic innovation for crop improvement. Nature 2017, 543, 346–354. [Google Scholar] [CrossRef]
  17. Garcia, S.; Leitch, I.J.; Anadon-Rosell, A.; Canela, M.Á.; Gálvez, F.; Garnatje, T.; Gras, A.; Hidalgo, O.; Johnston, E.; De Xaxars, G.M.; et al. Recent updates and developments to plant genome size databases. Nucleic Acids Res. 2013, 42, D1159–D1166. [Google Scholar] [CrossRef]
  18. Ricroch, A. Global developments of genome editing in agriculture. Transgenic Res. 2019, 28, 45–52. [Google Scholar] [CrossRef]
  19. The International Wheat Genome Sequencing Consortium (IWGSC); Appels, R.; Eversole, K.; Feuillet, C.; Keller, B.; Rogers, J.; Stein, N.; Pozniak, C.J.; Choulet, F.; Distelfeld, A.; et al. Shifting the limits in wheat research and breeding using a fully annotated reference genome. Science 2018, 361, eaar7191. [Google Scholar] [CrossRef] [Green Version]
  20. Goodwin, S.; McPherson, J.D.; McCombie, W.R. Coming of age: Ten years of next-generation sequencing technologies. Nat. Rev. Genet. 2016, 17, 333–351. [Google Scholar] [CrossRef] [PubMed]
  21. A Logsdon, G.; Vollger, M.R.; E Eichler, E. Long-read human genome sequencing and its applications. Nat. Rev. Genet. 2020, 21, 597. [Google Scholar] [CrossRef]
  22. Anandhakumar, C.; Kizaki, S.; Bando, T.; Pandian, G.; Sugiyama, H. Advancing Small-Molecule-Based Chemical Biology with Next-Generation Sequencing Technologies. ChemBioChem 2014, 16, 20–38. [Google Scholar] [CrossRef] [Green Version]
  23. Chen, W.; Gao, Y.; Xie, W.; Gong, L.; Lu, K.; Wang, W.; Li, Y.; Liu, X.; Zhang, H.; Dong, H.; et al. Genome-wide association analyses provide genetic and biochemical insights into natural variation in rice metabolism. Nat. Genet. 2014, 46, 714–721. [Google Scholar] [CrossRef] [PubMed]
  24. Furbank, R.T.; Tester, M. Phenomics–technologies to relieve the phenotyping bottleneck. Trends Plant Sci. 2011, 16, 635–644. [Google Scholar] [CrossRef] [PubMed]
  25. Zaidem, M.; Groen, S.C.; Purugganan, M.D. Evolutionary and ecological functional genomics, from lab to the wild. Plant J. 2019, 97, 40–55. [Google Scholar] [CrossRef] [Green Version]
  26. Tattaris, M.; Reynolds, M.P.; Chapman, S. A Direct Comparison of Remote Sensing Approaches for High-Throughput Phenotyping in Plant Breeding. Front. Plant Sci. 2016, 7, 1131. [Google Scholar] [CrossRef]
  27. Ma, C.; Zhang, H.H.; Wang, X. Machine learning for Big Data analytics in plants. Trends Plant Sci. 2014, 19, 798–808. [Google Scholar] [CrossRef]
  28. Esposito, S.; Carputo, D.; Cardi, T.; Tripodi, P. Applications and Trends of Machine Learning in Genomics and Phenomics for Next-Generation Breeding. Plants 2019, 9, 34. [Google Scholar] [CrossRef] [Green Version]
  29. Belhaj, K.; Chaparro-Garcia, A.; Kamoun, S.; Patron, N.J.; Nekrasov, V. Editing plant genomes with CRISPR/Cas. Curr. Opin. Biotechnol. 2015, 32, 76–84. [Google Scholar] [CrossRef]
  30. Chen, K.; Wang, Y.; Zhang, R.; Zhang, H.; Gao, C. CRISPR/Cas Genome Editing and Precision Plant Breeding in Agriculture. Annu. Rev. Plant Biol. 2019, 70, 667–697. [Google Scholar] [CrossRef]
  31. Sanger, F.; Nicklen, S.; Coulson, A.R. DNA sequencing with chain-terminating inhibitors. Proc. Natl. Acad. Sci. USA 1977, 74, 5463–5467. [Google Scholar] [CrossRef] [Green Version]
  32. Maxam, A.M.; Gilbert, W. A new method for sequencing DNA. Proc. Natl. Acad. Sci. USA 1977, 74, 560–564. [Google Scholar] [CrossRef] [Green Version]
  33. Devereux, H.L. Automated DNA sequencing. Methods Mol. Med. 1999, 31, 55–61. [Google Scholar] [CrossRef]
  34. Slatko, B.E.; Kieleczawa, J.; Ju, J.; Gardner, A.F.; Hendrickson, C.L.; Ausubel, F.M. “First Generation” Automated DNA Sequencing Technology. Curr. Protoc. Mol. Biol. 2011, 96, 7.2.1–7.2.28. [Google Scholar] [CrossRef]
  35. Chorley, B.N.; Wang, X.; Campbell, M.R.; Pittman, G.S.; Noureddine, M.A.; Bell, D.A. Discovery and verification of functional single nucleotide polymorphisms in regulatory genomic regions: Current and developing technologies. Mutat. Res. Mutat. Res. 2008, 659, 147–157. [Google Scholar] [CrossRef] [Green Version]
  36. Faber, K.; Glatting, K.-H.; Mueller, P.J.; Risch, A.; Hotz-Wagenblatt, A. Genome-wide prediction of splice-modifying SNPs in human genes using a new analysis pipeline called AASsites. BMC Bioinform. 2011, 12, S2. [Google Scholar] [CrossRef] [Green Version]
  37. NHGRI. The Cost of Sequencing a Human Genome; NHGRI: Bethesda, MD, USA, 2019.
  38. Liu, L.; Li, Y.; Li, S.; Hu, N.; He, Y.; Pong, R.; Lin, D.; Lu, L.; Law, M. Comparison of Next-Generation Sequencing Systems. J. Biomed. Biotechnol. 2012, 2012, 1–11. [Google Scholar] [CrossRef]
  39. Levy, S.; Sutton, G.; Ng, P.C.; Feuk, L.; Halpern, A.L.; Walenz, B.P.; Axelrod, N.; Huang, J.; Kirkness, E.F.; Denisov, G.; et al. The Diploid Genome Sequence of an Individual Human. PLoS Biol. 2007, 5, e254. [Google Scholar] [CrossRef] [Green Version]
  40. Michael, T.P.; Jackson, S. The First 50 Plant Genomes. Plant Genome 2013, 6, 2. [Google Scholar] [CrossRef] [Green Version]
  41. Kiechle, F.L.; Zhang, X. The postgenomic era: Implications for the clinical laboratory. Arch. Pathol. Lab. Med. 2002, 126, 255. [Google Scholar] [CrossRef]
  42. Michael, M.; Savin, K.W.; Maiko, S.; Pembleton, L.W.; Cogan Noel, O.I.; Shinozuka, K.; Forster, J.W. Transcriptome sequencing of lentil based on second-generation technology permits large-scale unigene assembly and SSR marker discovery. BMC Genom. 2011, 12, 265. [Google Scholar] [CrossRef] [Green Version]
  43. Wheeler, D.A.; Srinivasan, M.; Egholm, M.; Shen, Y.; Chen, L.; McGuire, A.; He, W.; Chen, Y.-J.; Makhijani, V.; Roth, G.T.; et al. The complete genome of an individual by massively parallel DNA sequencing. Nature 2008, 452, 872–876. [Google Scholar] [CrossRef]
  44. Yuan, Y.; Bayer, P.E.; Batley, J.; Edwards, D. Current status of structural variation studies in plants. Plant Biotechnol. J. 2021, 19, 2153–2163. [Google Scholar] [CrossRef]
  45. Thigpen, K.G. International sequencing consortium. Environ. Heal. Perspect. 2004, 112, A406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Hui, P. Next Generation Sequencing: Chemistry, Technology and Applications. Chem. Diagn. 2012, 336, 1–18. [Google Scholar] [CrossRef]
  47. Singh, I.; Wendeln, C.; Clark, A.W.; Cooper, J.M.; Ravoo, B.J.; Burley, G.A. Sequence-Selective Detection of Double-Stranded DNA Sequences Using Pyrrole–Imidazole Polyamide Microarrays. J. Am. Chem. Soc. 2013, 135, 3449–3457. [Google Scholar] [CrossRef]
  48. Park, P.J. ChIP–seq: Advantages and challenges of a maturing technology. Nat. Rev. Genet. 2009, 10, 669–680. [Google Scholar] [CrossRef] [Green Version]
  49. Hurd, P.J.; Nelson, C.J. Advantages of next-generation sequencing versus the microarray in epigenetic research. Brief. Funct. Genom. Proteom. 2009, 8, 174–183. [Google Scholar] [CrossRef] [Green Version]
  50. Unamba, C.I.N.; Nag, A.; Sharma, R.K. Next Generation Sequencing Technologies: The Doorway to the Unexplored Genomics of Non-Model Plants. Front. Plant Sci. 2015, 6, 1074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Fedoruk, M.J.; Vandenberg, A.; Bett, K.E. Quantitative Trait Loci Analysis of Seed Quality Characteristics in Lentil using Single Nucleotide Polymorphism Markers. Plant Genome 2013, 6. [Google Scholar] [CrossRef] [Green Version]
  52. Mardis, E.R. A decade’s perspective on DNA sequencing technology. Nature 2011, 470, 198–203. [Google Scholar] [CrossRef]
  53. Valouev, A.; Ichikawa, J.; Tonthat, T.; Stuart, J.; Ranade, S.; Peckham, H.; Zeng, K.; Malek, J.A.; Costa, G.; McKernan, K.; et al. A high-resolution, nucleosome position map of C. elegans reveals a lack of universal sequence-dictated positioning. Genome Res. 2008, 18, 1051–1063. [Google Scholar] [CrossRef] [Green Version]
  54. Margulies, M.; Egholm, M.; Altman, W.E.; Attiya, S.; Bader, J.S.; Bemben, L.A.; Berka, J.; Braverman, M.S.; Chen, Y.-J.; Chen, Z.J.N. Genome sequencing in microfabricated high-density picolitre reactors. Nature 2005, 437, 376–380. [Google Scholar] [CrossRef] [PubMed]
  55. Shendure, J.; Porreca, G.J.; Reppas, N.B.; Lin, X.; McCutcheon, J.P.; Rosenbaum, A.M.; Wang, M.D.; Zhang, K.; Mitra, R.D.; Church, G.M. Accurate Multiplex Polony Sequencing of an Evolved Bacterial Genome. Science 2005, 309, 1728–1732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Metzker, M.L. Sequencing technologies—The next generation. Nat. Rev. Genet. 2009, 11, 31–46. [Google Scholar] [CrossRef] [Green Version]
  57. Heather, J.M.; Chain, B. The sequence of sequencers: The history of sequencing DNA. Genomics 2016, 107, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Neale, D.B.; Wegrzyn, J.L.; A Stevens, K.; Zimin, A.V.; Puiu, D.; Crepeau, M.W.; Cardeno, C.; Koriabine, M.; E Holtz-Morris, A.; Liechty, J.D.; et al. Decoding the massive genome of loblolly pine using haploid DNA and novel assembly strategies. Genome Biol. 2014, 15, R59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Nystedt, B.; Street, N.R.; Wetterbom, A.; Zuccolo, A.; Lin, Y.-C.; Scofield, D.G.; Vezzi, F.; Delhomme, N.; Giacomello, S.; Alexeyenko, A.; et al. The Norway spruce genome sequence and conifer genome evolution. Nature 2013, 497, 579–584. [Google Scholar] [CrossRef] [Green Version]
