Search
2023 Volume 2
Article Contents
REVIEW   Open Access    

Synthetic apomixis: from genetic basis to agricultural application

More Information
  • Received Date: 19 May 2023
    Accepted Date: 04 July 2023
    Published Online: 04 September 2023
    Seed Biology  2 Article number: 10 (2023)  |  Cite this article
  • Apomixis in plants is a widely existing biological phenomenon, in which seeds are formed without egg cell and sperm uniting. Hybrid breeding exploits heterosis to obtain seeds with superior traits. However, segregation of traits in the offspring greatly limits the widespread use of hybrid vigor in agricultural production. Synthetic apomixis is considered a desired way of clonal propagation of heterozygous maternal parents, which bypasses laborious hybrid process. In recent years, with the increasing understanding of the molecular mechanisms of plant meiosis and double fertilization, scientists have introduced apomixis to rice and hybrid rice varieties by genetic engineering of genes that are involved in sexual reproduction. In this review article, we will summarize the recent research progress in the meiosis and double fertilization related to synthetic apomixis and provide perspectives on the potential application of synthetic apomixis in different crops and livestock pastures.
  • 加载中
  • [1]

    Hojsgaard D, Hörandl E. 2019. The rise of apomixis in natural plant populations. Frontiers in Plant Science 10:358

    doi: 10.3389/fpls.2019.00358

    CrossRef   Google Scholar

    [2]

    Wang N, Song X, Ye J, Zhang S, Cao Z, et al. 2022. Structural variation and parallel evolution of apomixis in citrus during domestication and diversification. National Science Review 9:nwac114

    doi: 10.1093/nsr/nwac114

    CrossRef   Google Scholar

    [3]

    Wang X, Xu Y, Zhang S, Cao L, Huang Y, et al. 2017. Genomic analyses of primitive, wild and cultivated citrus provide insights into asexual reproduction. Nature Genetics 49:765−72

    doi: 10.1038/ng.3839

    CrossRef   Google Scholar

    [4]

    Smith J. 1841. Notice of a plant which produces perfect seeds without any apparent action of pollen. Transactions of the Linnean Society of London 18:509−12

    doi: 10.1111/j.1095-8339.1838.tb00200.x

    CrossRef   Google Scholar

    [5]

    Hojsgaard D, Klatt S, Baier R, Carman JG, Hörandl E. 2014. Taxonomy and biogeography of apomixis in angiosperms and associated biodiversity characteristics. Critical Reviews in Plant Sciences 33:414−27

    doi: 10.1080/07352689.2014.898488

    CrossRef   Google Scholar

    [6]

    Liu C, He Z, Zhang Y, Hu F, Li M, et al. 2023. Synthetic apomixis enables stable transgenerational transmission of heterotic phenotypes in hybrid rice. Plant Communications 4:100470

    doi: 10.1016/j.xplc.2022.100470

    CrossRef   Google Scholar

    [7]

    Jiang C, Sun J, Li R, Yan S, Chen W, et al. 2022. A reactive oxygen species burst causes haploid induction in maize. Molecular Plant 15:943−55

    doi: 10.1016/j.molp.2022.04.001

    CrossRef   Google Scholar

    [8]

    Huang X, Yang S, Gong J, Zhao Q, Feng Q, et al. 2016. Genomic architecture of heterosis for yield traits in rice. Nature 537:629−33

    doi: 10.1038/nature19760

    CrossRef   Google Scholar

    [9]

    Ye J, Cui X. 2019. Next-generation crop breeding methods. Molecular Plant 12:470−71

    doi: 10.1016/j.molp.2019.03.007

    CrossRef   Google Scholar

    [10]

    Savidan Y. 2001. Gametophytic Apomixis. In Current Trends in the Embryology of Angiosperms, eds. Bhojwani SS, Soh WY. Netherlands: Springer Dordrecht. pp. 419−33. https://doi.org/10.1007/978-94-017-1203-3_16

    [11]

    Spillane C, Curtis MD, Grossniklaus U. 2004. Apomixis technology development—virgin births in farmers' fields? Nature Biotechnology 22:687−91

    doi: 10.1038/nbt976

    CrossRef   Google Scholar

    [12]

    Mieulet D, Jolivet S, Rivard M, Cromer L, Vernet A, et al. 2016. Turning rice meiosis into mitosis. Cell Research 26:1242−54

    doi: 10.1038/cr.2016.117

    CrossRef   Google Scholar

    [13]

    Mercier R, Mézard C, Jenczewski E, Macaisne N, Grelon M. 2015. The molecular biology of meiosis in plants. Annual Review of Plant Biology 66:297−327

    doi: 10.1146/annurev-arplant-050213-035923

    CrossRef   Google Scholar

    [14]

    Cheung AY, Duan Q, Li C, Liu MCJ, Wu HM. 2022. Pollen–pistil interactions: It takes two to tangle but a molecular cast of many to deliver. Current Opinion in Plant Biology 69:102279

    doi: 10.1016/j.pbi.2022.102279

    CrossRef   Google Scholar

    [15]

    Dresselhaus T, Sprunck S, Wessel GM. 2016. Fertilization mechanisms in flowering plants. Current Biology 26:R125−R139

    doi: 10.1016/j.cub.2015.12.032

    CrossRef   Google Scholar

    [16]

    d'Erfurth I, Jolivet S, Froger N, Catrice O, Novatchkova M, et al. 2009. Turning meiosis into mitosis. PLoS Biology 7:e1000124

    doi: 10.1371/journal.pbio.1000124

    CrossRef   Google Scholar

    [17]

    Wang C, Liu Q, Shen Y, Hua Y, Wang J, et al. 2019. Clonal seeds from hybrid rice by simultaneous genome engineering of meiosis and fertilization genes. Nature Biotechnology 37:283−86

    doi: 10.1038/s41587-018-0003-0

    CrossRef   Google Scholar

    [18]

    Khanday I, Skinner D, Yang B, Mercier R, Sundaresan V. 2019. A male-expressed rice embryogenic trigger redirected for asexual propagation through seeds. Nature 565:91−95

    doi: 10.1038/s41586-018-0785-8

    CrossRef   Google Scholar

    [19]

    de Massy B. 2013. Initiation of meiotic recombination: how and where? Conservation and specificities among eukaryotes Annual Review of Genetics 47:563−99

    doi: 10.1146/annurev-genet-110711-155423

    CrossRef   Google Scholar

    [20]

    Drouaud J, Khademian H, Giraut L, Zanni V, Bellalou S, et al. 2013. Contrasted patterns of crossover and non-crossover at Arabidopsis thaliana meiotic recombination hotspots. PLOS Genetics 9:e1003922

    doi: 10.1371/journal.pgen.1003922

    CrossRef   Google Scholar

    [21]

    Smeds L, Mugal CF, Qvarnström A, Ellegren H. 2016. High-resolution mapping of crossover and non-crossover recombination events by whole-genome re-sequencing of an avian pedigree. PLoS Genetics 12:e1006044

    doi: 10.1371/journal.pgen.1006044

    CrossRef   Google Scholar

    [22]

    Myers S, Spencer CCA, Auton A, Bottolo L, Freeman C, et al. 2006. The distribution and causes of meiotic recombination in the human genome. Biochemical Society Transactions 34:526−30

    doi: 10.1042/BST0340526

    CrossRef   Google Scholar

    [23]

    Comeron JM, Ratnappan R, Bailin S. 2012. The many landscapes of recombination in Drosophila melanogaster. PLoS Genetics 8:e1002905

    doi: 10.1371/journal.pgen.1002905

    CrossRef   Google Scholar

    [24]

    Benyahya F, Nadaud I, Da Ines O, Rimbert H, White C, et al. 2020. SPO11.2 is essential for programmed double-strand break formation during meiosis in bread wheat (Triticum aestivum L.). The Plant Journal 104:30−43

    doi: 10.1111/tpj.14903

    CrossRef   Google Scholar

    [25]

