Next Article in Journal
A Novel Role of Arrhythmia-Related Gene KCNQ1 Revealed by Multi-Omic Analysis: Theragnostic Value and Potential Mechanisms in Lung Adenocarcinoma
Next Article in Special Issue
The Plastome Sequences of Triticum sphaerococcum (ABD) and Triticum turgidum subsp. durum (AB) Exhibit Evolutionary Changes, Structural Characterization, Comparative Analysis, Phylogenomics and Time Divergence
Previous Article in Journal
Cytoskeleton Protein EB3 Contributes to Dendritic Spines Enlargement and Enhances Their Resilience to Toxic Effects of Beta-Amyloid
Previous Article in Special Issue
Cold Response Transcriptome Analysis of the Alternative Splicing Events Induced by the Cold Stress in D. catenatum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Expanding Gene-Editing Potential in Crop Improvement with Pangenomes

School of Biological Sciences, The University of Western Australia, Perth, WA 6009, Australia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(4), 2276; https://doi.org/10.3390/ijms23042276
Submission received: 17 January 2022 / Revised: 14 February 2022 / Accepted: 15 February 2022 / Published: 18 February 2022
(This article belongs to the Special Issue Plant Genomics and Genome Editing)

Abstract

:
Pangenomes aim to represent the complete repertoire of the genome diversity present within a species or cohort of species, capturing the genomic structural variance between individuals. This genomic information coupled with phenotypic data can be applied to identify genes and alleles involved with abiotic stress tolerance, disease resistance, and other desirable traits. The characterisation of novel structural variants from pangenomes can support genome editing approaches such as Clustered Regularly Interspaced Short Palindromic Repeats and CRISPR associated protein Cas (CRISPR-Cas), providing functional information on gene sequences and new target sites in variant-specific genes with increased efficiency. This review discusses the application of pangenomes in genome editing and crop improvement, focusing on the potential of pangenomes to accurately identify target genes for CRISPR-Cas editing of plant genomes while avoiding adverse off-target effects. We consider the limitations of applying CRISPR-Cas editing with pangenome references and potential solutions to overcome these limitations.

1. Introduction

The world’s population is predicted to increase to nearly 10 billion people by 2050 [1], coupled with a predicted increase of average surface temperature of 2 °C by 2043 [2] and more variable weather patterns. Hence, there is a need for more climate change-ready crops with increased yield [3,4]. Genome-editing technologies using nucleases such as Clustered Regularly Interspaced Short Palindromic Repeats and CRISPR associated protein Cas (CRISPR-Cas), Zinc finger nucleases (ZFNs) [5,6,7], and transcription activator-like effector-based nucleases (TALEN) [8,9,10,11] have already demonstrated their capacity in supporting crop resilience and yield improvements in several species [12,13,14,15,16], but the lack of knowledge of genome diversity and the polyploid nature of some crop genomes complicates the targeting of gene editing sites, leading to inefficiency in the traits that could be improved and the potential for adverse off-target effects from gene-editing experiments [12,17].
Modern crops have been subjected to breeding and selection, which often leads to large modifications in the genome and can reduce the efficiency of genome editing to improve specific functional traits or analyse traits through mutagenesis [18,19,20]. There is significant genome variation present between individuals, and this genomic variation can be associated with important functional traits [21,22,23,24,25,26], such as nucleotide deletions linked with embryo sac fertility and presence/absence variation of genes (PAV) linked with submergence tolerance, yield, and phosphorus deficiency tolerance in rice (Oryza sativa) [26]; PAV linked with silique length, seed weight, and flowering time in Brassica napus [27]; PAV associated with disease resistance, acyl lipid metabolism, and glucosinolate metabolism in B. napus [28]; and SNPs linked with number of branches, number of seeds per pod, number of pods per plant, plant height, seed weight, and seed yield in pigeon pea (Cajanus cajan) [29]. Gene editing to improve agronomically-significant traits is challenging if the target gene is not present in the reference genome sequence and the sequence cannot be used to tailor the genome editing experiment, and so single reference genomes are often inadequate for designing editing target sites.
The compilation of multiple genome sequences into a pangenome instead of a single reference genome provides a genomic sequence resource that can adequately represent genome diversity in different varieties or species. The advantages and disadvantages of CRISPR-Cas, TALENs, and ZFNs have been extensively reviewed, with CRISPR-Cas generally being easier to design and use and TALENs allowing higher specificity to targets [30,31,32,33], so here we focus on the CRISPR system for genome editing, including CRISPR/Cas9 and CRISPR/Cpf1 (previously known as CRISPR/Cas12a).

2. Pangenomes

Pangenomes were first introduced by Tettelin et al. [34] to describe gene diversity in Streptococcus agalactiae. Pangenomes can be constructed through the sequencing of individual genomes or survey of genetic variations within a species to describe the extensive repertoire of variation. The use of pangenomes removes the sample bias caused by using single reference genome assemblies, allowing the identification of structural variations (SVs) to assess the diversity within species. This diversity can encompass PAVs and non-genic regions. Genes in pangenomes can be classified as core, present in all individuals of the species, or dispensable, where they are absent in at least one individual, also known as accessory or variable genes. These gene classifications are sometimes extended to include private genes (present in 1% or fewer individuals) and near core/shell genes that are present in 99% or more of individuals [28,29,35,36,37].
Pangenomes are mostly constructed in one of three ways (Figure 1). The first, de novo sequencing and comparison, involves the sequencing, assembly, and comparison of multiple genomes to identify core and variable genes and genomic regions [38]. This approach reveals the physical position of genes and other genomic elements. However, errors in assembly and annotation may lead to the false calling of variation [28]. Furthermore, this approach is costly and requires high-quality data with high sequencing coverage, limiting the application to relatively few individuals. The second method for pangenome construction, iterative mapping, and assembly, uses a single reference genome as a base for the pangenome. Whole genome sequence data for multiple individuals is aligned to the reference genome and any non-aligning sequence reads are assembled and added to the reference to build a pangenome [28]. This approach is less expensive than de novo assembly and comparison as it requires much less data, and so permits the assessment of large numbers of individuals with relatively low sequencing coverage. After pangenome construction, gene PAVs can be called by realigning the sequencing data from each individual back to the final assembled pangenome. This approach usually only calls PAV within genes and requires further analysis to accurately place the non-reference contigs within a genomic context. However, a combination of de novo assembly for a small number of representative individuals together with iterative assembly using large numbers of low coverage individuals provides both genomic context and PAV at a population scale, permitting in depth diversity analysis [34,39].
The third way to assemble pangenomes uses graph-based approaches, including sequence and variation graphs (VGs) [27,40] and practical haplotype graphs [41]. Pangenome graphs can be constructed from whole genome assemblies or by de novo graph genome assembly. Instead of a single representative sequence, the graph represents genomic variants as multiple paths. Sequence regions shared between individuals are collapsed into a single path and SVs are added to the graph as a node at the genomic location of their discovery [42,43]. In doing so, variant information for dispensable regions are stored as unique paths through the graph, displaying genomic diversity and sequence conservation [27]. Graph-based pangenomes can provide gene position information, but are computationally intensive to assemble, and graph quality correlates directly with the quality of the input data. However, with further advancements in DNA sequencing and data processing, particularly the expansion of high quality long read data, pangenome graphs will become the standard approach to assemble pangenomes. Regardless of the manner of construction, pangenomes can provide comprehensive data resources that can be used in trait association and in guiding the CRISPR-Cas design, supporting efficient genome editing.
Due to falling sequencing costs and the increased acknowledgment of significant gene presence/absence variation in some species, pangenomes have expanded beyond bacteria to higher organisms such as chicken [44] and human [45] as well as many plant species, allowing the analysis of the large-scale PAV observed in plants [46,47]. Pangenomics in plants was first proposed by Morgante et al. in 2007 [48] and since then, pangenomes have been assembled for many crop plant species including soybean (Glycine max) [49,50], maize (Zea mays) [51], tomato (Solanum lycopersicum) [35], Brassica oleracea [39], Brassica napus [27,52], Brachypodium distachyon [53], barley (Hordeum vulgare) [54], rice [55], pigeon pea (Cajanus cajan) [29,56], apple (Malus domestica) [57], capsicum [25], sesame (Sesamum indicum) [58], sunflower (Helianthus annuus) [59], yuca (Manihot esculenta) [60], sorghum (Sorghum bicolor) [36,61], and bread wheat (Triticum aestivum) [62]. Pangenomes for non-food plant species such as Arabidopsis thaliana [63], Amborella trichopoda [64], cotton (Gossypium) [65], and barrel clover (Medicago truncatula) [66] have also been published (Table 1). These pangenomes are valuable resources for studying genomic variation within plant species, assisting with the association of genes with traits, and supporting accurate and specific guide RNA design. For example, the use of pangenomes has identified genes corresponding with disease resistance gene analogs (RGAs) in B. oleracea [67] and disease resistance gene loss in cotton [65] that can be further targeted.