  60. Salman-Minkov, A.; Sabath, N.; Mayrose, I. Whole-genome duplication as a key factor in crop domestication. Nat. Plants 2016, 2, 16115. [Google Scholar] [CrossRef]
  61. Peterson, T.W.; Nam, J.N.; Darby, A. Next gen sequencing survey. N. Am. Equity Res. 2010. [Google Scholar]
  62. Pei, S.; Liu, T.; Ren, X.; Li, W.; Chen, C.; Xie, Z. Benchmarking variant callers in next-generation and third-generation sequencing analysis. Brief. Bioinform. 2021, 22, 148. [Google Scholar] [CrossRef]
  63. Clark, T.A.; Spittle, K.E.; Turner, S.W.; Korlach, J. Direct Detection and Sequencing of Damaged DNA Bases. Genome Integr. 2011, 2, 10. [Google Scholar] [CrossRef] [Green Version]
  64. Harris, T.D.; Buzby, P.R.; Babcock, H.; Beer, E.; Bowers, J.; Braslavsky, I.; Causey, M.; Colonell, J.; DiMeo, J.; Efcavitch, J.W.; et al. Single-Molecule DNA Sequencing of a Viral Genome. Science 2008, 320, 106–109. [Google Scholar] [CrossRef] [Green Version]
  65. De Coster, W.; Weissensteiner, M.H.; Sedlazeck, F.J. Towards population-scale long-read sequencing. Nat. Rev. Genet. 2021, 22, 572–587. [Google Scholar] [CrossRef]
  66. Marx, V. Long road to long-read assembly. Nat. Methods 2021, 18, 125–129. [Google Scholar] [CrossRef] [PubMed]
  67. Tedersoo, L.; Albertsen, M.; Anslan, S.; Callahan, B. Perspectives and Benefits of High-Throughput Long-Read Sequencing in Microbial Ecology. Appl. Environ. Microbiol. 2021, 87, 0062621. [Google Scholar] [CrossRef]
  68. Wang, B.; Tseng, E.; Regulski, M.; Clark, T.A.; Hon, T.; Jiao, Y.; Lu, Z.; Olson, A.; Stein, J.C.; Ware, D. Unveiling the complexity of the maize transcriptome by single-molecule long-read sequencing. Nat. Commun. 2016, 7, 11708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Hu, T.; Chitnis, N.; Monos, D.; Dinh, A. Next-generation sequencing technologies: An overview. Hum. Immunol. 2021, 82, 801–811. [Google Scholar] [CrossRef]
  70. Bhat, J.A.; Yu, D.J.L.S. High-throughput NGS-based genotyping and phenotyping: Role in genomics-assisted breeding for soybean improvement. Legume Sci. 2021, 3, e81. [Google Scholar] [CrossRef]
  71. Bolisetty, M.T.; Rajadinakaran, G.; Graveley, B.R. Determining exon connectivity in complex mRNAs by nanopore sequencing. Genome Biol. 2015, 16, 1–12. [Google Scholar] [CrossRef] [Green Version]
  72. Thanaraj, T.A. ASD: The Alternative Splicing Database. Nucleic Acids Res. 2004, 32, 64–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Flint, J.; Mott, R. Finding the molecular basis of quatitative traits: Successes and pitfalls. Nat. Rev. Genet. 2001, 2, 437–445. [Google Scholar] [CrossRef]
  74. Abdel-Ghany, S.E.; Hamilton, M.; Jacobi, J.L.; Ngam, P.; Devitt, N.; Schilkey, F.; Ben-Hur, A.; Reddy, A.S.N. A survey of the sorghum transcriptome using single-molecule long reads. Nat. Commun. 2016, 7, 11706. [Google Scholar] [CrossRef] [Green Version]
  75. Steijger, T.; The RGASP Consortium; Abril, J.F.; Engström, P.; Kokocinski, F.; Hubbard, T.; Guigó, R.; Harrow, J.; Bertone, P. Assessment of transcript reconstruction methods for RNA-seq. Nat. Methods 2013, 10, 1177–1184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Garalde, D.R.; Snell, E.A.; Jachimowicz, D.; Sipos, B.; Lloyd, J.H.; Bruce, M.; Pantic, N.; Admassu, T.; James, P.; Warland, A.; et al. Highly parallel direct RNA sequencing on an array of nanopores. Nat. Methods 2018, 15, 201–206. [Google Scholar] [CrossRef] [PubMed]
  77. Weirather, J.L.; de Cesare, M.; Wang, Y.; Piazza, P.; Sebastiano, V.; Wang, X.-J.; Buck, D.; Au, K.F. Comprehensive comparison of Pacific Biosciences and Oxford Nanopore Technologies and their applications to transcriptome analysis. F1000Research 2017, 6, 100. [Google Scholar] [CrossRef] [PubMed]
  78. Montenegro, J.D.; Golicz, A.A.; Bayer, P.E.; Hurgobin, B.; Lee, H.; Chan, C.-K.K.; Visendi, P.; Lai, K.; Doležel, J.; Batley, J.; et al. The pangenome of hexaploid bread wheat. Plant J. 2017, 90, 1007–1013. [Google Scholar] [CrossRef] [Green Version]
  79. Zhao, Q.; Feng, Q.; Lu, H.; Li, Y.; Wang, A.; Tian, Q.; Zhan, Q.; Lu, Y.; Zhang, L.; Huang, T.J.N.G. Pan-genome analysis highlights the extent of genomic variation in cultivated and wild rice. Nat. Genet. 2018, 50, 278–284. [Google Scholar] [CrossRef] [Green Version]
  80. Barabaschi, D.; Guerra, D.; Lacrima, K.; Laino, P.; Michelotti, V.; Urso, S.; Valè, G.; Cattivelli, L. Emerging Knowledge from Genome Sequencing of Crop Species. Mol. Biotechnol. 2011, 50, 250–266. [Google Scholar] [CrossRef]
  81. Yeh, H.-M.; Liao, M.-H.; Chu, C.-L.; Lin, Y.-H.; Sun, W.-Z.; Lai, L.-P.; Chen, P.-L. Next-generation sequencing and bioinformatics to identify genetic causes of malignant hyperthermia. J. Formos. Med Assoc. 2021, 120, 883–892. [Google Scholar] [CrossRef]
  82. Blätke, M.-A.; Szymanski, J.J.; Gladilin, E.; Scholz, U.; Beier, S. Editorial: Advances in Applied Bioinformatics in Crops. Front. Plant Sci. 2021, 12, 12. [Google Scholar] [CrossRef]
  83. Ibrahim, M.A.; Shehata, M.A.; Nasry, N.N.; Fayez, M.S.; Bishay, S.K.; Aziz, M.M.; Ratib, N.R.; Ahmed, N.K.; Ali, S.M.; Ismail, E.; et al. Bioinformatics Approaches toward Plant Breeding Programs. Asian J. Res. Rev. Agric. 2021, 3, 5–14. [Google Scholar]
  84. Chu, C.; Wang, S.; Rudd, J.C.; Ibrahim, A.M.; Xue, Q.; Devkota, R.N.; Baker, J.A.; Baker, S.; Simoneaux, B.; Opena, G. A New Strategy for Using Historical Imbalanced Yield Data to Conduct Genome-Wide Association Studies and Develop Genomic Prediction Models for Wheat Breeding. Mol. Breed. 2021. [Google Scholar]
  85. Ariyadasa, R.; Stein, N. Advances in BAC-Based Physical Mapping and Map Integration Strategies in Plants. J. Biomed. Biotechnol. 2012, 2012, 1–11. [Google Scholar] [CrossRef] [Green Version]
  86. Zalapa, J.E.; Cuevas, H.; Zhu, H.; Steffan, S.; Senalik, U.; Zeldin, E.; McCown, B.; Harbut, R.; Simon, P. Using next-generation sequencing approaches to isolate simple sequence repeat (SSR) loci in the plant sciences. Am. J. Bot. 2012, 99, 193–208. [Google Scholar] [CrossRef] [Green Version]
  87. Gao, Q.; Yue, G.; Li, W.; Wang, J.; Xu, J.; Yin, Y. Recent Progress Using High-throughput Sequencing Technologies in Plant Molecular BreedingF. J. Integr. Plant Biol. 2012, 54, 215–227. [Google Scholar] [CrossRef]
  88. Paran, I.; Zamir, D. Quantitative traits in plants: Beyond the QTL. Trends Genet. 2003, 19, 303–306. [Google Scholar] [CrossRef]
  89. Davey, J.W.; Hohenlohe, P.A.; Etter, P.D.; Boone, J.Q.; Catchen, J.M.; Blaxter, M. Genome-wide genetic marker discovery and genotyping using next-generation sequencing. Nat. Rev. Genet. 2011, 12, 499–510. [Google Scholar] [CrossRef]
  90. Salvi, S.; Tuberosa, R. To clone or not to clone plant QTLs: Present and future challenges. Trends Plant Sci. 2005, 10, 297–304. [Google Scholar] [CrossRef]
  91. Liu, G.; Zhu, H.; Zhang, G.; Li, L.; Ye, G. Dynamic analysis of QTLs on tiller number in rice (Oryza sativa L.) with single segment substitution lines. Theor. Appl. Genet. 2012, 125, 143–153. [Google Scholar] [CrossRef] [PubMed]
  92. Hirsch, C.N.; Foerster, J.M.; Johnson, J.M.; Sekhon, R.S.; Muttoni, G.; Vaillancourt, B.; Peñagaricano, F.; Lindquist, E.; Pedraza, M.A.; Barry, K.; et al. Insights into the maize pan-genome and pan-transcriptome. Plant Cell 2014, 26, 121–135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Singh, D.; Singh, C.K.; Taunk, J.; Jadon, V.; Pal, M.; Gaikwad, K. Genome wide transcriptome analysis reveals vital role of heat responsive genes in regulatory mechanisms of lentil (Lens culinaris Medikus). Sci. Rep. 2019, 9, 1–19. [Google Scholar] [CrossRef]
  94. Zhou, Z.-X.; Zhang, M.-J.; Peng, X.; Takayama, Y.; Xu, X.-Y.; Huang, L.-Z.; Du, L.-L. Mapping genomic hotspots of DNA damage by a single-strand-DNA-compatible and strand-specific ChIP-seq method. Genome Res. 2013, 23, 705–715. [Google Scholar] [CrossRef] [Green Version]
  95. Schneeberger, K.; Weigel, D. Fast-forward genetics enabled by new sequencing technologies. Trends Plant Sci. 2011, 16, 282–288. [Google Scholar] [CrossRef]
  96. Hamblin, M.T.; Buckler, E.S.; Jannink, J.-L. Population genetics of genomics-based crop improvement methods. Trends Genet. 2011, 27, 98–106. [Google Scholar] [CrossRef] [PubMed]
  97. Spindel, J.; Begum, H.; Akdemir, D.; Virk, P.; Collard, B.; Redoña, E.; Atlin, G.; Jannink, J.-L.; McCouch, S.R. Genomic selection and association mapping in rice (Oryza sativa): Effect of trait genetic architecture, training population composition, marker number and statistical model on accuracy of rice genomic selection in elite, tropical rice breeding lines. PLoS Genet. 2019, 2, 1004982. [Google Scholar] [CrossRef]
  98. Meuwissen, T.H.E.; Hayes, B.J.; Goddard, M.E. Prediction of Total Genetic Value Using Genome-Wide Dense Marker Maps. Genetics 2001, 157, 1819–1829. [Google Scholar] [CrossRef]
  99. Mulder, H. Is GxE a burden or a blessing? Opportunities for genomic selection and big data. J. Anim. Breed. Genet. 2017, 134, 435–436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Kumar, J.; Gupta, D.S. Prospects of next generation sequencing in lentil breeding. Mol. Biol. Rep. 2020, 47, 9043–9053. [Google Scholar] [CrossRef]