    Stacey NJ, Kuromori T, Azumi Y, Roberts G, Breuer C, et al. 2006. Arabidopsis SPO11-2 functions with SPO11-1 in meiotic recombination. The Plant Journal 48:206−16

    doi: 10.1111/j.1365-313X.2006.02867.x

    CrossRef   Google Scholar

    [26]

    Fayos I, Meunier AC, Vernet A, Navarro-Sanz S, Portefaix M, et al. 2020. Assessment of the roles of SPO11-2 and SPO11-4 in meiosis in rice using CRISPR/Cas9 mutagenesis. Journal of Experimental Botany 71:7046−58

    doi: 10.1093/jxb/eraa391

    CrossRef   Google Scholar

    [27]

    Steckenborn S, Cuacos M, Ayoub MA, Feng C, Schubert V, et al. 2023. The meiotic topoisomerase VI B subunit (MTOPVIB) is essential for meiotic DNA double-strand break formation in barley (Hordeum vulgare L.). Plant Reproduction 36:1−15

    doi: 10.1007/s00497-022-00444-5

    CrossRef   Google Scholar

    [28]

    Jolivet S, Vezon D, Froger N, Mercier R. 2006. Non conservation of the meiotic function of the Ski8/Rec103 homolog in Arabidopsis. Genes Cells 11:615−22

    doi: 10.1111/j.1365-2443.2006.00972.x

    CrossRef   Google Scholar

    [29]

    Hyppa RW, Cromie GA, Smith GR. 2008. Indistinguishable landscapes of meiotic DNA breaks in rad50+ and rad50S strains of fission yeast revealed by a novel rad50+ recombination intermediate. PLoS Genetics 4:e1000267

    doi: 10.1371/journal.pgen.1000267

    CrossRef   Google Scholar

    [30]

    Puizina J, Siroky J, Mokros P, Schweizer D, Riha K. 2004. Mre11 deficiency in Arabidopsis is associated with chromosomal instability in somatic cells and Spo11-Dependent genome fragmentation during meiosis. The Plant Cell 16:1968−78

    doi: 10.1105/tpc.104.022749

    CrossRef   Google Scholar

    [31]

    Shi W, Ji J, Xue Z, Zhang F, Miao Y, et al. 2021. PRD1, a homologous recombination initiation factor, is involved in spindle assembly in rice meiosis. New Phytologist 230:585−600

    doi: 10.1111/nph.17178

    CrossRef   Google Scholar

    [32]

    Wang Z, Zhang Z, Zheng D, Zhang T, Li XL, et al. 2022. Efficient and genotype independent maize transformation using pollen transfected by DNA-coated magnetic nanoparticles. Journal of Integrative Plant Biology 64:1145−56

    doi: 10.1111/jipb.13263

    CrossRef   Google Scholar

    [33]

    De Muyt A, Vezon D, Gendrot G, Gallois JL, Stevens R, et al. 2007. AtPRD1 is required for meiotic double strand break formation in Arabidopsis thaliana. The EMBO Journal 26:4126−37

    doi: 10.1038/sj.emboj.7601815

    CrossRef   Google Scholar

    [34]

    De Muyt A, Pereira L, Vezon D, Chelysheva L, Gendrot G, et al. 2009. A high throughput genetic screen identifies new early meiotic recombination functions in Arabidopsis thaliana. PLOS Genetics 5:e1000654

    doi: 10.1371/journal.pgen.1000654

    CrossRef   Google Scholar

    [35]

    Zhang C, Song Y, Cheng Z, Wang Y, Zhu J, et al. 2012. The Arabidopsis thaliana DSB formation (AtDFO) gene is required for meiotic double-strand break formation. The Plant Journal 72:271−81

    doi: 10.1111/j.1365-313X.2012.05075.x

    CrossRef   Google Scholar

    [36]

    Nonomura KI, Nakano M, Fukuda T, Eiguchi M, Miyao A, et al. 2004. The novel gene HOMOLOGOUS PAIRING ABERRATION IN RICE MEIOSIS1 of rice encodes a putative coiled-coil protein required for homologous chromosome pairing in meiosis. The Plant Cell 16:1008−20

    doi: 10.1105/tpc.020701

    CrossRef   Google Scholar

    [37]

    Miao C, Tang D, Zhang H, Wang M, Li Y, et al. 2013. CENTRAL REGION COMPONENT1, a novel synaptonemal complex component, is essential for meiotic recombination initiation in rice. The Plant Cell 25:2998−3009

    doi: 10.1105/tpc.113.113175

    CrossRef   Google Scholar

    [38]

    Ji J, Tang D, Shen Y, Xue Z, Wang H, et al. 2016. P31comet, a member of the synaptonemal complex, participates in meiotic DSB formation in rice. PNAS 113:10577−82

    doi: 10.1073/pnas.1607334113

    CrossRef   Google Scholar

    [39]

    Wu Z, Ji J, Tang D, Wang H, Shen Y, et al. 2015. OsSDS is essential for DSB formation in rice meiosis. Frontiers in Plant Science 6:21

    doi: 10.3389/fpls.2015.00021

    CrossRef   Google Scholar

    [40]

    Casari E, Rinaldi C, Marsella A, Gnugnoli M, Colombo CV, et al. 2019. Processing of DNA Double-Strand Breaks by the MRX complex in a chromatin context. Frontiers in Molecular Biosciences 6:43

    doi: 10.3389/fmolb.2019.00043

    CrossRef   Google Scholar

    [41]

    Uanschou C, Siwiec T, Pedrosa-Harand A, Kerzendorfer C, Sanchez-Moran E, et al. 2007. A novel plant gene essential for meiosis is related to the human CtIP and the yeast COM1/SAE2 gene. The EMBO Journal 26:5061−70

    doi: 10.1038/sj.emboj.7601913

    CrossRef   Google Scholar

    [42]

    Neale MJ, Pan J, Keeney S. 2005. Endonucleolytic processing of covalent protein-linked DNA double-strand breaks. Nature 436:1053−57

    doi: 10.1038/nature03872

    CrossRef   Google Scholar

    [43]

    Hinch AG, Becker PW, Li T, Moralli D, Zhang G, et al. 2020. The configuration of RPA, RAD51, and DMC1 binding in meiosis reveals the nature of critical recombination intermediates. Molecular Cell 79:689−701. E10

    doi: 10.1016/j.molcel.2020.06.015

    CrossRef   Google Scholar

    [44]

    Berchowitz LE, Francis KE, Bey AL, Copenhaver GP. 2007. The role of AtMUS81 in interference-insensitive crossovers in A. thaliana. PLoS Genetics 3:e0030132

    doi: 10.1371/journal.pgen.0030132

    CrossRef   Google Scholar

    [45]

    Gutiérrez Pinzón Y, González Kise JK, Rueda P, Ronceret A. 2021. The formation of bivalents and the control of plant meiotic recombination. Frontiers in Plant Science 12:717423

    doi: 10.3389/fpls.2021.717423

    CrossRef   Google Scholar

    [46]

    Chelysheva L, Vezon D, Belcram K, Gendrot G, Grelon M. 2008. The Arabidopsis BLAP75/Rmi1 homologue plays crucial roles in meiotic double-strand break repair. PLoS Genetics 4:e1000309

    doi: 10.1371/journal.pgen.1000309

    CrossRef   Google Scholar

    [47]

    Hartung F, Suer S, Knoll A, Wurz-Wildersinn R, Puchta H. 2008. Topoisomerase 3α and RMI1 suppress somatic crossovers and are essential for resolution of meiotic recombination intermediates in Arabidopsis thaliana. PLoS Genetics 4:e1000285

    doi: 10.1371/journal.pgen.1000285

    CrossRef   Google Scholar

    [48]

    Girard C, Crismani W, Froger N, Mazel J, Lemhemdi A, et al. 2014. FANCM-associated proteins MHF1 and MHF2, but not the other Fanconi anemia factors, limit meiotic crossovers. Nucleic Acids Research 42:9087−95

    doi: 10.1093/nar/gku614

    CrossRef   Google Scholar

    [49]