3. Association Analysis Using Pangenomes Can Reveal Valuable Sites for Genome Editing

The regulation of gene expression and functional genome analysis using CRISPR-Cas systems has been widely demonstrated [37,69,70,71], including in the development of improved crops [71,72,73]. However, the editing efficiency achieved in plant studies can vary depending on the genotype and target site selected [72,73,74]. The inconsistency in CRISPR-Cas mutation rate can be partially attributed to target site GC content, target accessibility (due to chromatin state), and sgRNA secondary structure [37,69,75,76]. Successful editing can be even more challenging in polyploid plants because of the potential to edit multiple alleles or overcome genomic redundancy that may disguise the impact on the phenotype [77,78]. Detecting variant alleles and mapping their position in a pangenome allows for the design of allele-specific CRISPR sgRNA, as the alleles may have distinct effects in the plant phenotype. For example, the mlo gene discovered in barley (Hordeum vulgare) was used for decades in several crop species for inducing broad-spectrum resistance to powdery mildew. However, the pleiotropic effects of mlo can negatively affect yield [79]. To circumvent this issue, different mlo allele combinations can be used to modulate the degree of plant susceptibility to the pathogen and pleiotropic phenotype [80]. In wheat, mlo mutant plants also showed an allele-specific level of enhanced susceptibility to powdery mildew disease [79]. Numerous other examples of allele-specific phenotypes were observed to modulate crop disease resistance [59,81,82], abiotic stress tolerance [83], herbicide resistance [84,85], and yield in polyploid crops such as wheat [86] and camelina (Camelina sativa) [87]. In camelina, the selective mutagenesis of the three delta-12-desaturase genes (FAD2) showed reduced levels of polyunsaturated fatty acids and increased accumulation of oleic acid in the oil, corresponding with the different alleles for the three FAD2 loci [87]. The specificity of CRISPR-Cas opens new doors to testing the effects of individual small variants against a controlled genetic background. CRISPR-Cas can be used to study the effect of gene dosage by generating a series of allelic mutants through knock-out/down mutation of specific variant alleles [88,89]. A pangenome analysis associated with phenotypic information can assist the identification of these variant alleles and delimit CRISPR-Cas target sites, leading to the development of better performing varieties in the field. Structural variance uncovered by pangenomes can provide new alleles for genome functional analysis and also give detailed information about target allele location and accessibility in the genome (Figure 2).
A feature of pangenomes is the ability to show the impact of chromosomal inversions that can then be targeted by CRISPR-Cas editing for re-inversion (Figure 3). Chromosomal inversions can have a considerable impact on crop breeding as inverted regions are often prevented from crossing during recombination. In maize, one example of pangenomic analysis of chromosomal inversions is reported, using platinum-grade reference genomes from 66 maize key inbred lines. This analysis revealed several large (more than 100 kb) chromosomal rearrangements including insertions, deletions, and inversions on all 10 chromosomes, with the largest inversion spanning 75.5 Mb in the pericentric region of chromosome 2 [90]. The identification of this large structural variant by pangenome analysis and subsequent re-inversion of the genomic segment using CRISPR-Cas9 re-established the previous chromosomal state, allowing for the genes locked in this region to be accessed for recombination with the other inbred lines, which otherwise would be unfeasible [89]. This re-inversion in maize demonstrates the potential of pangenomes to identify chromosomal rearrangement boundaries with high precision and allows the editing of large regions of the chromosome within these chromosomal rearrangement boundaries using CRISPR-Cas.

4. Targeted Mutagenesis Guided by Pangenomes

Understanding the relationship between genotypes used in pangenome construction can assist in the identification of agronomically important SVs and subsequent targeted mutagenesis, particularly when comparing domesticated species to wild relatives. A recent study assembled a rice pangenome composed of 66 accessions that displayed green revolution phenotypes such as reduced height and early flowering phenotypes, with the aim of uncovering the underlying genes controlling these traits [91]. SV analysis of this rice pangenome identified 129 conserved gene loci potentially related to the shared phenotype. The analysis was followed by a subsequent CRISPR-Cas knock-out/down study uncovering 31 high yield-related genes, including six previously reported genes such as the sd1 semi-dwarf gene [91]. In a similar vein, the pangenome for medicinal cannabis (Cannabis sativa) was used to mine for cannabinoid biosynthesis genes, and 145 sgRNAs were generated for genes in the cannabinoid biosynthesis pathways [92]. These pedigree-based approaches for pangenome analysis take advantage of clear directional selection for finding conserved candidate genes among individuals with the desired phenotype, showing how pangenome-associated data can provide a powerful resource to uncover the role of genomic variations on the phenotype.
The inclusion of omics data such as transcriptome, metabolome, and proteomes can further support gene functional characterisation in pangenomes. A study in Brassica napus employed transcriptome-wide association studies in a constructed pangenome to identify QTLs related to regulating seed oil content in eight different environments across multiple years (2012–2018). Initially, 692 genes and four sets of coexpressed genes were significantly associated with seed oil content based on the seed transcriptome. A collection of genes more likely related to seed oil content were ranked using a gene prioritisation framework based on the multi-omics dataset. CRISPR-Cas9 and T-DNA mutants were employed to validate candidate gene function, revealing that two homologous genes (BnPMT6s) negatively regulate seed oil content [93]. Another study mapped 359 previously identified QTLs related to tomato flavour and aroma to the tomato pangenome, defining potential target regions for improving tomato lines [94]. The tomato pangenome was used to find promoter regions associated with QTLs related to tomato aroma, resulting in the identification of promoter alleles such as FLORAL4, which were present only in wild lines and have been lost during domestication [94]. These publications show that functional analysis of pangenome variable regions using omics data can broaden the understanding of QTL region conservation during domestication, assist in discovering structural variation and novel alleles, and map previously reported QTLs to support the selection of candidate genes for mutagenesis.
In silico association studies using small variants such as GWAS are usually limited to identifying a set of co-located, co-inherited loci linked within a haplotype [95]. However, in rice it was shown that up to 41.6% of trait-associated SNPs are located in presence/absence variable regions of the genome [96], which may be overlooked in a single reference genome. Pangenomes support comprehensive haplotyping by providing a genomic resource with variants across a diverse population of individuals, providing a full set of targets for modification with CRISPR-Cas [26]. In cases where the trait-associated haplotype is present in or near a gene, reverse genetics can be applied for inference of gene function through disruption of the promoter region with CRISPR-Cas [70,91]. However, traditional knockout experiments fail to characterise the specific effects of different small variants that could be contributing to a given trait of interest. Knockout experiments also rely on accurate gene annotations, which can be erroneous. Hence, characterising the individual and combinatorial effects of small variants linked within a trait-associated haplotype requires parallelizing sequence modifications at different genomic positions [89].
The first effective CRISPR-Cas toolkit for multiplexed editing in plants was developed in 2014 [97], though large-scale editing was not conducted until 2017 with the construction of mutant libraries for rice involving over 100,000 sites [98,99]. Since then, multiplexed editing has been applied to induce novel mutations in genic regions for rice, Brassica napus [100], soybean [101], and maize [37]. Whilst research with high throughput mutagenesis is promising, mutant libraries still do not functionally validate existing small variants present within and across plant populations. Small variant functional validation was accomplished in 2018 in humans using an approach called ‘saturation editing’, where CRISPR-Cas was used to assay 96.5% of all SNPs across 13 exons encoding for functional domains of the breast cancer susceptibility gene BRCA1 [102]. A potential solution to making small variant characterisation cost-effective in plants is to concentrate the established plant multiplexing methods to conduct saturation editing on specific trait-associated haplotypes from pangenomic datasets. Isolating the impact of specific small variants and allelic combinations identified in pangenomes will aid in elucidating the biochemical mechanisms involved in functional pathways. This understanding of variants and allelic combinations will provide a comprehensive understanding and precise control over specific variants underlying agronomic traits [101], enabling breeders to produce tailored crop varieties.

5. Off-Targets Effects in Multiplexed Editing

Off-target effects in CRISPR-Cas are often prevalent when employing multiplexed genome editing, particularly associated with assembly, expression, and processing of sgRNAs arrays and efficient delivery using current transformation methods. Multiplex genome editing involves simultaneously modifying multiple loci with multiple or single target-specific gRNA(s) [103]. The number of loci that can be edited by CRISPR-Cas in parallel is improving, but some technological challenges still remain in high-throughput mutagenesis. Beyond bottlenecks in throughput of sgRNA design and synthesis, multiplexing can lead to unintended interactions and competition between parallel CRISPR machinery [104,105] that can reduce binding specificity and efficiency as the number of simultaneously edited loci increases [106]. In addition, off-target binding remains a significant obstacle for guide design that scales with the number of targeted sites. Pangenome references are valuable tools for improving guide design as they can identify all potential off-target sites in a given population, which is necessary for cultivar-specific design of sgRNAs where variability may be present in the target sequence or protospacer adjacent motif (PAM) site within the population.
The use of pangenomes and associated knowledge of the gene content of the individual being modified can increase the reliability of genome editing technologies. The sequence of the CRISPR single guide RNA (sgRNA) is designed to match target sequences in the genome, within a specific distance up or downstream of a PAM site, which serves as the binding site for the Cas protein. The efficiency of the CRISPR-Cas system is impacted by the selection of the CRISPR target site that guides the Cas protein to a specific region within the genome of the individual. Potentially deleterious off-target activity can occur in regions of the genome that share sequence identity with the target site such as duplicated/repeated sequences [107]. Off-target effects are often undesirable and have been observed in rice (Oryza sativa) [108], grapevine (Vitis vinifera) [109], and cotton (Gossypium) [110]. Avoiding off-target effects requires detailed genomic information for the individual being modified [107]. Using pangenome references containing all variant data can help to avoid off-target effects because the gene editing design process will incorporate all available data and not just the data of the reference individual [111]. This comprehensive availability of data allows researchers to design specific sgRNAs that can accurately and precisely target the region of the gene (allele) and avoid mismatches due to sequence variation [12,17]. By targeting specific differences in allele sequences such as PAV and SNPs discovered through pangenomes, functional traits of crop species may be altered with great efficiency.