  101. Polanco, C.; de Miera, L.E.S.; González, A.I.; García, P.G.; Fratini, R.; Vaquero, F.; Vences, F.J.; De La Vega, M.P. Construction of a high-density interspecific (Lens culinaris x L. odemensis) genetic map based on functional markers for mapping morphological and agronomical traits, and QTLs affecting resistance to Ascochyta in lentil. PLoS ONE 2019, 14, e0214409. [Google Scholar] [CrossRef] [Green Version]
  102. Temel, H.Y.; Göl, D.; Akkale, H.B.K.; Kahriman, A.; Tanyolaç, M.B. Single nucleotide polymorphism discovery through Illumina-based transcriptome sequencing and mapping in lentil. Turk. J. Agric. For. 2015, 39, 470–488. [Google Scholar] [CrossRef]
  103. Lavin, M.; Herendeen, P.; Wojciechowski, M.F. Evolutionary Rates Analysis of Leguminosae Implicates a Rapid Diversification of Lineages during the Tertiary. Syst. Biol. 2005, 54, 575–594. [Google Scholar] [CrossRef] [Green Version]
  104. Sharpe, A.G.; Ramsay, L.; Sanderson, L.-A.; Fedoruk, M.J.; E Clarke, W.; Li, R.; Kagale, S.; Vijayan, P.; Vandenberg, A.; E Bett, K. Ancient orphan crop joins modern era: Gene-based SNP discovery and mapping in lentil. BMC Genom. 2013, 14, 192. [Google Scholar] [CrossRef] [Green Version]
  105. Brozynska, M.; Furtado, A.; Henry, R.J. Genomics of crop wild relatives: Expanding the gene pool for crop improvement. Plant Biotechnol. J. 2016, 14, 1070–1085. [Google Scholar] [CrossRef]
  106. Khazaei, H.; Caron, C.T.; Fedoruk, M.; Diapari, M.; Vandenberg, A.; Coyne, C.J.; McGee, R.; Bett, K.E. Genetic Diversity of Cultivated Lentil (Lens culinaris Medik.) and Its Relation to the World’s Agro-ecological Zones. Front. Plant Sci. 2016, 7, 1093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Horton, M.W.; Hancock, A.M.; Huang, Y.S.; Toomajian, C.; Atwell, S.; Auton, A.; Muliyati, N.W.; Platt, A.; Sperone, F.G.; Vilhjalmsson, B.J.; et al. Genome-wide patterns of genetic variation in worldwide Arabidopsis thaliana accessions from the RegMap panel. Nat. Genet. 2012, 44, 212–216. [Google Scholar] [CrossRef] [Green Version]
  108. Ferrero-Serrano, Á.; Assmann, S.M. Phenotypic and genome-wide association with the local environment of Arabidopsis. Nat. Ecol. Evol. 2019, 3, 274–285. [Google Scholar] [CrossRef]
  109. Lasky, J.R.; Marais, D.L.D.; McKay, J.K.; Richards, J.H.; Juenger, T.E.; Keitt, T.H. Data from: Characterizing genomic variation of Arabidopsis thaliana: The roles of geography and climate. Mol. Ecol. 2012, 21, 5512. [Google Scholar] [CrossRef] [PubMed]
  110. Gutaker, R.M.; Groen, S.C.; Bellis, E.; Choi, J.Y.; Pires, I.S.; Bocinsky, R.K.; Slayton, E.; Wilkins, O.; Castillo, C.C.; Negrão, S.; et al. Genomic history and ecology of the geographic spread of rice. Nat. Plants 2020, 6, 492–502. [Google Scholar] [CrossRef] [PubMed]
  111. Bilinski, P.; Albert, P.S.; Berg, J.J.; Birchler, J.A.; Grote, M.N.; Lorant, A.; Quezada, J.; Swarts, K.; Yang, J.; Ross-Ibarra, J. Parallel altitudinal clines reveal trends in adaptive evolution of genome size in Zea mays. PLoS Genet. 2018, 14, e1007162. [Google Scholar] [CrossRef]
  112. Lasky, J.R.; Upadhyaya, H.D.; Ramu, P.; Deshpande, S.; Hash, C.T.; Bonnette, J.; Juenger, T.E.; Hyma, K.; Acharya, C.; Mitchell, S.E.; et al. Genome-environment associations in sorghum landraces predict adaptive traits. Sci. Adv. 2015, 1, e1400218. [Google Scholar] [CrossRef] [Green Version]
  113. Rhoné, B.; Defrance, D.; Berthouly-Salazar, C.; Mariac, C.; Cubry, P.; Couderc, M.; Dequincey, A.; Assoumanne, A.; Kane, N.A.; Sultan, B.; et al. Pearl millet genomic vulnerability to climate change in West Africa highlights the need for regional collaboration. Nat. Commun. 2020, 11, 1–9. [Google Scholar] [CrossRef]
  114. Abrouk, M.; Ahmed, H.I.; Cubry, P.; Šimoníková, D.; Cauet, S.; Pailles, Y.; Bettgenhaeuser, J.; Gapa, L.; Scarcelli, N.; Couderc, M.; et al. Fonio millet genome unlocks African orphan crop diversity for agriculture in a changing climate. Nat. Commun. 2020, 11, 1–13. [Google Scholar] [CrossRef] [PubMed]
  115. Danilevicz, M.F.; Tay Fernandez, C.G.; Marsh, J.I.; Bayer, P.E.; Edwards, D. Plant pangenomics: Approaches, applications and advancements. Curr. Opin. Plant Biol. 2020, 54, 18–25. [Google Scholar] [CrossRef] [PubMed]
  116. Bayer, P.; Golicz, A.; Scheben, A.; Batley, J.; Edwards, D. Plant pan-genomes are the new reference. Nat. Plants 2020, 6, 914–920. [Google Scholar] [CrossRef] [PubMed]
  117. Hartmann, H.; Moura, C.F.; Anderegg, W.R.L.; Ruehr, N.K.; Salmon, Y.; Allen, C.D.; Arndt, S.K.; Breshears, D.D.; Davi, H.; Galbraith, D.; et al. Research frontiers for improving our understanding of drought-induced tree and forest mortality. New Phytol. 2018, 218, 15–28. [Google Scholar] [CrossRef] [Green Version]
  118. Choat, B.; Brodribb, T.; Brodersen, C.R.; Duursma, R.A.; López, R.; Medlyn, B. Triggers of tree mortality under drought. Nature 2018, 558, 531–539. [Google Scholar] [CrossRef]
  119. 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]
  120. Chen, F.; Song, Y.; Li, X.; Chen, J.; Mo, L.; Zhang, X.; Lin, Z.; Zhang, L. Genome sequences of horticultural plants: Past, present, and future. Hortic. Res. 2019, 6, 1–23. [Google Scholar] [CrossRef] [Green Version]
  121. Maroufi, A.; Van Bockstaele, E.; De Loose, M. Validation of reference genes for gene expression analysis in chicory (Cichorium intybus) using quantitative real-time PCR. BMC Mol. Biol. 2010, 11, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Vandesompele, J.; De Preter, K.; Pattyn, F.; Poppe, B.; Van Roy, N.; De Paepe, A.; Speleman, F. Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal control genes. Genome Biol. 2002, 3, research0034.1. [Google Scholar] [CrossRef] [Green Version]
  123. Nicot, N.; Hausman, J.-F.; Hoffmann, L.; Evers, D. Housekeeping gene selection for real-time RT-PCR normalization in potato during biotic and abiotic stress. J. Exp. Bot. 2005, 56, 2907–2914. [Google Scholar] [CrossRef] [PubMed]
  124. Guénin, S.; Mauriat, M.; Pelloux, J.; Van Wuytswinkel, O.; Bellini, C.; Gutierrez, L. Normalization of qRT-PCR data: The necessity of adopting a systematic, experimental conditions-specific, validation of references. J. Exp. Bot. 2009, 60, 487–493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Hellemans, J.; Mortier, G.; De Paepe, A.; Speleman, F.; Vandesompele, J. qBase relative quantification framework and software for management and automated analysis of real-time quantitative PCR data. Genome Biol. 2007, 8, R19. [Google Scholar] [CrossRef] [Green Version]
  126. McHale, L.K.; Haun, W.J.; Xu, W.W.; Bhaskar, P.B.; Anderson, J.; Hyten, D.; Gerhardt, D.J.; Jeddeloh, J.A.; Stupar, R.M. Structural Variants in the Soybean Genome Localize to Clusters of Biotic Stress-Response Genes. Plant Physiol. 2012, 159, 1295–1308. [Google Scholar] [CrossRef] [Green Version]
  127. Li, Y.-h.; Zhou, G.; Ma, J.; Jiang, W.; Jin, L.-g.; Zhang, Z.; Guo, Y.; Zhang, J.; Sui, Y.; Zheng, L.J.N.B. De novo assembly of soybean wild relatives for pan-genome analysis of diversity and agronomic traits. Nat. Biotechnol. 2014, 32, 1045–1052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Schatz, M.C.; Maron, L.G.; Stein, J.C.; Wences, A.H.; Gurtowski, J.; Biggers, E.; Lee, H.; Kramer, M.; Antoniou, E.; Ghiban, E.J.G.B. Whole genome de novo assemblies of three divergent strains of rice, Oryza sativa, document novel gene space of aus and indica. Genome Biol. 2014, 15, 506. [Google Scholar]
  129. Tao, Y.; Zhao, X.; Mace, E.; Henry, R.; Jordan, D. Exploring and Exploiting Pan-genomics for Crop Improvement. Mol. Plant 2019, 12, 156–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Lakušić, D.; Liber, Z.; Nikolić, T.; Surina, B.; Kovacic, S.; Bogdanović, S.; Stefanović, S. Molecular phylogeny of the Campanula pyramidalis species complex (Campanulaceae) inferred from chloroplast and nuclear non-coding sequences and its taxonomic implications. TAXON 2013, 62, 505–524. [Google Scholar] [CrossRef] [Green Version]
  131. Magdy, M.; Ou, L.; Yu, H.; Chen, R.; Zhou, Y.; Hassan, H.; Feng, B.; Taitano, N.; van der Knaap, E.; Zou, X.; et al. Pan-plastome approach empowers the assessment of genetic variation in cultivated Capsicum species. Hortic. Res. 2019, 6, 1–15. [Google Scholar] [CrossRef] [Green Version]
  132. Khan, A.W.; Garg, V.; Roorkiwal, M.; Golicz, A.A.; Edwards, D.; Varshney, R.K. Super-Pangenome by Integrating the Wild Side of a Species for Accelerated Crop Improvement. Trends Plant Sci. 2020, 25, 148–158. [Google Scholar] [CrossRef] [Green Version]
  133. Mascher, M.; Schreiber, M.; Scholz, U.; Graner, A.; Reif, J.C.; Stein, N. Genebank genomics bridges the gap between the conservation of crop diversity and plant breeding. Nat. Genet. 2019, 51, 1076–1081. [Google Scholar] [CrossRef]
  134. McCouch, S.; Navabi, Z.K.; Abberton, M.; Anglin, N.L.; Barbieri, R.L.; Baum, M.; Bett, K.; Booker, H.; Brown, G.L.; Bryan, G.J.; et al. Mobilizing Crop Biodiversity. Mol. Plant 2020, 13, 1341–1344. [Google Scholar] [CrossRef] [PubMed]
  135. Varshney, R.K.; Singh, V.K.; Kumar, A.; Powell, W.; Sorrells, M.E. Can genomics deliver climate-change ready crops? Curr. Opin. Plant Biol. 2018, 45, 205–211. [Google Scholar] [CrossRef] [PubMed]