    Crismani W, Girard C, Froger N, Pradillo M, Santos JL, et al. 2012. FANCM limits meiotic crossovers. Science 336:1588−90

    doi: 10.1126/science.1220381

    CrossRef   Google Scholar

    [50]

    Knoll A, Higgins JD, Seeliger K, Reha SJ, Dangel NJ, et al. 2012. The Fanconi anemia ortholog FANCM ensures ordered homologous recombination in both somatic and meiotic cells in Arabidopsis. The Plant Cell 24:1448−64

    doi: 10.1105/tpc.112.096644

    CrossRef   Google Scholar

    [51]

    Chelysheva L, Diallo S, Vezon D, Gendrot G, Vrielynck N, et al. 2005. AtREC8 and AtSCC3 are essential to the monopolar orientation of the kinetochores during meiosis. Journal of Cell Science 118:4621−32

    doi: 10.1242/jcs.02583

    CrossRef   Google Scholar

    [52]

    Shao T, Tang D, Wang K, Wang M, Che L, et al. 2011. OsREC8 is essential for chromatid cohesion and metaphase I monopolar orientation in rice meiosis. Plant Physiology 156:1386−96

    doi: 10.1104/pp.111.177428

    CrossRef   Google Scholar

    [53]

    Sun M, Nishino T, Marko JF. 2013. The SMC1-SMC3 cohesin heterodimer structures DNA through supercoiling-dependent loop formation. Nucleic Acids Research 41:6149−60

    doi: 10.1093/nar/gkt303

    CrossRef   Google Scholar

    [54]

    Zamariola L, de Storme N, Tiang CL, Armstrong SJ, Franklin FCH, et al. 2013. SGO1 but not SGO2 is required for maintenance of centromere cohesion in Arabidopsis thaliana meiosis. Plant Reproduction 26:197−208

    doi: 10.1007/s00497-013-0231-x

    CrossRef   Google Scholar

    [55]

    Wang M, Tang D, Luo Q, Jin Y, Shen Y, et al. 2012. BRK1, a Bub1-Related Kinase, is essential for generating proper tension between homologous kinetochores at metaphase I of rice meiosis. The Plant Cell 24:4961−73

    doi: 10.1105/tpc.112.105874

    CrossRef   Google Scholar

    [56]

    Marston AL, Amon A. 2004. Meiosis: cell-cycle controls shuffle and deal. Nature Reviews Molecular Cell Biology 5:983−97

    doi: 10.1038/nrm1526

    CrossRef   Google Scholar

    [57]

    Harashima H, Dissmeyer N, Schnittger A. 2013. Cell cycle control across the eukaryotic kingdom. Trends in Cell Biology 23:345−56

    doi: 10.1016/j.tcb.2013.03.002

    CrossRef   Google Scholar

    [58]

    Pesin JA, Orr-Weaver TL. 2008. Regulation of APC/C activators in mitosis and meiosis. Annual Review of Cell and Developmental Biology 24:475−99

    doi: 10.1146/annurev.cellbio.041408.115949

    CrossRef   Google Scholar

    [59]

    Fisher RP, Morgan DO. 1994. A novel cyclin associates with M015/CDK7 to form the CDK-activating kinase. Cell 78:713−24

    doi: 10.1016/0092-8674(94)90535-5

    CrossRef   Google Scholar

    [60]

    d'Erfurth I, Cromer L, Jolivet S, Girard C, Horlow C, et al. 2010. The CYCLIN-A CYCA1;2/TAM is required for the meiosis I to meiosis II transition and cooperates with OSD1 for the prophase to first meiotic division transition. PLoS Genetics 6:e1000989

    doi: 10.1371/journal.pgen.1000989

    CrossRef   Google Scholar

    [61]

    Cifuentes M, Jolivet S, Cromer L, Harashima H, Bulankova P, et al. 2016. TDM1 regulation determines the number of meiotic divisions. PLoS Genetics 12:e1005856

    doi: 10.1371/journal.pgen.1005856

    CrossRef   Google Scholar

    [62]

    Dissmeyer N, Nowack MK, Pusch S, Stals H, Inzé D, et al. 2007. T-Loop phosphorylation of Arabidopsis CDKA;1 is required for its function and can be partially substituted by an aspartate residue. The Plant Cell 19:972−85

    doi: 10.1105/tpc.107.050401

    CrossRef   Google Scholar

    [63]

    Cairo A, Vargova A, Shukla N, Capitao C, Mikulkova P, et al. 2022. Meiotic exit in Arabidopsis is driven by P-body–mediated inhibition of translation. Science 377:629−34

    doi: 10.1126/science.abo0904

    CrossRef   Google Scholar

    [64]

    Yao X, Hu W, Yang Z. 2022. The contributions of sporophytic tapetum to pollen formation. Seed Biology 1:5

    doi: 10.48130/seedbio-2022-0005

    CrossRef   Google Scholar

    [65]

    Yang W, Shi D, Chen Y. 2010. Female gametophyte development in flowering plants. Annual Review of Plant Biology 61:89−108

    doi: 10.1146/annurev-arplant-042809-112203

    CrossRef   Google Scholar

    [66]

    Hafidh S, Honys D. 2021. Reproduction multitasking: the male gametophyte. Annual Review of Plant Biology 72:581−614

    doi: 10.1146/annurev-arplant-080620-021907

    CrossRef   Google Scholar

    [67]

    Sprunck S. 2020. Twice the fun, double the trouble: gamete interactions in flowering plants. Current Opinion in Plant Biology 53:106−16

    doi: 10.1016/j.pbi.2019.11.003

    CrossRef   Google Scholar

    [68]

    Maruyama D, Ohtsu M, Higashiyama T. 2016. Cell fusion and nuclear fusion in plants. Seminars in Cell & Developmental Biology 60:127−35

    doi: 10.1016/j.semcdb.2016.07.024

    CrossRef   Google Scholar

    [69]

    Hamamura Y, Nishimaki M, Takeuchi H, Geitmann A, Kurihara D, et al. 2014. Live imaging of calcium spikes during double fertilization in Arabidopsis. Nature Communications 5:4722

    doi: 10.1038/ncomms5722

    CrossRef   Google Scholar

    [70]

    Mori T, Igawa T, Tamiya G, Miyagishima SY, Berger F. 2014. Gamete attachment requires GEX2 for successful fertilization inArabidopsis. Current Biology 24:170−75

    doi: 10.1016/j.cub.2013.11.030

    CrossRef   Google Scholar

    [71]

    Misamore MJ, Gupta S, Snell WJ. 2003. The Chlamydomonas Fus1 protein is present on the mating type plus fusion organelle and required for a critical membrane adhesion event during fusion with minus gametes. Molecular Biology of the Cell 14:2530−42

    doi: 10.1091/mbc.e02-12-0790

    CrossRef   Google Scholar

    [72]

    Hirohashi N, Kamei N, Kubo H, Sawada H, Matsumoto M, et al. 2008. Egg and sperm recognition systems during fertilization. Development, Growth & Differentiation 50:S221−S238

    doi: 10.1111/j.1440-169X.2008.01017.x

    CrossRef   Google Scholar

    [73]

    Pan J, Snell WJ. 2000. Signal transduction during fertilization in the unicellular green alga, Chlamydomonas. Current Opinion in Microbiology 3:596−602

    doi: 10.1016/S1369-5274(00)00146-6

    CrossRef   Google Scholar

    [74]

    Sprunck S, Rademacher S, Vogler F, Gheyselinck J, Grossniklaus U, et al. 2012. Egg cell-secreted EC1 triggers sperm cell activation during double fertilization. Science 338:1093−97

    doi: 10.1126/science.1223944

    CrossRef   Google Scholar

    [75]

    Johnson MA, von Besser K, Zhou Q, Smith E, Aux G, et al. 2004. Arabidopsis hapless mutations define essential gametophytic functions. Genetics 168:971−82

    doi: 10.1534/genetics.104.029447

    CrossRef   Google Scholar

    [76]