6. Future Applications of Pangenomes in Genome Editing through Super-Pangenome Guided CRISPR-Cas

A valuable target for genome editing is reintroducing agronomically beneficial genes that are lost in domesticated crop species but conserved in wild relatives. Genes can be lost in cultivars compared to wild types if they are selected against during domestication and breeding, both intentionally and unintentionally [20,112]. These lost genes can have agronomically beneficial functions such as disease resistance [113,114] or adaptations to extreme environments such as heat and drought tolerance or efficient nutrient use strategies [115]. In many crop species, genomic regions and functions unrelated to yield have been lost due to domestication selection including the Yr36 gene for rust resistance in bread wheat [116] and disease resistance genes from rice [117] and sorghum (Sorghum bicolor) [118]. Reintroduction of variable genes through wild introgression is possible as shown by the discovery of the TomLoxC promoter allele linked with flavour reintroduced into modern tomato cultivars [35]. However, wild introgression can potentially introduce deleterious alleles such as altered flowering time or reduced plant size [13,119,120]. Genome editing through CRISPR-Cas could allow the reintroduction of these traits without associated deleterious alleles by multiplexed editing of SVs linked to traits, but this requires thorough analysis of a wide gene pool of the species to increase specificity of the target and avoid off-target effects.
A super-pangenome aims to represent the genetic architecture of a group of taxa above the species level by combining different pangenomes from all species within that group [121]. By studying super-pangenomes, markers associated with desirable traits in wild relatives can be incorporated in domesticated cultivars. Mapping these gene PAVs through pangenomes allows for them to be compared with wild and exotic relatives to characterise advantageous traits that may have been lost as a result of selective breeding. Through CRISPR-Cas modification of targeted alleles based on wild relatives, these advantageous traits can be reintroduced in crops without also introducing associated deleterious alleles [122].
Super-pangenomes also show potential use in the future domestication of wild crop relatives as new food sources and improvement of modern crop varieties, primarily by linking PAVs and SVs to candidate domestication genes to guide CRISPR-Cas modification in a different species [121]. A benefit super-pangenomes provide for de novo domestication is a shared reference for direct comparison between all types of genomic variants present between advanced crops and wild species, in both core and dispensable regions. This allows inquiry into evolutionary divergence at loci of agronomic interest that has taken place since speciation, or over the course of domestication, highlighting specific targets for modification that may not be present at a single species level. Candidate domestication genes are genes strongly linked to traits such as increased yield, seed shatter resistance, flowering time, plant architecture, and climatic tolerance [13,123,124,125,126,127,128,129]. Modification of domestication genes through multiplexed CRISPR-Cas has been successful in wild crop relatives of tomato to produce new lines with increased fruit size, yield, and nutritional value, and greater abiotic and biotic stress tolerance than cultivated tomato lines [15,130,131,132]. Potential wild crop relatives for domestication through CRISPR-Cas include pennycress (Thlaspi arvense) to an oilseed crop with cold tolerance [133]; weeping grass (Microlaena stipoides) to a cereal crop with abiotic stress tolerance [134]; and common wild rice (Oryza rufipogon), wild emmer wheat (Triticum dicoccoides), and teosinte (Zea mays ssp. parviglumis) to new cereal crops with relatively high genetic diversity [135,136,137]. Furthermore, the transfer of traits between modern varieties is also possible, such as CRISPR-Cas modification of disease resistance genes in Brassica to other Brassicaceae species [138] or transfer of disease resistance genes in rust-resistant wheat varieties to other Poaceae species such as barley or sorghum [116]. Domestication of wild plants or improvement of modern crops based on candidate domestication genes in related species or genera would be a significant step in securing future food resources, as this will lead to varieties that are adapted to stressful environments or ecological niches [139], and the overall expansion of genetic diversity within agricultural systems [140,141].
Another application for super-pangenomes combined with CRISPR-Cas modification is the transfer of valuable adaptations for adverse environments across different plant genera or families through higher level super-pangenomes. While few pangenomes have been published above the species level, a super-pangenome of 10 different Poplar species was constructed [142,143] as well as a banana super-pangenome made of 15 different accessions [144]. Rapid advances in pangenome construction suggest that pangenomes spanning multiple eukaryotic genera and even families will become available in the near future. Analysis of desirable traits in genera and families related to crop species will allow multiplexed CRISPR-Cas systems to be designed to translate similar traits into crop varieties. Such adaptations could include increased photosynthetic efficiency in C3 plants such as rice via transporter adaptations of the carbon assimilation pathway in C4 plants [13,145], increased resilience to climate change based on SNPs across species associated with precipitation and thermal variability tolerance [146], and higher nutrient-use-efficiency based on phosphorus-saving strategies of the non-mycotrophic Proteaceae family in south-west Australia [147]. Development of new crop varieties with these traits through multiplexed CRISPR-Cas using super-pangenomics will be an important step to securing food resources against global food challenges such as climate change and new diseases [148].

7. Challenges and Considerations

A major limiting factor for the utilisation of pangenomes to guide CRISPR-Cas editing in crops is the size and number of parallel targets. The most common window sizes for base editors ranges from 4–5 bp to 50–150 bp, allowing for punctual modifications in the target [149,150,151]. Nonetheless, larger cassette insertions have been reported in plants through the use of different CRISPR methods such as non-homologous end joining (NHEJ) that have allowed insertions and deletions that are a few kilobases in size [152,153,154]. However, efficiency drops steeply as the size of the edit increases [155]. This drop in efficiency poses a significant constraint for editing SVs identified in pangenomes. Further approaches involve the development of novel Cas proteins requiring different or flexible PAM sequence sites that allow the targeting of a wider range of genomic sites, particularly when combined with multiplexed CRISPR-Cas [156]. Given the significant effect SVs have in crop evolution and diversity, future efforts should focus on developing methods to increase the size of the catalytic window for CRISPR-Cas editing to fully exploit advances in pangenomics.
With regards to shortcomings in pangenomics, iterative and de novo constructed pangenomes usually leave the chromosomal placement unspecified for variable regions that are not present in the original reference. These additional contigs can still theoretically be targeted with CRISPR for individuals in which they are present, but not knowing their genomic context can limit our understanding of regulatory elements that may influence their expression. Variation graph approaches overcome this limitation, however, they rely heavily on long-reads or deep-sequencing a large number of individuals to capture SVs at a high fidelity. Therefore, improvements in sequencing technology that allow the affordable identification of SVs are needed to enhance the information that pangenomes can provide for editing with CRISPR-Cas.

8. Conclusions

Pangenomes are valuable tools to identify agronomically important gene variants. To date, many plant pangenomes have been used to aid CRISPR studies in locating and targeting genomic regions of interest, broadening our understanding of QTL regions, supporting high throughput mutagenesis approaches, and linking important SVs to traits. Pangenomes have given way to a deeper understanding of functional traits that have been lost or were previously unknown in crop species. The study of pangenomics has also been used to reveal traits in wild crop species that can potentially be integrated into new or existing crops. Knowledge gained from pangenomic studies has already increased the effectiveness of the CRISPR-Cas tool to develop highly specific CRISPR target sequences, allowing for precise alteration of genome content and gene expression. While use of CRISPR-Cas systems guided by pangenomic studies is still limited by the size of catalytic windows needed for modification and the scale of SVs and PAVs linked to specific traits, pangenome construction technology is steadily advancing along with improvements in the accuracy of genome sequencing, annotation, and affordability of these processes. In the future, pangenomes will be key for thorough and effective design of genome editing experiments in crop varieties to increase global food security and resilience to climate change.

Author Contributions

C.G.T.F., B.J.N., J.I.M. and M.F.D. wrote and edited this manuscript. B.J.N. and M.F.D. constructed the figures. J.P., P.E.B., D.E. and J.B. reviewed and edited this manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work is funded by the Australia Research Council (Projects DP210100296, DP200100762, and DE210100398) and the Grains Research and Development Corporation (Projects 9177539 and 9177591). This work was supported by resources provided by the Pawsey Supercomputing Centre with funding from the Australian Government and the Government of Western Australia.