  136. Wang, W.; Mauleon, R.; Hu, Z.; Chebotarov, D.; Tai, S.; Wu, Z.; Li, M.; Zheng, T.; Fuentes, R.R.; Zhang, F.; et al. Genomic variation in 3,010 diverse accessions of Asian cultivated rice. Nature 2018, 557, 43–49. [Google Scholar] [CrossRef]
  137. Wing, R.A.; Purugganan, M.D.; Zhang, Q. The rice genome revolution: From an ancient grain to Green Super Rice. Nat. Rev. Genet. 2018, 19, 505–517. [Google Scholar] [CrossRef] [PubMed]
  138. Whitfield, P.; Kirwan, J. Metabolomics: An Emerging Post-genomic Tool for Nutrition. Genom. Proteom. Metab. Nutraceuticals Funct. Foods 2010, 92, 271–285. [Google Scholar] [CrossRef]
  139. 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]
  140. Kumar, R.; Bohra, A.; Pandey, A.K.; Pandey, M.K.; Kumar, A. Metabolomics for Plant Improvement: Status and Prospects. Front. Plant Sci. 2017, 8, 1302. [Google Scholar] [CrossRef] [Green Version]
  141. Dodds, P.N.; Rathjen, J.P. Plant immunity: Towards an integrated view of plant–pathogen interactions. Nat. Rev. Genet. 2010, 11, 539–548. [Google Scholar] [CrossRef]
  142. Tsuda, K.; Somssich, I.E. Transcriptional networks in plant immunity. New Phytol. 2015, 206, 932–947. [Google Scholar] [CrossRef]
  143. Parry, M.A.J.; Hawkesford, M. An Integrated Approach to Crop Genetic ImprovementF. J. Integr. Plant Biol. 2012, 54, 250–259. [Google Scholar] [CrossRef]
  144. Xavier, A.; Hall, B.; Hearst, A.A.; Cherkauer, K.A.; Rainey, K.M. Genetic Architecture of Phenomic-Enabled Canopy Coverage in Glycine max. Genetics 2017, 206, 1081–1089. [Google Scholar] [CrossRef] [Green Version]
  145. Wang, C.; Hu, S.; Gardner, C.; Lübberstedt, T. Emerging Avenues for Utilization of Exotic Germplasm. Trends Plant Sci. 2017, 22, 624–637. [Google Scholar] [CrossRef] [Green Version]
  146. Turner, M.F.; Heuberger, A.L.; Kirkwood, J.S.; Collins, C.C.; Wolfrum, E.J.; Broeckling, C.D.; Prenni, J.E.; Jahn, C.E. Non-targeted Metabolomics in Diverse Sorghum Breeding Lines Indicates Primary and Secondary Metabolite Profiles Are Associated with Plant Biomass Accumulation and Photosynthesis. Front. Plant Sci. 2016, 7, 953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Pilu, R.; Panzeri, D.; Cassani, E.; Cerino Badone, F.; Landoni, M.; Nielsen, E. A paramutation phenomenon is involved in the genetics of maize low phytic acid1-241 (lpa1-241) trait. Heredity 2009, 102, 236–245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Arbona, V.; Manzi, M.; De Ollas, C.; Gómez-Cadenas, A. Metabolomics as a Tool to Investigate Abiotic Stress Tolerance in Plants. Int. J. Mol. Sci. 2013, 14, 4885–4911. [Google Scholar] [CrossRef]
  149. Li, X.; Lawas, L.M.; Malo, R.; Glaubitz, U.; Erban, A.; Mauleon, R.; Heuer, S.; Zuther, E.; Kopka, J.; Hincha, D.K.; et al. Metabolic and transcriptomic signatures of rice floral organs reveal sugar starvation as a factor in reproductive failure under heat and drought stress. Plant Cell Environ. 2015, 38, 2171–2192. [Google Scholar] [CrossRef] [PubMed]
  150. 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]
  151. Diretto, G.; Al-Babili, S.; Tavazza, R.; Scossa, F.; Papacchioli, V.; Migliore, M.; Beyer, P.; Giuliano, G. Transcriptional-metabolic networks in beta-carotene-enriched potato tubers: The long and winding road to the Golden phenotype. Plant Physiol. 2010, 154, 899–912. [Google Scholar] [CrossRef] [Green Version]
  152. Paine, J.A.; Shipton, C.A.; Chaggar, S.; Howells, R.; Kennedy, M.J.; Vernon, G.; Wright, S.Y.; Hinchliffe, E.; Adams, J.L.; Silverstone, A.L.; et al. Improving the nutritional value of Golden Rice through increased pro-vitamin A content. Nat. Biotechnol. 2005, 23, 482–487. [Google Scholar] [CrossRef]
  153. Yang, L.; Wen, K.-S.; Ruan, X.; Zhao, Y.-X.; Wei, F.; Wang, Q. Response of Plant Secondary Metabolites to Environmental Factors. Molecules 2018, 23, 762. [Google Scholar] [CrossRef] [Green Version]
  154. Ramakrishna, A.; Ravishankar, G.A. Influence of abiotic stress signals on secondary metabolites in plants. Plant Signal. Behav. 2011, 6, 1720. [Google Scholar] [PubMed]
  155. Loskutov, I.; Shelenga, T.; Konarev, A.; Shavarda, A.; Blinova, E.; Dzubenko, N.R. The metabolomic approach to the comparative analysis of wild and cultivated species of oats (Avena L.). Russ. J. Genet. Appl. Res. 2017, 7, 501–508. [Google Scholar] [CrossRef]
  156. Sánchez-Martín, J.; Heald, J.; Kingston-Smith, A.; Winters, A.; Rubiales, D.; Sanz, M.; Mur, L.A.J.; Prats, E. A metabolomic study in oats (Avena sativa) highlights a drought tolerance mechanism based upon salicylate signalling pathways and the modulation of carbon, antioxidant and photo-oxidative metabolism. Plant Cell Environ. 2015, 38, 1434–1452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Gupta, S.; Rupasinghe, T.; Callahan, D.L.; Natera, S.H.; Smith, P.; Hill, C.B.; Roessner, U.; Boughton, B.S. Spatio-temporal metabolite and elemental profiling of salt stressed barley seeds during initial stages of germination by MALDI-MSI and µ-XRF spectrometry. Front. Plant Sci. 2019, 10, 1139. [Google Scholar] [CrossRef] [Green Version]
  158. Wang, Y.; Zeng, X.; Xu, Q.; Mei, X.; Yuan, H.; Jiabu, D.; Sang, Z.; Nyima, T. Metabolite profiling in two contrasting Tibetan hulless barley cultivars revealed the core salt-responsive metabolome and key salt-tolerance biomarkers. AoB Plants 2019, 11, plz021. [Google Scholar] [CrossRef] [Green Version]
  159. Piasecka, A.; Sawikowska, A.; Krajewski, P.; Kachlicki, P. Combined mass spectrometric and chromatographic methods for in-depth analysis of phenolic secondary metabolites in barley leaves. J. Mass Spectrom. 2015, 50, 513–532. [Google Scholar] [CrossRef] [PubMed]
  160. Kogel, K.-H.; Voll, L.M.; Schäfer, P.; Jansen, C.; Wu, Y.; Langen, G.; Imani, J.; Hofmann, J.; Schmiedl, A.; Sonnewald, S.; et al. Transcriptome and metabolome profiling of field-grown transgenic barley lack induced differences but show cultivar-specific variances. Proc. Natl. Acad. Sci. USA 2010, 107, 6198–6203. [Google Scholar] [CrossRef] [Green Version]
  161. Roessner, U.; Patterson, J.H.; Forbes, M.G.; Fincher, G.B.; Langridge, P.; Bacic, A. An Investigation of Boron Toxicity in Barley Using Metabolomics. Plant Physiol. 2006, 142, 1087–1101. [Google Scholar] [CrossRef] [Green Version]
  162. Song, E.-H.; Jeong, J.; Park, C.Y.; Kim, H.-Y.; Kim, E.-H.; Bang, E.; Hong, Y.-S. Metabotyping of rice (Oryza sativa L.) for understanding its intrinsic physiology and potential eating quality. Food Res. Int. 2018, 111, 20–30. [Google Scholar] [CrossRef]
  163. Yan, S.; Huang, W.; Gao, J.; Fu, H.; Liu, J. Comparative metabolomic analysis of seed metabolites associated with seed storability in rice (Oryza sativa L.) during natural aging. Plant Physiol. Biochem. 2018, 127, 590–598. [Google Scholar] [CrossRef]
  164. Gayen, D.; Ghosh, S.; Paul, S.; Sarkar, S.N.; Datta, S.K.; Datta, K. Metabolic Regulation of Carotenoid-Enriched Golden Rice Line. Front. Plant Sci. 2016, 7, 1622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Hu, C.; Shi, J.; Quan, S.; Cui, B.; Kleessen, S.; Nikoloski, Z.; Tohge, T.; Alexander, D.; Guo, L.; Lin, H.; et al. Metabolic variation between japonica and indica rice cultivars as revealed by non-targeted metabolomics. Sci. Rep. 2014, 4, 5067. [Google Scholar] [CrossRef] [Green Version]
  166. Zarei, I.; Luna, E.; Leach, J.E.; McClung, A.; Vilchez, S.; Koita, O.; Ryan, E.P. Comparative Rice Bran Metabolomics across Diverse Cultivars and Functional Rice Gene–Bran Metabolite Relationships. Metabolites 2018, 8, 63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Dhawi, F.; Datta, R.; Ramakrishna, W. Metabolomics, biomass and lignocellulosic total sugars analysis in foxtail millet (Setaria italica) inoculated with different combinations of plant growth promoting bacteria and mycorrhiza. Commun. Plant Sci. 2018, 8, 8. [Google Scholar] [CrossRef]
  168. Mareya, C.R.; Tugizimana, F.; Piater, L.A.; Madala, N.E.; Steenkamp, P.A.; Dubery, I.A. Untargeted Metabolomics Reveal Defensome-Related Metabolic Reprogramming in Sorghum bicolor against Infection by Burkholderia andropogonis. Metabolites 2019, 9, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Tugizimana, F.; Djami-Tchatchou, A.T.; Steenkamp, P.A.; Piater, L.A.; Dubery, I.A. Metabolomic Analysis of Defense-Related Reprogramming in Sorghum bicolor in Response to Colletotrichum sublineolum Infection Reveals a Functional Metabolic Web of Phenylpropanoid and Flavonoid Pathways. Front. Plant Sci. 2019, 9, 1840. [Google Scholar] [CrossRef] [Green Version]
  170. Ogbaga, C.C.; Stępień, P.; Dyson, B.C.; Rattray, N.J.W.; Ellis, D.I.; Goodacre, R.; Johnson, G.N. Biochemical Analyses of Sorghum Varieties Reveal Differential Responses to Drought. PLoS ONE 2016, 11, e0154423. [Google Scholar] [CrossRef] [PubMed]
  171. Michaletti, A.; Naghavi, M.R.; Toorchi, M.; Zolla, L.; Rinalducci, S. Metabolomics and proteomics reveal drought-stress responses of leaf tissues from spring-wheat. Sci. Rep. 2018, 8, 1–18. [Google Scholar] [CrossRef] [Green Version]
  172. Thomason, K.; Babar, A.; Erickson, J.E.; Mulvaney, M.; Beecher, C.; Macdonald, G. Comparative physiological and metabolomics analysis of wheat (Triticum aestivum L.) following post-anthesis heat stress. PLoS ONE 2018, 13, e0197919. [Google Scholar] [CrossRef] [PubMed]