    Cyprys P, Lindemeier M, Sprunck S. 2019. Gamete fusion is facilitated by two sperm cell-expressed DUF679 membrane proteins. Nature Plants 5:253−57

    doi: 10.1038/s41477-019-0382-3

    CrossRef   Google Scholar

    [77]

    Huang J, Ju Y, Wang X, Zhang Q, Sodmergen. 2015. A one-step rectification of sperm cell targeting ensures the success of double fertilization. Journal of Integrative Plant Biology 57:496−503

    doi: 10.1111/jipb.12322

    CrossRef   Google Scholar

    [78]

    Mori T, Kuroiwa H, Higashiyama T, Kuroiwa T. 2006. GENERATIVE CELL SPECIFIC 1 is essential for angiosperm fertilization. Nature Cell Biology 8:64−71

    doi: 10.1038/ncb1345

    CrossRef   Google Scholar

    [79]

    Liu Y, Tewari R, Ning J, Blagborough AM, Garbom S, et al. 2008. The conserved plant sterility gene HAP2 functions after attachment of fusogenic membranes in Chlamydomonas and Plasmodium gametes. Genes & Development 22:1051−68

    doi: 10.1101/gad.1656508

    CrossRef   Google Scholar

    [80]

    Fedry J, Forcina J, Legrand P, Péhau-Arnaudet G, Haouz A, et al. 2018. Evolutionary diversification of the HAP2 membrane insertion motifs to drive gamete fusion across eukaryotes. PLoS Biology 16:e2006357

    doi: 10.1371/journal.pbio.2006357

    CrossRef   Google Scholar

    [81]

    Kielian M, Rey FA. 2006. Virus membrane-fusion proteins: more than one way to make a hairpin. Nature Reviews Microbiology 4:67−76

    doi: 10.1038/nrmicro1326

    CrossRef   Google Scholar

    [82]

    Wang W, Xiong H, Zhao P, Peng X, Sun MX. 2022. DMP8 and 9 regulate HAP2/GCS1 trafficking for the timely acquisition of sperm fusion competence. PNAS 119:e2207608119

    doi: 10.1073/pnas.2207608119

    CrossRef   Google Scholar

    [83]

    Zhang J, Yin J, Luo J, Tang D, Zhu X, et al. 2022. Construction of homozygous diploid potato through maternal haploid induction. aBIOTECH 3:163−68

    doi: 10.1007/s42994-022-00080-7

    CrossRef   Google Scholar

    [84]

    Li Y, Li D, Xiao Q, Wang H, Wen J, et al. 2022. An in planta haploid induction system in Brassica napus. Journal of Integrative Plant Biology 64:1140−44

    doi: 10.1111/jipb.13270

    CrossRef   Google Scholar

    [85]

    Zhong Y, Wang Y, Chen B, Liu J, Wang D, et al. 2022. Establishment of a dmp based maternal haploid induction system for polyploid Brassica napus and Nicotiana tabacum. Journal of Integrative Plant Biology 64:1281−94

    doi: 10.1111/jipb.13244

    CrossRef   Google Scholar

    [86]

    Zhong Y, Chen B, Li M, Wang D, Jiao Y, et al. 2020. A DMP-triggered in vivo maternal haploid induction system in the dicotyledonous Arabidopsis. Nature Plants 6:466−72

    doi: 10.1038/s41477-020-0658-7

    CrossRef   Google Scholar

    [87]

    Zhong Y, Liu C, Qi X, Jiao Y, Wang D, et al. 2019. Mutation of ZmDMP enhances haploid induction in maize. Nature Plants 5:575−80

    doi: 10.1038/s41477-019-0443-7

    CrossRef   Google Scholar

    [88]

    Wang N, Xia X, Jiang T, Li L, Zhang P, et al. 2022. In planta haploid induction by genome editing of DMP in the model legume Medicago truncatula. Plant Biotechnology Journal 20:22−24

    doi: 10.1111/pbi.13740

    CrossRef   Google Scholar

    [89]

    Chen X, Li Y, Ai G, Chen J, Guo D, et al. 2023. Creation of a watermelon haploid inducer line via ClDMP3-mediated single fertilization of the central cell. Horticulture Research 10:uhad081

    doi: 10.1093/hr/uhad081

    CrossRef   Google Scholar

    [90]

    Zhang X, Zhang L, Zhang J, Jia M, Cao L, et al. 2022. Haploid induction in allotetraploid tobacco using DMPs mutation. Planta 255:98

    doi: 10.1007/s00425-022-03877-4

    CrossRef   Google Scholar

    [91]

    Zhao X, Yuan K, Liu Y, Zhang N, Yang L, et al. 2022. In vivo maternal haploid induction based on genome editing of DMP in Brassica oleracea. Plant Biotechnology Journal 20:2242−44

    doi: 10.1111/pbi.13934

    CrossRef   Google Scholar

    [92]

    Yu X, Zhang X, Zhao P, Peng X, Chen H, et al. 2021. Fertilized egg cells secrete endopeptidases to avoid polytubey. Nature 592:433−37

    doi: 10.1038/s41586-021-03387-5

    CrossRef   Google Scholar

    [93]

    Zhang X, Shi C, Li S, Zhang B, Luo P, et al. 2023. A female in vivo haploid-induction system via mutagenesis of egg cell-specific peptidases. Molecular Plant 16:471−80

    doi: 10.1016/j.molp.2023.01.001

    CrossRef   Google Scholar

    [94]

    Jiang J, Stührwohldt N, Liu T, Huang Q, Li L, et al. 2022. Egg cell-secreted aspartic proteases ECS1/2 promote gamete attachment to prioritize the fertilization of egg cells over central cells in Arabidopsis. Journal of Integrative Plant Biology 64:2047−59

    doi: 10.1111/jipb.13371

    CrossRef   Google Scholar

    [95]

    Zhang X, Yu X, Shi C, Dresselhaus T, Sun MX. 2023. Do egg cell-secreted aspartic proteases promote gamete attachment? Journal of Integrative Plant Biology 65:3−6

    doi: 10.1111/jipb.13447

    CrossRef   Google Scholar

    [96]

    Jiang J, Qu LJ. 2023. Response to Zhang et al., 'do egg cell-secreted aspartic proteases promote gamete attachment?' Journal of Integrative Plant Biology 65:7−9

    doi: 10.1111/jipb.13448

    CrossRef   Google Scholar

    [97]

    Marimuthu MPA, Jolivet S, Ravi M, Pereira L, Davda JN, et al. 2011. Synthetic clonal reproduction through seeds. Science 331:876−76

    doi: 10.1126/science.1199682

    CrossRef   Google Scholar

    [98]

    Xiong J, Hu F, Ren J, Huang Y, Liu C, Wang K. 2023. Synthetic apomixis: the beginning of a new era. Current Opinion in Biotechnology 79:102877

    doi: 10.1016/j.copbio.2022.102877

    CrossRef   Google Scholar

    [99]

    Ravi M, Chan SWL. 2010. Haploid plants produced by centromere-mediated genome elimination. Nature 464:615−18

    doi: 10.1038/nature08842

    CrossRef   Google Scholar

    [100]

    Wang N, Dawe RK. 2018. Centromere size and its relationship to haploid formation in plants. Molecular Plant 11:398−406

    doi: 10.1016/j.molp.2017.12.009

    CrossRef   Google Scholar

    [101]

    Lv J, Yu K, Wei J, Gui H, Liu C, et al. 2020. Generation of paternal haploids in wheat by genome editing of the centromeric histone CENH3. Nature Biotechnology 38:1397−401

    doi: 10.1038/s41587-020-0728-4

    CrossRef   Google Scholar

    [102]

    Maheshwari S, Tan EH, West A, Franklin FCH, Comai L, Chan SWL. 2015. Naturally occurring differences in CENH3 affect chromosome segregation in zygotic mitosis of hybrids. PLoS Genetics 11:e1004970

    doi: 10.1371/journal.pgen.1004970

    CrossRef   Google Scholar

    [103]