Acknowledgments

Cassandria G. Tay Fernandez, Benjamin J. Nestor, and Monica F. Danilevicz are supported by Research Training Program scholarships. Benjamin J. Nestor is supported by a university postgraduate award at The University of Western Australia. Monica F. Danilevicz receives further support from the Forrest Research Foundation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roser, M. Future Population Growth. Available online: https://ourworldindata.org/future-population-growth (accessed on 11 June 2021).
  2. Masson-Delmotte, V.; Zhai, P.; Pirani, A.; Connors, S.L.; Péan, C.; Berger, S.; Caud, N.; Chen, Y.; Goldfarb, L.; Gomis, M.I.; et al. Climate Change 2021: The Physical Science Basis. In Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, MA, USA, 2021. [Google Scholar]
  3. Abberton, M.; Batley, J.; Bentley, A.; Bryant, J.; Cai, H.; Cockram, J.; de Oliveira, A.C.; Cseke, L.J.; Dempewolf, H.; De Pace, C.; et al. Global agricultural intensification during climate change: A role for genomics. Plant Biotechnol. J. 2016, 14, 1095–1098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Anderson, R.; Bayer, P.E.; Edwards, D. Climate change and the need for agricultural adaptation. Curr. Opin. Plant Biol. 2020, 56, 197–202. [Google Scholar] [CrossRef] [PubMed]
  5. Cai, C.Q.; Doyon, Y.; Ainley, W.M.; Miller, J.C.; Dekelver, R.C.; Moehle, E.A.; Rock, J.M.; Lee, Y.L.; Garrison, R.; Schulenberg, L.; et al. Targeted transgene integration in plant cells using designed zinc finger nucleases. Plant Mol. Biol. 2009, 69, 699–709. [Google Scholar] [CrossRef] [PubMed]
  6. De Pater, S.; Neuteboom, L.W.; Pinas, J.E.; Hooykaas, P.J.; van der Zaal, B.J. ZFN-induced mutagenesis and gene-targeting in Arabidopsis through Agrobacterium-mediated floral dip transformation. Plant Biotechnol. J. 2009, 7, 821–835. [Google Scholar] [CrossRef] [PubMed]
  7. Lloyd, A.; Plaisier, C.L.; Carroll, D.; Drews, G.N. Targeted mutagenesis using zinc-finger nucleases in Arabidopsis. Proc. Natl. Acad. Sci. USA 2005, 102, 2232–2237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Zhang, Y.; Zhang, F.; Li, X.; Baller, J.A.; Qi, Y.; Starker, C.G.; Bogdanove, A.J.; Voytas, D.F. Transcription activator-like effector nucleases enable efficient plant genome engineering. Plant Physiol. 2013, 161, 20–27. [Google Scholar] [CrossRef] [Green Version]
  9. Forsyth, A.; Weeks, T.; Richael, C.; Duan, H. Transcription Activator-Like Effector Nucleases (TALEN)-Mediated Targeted DNA Insertion in Potato Plants. Front. Plant Sci. 2016, 7, 1572. [Google Scholar] [CrossRef] [Green Version]
  10. Li, T.; Liu, B.; Spalding, M.H.; Weeks, D.P.; Yang, B. High-efficiency TALEN-based gene editing produces disease-resistant rice. Nat. Biotechnol. 2012, 30, 390–392. [Google Scholar] [CrossRef]
  11. Cong, L.; Ran, F.A.; Cox, D.; Lin, S.; Barretto, R.; Habib, N.; Hsu, P.D.; Wu, X.; Jiang, W.; Marraffini, L.A.; et al. Multiplex Genome Engineering Using CRISPR/Cas Systems. Science 2013, 339, 819. [Google Scholar] [CrossRef] [Green Version]
  12. Scheben, A.; Edwards, D. Genome editors take on crops. Science 2017, 355, 1122. [Google Scholar] [CrossRef]
  13. Scheben, A.; Wolter, F.; Batley, J.; Puchta, H.; Edwards, D. Towards CRISPR/Cas crops—Bringing together genomics and genome editing. New Phytologist. 2017, 216, 682–698. [Google Scholar] [CrossRef] [Green Version]
  14. Shi, J.; Gao, H.; Wang, H.; Lafitte, H.R.; Archibald, R.L.; Yang, M.; Hakimi, S.M.; Mo, H.; Habben, J.E. ARGOS8 variants generated by CRISPR-Cas9 improve maize grain yield under field drought stress conditions. Plant Biotechnol. J. 2017, 15, 207–216. [Google Scholar] [CrossRef] [Green Version]
  15. Wang, T.; Zhang, H.; Zhu, H. CRISPR technology is revolutionizing the improvement of tomato and other fruit crops. Hortic. Res. 2019, 6, 77. [Google Scholar] [CrossRef] [Green Version]
  16. Zeng, Y.; Wen, J.; Zhao, W.; Wang, Q.; Huang, W. Rational Improvement of Rice Yield and Cold Tolerance by Editing the Three Genes OsPIN5b, GS3, and OsMYB30 With the CRISPR–Cas9 System. Front. Plant Sci. 2020, 10, 1663. [Google Scholar] [CrossRef] [Green Version]
  17. Bayer, P.E.; Golicz, A.A.; Scheben, A.; Batley, J.; Edwards, D. Plant pan-genomes are the new reference. Nat. Plants 2020, 6, 914–920. [Google Scholar] [CrossRef]
  18. 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]
  19. Gabriel, R.; von Kalle, C.; Schmidt, M. Mapping the precision of genome editing. Nat. Biotechnol. 2015, 33, 150–152. [Google Scholar] [CrossRef]
  20. Doebley, J.F.; Gaut, B.S.; Smith, B.D. The Molecular Genetics of Crop Domestication. Cell 2006, 127, 1309–1321. [Google Scholar] [CrossRef] [Green Version]
  21. 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] [Green Version]
  22. The Rice, C.; Sequencing, C. The sequence of rice chromosomes 11 and 12, rich in disease resistance genes and recent gene duplications. BMC Biol. 2005, 3, 20. [Google Scholar] [CrossRef] [Green Version]
  23. Woodhouse, M.R.; Schnable, J.C.; Pedersen, B.S.; Lyons, E.; Lisch, D.; Subramaniam, S.; Freeling, M. Following tetraploidy in maize, a short deletion mechanism removed genes preferentially from one of the two homologs. PLoS Biol. 2010, 8, e1000409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Zhao, D.; Ferguson, A.A.; Jiang, N. What makes up plant genomes: The vanishing line between transposable elements and genes. Biochim. Biophys. Acta (BBA) Gene Regul. Mech. 2016, 1859, 366–380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Ou, L.; Li, D.; Lv, J.; Chen, W.; Zhang, Z.; Li, X.; Yang, B.; Zhou, S.; Yang, S.; Li, W.; et al. Pan-genome of cultivated pepper (Capsicum) and its use in gene presence–absence variation analyses. New Phytol. 2018, 220, 360–363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Schatz, M.C.; Maron, L.G.; Stein, J.C.; Wences, A.H.; Gurtowski, J.; Biggers, E.; Lee, H.; Kramer, M.; Antoniou, E.; Ghiban, E.; et al. 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] [CrossRef] [Green Version]
  27. Song, J.-M.; Guan, Z.; Hu, J.; Guo, C.; Yang, Z.; Wang, S.; Liu, D.; Wang, B.; Lu, S.; Zhou, R.; et al. Eight high-quality genomes reveal pan-genome architecture and ecotype differentiation of Brassica napus. Nat. Plants 2020, 6, 34–45. [Google Scholar] [CrossRef]
  28. Hurgobin, B.; Golicz, A.A.; Bayer, P.E.; Chan, C.-K.K.; Tirnaz, S.; Dolatabadian, A.; Schiessl, S.V.; Samans, B.; Montenegro, J.D.; Parkin, I.A.P.; et al. Homoeologous exchange is a major cause of gene presence/absence variation in the amphidiploid Brassica napus. Plant Biotechnol. J. 2018, 16, 1265–1274. [Google Scholar] [CrossRef] [Green Version]
  29. Zhao, J.; Bayer, P.E.; Ruperao, P.; Saxena, R.K.; Khan, A.W.; Golicz, A.A.; Nguyen, H.T.; Batley, J.; Edwards, D.; Varshney, R.K. Trait associations in the pangenome of pigeon pea (Cajanus cajan). Plant Biotechnol. J. 2020, 18, 1946–1954. [Google Scholar] [CrossRef] [Green Version]
  30. Malzahn, A.; Lowder, L.; Qi, Y. Plant genome editing with TALEN and CRISPR. Cell Biosci. 2017, 7, 21. [Google Scholar] [CrossRef] [Green Version]
  31. Bhardwaj, A.; Nain, V. TALENs—an indispensable tool in the era of CRISPR: A mini review. J. Genet. Eng. Biotechnol. 2021, 19, 125. [Google Scholar] [CrossRef]
  32. Alok, A.; Sandhya, D.; Jogam, P.; Rodrigues, V.; Bhati, K.K.; Sharma, H.; Kumar, J. The Rise of the CRISPR/Cpf1 System for Efficient Genome Editing in Plants. Front. Plant Sci. 2020, 11, 264. [Google Scholar] [CrossRef]
  33. Gaj, T.; Gersbach, C.A.; Barbas, C.F. ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering. Trends Biotechnol. 2013, 31, 397–405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Tettelin, H.; Masignani, V.; Cieslewicz, M.J.; Donati, C.; Medini, D.; Ward, N.L.; Angiuoli, S.V.; Crabtree, J.; Jones, A.L.; Durkin, A.S.; et al. Genome analysis of multiple pathogenic isolates of Streptococcus agalactiae: Implications for the microbial “pan-genome”. Proc. Natl. Acad. Sci. USA 2005, 102, 13950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Gao, L.; Gonda, I.; Sun, H.; Ma, Q.; Bao, K.; Tieman, D.M.; Burzynski-Chang, E.A.; Fish, T.L.; Stromberg, K.A.; Sacks, G.L.; et al. The tomato pan-genome uncovers new genes and a rare allele regulating fruit flavor. Nat. Genet. 2019, 51, 1044–1051. [Google Scholar] [CrossRef] [PubMed]
  36. Ruperao, P.; Thirunavukkarasu, N.; Gandham, P.; Selvanayagam, S.; Govindaraj, M.; Nebie, B.; Manyasa, E.; Gupta, R.; Das, R.R.; Gandhi, H.; et al. Sorghum pan-genome explores the functional utility to accelerate the genetic gain. Front. Plant Sci. 2021, 12, 963. [Google Scholar] [CrossRef]
  37. Liu, H.-J.; Jian, L.; Xu, J.; Zhang, Q.; Zhang, M.; Jin, M.; Peng, Y.; Yan, J.; Han, B.; Liu, J.; et al. High-Throughput CRISPR/Cas9 Mutagenesis Streamlines Trait Gene Identification in Maize[OPEN]. Plant Cell 2020, 32, 1397–1413. [Google Scholar] [CrossRef] [Green Version]