  173. Shi, T.; Zhu, A.; Jia, J.; Hu, X.; Chen, J.; Liu, W.; Ren, X.; Sun, D.; Fernie, A.R.; Cui, F.; et al. Metabolomics analysis and metabolite-agronomic trait associations using kernels of wheat (Triticum aestivum) recombinant inbred lines. Plant J. 2020, 103, 279–292. [Google Scholar] [CrossRef] [Green Version]
  174. Shewry, P.R.; Corol, D.I.; Jones, H.; Beale, M.H.; Ward, J.L. Defining genetic and chemical diversity in wheat grain by 1H-NMR spectroscopy of polar metabolites. Mol. Nutr. Food Res. 2017, 61, 1600807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Matthews, S.B.; Santra, M.; Mensack, M.M.; Wolfe, P.; Byrne, P.F.; Thompson, H.J. Metabolite Profiling of a Diverse Collection of Wheat Lines Using Ultraperformance Liquid Chromatography Coupled with Time-of-Flight Mass Spectrometry. PLoS ONE 2012, 7, e44179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Chen, M.; Rao, R.S.P.; Zhang, Y.; Zhong, C.; Thelen, J.J. Metabolite variation in hybrid corn grain from a large-scale multisite study. Crop. J. 2016, 4, 177–187. [Google Scholar] [CrossRef] [Green Version]
  177. Rao, J.; Cheng, F.; Hu, C.; Quan, S.; Lin, H.; Wang, J.; Chen, G.; Zhao, X.; Alexander, D.; Guo, L.; et al. Metabolic map of mature maize kernels. Metabolomics 2014, 10, 775–787. [Google Scholar] [CrossRef]
  178. Barros, E.; Lezar, S.; Anttonen, M.J.; Van Dijk, J.P.; Röhlig, R.M.; Kok, E.; Engel, K.-H. Comparison of two GM maize varieties with a near-isogenic non-GM variety using transcriptomics, proteomics and metabolomics. Plant Biotechnol. J. 2010, 8, 436–451. [Google Scholar] [CrossRef]
  179. Lamari, N.; Zhendre, V.; Urrutia, M.; Bernillon, S.; Maucourt, M.; Deborde, C.; Prodhomme, D.; Jacob, D.; Ballias, P.; Rolin, D.; et al. Metabotyping of 30 maize hybrids under early-sowing conditions reveals potential marker-metabolites for breeding. Metabolomics 2018, 14, 1–15. [Google Scholar] [CrossRef] [Green Version]
  180. Xu, G.; Cao, J.; Wang, X.; Chen, Q.; Jin, W.; Li, Z.; Tian, F. Evolutionary Metabolomics Identifies Substantial Metabolic Divergence between Maize and Its Wild Ancestor, Teosinte. Plant Cell 2019, 31, 1990–2009. [Google Scholar] [CrossRef] [Green Version]
  181. Jin, M.; Zhang, X.; Zhao, M.; Deng, M.; Du, Y.; Zhou, Y.; Wang, S.; Tohge, T.; Fernie, A.R.; Willmitzer, L.; et al. Integrated genomics-based mapping reveals the genetics underlying maize flavonoid biosynthesis. BMC Plant Biol. 2017, 17, 17. [Google Scholar] [CrossRef] [Green Version]
  182. Langridge, P.; Fleury, D. Making the most of ’omics’ for crop breeding. Trends Biotechnol. 2011, 29, 33. [Google Scholar] [CrossRef]
  183. Jayawardena, R.S.; Hyde, K.D.; de Farias, A.R.G.; Bhunjun, C.S.; Ferdinandez, H.S.; Manamgoda, D.S.; Udayanga, D.; Herath, I.S.; Thambugala, K.M.; Manawasinghe, I.S.; et al. What is a species in fungal plant pathogens? Fungal Divers. 2021, 109, 239–266. [Google Scholar] [CrossRef]
  184. Gougherty, A.V.; Davies, T.J. Towards a phylogenetic ecology of plant pests and pathogens. Philos Trans. R. Soc. Lond B Biol. Sci. 2021, 376, 20200359. [Google Scholar] [CrossRef] [PubMed]
  185. Xie, J.; Jiang, T.; Li, Z.; Li, X.; Fan, Z.; Zhou, T. Sugarcane mosaic virus remodels multiple intracellular organelles to form genomic RNA replication sites. Arch. Virol. 2021, 166, 1921–1930. [Google Scholar] [CrossRef] [PubMed]
  186. Gao, X.; Chen, Y.; Luo, X.; Du, Z.; Hao, K.; An, M.; Xia, Z.; Wu, Y. Recombinase Polymerase Amplification Assay for Simultaneous Detection of Maize Chlorotic Mottle Virus and Sugarcane Mosaic Virus in Maize. ACS Omega 2021, 6, 18008–18013. [Google Scholar] [CrossRef] [PubMed]
  187. Taş, N.; de Jong, A.E.; Li, Y.; Trubl, G.; Xue, Y.; Dove, N.C. Metagenomic tools in microbial ecology research. Curr. Opin. Biotechnol. 2021, 67, 184–191. [Google Scholar] [CrossRef] [PubMed]
  188. Wuolikainen, A.; Jonsson, P.; Ahnlund, M.; Antti, H.; Marklund, S.L.; Moritz, T.; Forsgren, L.; Andersen, P.M.; Trupp, M. Multi-platform mass spectrometry analysis of the CSF and plasma metabolomes of rigorously matched amyotrophic lateral sclerosis, Parkinson’s disease and control subjects. Mol. BioSyst. 2016, 12, 1287–1298. [Google Scholar] [CrossRef]
  189. van Dam, N.M.; Bouwmeester, H.J. Metabolomics in the Rhizosphere: Tapping into Belowground Chemical Communication. Trends Plant Sci. 2016, 21, 256. [Google Scholar] [CrossRef]
  190. Fernie, A.R. Review: Metabolome characterisation in plant system analysis. Funct. Plant Biol. 2003, 30, 111–120. [Google Scholar] [CrossRef]
  191. Sakakibara, K.Y.; Saito, K. Review: Genetically modified plants for the promotion of human health. Biotechnol. Lett. 2006, 28, 1983–1991. [Google Scholar] [CrossRef]
  192. Kusano, M.; Saito, K. Role of Metabolomics in Crop Improvement. J. Plant Biochem. Biotechnol. 2012, 21, 24–31. [Google Scholar] [CrossRef]
  193. Butelli, E.; Titta, L.; Giorgio, M.; Mock, H.-P.; Matros, A.; Peterek, S.; Schijlen, E.G.W.M.; Hall, R.D.; Bovy, A.G.; Luo, J.; et al. Enrichment of tomato fruit with health-promoting anthocyanins by expression of select transcription factors. Nat. Biotechnol. 2008, 26, 1301–1308. [Google Scholar] [CrossRef]
  194. Oikawa, A.; Matsuda, F.; Kusano, M.; Okazaki, Y.; Saito, K. Rice Metabolomics. Rice 2008, 1, 63–71. [Google Scholar] [CrossRef] [Green Version]
  195. Fernie, A.R.; Schauer, N. Metabolomics-assisted breeding: A viable option for crop improvement? Trends Genet. 2009, 25, 39–48. [Google Scholar] [CrossRef] [PubMed]
  196. Simó, C.; Ibáez, C.; Valdes, A.; Cifuentes, A.; García-Cañas, V.; Ibáñez, C. Metabolomics of Genetically Modified Crops. Int. J. Mol. Sci. 2014, 15, 18941–18966. [Google Scholar] [CrossRef] [Green Version]
  197. Rao, M.J.; Wang, L. CRISPR/Cas9 technology for improving agronomic traits and future prospective in agriculture. Planta 2021, 254, 1–16. [Google Scholar] [CrossRef] [PubMed]
  198. Lobell, D.B.; Schlenker, W.; Costa-Roberts, J. Climate Trends and Global Crop Production Since. Science 2011, 333, 616–620. [Google Scholar] [CrossRef] [Green Version]
  199. Ortiz-Bobea, A.; Knippenberg, E.; Chambers, R.G. Growing climatic sensitivity of U.S. agriculture linked to technological change and regional specialization. Sci. Adv. 2018, 4, eaat4343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Anderegg, W.R.L.; Trugman, A.T.; Badgley, G.; Anderson, C.M.; Bartuska, A.; Ciais, P.; Cullenward, D.; Field, C.B.; Freeman, J.; Goetz, S.J.; et al. Climate-driven risks to the climate mitigation potential of forests. Science 2020, 368, 6497. [Google Scholar] [CrossRef]
  201. Allen, C.D.; Breshears, D.D.; McDowell, N.G. On underestimation of global vulnerability to tree mortality and forest die-off from hotter drought in the Anthropocene. Ecosphere 2015, 6, art129. [Google Scholar] [CrossRef]
  202. Seidl, R.; Thom, D.; Kautz, M.; Martin-Benito, D.; Peltoniemi, M.; Vacchiano, G.; Wild, J.; Ascoli, D.; Petr, M.; Honkaniemi, J.; et al. Forest disturbances under climate change. Nat. Clim. Chang. 2017, 7, 395–402. [Google Scholar] [CrossRef] [Green Version]
  203. Klein, T.; Hartmann, H. Climate change drives tree mortality. Science 2018, 362, 758. [Google Scholar] [CrossRef]
  204. Zhao, H.; Zhao, S.; Fei, B.; Liu, H.; Yang, H.; Dai, H.; Wang, D.; Jin, W.; Tang, F.; Gao, Q.; et al. Announcing the Genome Atlas of Bamboo and Rattan (GABR) project: Promoting research in evolution and in economically and ecologically beneficial plants. GigaScience 2017, 6, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Canavan, S.; Richardson, D.; Visser, V.; Le Roux, J.; Vorontsova, M.S.; Wilson, J. The global distribution of bamboos: Assessing correlates of introduction and invasion. AoB Plants 2016, 9, 078. [Google Scholar] [CrossRef] [Green Version]
  206. Hou, D.; Zhao, Z.; Hu, Q.; Li, L.; Vasupalli, N.; Zhuo, J.; Zeng, W.; Wu, A.; Lin, X. PeSNAC-1 a NAC transcription factor from moso bamboo (Phyllostachys edulis) confers tolerance to salinity and drought stress in transgenic rice. Tree Physiol. 2020, 40, 1792–1806. [Google Scholar] [CrossRef] [PubMed]
  207. Guo, Z.-H.; Ma, P.-F.; Yang, G.-Q.; Hu, J.-Y.; Liu, Y.-L.; Xia, E.-H.; Zhong, M.-C.; Zhao, L.; Sun, G.-L.; Xu, Y.-X.; et al. Genome Sequences Provide Insights into the Reticulate Origin and Unique Traits of Woody Bamboos. Mol. Plant 2019, 12, 1353–1365. [Google Scholar] [CrossRef]
  208. Cui, K.; He, C.-Y.; Zhang, J.-G.; Duan, A.-G.; Zeng, Y.-F. Temporal and Spatial Profiling of Internode Elongation-Associated Protein Expression in Rapidly Growing Culms of Bamboo. J. Proteome Res. 2012, 11, 2492–2507. [Google Scholar] [CrossRef]
  209. Zhao, Z.; Hou, D.; Hu, Q.; Wei, H.; Zheng, Y.; Lin, X.J.J.A.B. Cloning and expression analysis of PeNAC047 gene from Phyllostachys edulis. Int. J. Agric. Biotechnol. 2020, 28, 58–71. [Google Scholar]
  210. Zhao, H.; Gao, Z.; Wang, L.; Wang, J.; Wang, S.; Fei, B.; Chen, C.; Shi, C.; Liu, X.; Zhang, H.; et al. Chromosome-level reference genome and alternative splicing atlas of moso bamboo (Phyllostachys edulis). GigaScience 2018, 7, 115. [Google Scholar] [CrossRef] [PubMed]