    Kuppu S, Tan EH, Nguyen H, Rodgers A, Comai L, et al. 2015. Point mutations in centromeric histone induce post-zygotic incompatibility and uniparental inheritance. PLoS Genetics 11:e1005494

    doi: 10.1371/journal.pgen.1005494

    CrossRef   Google Scholar

    [104]

    Kelliher T, Starr D, Wang W, McCuiston J, Zhong H, et al. 2016. Maternal haploids are preferentially induced by CENH3-tailswap transgenic complementation in maize. Frontiers in Plant Science 7:414

    doi: 10.3389/fpls.2016.00414

    CrossRef   Google Scholar

    [105]

    Wang Z, Chen M, Yang H, Hu Z, Yu Y, et al. 2023. A simple and highly efficient strategy to induce both paternal and maternal haploids through temperature manipulation. Nature Plants 9:699−705

    doi: 10.1038/s41477-023-01389-x

    CrossRef   Google Scholar

    [106]

    Inagaki MN, Hash CT. 1998. Production of haploids in bread wheat, durum wheat and hexaploid triticale crossed with pearl millet. Plant Breeding 117:485−87

    doi: 10.1111/j.1439-0523.1998.tb01978.x

    CrossRef   Google Scholar

    [107]

    Inagaki M, Mujeeb-Kazi A. 1995. Comparison of polyhaploid production frequencies in crosses of hexaploid wheat with maize, pearl millet and sorghuml. Japanese Journal of Breeding 45:157−61

    doi: 10.1270/jsbbs1951.45.157

    CrossRef   Google Scholar

    [108]

    Kasha KJ, Kao KN. 1970. High frequency haploid production in barley (Hordeum vulgare L.). Nature 225:874−76

    doi: 10.1038/225874a0

    CrossRef   Google Scholar

    [109]

    Laurie DA. 1989. The frequency of fertilization in wheat × pearl millet crosses. Genome 32:1063−67

    doi: 10.1139/g89-554

    CrossRef   Google Scholar

    [110]

    Gernand D, Rutten T, Varshney A, Rubtsova M, Prodanovic S, et al. 2005. Uniparental chromosome elimination at mitosis and interphase in wheat and pearl millet crosses involves micronucleus formation, progressive heterochromatinization, and DNA fragmentation. The Plant Cell 17:2431−38

    doi: 10.1105/tpc.105.034249

    CrossRef   Google Scholar

    [111]

    Mochida K, Tsujimoto H, Sasakuma T. 2004. Confocal analysis of chromosome behavior in wheat × maize zygotes. Genome 47:199−205

    doi: 10.1139/g03-123

    CrossRef   Google Scholar

    [112]

    Coe EH Jr. 1959. A line of maize with high haploid frequency. The American Naturalist 93:381−82

    doi: 10.1086/282098

    CrossRef   Google Scholar

    [113]

    Liu C, Li X, Meng D, Zhong Y, Chen C, et al. 2017. A 4-bp insertion at ZmPLA1 encoding a putative phospholipase a generates haploid induction in maize. Molecular Plant 10:520−22

    doi: 10.1016/j.molp.2017.01.011

    CrossRef   Google Scholar

    [114]

    Gilles LM, Khaled A, Laffaire JB, Chaignon S, Gendrot G, et al. 2017. Loss of pollen-specific phospholipase NOT LIKE DAD triggers gynogenesis in maize. The EMBO Journal 36:707−17

    doi: 10.15252/embj.201796603

    CrossRef   Google Scholar

    [115]

    Dong X, Xu X, Miao J, Li L, Zhang D, et al. 2013. Fine mapping of qhir1 influencing in vivo haploid induction in maize. Theoretical and Applied Genetics 126:1713−20

    doi: 10.1007/s00122-013-2086-9

    CrossRef   Google Scholar

    [116]

    Kelliher T, Starr D, Richbourg L, Chintamanani S, Delzer B, et al. 2017. MATRILINEAL, a sperm-specific phospholipase, triggers maize haploid induction. Nature 542:105−9

    doi: 10.1038/nature20827

    CrossRef   Google Scholar

    [117]

    Li X, Meng D, Chen S, Luo H, Zhang Q, et al. 2017. Single nucleus sequencing reveals spermatid chromosome fragmentation as a possible cause of maize haploid induction. Nature Communications 8:991

    doi: 10.1038/s41467-017-00969-8

    CrossRef   Google Scholar

    [118]

    Liu H, Wang K, Jia Z, Gong Q, Lin Z, et al. 2019. Efficient induction of haploid plants in wheat by editing of TaMTL using an optimized Agrobacterium-mediated CRISPR system. Journal of Experimental Botany 71:1337−49

    doi: 10.1093/jxb/erz529

    CrossRef   Google Scholar

    [119]

    Cheng Z, Sun Y, Yang S, Zhi H, Yin T, et al. 2021. Establishing in planta haploid inducer line by edited SiMTL in foxtail millet (Setaria italica). Plant Biotechnology Journal 19:1089−91

    doi: 10.1111/pbi.13584

    CrossRef   Google Scholar

    [120]

    La Camera S, Geoffroy P, Samaha H, Ndiaye A, Rahim G, et al. 2005. A pathogen-inducible patatin-like lipid acyl hydrolase facilitates fungal and bacterial host colonization in Arabidopsis. The Plant Journal 44:810−25

    doi: 10.1111/j.1365-313X.2005.02578.x

    CrossRef   Google Scholar

    [121]

    Jang JH, Seo HS, Widiez T, Lee OR. 2023. Loss-of-function of gynoecium-expressed phospholipase pPLAIIγ triggers maternal haploid induction in Arabidopsis. New Phytologist 238:1813−24

    doi: 10.1111/nph.18898

    CrossRef   Google Scholar

    [122]

    Li Y, Lin Z, Yue Y, Zhao H, Fei X, et al. 2021. Loss-of-function alleles of ZmPLD3 cause haploid induction in maize. Nature Plants 7:1579−88

    doi: 10.1038/s41477-021-01037-2

    CrossRef   Google Scholar

    [123]

    Conner JA, Mookkan M, Huo H, Chae K, Ozias-Akins P. 2015. A parthenogenesis gene of apomict origin elicits embryo formation from unfertilized eggs in a sexual plant. PNAS 112:11205−10

    doi: 10.1073/pnas.1505856112

    CrossRef   Google Scholar

    [124]

    Chen B, Maas L, Figueiredo D, Zhong Y, Reis R, et al. 2022. BABY BOOM regulates early embryo and endosperm development. PNAS 119:e2201761119

    doi: 10.1073/pnas.2201761119

    CrossRef   Google Scholar

    [125]

    Horstman A, Willemsen V, Boutilier K, Heidstra R. 2014. AINTEGUMENTA-LIKE proteins: hubs in a plethora of networks. Trends in Plant Science 19:146−57

    doi: 10.1016/j.tplants.2013.10.010

    CrossRef   Google Scholar

    [126]

    Akiyama Y, Goel S, Conner JA, Hanna WW, Yamada-Akiyama H, Ozias-Akins P. 2011. Evolution of the apomixis transmitting chromosome in Pennisetum. BMC Evolutionary Biology 11:289

    doi: 10.1186/1471-2148-11-289

    CrossRef   Google Scholar

    [127]

    Conner JA, Goel S, Gunawan G, Cordonnier-Pratt MM, Johnson VE, et al. 2008. Sequence analysis of bacterial artificial chromosome clones from the apospory-specific genomic region of Pennisetum and Cenchrus. Plant Physiology 147:1396−411

    doi: 10.1104/pp.108.119081

    CrossRef   Google Scholar

    [128]

    Conner JA, Ozias-Akins P. 2017. Apomixis: engineering the ability to harness hybrid vigor in crop plants. Methods in Molecular Biology 1669:17−34

    doi: 10.1007/978-1-4939-7286-9_2

    CrossRef   Google Scholar

    [129]