  38. Ossowski, S.; Schneeberger, K.; Clark, R.M.; Lanz, C.; Warthmann, N.; Weigel, D. Sequencing of natural strains of Arabidopsis thaliana with short reads. Genome Res. 2008, 18, 2024–2033. [Google Scholar] [CrossRef] [Green Version]
  39. Golicz, A.A.; Bayer, P.E.; Barker, G.C.; Edger, P.P.; Kim, H.; Martinez, P.A.; Chan, C.K.K.; Severn-Ellis, A.; McCombie, W.R.; Parkin, I.A.P.; et al. The pangenome of an agronomically important crop plant Brassica oleracea. Nat. Commun. 2016, 7, 13390. [Google Scholar] [CrossRef]
  40. Della Coletta, R.; Qiu, Y.; Ou, S.; Hufford, M.B.; Hirsch, C.N. How the pan-genome is changing crop genomics and improvement. Genome Biol. 2021, 22, 3. [Google Scholar] [CrossRef]
  41. Sirén, J.; Garrison, E.; Novak, A.M.; Paten, B.; Durbin, R. Haplotype-aware graph indexes. Bioinformatics 2020, 36, 400–407. [Google Scholar] [CrossRef]
  42. Garrison, E.; Sirén, J.; Novak, A.M.; Hickey, G.; Eizenga, J.M.; Dawson, E.T.; Jones, W.; Garg, S.; Markello, C.; Lin, M.F.; et al. Variation graph toolkit improves read mapping by representing genetic variation in the reference. Nat. Biotechnol. 2018, 36, 875–879. [Google Scholar] [CrossRef]
  43. Rakocevic, G.; Semenyuk, V.; Lee, W.-P.; Spencer, J.; Browning, J.; Johnson, I.J.; Arsenijevic, V.; Nadj, J.; Ghose, K.; Suciu, M.C.; et al. Fast and accurate genomic analyses using genome graphs. Nat. Genet. 2019, 51, 354–362. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, K.; Hu, H.; Tian, Y.; Li, J.; Scheben, A.; Zhang, C.; Li, Y.; Wu, J.; Yang, L.; Fan, X.; et al. The Chicken Pan-Genome Reveals Gene Content Variation and a Promoter Region Deletion in IGF2BP1 Affecting Body Size. Mol. Biol. Evol. 2021, 38, 5066–5081. [Google Scholar] [CrossRef] [PubMed]
  45. Lander, E.S.; Linton, L.M.; Birren, B.; Nusbaum, C.; Zody, M.C.; Baldwin, J.; Devon, K.; Dewar, K.; Doyle, M.; FitzHugh, W.; et al. Initial sequencing and analysis of the human genome. Nature 2001, 409, 860–921. [Google Scholar] [CrossRef] [Green Version]
  46. Gerdol, M.; Moreira, R.; Cruz, F.; Gómez-Garrido, J.; Vlasova, A.; Rosani, U.; Venier, P.; Naranjo-Ortiz, M.A.; Murgarella, M.; Greco, S.; et al. Massive gene presence-absence variation shapes an open pan-genome in the Mediterranean mussel. Genome Biol. 2020, 21, 275. [Google Scholar] [CrossRef]
  47. Li, R.; Li, Y.; Zheng, H.; Luo, R.; Zhu, H.; Li, Q.; Qian, W.; Ren, Y.; Tian, G.; Li, J.; et al. Building the sequence map of the human pan-genome. Nat. Biotechnol. 2010, 28, 57–63. [Google Scholar] [CrossRef]
  48. Morgante, M.; De Paoli, E.; Radovic, S. Transposable elements and the plant pan-genomes. Curr. Opin. Plant Biol. 2007, 10, 149–155. [Google Scholar] [CrossRef]
  49. Li, Y.-h.; Zhou, G.; Ma, J.; Jiang, W.; Jin, L.-g.; Zhang, Z.; Guo, Y.; Zhang, J.; Sui, Y.; Zheng, L.; et al. 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] [Green Version]
  50. Liu, Y.; Du, H.; Li, P.; Shen, Y.; Peng, H.; Liu, S.; Zhou, G.-A.; Zhang, H.; Liu, Z.; Shi, M.; et al. Pan-Genome of Wild and Cultivated Soybeans. Cell 2020, 182, 162–176.e13. [Google Scholar] [CrossRef]
  51. Lu, F.; Romay, M.C.; Glaubitz, J.C.; Bradbury, P.J.; Elshire, R.J.; Wang, T.; Li, Y.; Li, Y.; Semagn, K.; Zhang, X.; et al. High-resolution genetic mapping of maize pan-genome sequence anchors. Nat. Commun. 2015, 6, 6914. [Google Scholar] [CrossRef] [Green Version]
  52. Dolatabadian, A.; Bayer, P.E.; Tirnaz, S.; Hurgobin, B.; Edwards, D.; Batley, J. Characterization of disease resistance genes in the Brassica napus pangenome reveals significant structural variation. Plant Biotechnol. J. 2020, 18, 969–982. [Google Scholar] [CrossRef] [Green Version]
  53. Gordon, S.P.; Contreras-Moreira, B.; Woods, D.P.; Des Marais, D.L.; Burgess, D.; Shu, S.; Stritt, C.; Roulin, A.C.; Schackwitz, W.; Tyler, L.; et al. Extensive gene content variation in the Brachypodium distachyon pan-genome correlates with population structure. Nat. Commun. 2017, 8, 2184. [Google Scholar] [CrossRef]
  54. Jayakodi, M.; Padmarasu, S.; Haberer, G.; Bonthala, V.S.; Gundlach, H.; Monat, C.; Lux, T.; Kamal, N.; Lang, D.; Himmelbach, A.; et al. The barley pan-genome reveals the hidden legacy of mutation breeding. Nature 2020, 588, 284–289. [Google Scholar] [CrossRef]
  55. Zhao, Q.; Feng, Q.; Lu, H.; Li, Y.; Wang, A.; Tian, Q.; Zhan, Q.; Lu, Y.; Zhang, L.; Huang, T.; et al. 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]
  56. Varshney, R.K.; Roorkiwal, M.; Sun, S.; Bajaj, P.; Chitikineni, A.; Thudi, M.; Singh, N.P.; Du, X.; Upadhyaya, H.D.; Khan, A.W.; et al. A chickpea genetic variation map based on the sequencing of 3366 genomes. Nature 2021, 599, 622–627. [Google Scholar] [CrossRef]
  57. Sun, X.; Jiao, C.; Schwaninger, H.; Chao, C.T.; Ma, Y.; Duan, N.; Khan, A.; Ban, S.; Xu, K.; Cheng, L.; et al. Phased diploid genome assemblies and pan-genomes provide insights into the genetic history of apple domestication. Nat. Genet. 2020, 52, 1423–1432. [Google Scholar] [CrossRef]
  58. Yu, J.; Golicz, A.A.; Lu, K.; Dossa, K.; Zhang, Y.; Chen, J.; Wang, L.; You, J.; Fan, D.; Edwards, D.; et al. Insight into the evolution and functional characteristics of the pan-genome assembly from sesame landraces and modern cultivars. Plant Biotechnol. J. 2019, 17, 881–892. [Google Scholar] [CrossRef] [Green Version]
  59. Hübner, S.; Bercovich, N.; Todesco, M.; Mandel, J.R.; Odenheimer, J.; Ziegler, E.; Lee, J.S.; Baute, G.J.; Owens, G.L.; Grassa, C.J.; et al. Sunflower pan-genome analysis shows that hybridization altered gene content and disease resistance. Nat. Plants 2019, 5, 54–62. [Google Scholar] [CrossRef]
  60. Long, E.; Bradbury, P.; Romay, M.; Buckler, E.; Robbins, K. Genome-wide Imputation Using the Practical Haplotype Graph in the Heterozygous Crop Cassava. Genes Genomes Genet. 2021, 12, jkab383. [Google Scholar] [CrossRef]
  61. Jensen, S.; Charles, J.R.; Muleta, K.; Bradbury, P.; Casstevens, T.; Deshpande, S.P.; Gore, M.A.; Gupta, R.; Ilut, D.C.; Johnson, L.; et al. A sorghum Practical Haplotype Graph facilitates genome-wide imputation and cost-effective genomic prediction. Plant Genome 2020, 13, e20009. [Google Scholar] [CrossRef] [Green Version]
  62. 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]
  63. Jiao, W.-B.; Schneeberger, K. Chromosome-level assemblies of multiple Arabidopsis genomes reveal hotspots of rearrangements with altered evolutionary dynamics. Nat. Commun. 2020, 11, 989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Hu, H.; Scheben, A.; Verpaalen, B.; Tirnaz, S.; Bayer, P.E.; Hodel, R.G.J.; Batley, J.; Soltis, D.E.; Soltis, P.S.; Edwards, D. Amborella gene presence/absence variation is associated with abiotic stress responses that may contribute to environmental adaptation. New Phytol. 2021, 233, 1548–1555. [Google Scholar] [CrossRef]
  65. Li, J.; Yuan, D.; Wang, P.; Wang, Q.; Sun, M.; Liu, Z.; Si, H.; Xu, Z.; Ma, Y.; Zhang, B.; et al. Cotton pan-genome retrieves the lost sequences and genes during domestication and selection. Genome Biol. 2021, 22, 119. [Google Scholar] [CrossRef]
  66. Zhou, P.; Silverstein, K.A.T.; Ramaraj, T.; Guhlin, J.; Denny, R.; Liu, J.; Farmer, A.D.; Steele, K.P.; Stupar, R.M.; Miller, J.R.; et al. Exploring structural variation and gene family architecture with De Novo assemblies of 15 Medicago genomes. BMC Genom. 2017, 18, 261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Bayer, P.E.; Golicz, A.A.; Tirnaz, S.; Chan, C.-K.K.; Edwards, D.; Batley, J. Variation in abundance of predicted resistance genes in the Brassica oleracea pangenome. Plant Biotechnol. J. 2019, 17, 789–800. [Google Scholar] [CrossRef] [Green Version]
  68. Bayer, P.E.; Valliyodan, B.; Hu, H.; Marsh, J.I.; Yuan, Y.; Vuong, T.D.; Patil, G.; Song, Q.; Batley, J.; Varshney, R.K.; et al. Sequencing the USDA core soybean collection reveals gene loss during domestication and breeding. Plant Genome 2021, e20109. [Google Scholar] [CrossRef] [PubMed]
  69. Ali, Z.; Mahfouz, M.M.; Mansoor, S. CRISPR-TSKO: A Tool for Tissue-Specific Genome Editing in Plants. Trends Plant Sci. 2020, 25, 123–126. [Google Scholar] [CrossRef]
  70. Li, Q.; Wu, G.; Zhao, Y.; Wang, B.; Zhao, B.; Kong, D.; Wei, H.; Chen, C.; Wang, H. CRISPR/Cas9-mediated knockout and overexpression studies reveal a role of maize phytochrome C in regulating flowering time and plant height. Plant Biotechnol. J. 2020, 18, 2520–2532. [Google Scholar] [CrossRef]
  71. Ueta, R.; Abe, C.; Watanabe, T.; Sugano, S.S.; Ishihara, R.; Ezura, H.; Osakabe, Y.; Osakabe, K. Rapid breeding of parthenocarpic tomato plants using CRISPR/Cas9. Sci. Rep. 2017, 7, 507. [Google Scholar] [CrossRef] [Green Version]
  72. Li, X.; Wang, Y.; Chen, S.; Tian, H.; Fu, D.; Zhu, B.; Luo, Y.; Zhu, H. Lycopene Is Enriched in Tomato Fruit by CRISPR/Cas9-Mediated Multiplex Genome Editing. Front. Plant Sci. 2018, 9, 559. [Google Scholar] [CrossRef]