  211. Peng, Z.; Lu, Y.; Li, L.; Zhao, Q.; Feng, Q.; Gao, Z.; Lu, H.; Hu, T.; Yao, N.; Liu, K.; et al. The draft genome of the fast-growing non-timber forest species moso bamboo (Phyllostachys heterocycla). Nat. Genet. 2013, 45, 456–461. [Google Scholar] [CrossRef] [Green Version]
  212. Hou, D.; Lu, H.; Zhao, Z.; Pei, J.; Yang, H.; Wu, A.; Yu, X.; Lin, X. Integrative transcriptomic and metabolomic data provide insights into gene networks associated with lignification in postharvest Lei bamboo shoots under low temperature. Food Chem. 2021, 368, 130822. [Google Scholar] [CrossRef]
  213. Huang, R.; Gao, H.; Liu, J.; Li, X. WRKY transcription factors in moso bamboo that are responsive to abiotic stresses. J. Plant Biochem. Biotechnol. 2021, 1–8. [Google Scholar] [CrossRef]
  214. Liu, J.; Cheng, Z.; Li, X.; Xie, L.; Bai, Y.; Peng, L.; Li, J.; Gao, J. Expression Analysis and Regulation Network Identification of the CONSTANS-Like Gene Family in Moso Bamboo (Phyllostachys edulis) Under Photoperiod Treatments. DNA Cell Biol. 2019, 38, 607–626. [Google Scholar] [CrossRef]
  215. Li, X.; Chang, Y.; Ma, S.; Shen, J.; Hu, H.; Xiong, L. Genome-Wide Identification of SNAC1-Targeted Genes Involved in Drought Response in Rice. Front. Plant Sci. 2019, 10, 982. [Google Scholar] [CrossRef] [Green Version]
  216. Huang, B.; Huang, Z.; Ma, R.; Chen, J.; Zhang, Z.; Yrjälä, K. Genome-wide identification and analysis of the heat shock transcription factor family in moso bamboo (Phyllostachys edulis). Sci. Rep. 2021, 11, 1–19. [Google Scholar] [CrossRef] [PubMed]
  217. Zheng, W.; Zhang, Y.; Zhang, Q.; Wu, R.; Wang, X.; Feng, S.; Chen, S.; Lu, C.; Du, L. Genome-Wide Identification and Characterization of Hexokinase Genes in Moso Bamboo (Phyllostachys edulis). Front. Plant Sci. 2020, 11, 600. [Google Scholar] [CrossRef] [PubMed]
  218. Shan, X.; Yang, K.; Xu, X.; Zhu, C.; Gao, Z. Genome-Wide Investigation of the NAC Gene Family and Its Potential Association with the Secondary Cell Wall in Moso Bamboo. Biomolecules 2019, 9, 609. [Google Scholar] [CrossRef] [Green Version]
  219. Chen, D.; Chen, Z.; Wu, M.; Wang, Y.; Wang, Y.; Yan, H.; Xiang, Y. Genome-Wide Identification and Expression Analysis of the HD-Zip Gene Family in Moso Bamboo (Phyllostachys edulis). J. Plant Growth Regul. 2017, 36, 323–337. [Google Scholar] [CrossRef]
  220. Sun, H.; Li, L.; Lou, Y.; Zhao, H.; Gao, Z. Genome-wide identification and characterization of aquaporin gene family in moso bamboo (Phyllostachys edulis). Mol. Biol. Rep. 2016, 43, 437–450. [Google Scholar] [CrossRef]
  221. Wu, R.; Shi, Y.; Zhang, Q.; Zheng, W.; Chen, S.; Du, L.; Lu, C. Genome-Wide Identification and Characterization of the UBP Gene Family in Moso Bamboo (Phyllostachys edulis). Int. J. Mol. Sci. 2019, 20, 4309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Gao, Y.; Liu, H.; Wu, L.; Xiong, R.; Shi, Y.; Xiang, Y. Systematic Identification and Analysis of NAC Gene Family in Moso Bamboo (Phyllostachys edulis); Research Square: Durham, NC, USA, 2020. [Google Scholar]
  223. Liu, H.; Wu, M.; Zhu, D.; Pan, F.; Wang, Y.; Wang, Y.; Xiang, Y. Genome-Wide analysis of the AAAP gene family in moso bamboo (Phyllostachys edulis). BMC Plant Biol. 2017, 17, 1–18. [Google Scholar] [CrossRef] [Green Version]
  224. Ma, R.; Huang, B.; Chen, J.; Huang, Z.; Yu, P.; Ruan, S.; Zhang, Z. Genome-wide identification and expression analysis of dirigent-jacalin genes from plant chimeric lectins in Moso bamboo (Phyllostachys edulis). PLoS ONE 2021, 16, e0248318. [Google Scholar] [CrossRef]
  225. Gao, Y.; Liu, H.; Zhang, K.; Li, F.; Wu, M.; Xiang, Y. A moso bamboo transcription factor, Phehdz1, positively regulates the drought stress response of transgenic rice. Plant Cell Rep. 2021, 40, 187–204. [Google Scholar] [CrossRef]
  226. Lan, Y.; Wu, L.; Wu, M.; Liu, H.; Gao, Y.; Zhang, K.; Xiang, Y. Transcriptome analysis reveals key genes regulating signaling and metabolic pathways during the growth of moso bamboo (Phyllostachys edulis) shoots. Physiol. Plant. 2021, 172, 91–105. [Google Scholar] [CrossRef]
  227. Li, L.; Sun, H.; Lou, Y.; Yang, Y.; Zhao, H.; Gao, Z. Cloning and expression analysis of PeLAC in Phyllostachys edulis. Plant Sci. J. 2017, 35, 252–259. [Google Scholar]
  228. Sun, H.; Lou, Y.; Li, L.; Zhao, H.; Gao, Z. Tissue expression pattern analysis of TIPs genes in Phyllostachys edulis. For. Res. 2016, 29, 521–528. [Google Scholar]
  229. Wu, M.; Zhang, K.; Xu, Y.; Wang, L.; Liu, H.; Qin, Z.; Xiang, Y. The moso bamboo WRKY transcription factor, PheWRKY86, regulates drought tolerance in transgenic plants. Plant Physiol. Biochem. 2021, 170, 180. [Google Scholar] [CrossRef] [PubMed]
  230. Yu, F.; Ding, Y. Ultracytochemical localization of Ca2+ during the phloem ganglion development in Phyllostachys edulis. Front. Biol. China 2006, 1, 219–224. [Google Scholar] [CrossRef]
  231. Cushman, K.R.; Pabuayon, I.C.M.; Hinze, L.L.; Sweeney, M.E.; Reyes, B.G.D.L. Networks of Physiological Adjustments and Defenses, and Their Synergy With Sodium (Na+) Homeostasis Explain the Hidden Variation for Salinity Tolerance Across the Cultivated Gossypium hirsutum Germplasm. Front. Plant Sci. 2020, 11, 588854. [Google Scholar] [CrossRef]
  232. Afzal, M.Z.; Jia, Q.; Ibrahim, A.K.; Niyitanga, S.; Zhang, L. Mechanisms and Signaling Pathways of Salt Tolerance in Crops: Understanding from the Transgenic Plants. Trop. Plant Biol. 2020, 13, 297–320. [Google Scholar] [CrossRef]
  233. Wang, W.; Gu, L.; Ye, S.; Zhang, H.; Cai, C.; Xiang, M.; Gao, Y.; Wang, Q.; Lin, C.; Zhu, Q. Genome-wide analysis and transcriptomic profiling of the auxin biosynthesis, transport and signaling family genes in moso bamboo (Phyllostachys heterocycla). BMC Genom. 2017, 18, 870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Wang, S.; Sun, H.Y.; Li, L.C.; Yang, Y.H.; Xu, H.; Zhao, H.; Gao, Z. Cloning and expression pattern analysis of PeDWF4 Gene in Moso bamboo (Phyllostachys edulis). For. Res. 2018, 31, 50–56. [Google Scholar] [CrossRef]
  235. Zhang, Z.; Yang, X.; Cheng, L.; Guo, Z.; Wang, H.; Wu, W.; Shin, K.; Zhu, J.; Zheng, X.; Bian, J.; et al. Physiological and transcriptomic analyses of brassinosteroid function in moso bamboo (Phyllostachys edulis) seedlings. Planta 2020, 252, 1–13. [Google Scholar] [CrossRef] [PubMed]
  236. Sarkanen, K. Formation, structure; reactions. Classification and distribution. J. Polym. Sci. Part B Polym. Lett. 1971, 43–94. [Google Scholar]
  237. Shimada, N.; Munekata, N.; Tsuyama, T.; Matsushita, Y.; Fukushima, K.; Kijidani, Y.; Takabe, K.; Yazaki, K.; Kamei, I. Active Transport of Lignin Precursors into Membrane Vesicles from Lignifying Tissues of Bamboo. Plants 2021, 10, 2237. [Google Scholar] [CrossRef]
  238. Ishihara, A.; Hashimoto, Y.; Miyagawa, H.; Wakasa, K. Induction of serotonin accumulation by feeding of rice striped stem borer in rice leaves. Plant Signal. Behav. 2008, 3, 714–716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Ishihara, A.; Hashimoto, Y.; Tanaka, C.; Dubouzet, J.G.; Nakao, T.; Matsuda, F.; Nishioka, T.; Miyagawa, H.; Wakasa, K. The tryptophan pathway is involved in the defense responses of rice against pathogenic infection via serotonin production. Plant J. 2008, 54, 481–495. [Google Scholar] [CrossRef] [PubMed]
  240. Riley, R.G.; Kolattukudy, P.E. Evidence for Covalently Attached p-Coumaric Acid and Ferulic Acid in Cutins and Suberins. Plant Physiol. 1975, 56, 650–654. [Google Scholar] [CrossRef] [Green Version]
  241. Bernards, M.A.; Lopez, M.L.; Zajicek, J.; Lewis, N. Hydroxycinnamic Acid-derived Polymers Constitute the Polyaromatic Domain of Suberin. J. Biol. Chem. 1995, 270, 7382–7386. [Google Scholar] [CrossRef] [Green Version]
  242. Rozema, J.; Broekman, R.; Blokker, P.; Meijkamp, B.B.; de Bakker, N.; van de Staaij, J.; van Beem, A.; Ariese, F.; Kars, S.M. UV-B absorbance and UV-B absorbing compounds (para-coumaric acid) in pollen and sporopollenin: The perspective to track historic UV-B levels. J. Photochem. Photobiol. B Biol. 2001, 62, 108–117. [Google Scholar] [CrossRef]
  243. Liu, C.-J. Biosynthesis of hydroxycinnamate conjugates: Implications for sustainable biomass and biofuel production. Biofuels 2010, 1, 745–761. [Google Scholar] [CrossRef]
  244. Turnbull, C.; Lillemo, M.; Hvoslef-Eide, T.A.K. Global Regulation of Genetically Modified Crops Amid the Gene Edited Crop Boom–A Review. Front. Plant Sci. 2021, 12, 630396. [Google Scholar] [CrossRef]
  245. Ye, S.; Chen, G.; Kohnen, M.V.; Wang, W.; Cai, C.; Ding, W.; Wu, C.; Gu, L.; Zheng, Y.; Ma, X.; et al. Robust CRISPR/Cas9 mediated genome editing and its application in manipulating plant height in the first generation of hexaploid Ma bamboo (Dendrocalamus latiflorus Munro). Plant Biotechnol. J. 2019, 18, 1501–1503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Yuan, J.-L.; Yue, J.-J.; Wu, X.-L.; Gu, X.-P. Protocol for Callus Induction and Somatic Embryogenesis in Moso Bamboo. PLoS ONE 2013, 8, e81954. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The key challenges for sustainable agriculture production. Four major challenges for agriculture and raising serious issues across the world.