    Zhang Z, Conner J, Guo Y, Ozias-Akins P. 2020. Haploidy in tobacco induced by PsASGR-BBML transgenes via parthenogenesis. Genes 11:1072

    doi: 10.3390/genes11091072

    CrossRef   Google Scholar

    [130]

    Conner JA, Podio M, Ozias-Akins P. 2017. Haploid embryo production in rice and maize induced by PsASGR-BBML transgenes. Plant Reproduction 30:41−52

    doi: 10.1007/s00497-017-0298-x

    CrossRef   Google Scholar

    [131]

    Horstman A, Bemer M, Boutilier K. 2017. A transcriptional view on somatic embryogenesis. Regeneration 4:201−16

    doi: 10.1002/reg2.91

    CrossRef   Google Scholar

    [132]

    Khanday I, Santos-Medellín C, Sundaresan V. 2023. Somatic embryo initiation by rice BABY BOOM1 involves activation of zygote-expressed auxin biosynthesis genes. New Phytologist 238:673−87

    doi: 10.1111/nph.18774

    CrossRef   Google Scholar

    [133]

    Vernet A, Meynard D, Lian Q, Mieulet D, Gibert O, et al. 2022. High-frequency synthetic apomixis in hybrid rice. Nature Communications 13:7963

    doi: 10.1038/s41467-022-35679-3

    CrossRef   Google Scholar

    [134]

    Wei X, Liu C, Chen X, Lu H, Wang J, et al. 2023. Synthetic apomixis with normal hybrid rice seed production. Molecular Plant 16:489−92

    doi: 10.1016/j.molp.2023.01.005

    CrossRef   Google Scholar

    [135]

    Underwood CJ, Vijverberg K, Rigola D, Okamoto S, Oplaat C, et al. 2022. A PARTHENOGENESIS allele from apomictic dandelion can induce egg cell division without fertilization in lettuce. Nature Genetics 54:84−93

    doi: 10.1038/s41588-021-00984-y

    CrossRef   Google Scholar

    [136]

    Borges F, Gomes G, Gardner R, Moreno N, McCormick S, et al. 2008. Comparative transcriptomics of Arabidopsis sperm cells. Plant Physiology 148:1168−81

    doi: 10.1104/pp.108.125229

    CrossRef   Google Scholar

    [137]

    Hand ML, Koltunow AMG. 2014. The genetic control of apomixis: asexual seed formation. Genetics 197:441−50

    doi: 10.1534/genetics.114.163105

    CrossRef   Google Scholar

    [138]

    Noyes RD, Rieseberg LH. 2000. Two independent loci control agamospermy (apomixis) in the triploid flowering plant Erigeron annuus. Genetics 155:379−90

    doi: 10.1093/genetics/155.1.379

    CrossRef   Google Scholar

    [139]

    Bicknell RA, Borst NK, Koltunow AM. 2000. Monogenic inheritance of apomixis in two Hieracium species with distinct developmental mechanisms. Heredity 84:228−37

    doi: 10.1046/j.1365-2540.2000.00663.x

    CrossRef   Google Scholar

    [140]

    Tas ICQ, van Dijk PJ. 1999. Crosses between sexual and apomictic dandelions (Taraxacum). I. The inheritance of apomixis. Heredity 83(Pt 6):707−14

    doi: 10.1046/j.1365-2540.1999.00619.x

    CrossRef   Google Scholar

    [141]

    Henderson ST, Johnson SD, Eichmann J, Koltunow AMG. 2017. Genetic analyses of the inheritance and expressivity of autonomous endosperm formation in Hieracium with different modes of embryo sac and seed formation. Annals of Botany 119:1001−10

    doi: 10.1093/aob/mcw262

    CrossRef   Google Scholar

    [142]

    Guitton AE, Page DR, Chambrier P, Lionnet C, Faure JE, et al. 2004. Identification of new members of Fertilisation Independent Seed Polycomb Group pathway involved in the control of seed development in Arabidopsis thaliana. Development 131:2971−81

    doi: 10.1242/dev.01168

    CrossRef   Google Scholar

    [143]

    Grossniklaus U, Vielle-Calzada JP, Hoeppner MA, Gagliano WB. 1998. Maternal control of embryogenesis by MEDEA, a Polycomb group gene in Arabidopsis. Science 280:446−50

    doi: 10.1126/science.280.5362.446

    CrossRef   Google Scholar

    [144]

    Luo M, Bilodeau P, Koltunow A, Dennis ES, Peacock WJ, et al. 1999. Genes controlling fertilization-independent seed development in Arabidopsis thaliana. Proceedings of the National Academy of Sciences 96:296−301

    doi: 10.1073/pnas.96.1.296

    CrossRef   Google Scholar

    [145]

    Ohad N, Yadegari R, Margossian L, Hannon M, Michaeli D, et al. 1999. Mutations in FIE, a WD polycomb group gene, allow endosperm development without fertilization. The Plant Cell 11:407−15

    doi: 10.1105/tpc.11.3.407

    CrossRef   Google Scholar

    [146]

    Köhler C, Hennig L, Bouveret R, Gheyselinck J, Grossniklaus U, et al. 2003. Arabidopsis MSI1 is a component of the MEA/FIE Polycomb group complex and required for seed development. The EMBO Journal 22:4804−14

    doi: 10.1093/emboj/cdg444

    CrossRef   Google Scholar

    [147]

    Luo M, Bilodeau P, Dennis ES, Peacock WJ, Chaudhury A. 2000. Expression and parent-of-origin effects for FIS2, MEA, and FIE in the endosperm and embryo of developing Arabidopsis seeds. PNAS 97:10637−42

    doi: 10.1073/pnas.170292997

    CrossRef   Google Scholar

    [148]

    Chaudhury AM, Ming L, Miller C, Craig S, Dennis ES, et al. 1997. Fertilization-independent seed development in Arabidopsis thaliana. PNAS 94:4223−28

    doi: 10.1073/pnas.94.8.4223

    CrossRef   Google Scholar

    [149]

    Cheng X, Pan M, E Z, Zhou Y, Niu B, et al. 2020. Functional divergence of two duplicated Fertilization Independent Endosperm genes in rice with respect to seed development. The Plant Journal 104:124−37

    doi: 10.1111/tpj.14911

    CrossRef   Google Scholar

    [150]

    Tonosaki K, Ono A, Kunisada M, Nishino M, Nagata H, et al. 2020. Mutation of the imprinted gene OsEMF2a induces autonomous endosperm development and delayed cellularization in rice. The Plant Cell 33:85−103

    doi: 10.1093/plcell/koaa006

    CrossRef   Google Scholar

    [151]

    Springer NM, Danilevskaya ON, Hermon P, Helentjaris TG, Phillips RL, et al. 2002. Sequence relationships, conserved domains, and expression patterns for maize homologs of the polycomb group genes E(z), esc, and E(Pc). Plant Physiology 128:1332−45

    doi: 10.1104/pp.010742

    CrossRef   Google Scholar

    [152]

    Luo M, Platten D, Chaudhury A, Peacock WJ, Dennis ES. 2009. Expression, imprinting, and evolution of rice homologs of the polycomb group genes. Molecular Plant 2:711−23

    doi: 10.1093/mp/ssp036

    CrossRef   Google Scholar

    [153]

    Kapazoglou A, Tondelli A, Papaefthimiou D, Ampatzidou H, Francia E, et al. 2010. Epigenetic chromatin modifiers in barley: IV. The study of barley Polycomb group (PcG) genes during seed development and in response to external ABA. BMC Plant Biology 10:73

    doi: 10.1186/1471-2229-10-73

    CrossRef   Google Scholar

    [154]

    Li S, Zhou B, Peng X, Kuang Q, Huang X, et al. 2014. OsFIE2 plays an essential role in the regulation of rice vegetative and reproductive development. New Phytologist 201:66−79

    doi: 10.1111/nph.12472

    CrossRef   Google Scholar

    [155]