  73. Yu, Q.-h.; Wang, B.; Li, N.; Tang, Y.; Yang, S.; Yang, T.; Xu, J.; Guo, C.; Yan, P.; Wang, Q.; et al. CRISPR/Cas9-induced Targeted Mutagenesis and Gene Replacement to Generate Long-shelf Life Tomato Lines. Sci. Rep. 2017, 7, 11874. [Google Scholar] [CrossRef]
  74. Zhang, N.; Roberts, H.M.; Van Eck, J.; Martin, G.B. Generation and Molecular Characterization of CRISPR/Cas9-Induced Mutations in 63 Immunity-Associated Genes in Tomato Reveals Specificity and a Range of Gene Modifications. Front. Plant Sci. 2020, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Decaestecker, W.; Buono, R.A.; Pfeiffer, M.L.; Vangheluwe, N.; Jourquin, J.; Karimi, M.; Van Isterdael, G.; Beeckman, T.; Nowack, M.K.; Jacobs, T.B. CRISPR-TSKO: A Technique for Efficient Mutagenesis in Specific Cell Types, Tissues, or Organs in Arabidopsis. Plant Cell 2019, 31, 2868–2887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Ma, X.; Zhang, Q.; Zhu, Q.; Liu, W.; Chen, Y.; Qiu, R.; Wang, B.; Yang, Z.; Li, H.; Lin, Y.; et al. A Robust CRISPR/Cas9 System for Convenient, High-Efficiency Multiplex Genome Editing in Monocot and Dicot Plants. Mol. Plant 2015, 8, 1274–1284. [Google Scholar] [CrossRef]
  77. Borrill, P.; Harrington, S.A.; Uauy, C. Applying the latest advances in genomics and phenomics for trait discovery in polyploid wheat. Plant J. 2019, 97, 56–72. [Google Scholar] [CrossRef] [PubMed]
  78. Li, Q.; Sapkota, M.; van der Knaap, E. Perspectives of CRISPR/Cas-mediated cis-engineering in horticulture: Unlocking the neglected potential for crop improvement. Hortic. Res. 2020, 7, 36. [Google Scholar] [CrossRef] [Green Version]
  79. Gruner, K.; Esser, T.; Acevedo-Garcia, J.; Freh, M.; Habig, M.; Strugala, R.; Stukenbrock, E.; Schaffrath, U.; Panstruga, R. Evidence for Allele-Specific Levels of Enhanced Susceptibility of Wheat mlo Mutants to the Hemibiotrophic Fungal Pathogen Magnaporthe oryzae pv. Triticum. Genes 2020, 11, 517. [Google Scholar] [CrossRef]
  80. Consonni, C.; Humphry, M.E.; Hartmann, H.A.; Livaja, M.; Durner, J.; Westphal, L.; Vogel, J.; Lipka, V.; Kemmerling, B.; Schulze-Lefert, P.; et al. Conserved requirement for a plant host cell protein in powdery mildew pathogenesis. Nat. Genet. 2006, 38, 716–720. [Google Scholar] [CrossRef]
  81. Tian, D.; Chen, Z.; Chen, Z.; Zhou, Y.; Wang, Z.; Wang, F.; Chen, S. Allele-specific marker-based assessment revealed that the rice blast resistance genes Pi2 and Pi9 have not been widely deployed in Chinese indica rice cultivars. Rice 2016, 9, 19. [Google Scholar] [CrossRef] [Green Version]
  82. Chai, L.; Zhang, J.; Fernando, W.G.D.; Li, H.; Huang, X.; Cui, C.; Jiang, J.; Zheng, B.; Liu, Y.; Jiang, L. Detection of Blackleg Resistance Gene Rlm1 in Double-Low Rapeseed Accessions from Sichuan Province, by Kompetitive Allele-Specific PCR. Plant Pathol. J. 2021, 37, 194–199. [Google Scholar] [CrossRef]
  83. Van Bezouw, R.F.H.M.; Janssen, E.M.; Ashrafuzzaman, M.; Ghahramanzadeh, R.; Kilian, B.; Graner, A.; Visser, R.G.F.; van der Linden, C.G. Shoot sodium exclusion in salt stressed barley (Hordeum vulgare L.) is determined by allele specific increased expression of HKT1;5. J. Plant Physiol. 2019, 241, 153029. [Google Scholar] [CrossRef] [PubMed]
  84. Cai, M.; Lin, J.; Li, Z.; Lin, Z.; Ma, Y.; Wang, Y.; Ming, R. Allele specific expression of Dof genes responding to hormones and abiotic stresses in sugarcane. PLoS ONE 2020, 15, e0227716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Giacomini, D.A.; Patterson, E.L.; Küpper, A.; Beffa, R.; Gaines, T.A.; Tranel, P.J. Coexpression Clusters and Allele-Specific Expression in Metabolism-Based Herbicide Resistance. Genome Biol. Evol. 2020, 12, 2267–2278. [Google Scholar] [CrossRef] [PubMed]
  86. Li, F.; Wen, W.; Liu, J.; Zhang, Y.; Cao, S.; He, Z.; Rasheed, A.; Jin, H.; Zhang, C.; Yan, J.; et al. Genetic architecture of grain yield in bread wheat based on genome-wide association studies. BMC Plant Biol. 2019, 19, 168. [Google Scholar] [CrossRef]
  87. Morineau, C.; Bellec, Y.; Tellier, F.; Gissot, L.; Kelemen, Z.; Nogué, F.; Faure, J.-D. Selective gene dosage by CRISPR-Cas9 genome editing in hexaploid Camelina sativa. Plant Biotechnol. J. 2017, 15, 729–739. [Google Scholar] [CrossRef] [Green Version]
  88. Schaart, J.G.; van de Wiel, C.C.M.; Smulders, M.J.M. Genome editing of polyploid crops: Prospects, achievements and bottlenecks. Transgenic Res. 2021, 30, 337–351. [Google Scholar] [CrossRef]
  89. Zaman, Q.U.; Li, C.; Cheng, H.; Hu, Q. Genome editing opens a new era of genetic improvement in polyploid crops. Crop J. 2019, 7, 141–150. [Google Scholar] [CrossRef]
  90. Schwartz, C.; Lenderts, B.; Feigenbutz, L.; Barone, P.; Llaca, V.; Fengler, K.; Svitashev, S. CRISPR–Cas9-mediated 75.5-Mb inversion in maize. Nat. Plants 2020, 6, 1427–1431. [Google Scholar] [CrossRef]
  91. Huang, J.; Li, J.; Zhou, J.; Wang, L.; Yang, S.; Hurst, L.D.; Li, W.-H.; Tian, D. Identifying a large number of high-yield genes in rice by pedigree analysis, whole-genome sequencing, and CRISPR-Cas9 gene knockout. Proc. Natl. Acad. Sci. USA 2018, 115, E7559. [Google Scholar] [CrossRef] [Green Version]
  92. Matchett-Oates, L.; Braich, S.; Spangenberg, G.C.; Rochfort, S.; Cogan, N.O.I. In silico analysis enabling informed design for genome editing in medicinal cannabis; gene families and variant characterisation. PLoS ONE 2021, 16, e0257413. [Google Scholar] [CrossRef]
  93. Tang, S.; Zhao, H.; Lu, S.; Yu, L.; Zhang, G.; Zhang, Y.; Yang, Q.-Y.; Zhou, Y.; Wang, X.; Ma, W.; et al. Genome- and transcriptome-wide association studies provide insights into the genetic basis of natural variation of seed oil content in Brassica napus. Mol. Plant 2021, 14, 470–487. [Google Scholar] [CrossRef] [PubMed]
  94. Martina, M.; Tikunov, Y.; Portis, E.; Bovy, A.G. The Genetic Basis of Tomato Aroma. Genes 2021, 12, 226. [Google Scholar] [CrossRef] [PubMed]
  95. Hirschhorn, J.N.; Daly, M.J. Genome-wide association studies for common diseases and complex traits. Nat. Rev. Genet. 2005, 6, 95–108. [Google Scholar] [CrossRef] [PubMed]
  96. Yao, W.; Li, G.; Zhao, H.; Wang, G.; Lian, X.; Xie, W. Exploring the rice dispensable genome using a metagenome-like assembly strategy. Genome Biol. 2015, 16, 187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Xing, H.-L.; Dong, L.; Wang, Z.-P.; Zhang, H.-Y.; Han, C.-Y.; Liu, B.; Wang, X.-C.; Chen, Q.-J. A CRISPR/Cas9 toolkit for multiplex genome editing in plants. BMC Plant Biol. 2014, 14, 327. [Google Scholar] [CrossRef] [Green Version]
  98. Lu, Y.; Ye, X.; Guo, R.; Huang, J.; Wang, W.; Tang, J.; Tan, L.; Zhu, J.K.; Chu, C.; Qian, Y. Genome-wide Targeted Mutagenesis in Rice Using the CRISPR/Cas9 System. Mol. Plant 2017, 10, 1242–1245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Meng, X.; Yu, H.; Zhang, Y.; Zhuang, F.; Song, X.; Gao, S.; Gao, C.; Li, J. Construction of a Genome-Wide Mutant Library in Rice Using CRISPR/Cas9. Mol. Plant 2017, 10, 1238–1241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Li, C.; Hao, M.; Wang, W.; Wang, H.; Chen, F.; Chu, W.; Zhang, B.; Mei, D.; Cheng, H.; Hu, Q. An Efficient CRISPR/Cas9 Platform for Rapidly Generating Simultaneous Mutagenesis of Multiple Gene Homoeologs in Allotetraploid Oilseed Rape. Front. Plant Sci. 2018, 9, 442. [Google Scholar] [CrossRef]
  101. Bai, M.; Yuan, J.; Kuang, H.; Gong, P.; Li, S.; Zhang, Z.; Liu, B.; Sun, J.; Yang, M.; Yang, L.; et al. Generation of a multiplex mutagenesis population via pooled CRISPR-Cas9 in soya bean. Plant Biotechnol. J. 2020, 18, 721–731. [Google Scholar] [CrossRef] [Green Version]
  102. Findlay, G.M.; Daza, R.M.; Martin, B.; Zhang, M.D.; Leith, A.P.; Gasperini, M.; Janizek, J.D.; Huang, X.; Starita, L.M.; Shendure, J. Accurate classification of BRCA1 variants with saturation genome editing. Nature 2018, 562, 217–222. [Google Scholar] [CrossRef]
  103. Hashimoto, R.; Ueta, R.; Abe, C.; Osakabe, Y.; Osakabe, K. Efficient Multiplex Genome Editing Induces Precise, and Self-Ligated Type Mutations in Tomato Plants. Front. Plant Sci. 2018, 9, 916. [Google Scholar] [CrossRef] [PubMed]
  104. Qian, Y.; Huang, H.-H.; Jimenez, J.; Del Vecchio, D. Resource Competition Shapes the Response of Genetic Circuits. ACS Synth. Biol. 2017, 6, 1263–1272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zhang, S.; Voigt, C.A. Engineered dCas9 with reduced toxicity in bacteria: Implications for genetic circuit design. Nucleic Acids Res. 2018, 46, 11115–11125. [Google Scholar] [CrossRef] [Green Version]
  106. McCarty, N.S.; Graham, A.E.; Studená, L.; Ledesma-Amaro, R. Multiplexed CRISPR technologies for gene editing and transcriptional regulation. Nat. Commun. 2020, 11, 1281. [Google Scholar] [CrossRef] [PubMed]
  107. Zhang, X.-H.; Tee, L.Y.; Wang, X.-G.; Huang, Q.-S.; Yang, S.-H. Off-target Effects in CRISPR/Cas9-mediated Genome Engineering. Mol. Ther. Nucleic Acids 2015, 4, e264. [Google Scholar] [CrossRef]