Figure 1. The key challenges for sustainable agriculture production. Four major challenges for agriculture and raising serious issues across the world.
Ijms 23 00651 g001
Figure 2. Timeline of the sequenced organisms. Sequencing technologies generated genomic information bank of various organisms and suggested an exciting research forum for further innovation. In the figure, genome size in mega-bases [Mb]-GS.
Figure 2. Timeline of the sequenced organisms. Sequencing technologies generated genomic information bank of various organisms and suggested an exciting research forum for further innovation. In the figure, genome size in mega-bases [Mb]-GS.
Ijms 23 00651 g002
Figure 3. Sequencing forums/technologies, metabolome data generating methods, and plant growth under stressors. In the figure, second-generation sequencing short-read forum—SGS-SR; third-generation sequencing long-read forum—TGS-LR; single-molecule real-time sequencing—SMART; thin-layer chromatography—TLC; gas/liquid chromatography mass spectrometry—GC/LC-MS; LC-electrochemistry-MS—LC-EC-MS; nuclear magnetic resonance—NMR; direct infusion mass spectrometry—DIMS; Fourier-transform infrared—FT-IR; capillary electrophoresis-LC-MS—CE-MS.
Figure 3. Sequencing forums/technologies, metabolome data generating methods, and plant growth under stressors. In the figure, second-generation sequencing short-read forum—SGS-SR; third-generation sequencing long-read forum—TGS-LR; single-molecule real-time sequencing—SMART; thin-layer chromatography—TLC; gas/liquid chromatography mass spectrometry—GC/LC-MS; LC-electrochemistry-MS—LC-EC-MS; nuclear magnetic resonance—NMR; direct infusion mass spectrometry—DIMS; Fourier-transform infrared—FT-IR; capillary electrophoresis-LC-MS—CE-MS.
Ijms 23 00651 g003
Figure 4. Various Growth and development phases and stress tolerance mechanism in bamboos. Different genes/TFs and non-coding RNAs participate in functional regulation during the growth and development phase I to perform the specific or multiple roles. The biochemical alteration also defines the growth transition (e.g., culm development under an altering gradient of hormones as the GA, IAA, ABA, zeatin (ZT). Genes-responsible for the lignin (PvNST1/2–1, PvC3H-2/3, PvC4H-2/4, PvCADs, PvCCR-2/4, PvHCT-2/5/8, PvPAL-2/4/6) and JA (PvOPR2, PvPEX5, PvJAZ-4) synthesis; early flowering (PeMADS2), etc. During phase II, plants regulate internal adjustments to cope with stressors such as hormone alterations (e.g., rhizome generates new shoot under an altering gradient of hormones), RNA metabolism, epigenetic modifications, and accumulation of various plant metabolites. Phase III is considered acclimatization and evolution; many evolution events take place during evolving plants to produce multiple copies of transcripts as compared to the ancestral donors to flourish new generations of plants under consistent overwhelming environments. Abbreviations in figure: regulation—reg.; jasmonic acid—JA; gibberellic acid—GA; indole-3-acetic acid—IAA; abscisic acid—ABA; zeatin—ZT; long-noncoding ribonucleic acid—lnc RNAs, small noncoding RNAs—sRNAs; short interfering RNAs—siRNAs; microRNAs—miRNAs; transfer RNAs—tRNAs; RNA-directed DNA methylation—RdDM; gibberellic acid—GA; indole-3-acetic acid—IAA; abscisic acid—ABA; zeatin—ZT; low-temperature—LT; peroxidase—POD; phenylalanine-ammonia-lyase- PAL; and 4-coumarate responsive ligase—4CL.
Figure 4. Various Growth and development phases and stress tolerance mechanism in bamboos. Different genes/TFs and non-coding RNAs participate in functional regulation during the growth and development phase I to perform the specific or multiple roles. The biochemical alteration also defines the growth transition (e.g., culm development under an altering gradient of hormones as the GA, IAA, ABA, zeatin (ZT). Genes-responsible for the lignin (PvNST1/2–1, PvC3H-2/3, PvC4H-2/4, PvCADs, PvCCR-2/4, PvHCT-2/5/8, PvPAL-2/4/6) and JA (PvOPR2, PvPEX5, PvJAZ-4) synthesis; early flowering (PeMADS2), etc. During phase II, plants regulate internal adjustments to cope with stressors such as hormone alterations (e.g., rhizome generates new shoot under an altering gradient of hormones), RNA metabolism, epigenetic modifications, and accumulation of various plant metabolites. Phase III is considered acclimatization and evolution; many evolution events take place during evolving plants to produce multiple copies of transcripts as compared to the ancestral donors to flourish new generations of plants under consistent overwhelming environments. Abbreviations in figure: regulation—reg.; jasmonic acid—JA; gibberellic acid—GA; indole-3-acetic acid—IAA; abscisic acid—ABA; zeatin—ZT; long-noncoding ribonucleic acid—lnc RNAs, small noncoding RNAs—sRNAs; short interfering RNAs—siRNAs; microRNAs—miRNAs; transfer RNAs—tRNAs; RNA-directed DNA methylation—RdDM; gibberellic acid—GA; indole-3-acetic acid—IAA; abscisic acid—ABA; zeatin—ZT; low-temperature—LT; peroxidase—POD; phenylalanine-ammonia-lyase- PAL; and 4-coumarate responsive ligase—4CL.
Ijms 23 00651 g004
Figure 5. Regulation of plant tolerance through transporters and/or transcription factors. Osmatic regulation is attained in plants by opening and closing of channels using transporters (Trptrs) related to the cations (Ca2+, Na+, and H+). Furthermore transcription factors (TFs) play a crucial role in operating plant tolerance (e.g., SnRKs interact with other TFs and/or genes by phosphorylating and activating more genetic factors to help plants build food reverse that can be utilized under the stress condition. In the figure, HKT: high affinity K+ transporters; SOS: salt overly sensitive 1; (X): various transporters/genes such as 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 12, 14, 16, 18, and 20; CNGC: cyclic-nucleotide-gated channels; MOCA1: monocation-induced Ca2+ enhancer 1; ANN: Annexin; IPUT: inositol-phosphorylceramide glucuronosyltransferase; GlyIPC: glycosyl-inositol-phosphorylceramide; CA: cation exchanger; NH: Na+/H+ exchangers; cNMP: cyclic-nucleotide monophosphate; and CaM: calmodulin.
Figure 5. Regulation of plant tolerance through transporters and/or transcription factors. Osmatic regulation is attained in plants by opening and closing of channels using transporters (Trptrs) related to the cations (Ca2+, Na+, and H+). Furthermore transcription factors (TFs) play a crucial role in operating plant tolerance (e.g., SnRKs interact with other TFs and/or genes by phosphorylating and activating more genetic factors to help plants build food reverse that can be utilized under the stress condition. In the figure, HKT: high affinity K+ transporters; SOS: salt overly sensitive 1; (X): various transporters/genes such as 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 12, 14, 16, 18, and 20; CNGC: cyclic-nucleotide-gated channels; MOCA1: monocation-induced Ca2+ enhancer 1; ANN: Annexin; IPUT: inositol-phosphorylceramide glucuronosyltransferase; GlyIPC: glycosyl-inositol-phosphorylceramide; CA: cation exchanger; NH: Na+/H+ exchangers; cNMP: cyclic-nucleotide monophosphate; and CaM: calmodulin.
Ijms 23 00651 g005
Table 1. Information of the sequenced species provides a source of genetic manipulation and understanding of the domestication process.
Table 1. Information of the sequenced species provides a source of genetic manipulation and understanding of the domestication process.