    Derkacheva M, Hennig L. 2013. Variations on a theme: Polycomb group proteins in plants. Journal of Experimental Botany 65:2769−84

    doi: 10.1093/jxb/ert410

    CrossRef   Google Scholar

    [156]

    Hsieh TF, Shin J, Uzawa R, Silva P, Cohen S, et al. 2011. Regulation of imprinted gene expression in Arabidopsis endosperm. Proceedings of the National Academy of Sciences 108:1755−62

    doi: 10.1073/pnas.1019273108

    CrossRef   Google Scholar

    [157]

    Figueiredo DD, Batista RA, Roszak PJ, Köhler C. 2015. Auxin production couples endosperm development to fertilization. Nature Plants 1:15184

    doi: 10.1038/nplants.2015.184

    CrossRef   Google Scholar

    [158]

    Xu X, E Z, Zhang D, Yun Q, Zhou Y, et al. 2020. OsYUC11-mediated auxin biosynthesis is essential for endosperm development of rice. Plant Physiology 185:934−50

    doi: 10.1093/plphys/kiaa057

    CrossRef   Google Scholar

    [159]

    Sun X, Ling S, Lu Z, Ouyang Y, Liu S, Yao J. 2014. OsNF-YB1, a rice endosperm-specific gene, is essential for cell proliferation in endosperm development. Gene 551:214−21

    doi: 10.1016/j.gene.2014.08.059

    CrossRef   Google Scholar

    [160]

    Bernardi J, Lanubile A, Li Q-B, Kumar D, Kladnik A, et al. 2012. Impaired auxin biosynthesis in the defective endosperm18 mutant is due to mutational loss of expression in the ZmYuc1 gene encoding endosperm-specific YUCCA1 protein in maize. Plant Physiology 160:1318−28

    doi: 10.1104/pp.112.204743

    CrossRef   Google Scholar

    [161]

    Luo M, Wu X, Xie L, Sun X, Wang N, et al. 2023. Polycomb Repressive Complex 2 (PRC2) suppresses asexual embryo and autonomous endosperm formation in rice. Research Square Preprint

    doi: 10.21203/rs.3.rs-1087314/v1

    CrossRef   Google Scholar

    [162]

    Xiong H, Wang W, Sun MX. 2021. Endosperm development is an autonomously programmed process independent of embryogenesis. The Plant Cell 33:1151−60

    doi: 10.1093/plcell/koab007

    CrossRef   Google Scholar

    [163]

    Zhang T. 2021. Autonomous endosperm development in embryo-free seeds. The Plant Cell 33:1091−92

    doi: 10.1093/plcell/koab009

    CrossRef   Google Scholar

    [164]

    Duan B, Zhou C, Zhu C, Yu Y, Li G, et al. 2019. Model-based understanding of single-cell CRISPR screening. Nature Communications 10:2233

    doi: 10.1038/s41467-019-10216-x

    CrossRef   Google Scholar

    [165]

    Lo A, Qi L. 2017. Genetic and epigenetic control of gene expression by CRISPR-Cas systems. F1000Research 6:747

    doi: 10.12688/f1000research.11113.1

    CrossRef   Google Scholar

  • Cite this article

    Li S, Wang J, Jia S, Wang K, Li H. 2023. Synthetic apomixis: from genetic basis to agricultural application. Seed Biology 2:10 doi: 10.48130/SeedBio-2023-0010
    Li S, Wang J, Jia S, Wang K, Li H. 2023. Synthetic apomixis: from genetic basis to agricultural application. Seed Biology 2:10 doi: 10.48130/SeedBio-2023-0010

Figures(3)

Article Metrics

Article views(2804) PDF downloads(614)

REVIEW   Open Access    

Synthetic apomixis: from genetic basis to agricultural application

Seed Biology  2 Article number: 10  (2023)  |  Cite this article

Abstract: Apomixis in plants is a widely existing biological phenomenon, in which seeds are formed without egg cell and sperm uniting. Hybrid breeding exploits heterosis to obtain seeds with superior traits. However, segregation of traits in the offspring greatly limits the widespread use of hybrid vigor in agricultural production. Synthetic apomixis is considered a desired way of clonal propagation of heterozygous maternal parents, which bypasses laborious hybrid process. In recent years, with the increasing understanding of the molecular mechanisms of plant meiosis and double fertilization, scientists have introduced apomixis to rice and hybrid rice varieties by genetic engineering of genes that are involved in sexual reproduction. In this review article, we will summarize the recent research progress in the meiosis and double fertilization related to synthetic apomixis and provide perspectives on the potential application of synthetic apomixis in different crops and livestock pastures.

    • Apomixis is a natural phenomenon in some flowering plants that results in clonal seeds through bypassing the normal meiosis and double fertilization[1]. In apomicts (the species able to propagate through apomixis), the seeds can develop from the embryo sac formed by somatic ovule cell (apospory) or megaspore mother cell (MMC) without meiotic division (diplospory)[1]. Apospory and diplospory are collectively referred to as gametophytic apomixis[1]. In Citrus and Fortunella, the nucellus or integument of ovule can develop into adventitious embryo (sporophytic apomixis)[2,3]. These asexual reproduction modes result in offspring that are genetically identical to the maternal parent plant (Fig. 1). Since the first documentation of apomixis in Alchornea ilicifolia by Smith in 1841[4], apomixis has now been found in more than 400 species, most of which are tropical and temperate plants[5]. However, no staple crop species have been found to propagate in the way of apomixis[1,6,7] .

      Figure 1. 

      Schematic representation of sexual and apomictic seed formation in flowering plants. Apomixis deviates from sexual propagation mainly in meiosis, mitosis and double fertilization, which are the primary developmental stages in forming seeds. This diagram depicts the (a) normal sexual reproduction process, and the asexual reproduction processes of apomixis, (b) including apospory, (c) diplospory, and (d) adventitious embryony in flowering plants. Arrows indicate the developmental sequences for seed production. Red crosses mark the bypass of double fertilization in gametophytic apomixis (apospory and diplospory). In gametophytic apomixis (b & c), diplospory produces diploid embryo sac through mitotic division from MMC (red), while the unreduced embryo sac in apospory initiates from a different ovule precursor cell (blue) that divides mitotically, and the unreduced egg cell (b & c) can further develop into embryo without gamete fusion. However, the endosperm formation of apospory and diplospory may need central cell fertilization. In (d) adventitious embryony, embryos initiate directly from somatic cells of the nucellus or integument that are adjacent to the embryo sac, and endosperm develops from the fertilized central cell.

      Nowadays, heterosis is widely exploited in modern plant breeding[8], however, the trait segregation of superior hybrid seeds in the progeny greatly limits the agricultural application of hybrid vigor. The breeders need to perform laborious cross-breeding yearly to harvest the F1 hybrid seeds with the desired characteristics. Considering the clonal-propagation properties of apomixis, apomixis has long been the research hotspot, which is thought to be the ideal propagation manner in agricultural practice allowing the instant fixation of superior traits[9]. In this context, great efforts have been exerted in dissecting the underlying molecular mechanism of apomixis and introducing the apomixis to the staple crop species.

      The introgression strategy was the early method used to introduce apomixis from the wild relatives to staple crops. However, attempts in developing apomictic maize through introgression of Tripsacum dactyloides to Zea mays has yielded few successes despite of the elaborate and laborious hybridizing and backcrossing[10]. Therefore, another strategy termed synthetic clonal reproduction (also called synthetic apomixis), has been proposed[11]. In synthetic apomixis, the normal meiosis is circumvented to produce unrecombined and unreduced diploid egg cells, and double fertilization is replaced by parthenogenetic development[11,12].

      In the last decades, great advances have been made in understanding the molecular mechanisms of plant sexual reproduction, particularly the meiosis and double fertilization in plants[1315]. Apomeiosis, as one key hallmark of apomixis, has been realized through combining several meiosis-specific gene mutations in Arabidopsis and rice[12,16]. Recently, synthetic apomixis has been engineered in rice by combining Mitosis instead of Meiosis (MiMe) with the mutation of a sperm-specific gene MATRILINEAL (MTL) or ectopic expression of paternal gene BABY BOOM1 (BBM1) in the egg cell, which enables the clonal reproduction of F1 hybrids through seeds and stable transmission of heterotic phenotypes over generations[17,18]. These revolutionary achievements in rice implies bright prospect in applying synthetic apomixis to other staple crops.