  108. Xu, K.; Xu, X.; Fukao, T.; Canlas, P.; Maghirang-Rodriguez, R.; Heuer, S.; Ismail, A.M.; Bailey-Serres, J.; Ronald, P.C.; Mackill, D.J. Sub1A is an ethylene-response-factor-like gene that confers submergence tolerance to rice. Nature 2006, 442, 705–708. [Google Scholar] [CrossRef] [Green Version]
  109. Wang, X.; Tu, M.; Wang, Y.; Yin, W.; Zhang, Y.; Wu, H.; Gu, Y.; Li, Z.; Xi, Z.; Wang, X. Whole-genome sequencing reveals rare off-target mutations in CRISPR/Cas9-edited grapevine. Hortic. Res. 2021, 8, 114. [Google Scholar] [CrossRef]
  110. Li, J.; Manghwar, H.; Sun, L.; Wang, P.; Wang, G.; Sheng, H.; Zhang, J.; Liu, H.; Qin, L.; Rui, H.; et al. Whole genome sequencing reveals rare off-target mutations and considerable inherent genetic or/and somaclonal variations in CRISPR/Cas9-edited cotton plants. Plant Biotechnol. J. 2019, 17, 858–868. [Google Scholar] [CrossRef]
  111. Grohmann, L.; Keilwagen, J.; Duensing, N.; Dagand, E.; Hartung, F.; Wilhelm, R.; Bendiek, J.; Sprink, T. Detection and Identification of Genome Editing in Plants: Challenges and Opportunities. Front. Plant Sci. 2019, 10, 236. [Google Scholar] [CrossRef] [Green Version]
  112. Golicz, A.A.; Batley, J.; Edwards, D. Towards plant pangenomics. Plant Biotechnol. J. 2016, 14, 1099–1105. [Google Scholar] [CrossRef]
  113. Schouten, H.J.; Tikunov, Y.; Verkerke, W.; Finkers, R.; Bovy, A.; Bai, Y.; Visser, R.G.F. Breeding Has Increased the Diversity of Cultivated Tomato in The Netherlands. Front. Plant Sci. 2019, 10, 1606. [Google Scholar] [CrossRef] [PubMed]
  114. Zhou, Z.; Jiang, Y.; Wang, Z.; Gou, Z.; Lyu, J.; Li, W.; Yu, Y.; Shu, L.; Zhao, Y.; Ma, Y.; et al. Resequencing 302 wild and cultivated accessions identifies genes related to domestication and improvement in soybean. Nat. Biotechnol. 2015, 33, 408–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Warschefsky, E.; Penmetsa, R.V.; Cook, D.R.; von Wettberg, E.J.B. Back to the wilds: Tapping evolutionary adaptations for resilient crops through systematic hybridization with crop wild relatives. Am. J. Bot. 2014, 101, 1791–1800. [Google Scholar] [CrossRef] [PubMed]
  116. Krattinger, S.G.; Keller, B. Molecular genetics and evolution of disease resistance in cereals. New Phytol. 2016, 212, 320–332. [Google Scholar] [CrossRef] [Green Version]
  117. Sakai, H.; Itoh, T. Massive gene losses in Asian cultivated rice unveiled by comparative genome analysis. BMC Genom. 2010, 11, 121. [Google Scholar] [CrossRef] [Green Version]
  118. Mbuvi, D.; Masiga, C.; Kuria, E.; Masanga, J.; Wamalwa, M.; Mohamed, A.; Odeny, D.; Hamza, N.; Timko, M.; Runo, S. Novel Sources of Witchweed (Striga) Resistance from Wild Sorghum Accessions. Front. Plant Sci. 2017, 8, 116. [Google Scholar] [CrossRef] [Green Version]
  119. Yang, J.; Mezmouk, S.; Baumgarten, A.; Buckler, E.S.; Guill, K.E.; McMullen, M.D.; Mumm, R.H.; Ross-Ibarra, J. Incomplete dominance of deleterious alleles contributes substantially to trait variation and heterosis in maize. PLoS Genet. 2017, 13, e1007019. [Google Scholar] [CrossRef] [Green Version]
  120. Soyk, S.; Müller, N.A.; Park, S.J.; Schmalenbach, I.; Jiang, K.; Hayama, R.; Zhang, L.; Van Eck, J.; Jiménez-Gómez, J.M.; Lippman, Z.B. Variation in the flowering gene SELF PRUNING 5G promotes day-neutrality and early yield in tomato. Nat. Genet. 2017, 49, 162–168. [Google Scholar] [CrossRef]
  121. 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]
  122. Scheben, A.; Edwards, D. Bottlenecks for genome-edited crops on the road from lab to farm. Genome Biol. 2018, 19, 178. [Google Scholar] [CrossRef] [Green Version]
  123. Østerberg, J.T.; Xiang, W.; Olsen, L.I.; Edenbrandt, A.K.; Vedel, S.E.; Christiansen, A.; Landes, X.; Andersen, M.M.; Pagh, P.; Sandøe, P.; et al. Accelerating the Domestication of New Crops: Feasibility and Approaches. Trends Plant Sci. 2017, 22, 373–384. [Google Scholar] [CrossRef] [PubMed]
  124. Komatsuda, T.; Pourkheirandish, M.; He, C.; Azhaguvel, P.; Kanamori, H.; Perovic, D.; Stein, N.; Graner, A.; Wicker, T.; Tagiri, A.; et al. Six-rowed barley originated from a mutation in a homeodomain-leucine zipper I-class homeobox gene. Proc. Natl. Acad. Sci. USA 2007, 104, 1424–1429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Wang, H.; Nussbaum-Wagler, T.; Li, B.; Zhao, Q.; Vigouroux, Y.; Faller, M.; Bomblies, K.; Lukens, L.; Doebley, J.F. The origin of the naked grains of maize. Nature 2005, 436, 714–719. [Google Scholar] [CrossRef] [PubMed]
  126. Jin, J.; Huang, W.; Gao, J.-P.; Yang, J.; Shi, M.; Zhu, M.-Z.; Luo, D.; Lin, H.-X. Genetic control of rice plant architecture under domestication. Nat. Genet. 2008, 40, 1365–1369. [Google Scholar] [CrossRef] [PubMed]
  127. Lin, Z.; Li, X.; Shannon, L.M.; Yeh, C.-T.; Wang, M.L.; Bai, G.; Peng, Z.; Li, J.; Trick, H.N.; Clemente, T.E.; et al. Parallel domestication of the Shattering1 genes in cereals. Nat. Genet. 2012, 44, 720–724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Chen, K.-Y.; Cong, B.; Wing, R.; Vrebalov, J.; Tanksley, S.D. Changes in Regulation of a Transcription Factor lead to Autogamy in Cultivated Tomatoes. Science 2007, 318, 643. [Google Scholar] [CrossRef]
  129. Zsögön, A.; Cermak, T.; Voytas, D.; Peres, L.E.P. Genome editing as a tool to achieve the crop ideotype and de novo domestication of wild relatives: Case study in tomato. Plant Sci. 2017, 256, 120–130. [Google Scholar] [CrossRef]
  130. Khan, M.Z.; Zaidi, S.S.-e.-A.; Amin, I.; Mansoor, S. A CRISPR Way for Fast-Forward Crop Domestication. Trends Plant Sci. 2019, 24, 293–296. [Google Scholar] [CrossRef]
  131. Zsögön, A.; Čermák, T.; Naves, E.R.; Notini, M.M.; Edel, K.H.; Weinl, S.; Freschi, L.; Voytas, D.F.; Kudla, J.; Peres, L.E.P. De novo domestication of wild tomato using genome editing. Nat. Biotechnol. 2018, 36, 1211–1216. [Google Scholar] [CrossRef] [Green Version]
  132. Li, T.; Yang, X.; Yu, Y.; Si, X.; Zhai, X.; Zhang, H.; Dong, W.; Gao, C.; Xu, C. Domestication of wild tomato is accelerated by genome editing. Nat. Biotechnol. 2018, 36, 1160–1163. [Google Scholar] [CrossRef]
  133. Sedbrook, J.C.; Phippen, W.B.; Marks, M.D. New approaches to facilitate rapid domestication of a wild plant to an oilseed crop: Example pennycress (Thlaspi arvense L.). Plant Sci. 2014, 227, 122–132. [Google Scholar] [CrossRef]
  134. Shapter, F.M.; Cross, M.; Ablett, G.; Malory, S.; Chivers, I.H.; King, G.J.; Henry, R.J. High-Throughput Sequencing and Mutagenesis to Accelerate the Domestication of Microlaena stipoides as a New Food Crop. PLoS ONE 2013, 8, e82641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Liu, Y.; Merrick, P.; Zhang, Z.; Ji, C.; Yang, B.; Fei, S.-Z. Targeted mutagenesis in tetraploid switchgrass (Panicum virgatum L.) using CRISPR/Cas9. Plant Biotechnol. J. 2018, 16, 381–393. [Google Scholar] [CrossRef] [Green Version]
  136. Zhu, Y.; Lin, Y.; Chen, S.; Liu, H.; Chen, Z.; Fan, M.; Hu, T.; Mei, F.; Chen, J.; Chen, L.; et al. CRISPR/Cas9-mediated functional recovery of the recessive rc allele to develop red rice. Plant Biotechnol. J. 2019, 17, 2096–2105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Liu, H.; Wang, K.; Tang, H.; Gong, Q.; Du, L.; Pei, X.; Ye, X. CRISPR/Cas9 editing of wheat TaQ genes alters spike morphogenesis and grain threshability. J. Genet. Genom. 2020, 47, 563–575. [Google Scholar] [CrossRef]
  138. Zhang, Y.; Thomas, W.; Bayer, P.E.; Edwards, D.; Batley, J. Frontiers in Dissecting and Managing Brassica Diseases: From Reference-Based RGA Candidate Identification to Building Pan-RGAomes. Int. J. Mol. Sci. 2020, 21, 8964. [Google Scholar] [CrossRef] [PubMed]
  139. Zhang, H.; Li, Y.; Zhu, J.-K. Developing naturally stress-resistant crops for a sustainable agriculture. Nat. Plants 2018, 4, 989–996. [Google Scholar] [CrossRef]
  140. Kole, C. Wild Crop Relatives: Genomic and Breeding Resources: Cereals; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2011; Volume 1. [Google Scholar]
  141. Smýkal, P.; Nelson, M.N.; Berger, J.D.; Von Wettberg, E.J. The impact of genetic changes during crop domestication. Agronomy 2018, 8, 119. [Google Scholar] [CrossRef] [Green Version]
  142. Pinosio, S.; Giacomello, S.; Faivre-Rampant, P.; Taylor, G.; Jorge, V.; Le Paslier, M.C.; Zaina, G.; Bastien, C.; Cattonaro, F.; Marroni, F.; et al. Characterization of the Poplar Pan-Genome by Genome-Wide Identification of Structural Variation. Mol. Biol. Evol. 2016, 33, 2706–2719. [Google Scholar] [CrossRef] [Green Version]
  143. Zhang, B.; Zhu, W.; Diao, S.; Wu, X.; Lu, J.; Ding, C.; Su, X. The poplar pangenome provides insights into the evolutionary history of the genus. Commun. Biol. 2019, 2, 215. [Google Scholar] [CrossRef]
  144. Rijzaani, H.; Bayer, P.E.; Rouard, M.; Doležel, J.; Batley, J.; Edwards, D. The pangenome of banana highlights differences between genera and genomes. Plant Genome 2021, 11, e20100. [Google Scholar] [CrossRef]
  145. Sharwood, R.E. Engineering chloroplasts to improve Rubisco catalysis: Prospects for translating improvements into food and fiber crops. New Phytol. 2017, 213, 494–510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Franks, S.J.; Hoffmann, A.A. Genetics of Climate Change Adaptation. Annu. Rev. Genet. 2012, 46, 185–208. [Google Scholar] [CrossRef] [PubMed]