Family NameOrganismGSNPTsOnline Accessible Links
AmaranthaceaeBeta vulgaris (sugar beet), spp. vulgaris var. cicla)604 Mbp34,521https://bvseq.boku.ac.at/
Suaeda aralocaspica (shrubby sea-blite)467 Mbp29,604https://www.ncbi.nlm.nih.gov/bioproject/?term=PRJNA428881
ArecaceaeElaeis guineensis (African oil palm)1800 Mbp25,405https://www.ncbi.nlm.nih.gov/genome/?term=txid51953[orgn]
Phoenis dactylifera (date palm), an elite variety (Khalas)605.4 Mbp~41,660https://pubmed.ncbi.nlm.nih.gov/23917264/
BrassicaceaeArabidopsis thaliana (Arabidopsis)125 Mbp~27,025https://www.arabidopsis.org/ and https://www.nature.com/articles/ng.807
Arabidopsis lyrata (Arabidopsis)207 Mbp~32,670https://www.arabidopsis.org/ and https://www.nature.com/articles/ng.807
Capsella rubella (pink shepherd’s-purse)134.8 Mbp~28,447https://www.nature.com/articles/ng.2669
Eruca sativa (salad rocket)∼851 Mbp45,438https://www.frontiersin.org/articles/10.3389/fpls.2020.525102/full
Eutrema salsugineum (saltwater cress)241 Mbp26,531https://www.frontiersin.org/articles/10.3389/fpls.2013.00046/full
CannabaceaeCannabis sativa (hemp)808 Mbp38,828https://www.nature.com/articles/s41438-020-0295-3
CactaceaeCarnegiea gigantea (saguaro)1.40 GB28,292https://www.pnas.org/content/114/45/12003
CucurbitaceaeCucumis melo (musk melon), doubled-haploid line DHL92375 Mbp27,427https://www.pnas.org/content/109/29/11872#abstract-1
Cucumis sativus (cucumber), ‘Chinese long’ inbred line 9930226.2 Mbp26,682https://academic.oup.com/gigascience/article/8/6/giz072/5520540
DioscoreaceaeDioscorea rotundata (Yam)594 Mbp26,198https://bmcbiol.biomedcentral.com/articles/10.1186/s12915-017-0419-x
EuphorbiaceaeManihot esculenta (cassava), domesticated KU50495 Mbp37,592https://www.nature.com/articles/ncomms6110#Sec8
FabaceaeCajanus cajan (pigeon pea)833.07 Mbp48,680https://www.nature.com/articles/nbt.2022
Cicer arietinum (chickpea)∼738 Mbp28,269https://www.nature.com/articles/nbt.2491
Glycine max (soybean), cultivar Williams 82 969.6 Mbp46,430https://www.nature.com/articles/nature08670#Sec9
Medicago turncatula (medick or burclover) ~330 Mbp50,894http://europepmc.org/article/MED/24767513
Vigna unguiculata (cowpea)640.6 Mbp29,773https://onlinelibrary.wiley.com/doi/full/10.1111/tpj.14349
GinkgoaceaeGinkgo biloba (ginkgo)10.61 Gb41,840https://gigascience.biomedcentral.com/articles/10.1186/s13742-016-0154-1
MusaceaeMusa acuminata (Banana) spp. Malaccensis523 Mbp36,542https://www.nature.com/articles/nature11241
PinaceaePicea abies (Norway spruce)20 GB28,354https://www.nature.com/articles/nature12211
PoaceaeHordeum vulgare (barley)5.1 GB26,159https://www.nature.com/articles/nature11543
Oryza sativa (rice)373.2 Mbp3475https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5395016/
Phyllostachys heterocycla var. pubescens2.05 Gb31,987https://www.nature.com/articles/ng.2569
Phyllostachys edulis1.91 GB51,074https://academic.oup.com/gigascience/article/7/10/giy115/5092772
Raddia distichophylla (Schrad. ex Nees) Chase589 Mbp30,763https://academic.oup.com/g3journal/article/11/2/jkaa049/6066164
Sorghum bicolor (sorghum), Rio genetic material729.4 Mbp35,467https://bmcgenomics.biomedcentral.com/articles/10.1186/s12864-019-5734-x#Sec6
Triticum urartu (einkorn wheat), accession G1812 (PI428198)~4.94 GB34,879https://www.nature.com/articles/nature11997
Zea mays (maize), B73 inbred maize line2.3 GB>32,000https://pubmed.ncbi.nlm.nih.gov/19965430/
SalicaceaePopulus tricchocarpa (poplar)380 Mbp37,238https://www.pnas.org/content/115/46/E10970#sec-1
SolanaceaeCapsicum annuum (pepper)3.06 GB34,903https://www.nature.com/articles/ng.2877#Sec10
Nicotiana benthamiana (tobacco)3.1 GB42,855https://www.biorxiv.org/content/10.1101/373506v2
Solanum lycopersicum (tomato), cv. Heinz 1706799.09 Mbp34,384https://www.biorxiv.org/content/10.1101/2021.05.04.441887v1.full.pdf
Solanum tuberosum (potato)844 Mbp39,031https://www.nature.com/articles/nature10158/
ArecaceaeElaeis guineensis (African oil palm)1.8 GB~34,802https://www.nature.com/articles/nature12309
Phoenis dactylifera (date palm), an elite variety (Khalas)605.4 Mbp~41,660https://europepmc.org/article/PMC/3741641
RosaceaePrumus persica (peach)247.33 Mbp26,335https://onlinelibrary.wiley.com/doi/10.1111/tpj.15439?af=R
VitaceaeVitis sylvestris (grape), accession of Sylvestris C1-2 469 Mbp39,031https://genomebiology.biomedcentral.com/articles/10.1186/s13059-020-02131-y#Sec2
Genome size—GS; number of predicted transcripts/proteins—NPTs.
Table 2. Different plants produce various kinds of plant metabolites at varying developmental stages under stress conditions by regulating primary and/or secondary metabolism.
Table 2. Different plants produce various kinds of plant metabolites at varying developmental stages under stress conditions by regulating primary and/or secondary metabolism.
PlantEStage and Specific OrganMetabolitesRefs.
Avena sativa (oats)E1Not specified using grainsPM **: malic, gluconic, and galacturonic acids, fatty acids (FAs), palmitic acid and linoleic acid.[155]
E2Seedling stage (three weeks old) Leaves PMS **: Ascorbate, aldarate phenylpropanoids.[156]
Hordeum vulgare (barley)E1Germination using seedsPM *: glycero(phospho)lipids, prenol lipids, sterol lipids, methylation.
SM *: polyketides.
[157]
E2Two-leaf stage seedlings using leavesPM **: organic acids (OAs), amino acids (AAs), nucleotides, and derivatives.
SM *: flavonoids, absiscic acid.
[158]
E3Three-leaf stage using leavesSM **: chlorogenic acids, hydrocinnamic acid derivatives, and hordatines and their glycosides.[119]
E4Three-leaf stage and flag leaf stage using leavesSM *: flavonoids, hydroxycinnamic acid, phenolics, glycosides, esters, and amides.[159]
E5During grain filling using seedsPM *: Tricarboxylic acid (TCA), OAs, aldehydes, alcohols, polyols, FAs, carbohydrates, mevalonate.
SM **: phenolic compounds, flavonoids.
[10]
E6Four weeks old using leavesPM *: carbohydrates, free AAs, carboxylates, phosphorylated intermediates, antioxidants, carotenoids.[160]
E71–3 weeks old using leaves and rootsPM **: AAs, sugars, OAs as fumaric acid, malic acid, glyceric acid[161]
Oryza sativa L. (rice)E1Flowering and early grain filling stages using leaves, spikelets, seedsPSM **: isoleucine, 3-cyano-alanine, phenylalanine, spermidine, polyamine, ornithine[161]
E2At reproductive stage using leaves and grains ripe stagesPM **: saturated and unsaturated FAs, AAs, sugars, and OAs.[162]
E324 months old seeds usedPM **: sugar synthesis related compounds, AAs, free FAs, TCA cycle intermediates.[163]
E4Not specified using grainPM *: aromatic AAs, carbohydrates, cofactors and vitamins, lipids, oxylipins, nucleotides.
SM *: benzenoids.
[164]
E5Maturation using mature seedPM *: carbohydrates, lipids, cofactors, prosthetic groups, electron carriers, nucleotides.
SM *: benzenoids.
[165]
E6Maturation using mature seedPM *: carbohydrates and lipids.
SM *: α-carotene, β-carotene, and lutein.
[166]
E7Six weeks old using leavesPM *: AAs (arginine, ornithine, citrulline, tyrosine, phenylalanine and lysine), FAs and lipids, glutathione, carbohydrates.
SM *: rutin, acetophenone, alkaloids.
[157]
Setaria italica (foxtail millet) E160 days using shootsPM *: fructose, glucose, gluconate, formate, threonine, 4-aminobutyrate, 2-hydroxyvalerate, sarcosine, betaine, choline, isovalerate, acetate, pyruvate, TCA-OAs, and uridine.[167]
E23–5 leaves stages using leavesPM *: glycerophospholipids, AAs, OAs.
SM: flavonoids, hydroxycinnamic acids, phenolamides, and vitamin-related compounds.
[167]
Sorghum bicolor (sorghum)E1Four-leaf stage using leavesPM *: AAs, carboxylic acids, FAs.
SM: cyanogenic glycosides, flavonoids, hydroxycinnamic acids, indoles, benzoates, phytohormones, and shikimates.
[168]
E2Four-leaf stage using leavesSM *: 3-Deoxyanthocyanidins, phenolics, flavonoids, phytohormones, luteolinidin, apigeninidin, riboflavin.[169]
E3Around 26 days using roots and leavesPM *: sugars, sugar alcohols, AAs, and OAs.[170]
E4Four weeks old using grain and biomassPM **: OAs.
SM **: phenylpropanoids.
[146]
Triticum aestivum (wheat)E1NAS using leavesPM *: sugars, glycolysis and gluconeogenesis intermediates, AAs, nucleic acid precursors, and intermediates.
SM *: chorismate, polyamines, L-pipecolate, amino-adipic acid, phenylpropanoids, terpene skeleton, and ubiquinone.
[171]
E2Physiological maturity using leavesPM *: AAs metabolism, sugar alcohols, purine metabolism, glycerolipids, and guanine.
SM *: shikimates, anthranilate, absiscic acid.
[172]
E3Maturation using matured kernelsPM *: FAs, sugar, nucleic acids and derivatives.
SM *: phenolamides, flavonoids, polyphenols, vitamins, OAs, AAs, phytohormones, and derivatives.
[173]
E4Not specified using grainPM *: osmolytes, glycine betaine, choline, and asparagine.[174]
E5Not specified using seedsPM *: sterols, FAs, long chain FAs derivatives, glycerol (phospho) lipids.
SM *: polyketides.
[175]
Zea mays (maize) E1R6 stage using grains PM **: sugars, sucrose, glucose, and fructose.[176]
E2Physiological maturity using kernelsPM *: glycolysis, TCA cycle, starch, amino acids.
SM: alkaloids, benzenoids, fatty acid and sugar derivatives, flavonoids, phenylpropanoids, and terpenoids.
[177]
E38 months using kernelsPM *: glucose, fructose, sucrose, tocopherol, phytosterol, inositol, asparagine, glutamic acid, pyroglutamic acid.[178]
E4Eight-visible-leaf stage using leavesPM *: choline, inositol, sugars, raffinose, rhamnose, TCA cycle, AAs, trigonelline, putrescine, quinate, shikimate.
SM *: flavonoids, and benzoxazinoids.
[179]
E5Seedling stage using entire seedlingPM *: amino acids, lipids, carboxylic acid.
SM *: alkaloids, terpenoids, flavonoids, alkaloids, and benzenoids.
[180]
E6Physiological maturity using kernelsSM *: flavanones, flavones, anthocyanins, and methoxylated flavonoids.[181]
E—experiment; **—upregulation/significant contents; *—difference examined as compare to control/mock; primary and secondary metabolism/metabolites—PSM; primary metabolism/metabolites—PM; secondary metabolism/metabolites—SM; not available stage—NAS.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ashraf, M.F.; Hou, D.; Hussain, Q.; Imran, M.; Pei, J.; Ali, M.; Shehzad, A.; Anwar, M.; Noman, A.; Waseem, M.; et al. Entailing the Next-Generation Sequencing and Metabolome for Sustainable Agriculture by Improving Plant Tolerance. Int. J. Mol. Sci. 2022, 23, 651. https://doi.org/10.3390/ijms23020651

AMA Style

Ashraf MF, Hou D, Hussain Q, Imran M, Pei J, Ali M, Shehzad A, Anwar M, Noman A, Waseem M, et al. Entailing the Next-Generation Sequencing and Metabolome for Sustainable Agriculture by Improving Plant Tolerance. International Journal of Molecular Sciences. 2022; 23(2):651. https://doi.org/10.3390/ijms23020651

Chicago/Turabian Style

Ashraf, Muhammad Furqan, Dan Hou, Quaid Hussain, Muhammad Imran, Jialong Pei, Mohsin Ali, Aamar Shehzad, Muhammad Anwar, Ali Noman, Muhammad Waseem, and et al. 2022. "Entailing the Next-Generation Sequencing and Metabolome for Sustainable Agriculture by Improving Plant Tolerance" International Journal of Molecular Sciences 23, no. 2: 651. https://doi.org/10.3390/ijms23020651

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