  147. Lambers, H.; Martinoia, E.; Renton, M. Plant adaptations to severely phosphorus-impoverished soils. Curr. Opin. Plant Biol. 2015, 25, 23–31. [Google Scholar] [CrossRef] [Green Version]
  148. Cole, M.B.; Augustin, M.A.; Robertson, M.J.; Manners, J.M. The science of food security. Npj Sci. Food 2018, 2, 14. [Google Scholar] [CrossRef] [PubMed]
  149. Hess, G.T.; Frésard, L.; Han, K.; Lee, C.H.; Li, A.; Cimprich, K.A.; Montgomery, S.B.; Bassik, M.C. Directed evolution using dCas9-targeted somatic hypermutation in mammalian cells. Nat. Methods 2016, 13, 1036–1042. [Google Scholar] [CrossRef] [Green Version]
  150. Mishra, R.; Joshi, R.K.; Zhao, K. Base editing in crops: Current advances, limitations and future implications. Plant Biotechnol. J. 2020, 18, 20–31. [Google Scholar] [CrossRef] [PubMed]
  151. Rees, H.A.; Liu, D.R. Base editing: Precision chemistry on the genome and transcriptome of living cells. Nat. Rev. Genet. 2018, 19, 770–788. [Google Scholar] [CrossRef]
  152. Dong, O.X.; Yu, S.; Jain, R.; Zhang, N.; Duong, P.Q.; Butler, C.; Li, Y.; Lipzen, A.; Martin, J.A.; Barry, K.W.; et al. Marker-free carotenoid-enriched rice generated through targeted gene insertion using CRISPR-Cas9. Nat. Commun. 2020, 11, 1178. [Google Scholar] [CrossRef] [Green Version]
  153. Hoppe, C.; Ashe, H.L. CRISPR-Cas9 strategies to insert MS2 stem-loops into endogenous loci in Drosophila embryos. STAR Protoc. 2021, 2, 100380. [Google Scholar] [CrossRef]
  154. Lu, Y.; Tian, Y.; Shen, R.; Yao, Q.; Wang, M.; Chen, M.; Dong, J.; Zhang, T.; Li, F.; Lei, M.; et al. Targeted, efficient sequence insertion and replacement in rice. Nat. Biotechnol. 2020, 38, 1402–1407. [Google Scholar] [CrossRef] [PubMed]
  155. Poernbacher, I.; Crossman, S.; Kurth, J.; Nojima, H.; Baena-Lopez, A.; Alexandre, C.; Vincent, J.-P. Lessons in genome engineering: Opportunities, tools and pitfalls. bioRxiv 2019, 710871. [Google Scholar] [CrossRef]
  156. Hendriks, D.; Clevers, H.; Artegiani, B. CRISPR-Cas Tools and Their Application in Genetic Engineering of Human Stem Cells and Organoids. Cell Stem Cell 2020, 27, 705–731. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Diagram of pangenome construction methods based on genome sequencing data. Genome sequencing reads for genomes A, B and C are shown at the top of the image, each colour represents a gene in the genome. The genome sequencing reads may be assembled into a pangenome using de novo, iterative and graph-based assemblies which may influence the positioning of the assembled genes. The * indicates that genome A is used as reference genome in the iterative assembly method.
Figure 1. Diagram of pangenome construction methods based on genome sequencing data. Genome sequencing reads for genomes A, B and C are shown at the top of the image, each colour represents a gene in the genome. The genome sequencing reads may be assembled into a pangenome using de novo, iterative and graph-based assemblies which may influence the positioning of the assembled genes. The * indicates that genome A is used as reference genome in the iterative assembly method.
Ijms 23 02276 g001
Figure 2. (A). Representation of a pangenome assembly composed of genomes from six individuals sourced from two populations. The core and variable regions of the pangenome are highlighted in this representation, in which the genetic diversity observed in the variable region can be caused by chromosome inversion or copy number variation (CNV). (B). Potential benefits of using pangenome reference for genetic modification, as the genetic diversity analysis can be used to define target sites in variant alleles, identify CNV that influence CRISPR-Cas mutation effectiveness and discover novel target alleles.
Figure 2. (A). Representation of a pangenome assembly composed of genomes from six individuals sourced from two populations. The core and variable regions of the pangenome are highlighted in this representation, in which the genetic diversity observed in the variable region can be caused by chromosome inversion or copy number variation (CNV). (B). Potential benefits of using pangenome reference for genetic modification, as the genetic diversity analysis can be used to define target sites in variant alleles, identify CNV that influence CRISPR-Cas mutation effectiveness and discover novel target alleles.
Ijms 23 02276 g002
Figure 3. Reversal of inversion through CRISPR to allow crossing of inverted genes. A pangenome is used to identify a non-recombinant inversion in individual A compared to individuals B and C. CRISPR-Cas proteins are then used to induce double-stranded breaks at specific target sites in the inverted region, leading to re-inversion of the genomic segment and accessibility of the locked genes for recombination. The previously inverted genes in individual A can then be crossed with other individuals in the population.
Figure 3. Reversal of inversion through CRISPR to allow crossing of inverted genes. A pangenome is used to identify a non-recombinant inversion in individual A compared to individuals B and C. CRISPR-Cas proteins are then used to induce double-stranded breaks at specific target sites in the inverted region, leading to re-inversion of the genomic segment and accessibility of the locked genes for recombination. The previously inverted genes in individual A can then be crossed with other individuals in the population.
Ijms 23 02276 g003
Table 1. A summary of plant pangenome studies.
Table 1. A summary of plant pangenome studies.
Species NameAccessionsSize (Individuals)AnalysisReference
Glycine maxUSDA collection1110PAV, GO, SNP discovery and population genetics analysis[68]
Chinese population26Synteny, SV, genetic variation and gene expression analysis[50]
Zea maysUSDA Collection14,129GBS tagging, GWAS mapping, and PAV analysis[51]
Solanum lycopersicumNCBI SRA database725SNP calling, QTL mapping, expression analysis, and PAV analysis[35]
Brassica oleraceaChinese Kale/TO10010Gene clustering, TE annotation, SNP calling, phylogenetic, PAV, and GO analysis[39]
Brassica napusDiversity set8Phylogenetic, SNP, InDels, SV, PAV, population analysis[27]
Diversity set53Candidate identification, QTL, and SNP analysis[52]
Brassica distachyonDiversity set54Pan-gene clustering, variant calling, TE, and indel phylogenetic analysis[53]
Hordeum vulgareDiversity set20GWAS, inversion calling, SNP calling, QTL mapping, and PAV analysis[54]
Oryza sativaChina National Rice Research Institute andNational Institute of Genetics in Japan1529Evolutionarily and PAV analysis[55]
Cajanus cajanICRISAT89SNP and PAV analysis[29]
Diversity set3366SNP, SV, CNV, phylogenetic, GWAS analysis, and genomeprediction[56]
Malus domesticaPlant Genetic Resources Unit91Gene prediction, comparative analysis, and PAV/variant calling[57]
Sesamum indicumDiversity set5PAV and evolutionary analysis[58]
Helianthus annuusUSDA Collection493SNP calling, genome positioning, and GO and GWAS analysis[59]
Manihot esculentaDiversity set57Haplotype sampling and genomic prediction[60]
Sorghum bicolorDiversity set354PAV, SNP, GWAS, diversity, and population analysis[36]
Chibas sorghum breeding program24Genotype prediction, haplotype sampling, and WGS[61]
Triticum aestivumChinese Spring18PAV and SNP analysis[62]
Arabidopsis thalianaMPI for Plant Breeding Research7Pangenomic, CNV, and synteny analysis[63]
Amborella trichopodaAmbroella Genome Project10PAV, GO, candidate gene, phylogenetic, and SNP analysis[64]
Medicago truncatulaDiversity set15Comparative genomic analysis, protein orthlog, diversity, and SV analysis[66]
GossypiumNCBI database1961InDel, population structure, LD, SV, CNV, PAV, and metagenome association analysis[65]
CapsicumDiversity set5PAV and GWAS analysis[25]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tay Fernandez, C.G.; Nestor, B.J.; Danilevicz, M.F.; Marsh, J.I.; Petereit, J.; Bayer, P.E.; Batley, J.; Edwards, D. Expanding Gene-Editing Potential in Crop Improvement with Pangenomes. Int. J. Mol. Sci. 2022, 23, 2276. https://doi.org/10.3390/ijms23042276

AMA Style

Tay Fernandez CG, Nestor BJ, Danilevicz MF, Marsh JI, Petereit J, Bayer PE, Batley J, Edwards D. Expanding Gene-Editing Potential in Crop Improvement with Pangenomes. International Journal of Molecular Sciences. 2022; 23(4):2276. https://doi.org/10.3390/ijms23042276

Chicago/Turabian Style

Tay Fernandez, Cassandria G., Benjamin J. Nestor, Monica F. Danilevicz, Jacob I. Marsh, Jakob Petereit, Philipp E. Bayer, Jacqueline Batley, and David Edwards. 2022. "Expanding Gene-Editing Potential in Crop Improvement with Pangenomes" International Journal of Molecular Sciences 23, no. 4: 2276. https://doi.org/10.3390/ijms23042276

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