      In this review, we will briefly summarize the characterized genes in meiosis, double fertilization and chromosome stability that could be engineered for synthetic apomixis. Moreover, we will discuss the potential opportunities and challenges in the future application of synthetic apomixis in crops.

    • Sexual reproduction is the ubiquitous and dominant propagation mode in angiosperms[15]. In higher plants, sexual reproduction refers to the formation of seeds from the zygote through the union of haploid male and female gametes[15]. To achieving this, sequential events, including meiosis-dependent haploid gametophyte formation, pollen-pistil interaction and the double fertilization, need to occur in an error-free manner[1315].

    • Meiosis is a specialized cell division that halves the ploidy of the germ cells (Fig. 2a). During meiosis (meiosis I and II), the chromosomes of MMC undergoes one round of replication but two rounds of chromosome segregation. The segregation of paired homologous chromosomes during meiosis I and then the following segregation of sister chromatids during meiosis II result in the formation of four haploid cells after cytokinesis (Fig. 2a). The core process of meiosis is characterized by the involvement of many genes (for reviews see[13,19]). In this context, only the pathways and genes of meiotic recombination and segregation of chromosomes that are directly associated with synthetic apomixis will be focused.

      Figure 2. 

      Schematic representation of meiosis in wild type and MiMe plants. Meiosis in wild type plant is depicted in (a), and (b) exhibits the altered meiosis in MiMe plants. In normal meiosis, the chromosome undergoes one replication, but twice divisions (meiosis I and II). During meiotic prophase I, the homologous recombination takes places with exchange of genetical information between homologous chromosomes. After meiosis II, four haploid recombinational gametes are formed from one megasporocyte (a). However, only two diploid clonal gametes are formed from one megasporocyte in MiMe system without recombination event and meiosis II (b).

    • The meiotic recombination is a highly complicated biological event that is characterized as a repair mechanism for programmed DNA double-strand-breaks (DSBs) at chromosomes during meiosis. Meiotic recombination initiates with the formation of DSBs, which is repaired by homologous recombination to form crossover (CO) or non-crossover (NCO) products[20].

      Meiotic recombination is regulated by many conserved genes in a sequential way[2023]. Sporulation-defective11 (SPO11) has been identified as the primary gene responsible for the formation of DSB in eukaryotes. SPO11 shares sequence similarity with the A subunit of Topoisomerase VI in Archaea[24]. The Arabidopsis genome harbors three homologs of SPO11, namely SPO11-1, SPO11-2, and SPO11-3, with SPO11-1 and SPO11-2 being involved in DSB formation, while SPO11-3 does not contribute to this process[25]. In rice, four homologs of SPO11 have been identified, with SPO11-1, SPO11-2, and SPO11-4 being implicated in DSB formation[26]. Moreover, one another meiotic topoisomerase VIB-like (MTOPVIB) protein that shows structural similarity with the Archaeal topo-VIB subunit, is also highly conserved in plants in regulating the DSB formation[27]. Similar with topoisomerase VI that functions in A2B2 heterotetramer formation, AtSPO11-1 and its homologs AtSPO11-2 in Arabidopsis are thought to form complex with MTOPVIB to catalyze the linkage of 5'-phosphotyrosine to induce DSBs[25].

      In Saccharomyces cerevisiae, function of SPO11 requires various proteins for the generation of DSBs, including RAD50, XRS2, REC102, SKI8/REC103, REC104, REC114, MER2, MEI4, and MRE[19]. In contrast to SPO11, these proteins exhibit limited conservation at the sequence level among different kingdoms. Moreover, even when the protein sequences are conserved, functional divergence is frequently observed. For instance, the meiotic role of SKI8/REC103 is conserved in S. cerevisiae and several fungi[19,28]. However, the homologs of SKI8/REC103 in Arabidopsis do not exhibit any meiotic functionality[28]. Although RAD50 plays a crucial role in DSB formation in S. cerevisiae, the orthologs of RAD50 in Saccharomyces pombe are not necessary for the generation of DSBs[29]. Moreover, MRE11 is essential for DSB processing but not for the initiation of SPO11-dependent meiotic DNA breaks in Arabidopsis[30]. Additionally, diverse DSB formation-related genes have also been identified in different species via genetic screening, like PUTATIVE RECOMBINATION INITIATION DEFECT 1, 2 and 3 (PRD1, 2 and 3), and DSB forming (DFO) in Arabidopsis[3135], and HOMOLOGOUS PAIRING ABERRATION IN RICE MEIOSIS1 (PAIR1), also called OsPRD3, Central Region Component 1 (CRC1), P31comet, and SOLO DANCERS (SDS) in rice[3639]. Mutants impaired in any of these core recombinant-specific genes display defects in DSB formation, indicating their indispensable role in meiotic recombination[13].

      After creation of DSBs by the SPO11 dimer, the nuclease MRE11 recruits RAD50 and XRS2/NBS1 to form MRX/MRN complexes, which associate with DSBs in eukaryotes[13,40]. In the presence of SAE2/COM1, MRX/MRN complexes process DSBs via endonucleolytic cleavage to produce asymmetrically spaced nicks flanking DSBs and exonucleolytic resection, yielding 3′-OH single-stranded DNA (ssDNA) ends[41]. Meanwhile, SPO11 is removed from the chromosome[42]. Replication Protein A (RPA) complexes bind to these ssDNA ends, which are then replaced by RAD51 and DMC1 to form nucleofilaments[43]. Subsequently, RAD51 and DMC1 targets the nucleofilaments for homology search and heteroduplex formation[43]. In most eukaryotes, there are two pathways involved in CO formation: ZMM pathway and non-ZMM pathway that form class I and II CO, respectively[13]. ZMM is the primary pathway, which consists of ZIP1, ZIP2, ZIP3, ZIP4 MER3, MSH4, and MSH5 in S. cerevisiae, and MSH4, MSH5, SHOC1, HEI10, and ZIP4 in Arabidopsis[13]. ZMM pathways account for more than 85% CO formation in eukaryotes. However, there are many other genes identified in yeast and plants to be implicated in CO formation, but not classified in ZMM pathways (for review see[13]). To date, less is known in regarding to class II CO formation. MUS81 is characterized in non-ZMM CO pathways in plants[44]. Lesion in MUS81 reduces CO frequency up to 10% in a wild-type background[44]. Collectively, COs generate genetic diversity in offspring, which is crucial for the survival and adaptation of a species[13]. The presence of the bivalent structure also allows homologous chromosomes to be equally separated during the first meiotic division[45]. When bivalent structure is absent between a pair of homologous chromosomes, unequilibrated segregation of the chromosomes may occur during the mid-prophase I to late prophase I of the first meiotic division[45]. However, meiotic recombination is detrimental to fix the superior traits in heterozygous hybrid seeds.

      In addition to COs, NCOs also contribute to meiotic DSB repair. Unlike COs, NCOs involve the copying of a small patch of intact homologous chromosome to the broken chromosome without reciprocal exchange[13]. In Arabidopsis, several components inhibiting CO formation have been identified, which participate in various NCO pathways. Generally, the successful completion of recombination requires the involvement of topoisomerase IIIα and its associated protein, AtBLAP75/AtRMI1[46,47]. Mutations in these genes result in SPO11-dependent connections between homologous chromosomes during metaphase I, independent of ZMM proteins[46,47]. In addition to topoisomerases, the helicase FANCM, along with its cofactors MHF1 and MHF2, as well as the RECQ4 helicase (a homolog of BLOOM/Sgs1), also acts to suppress CO formation[13,4850]. Notably, COs frequency increases dramatically in either fancm or recq4a/recq4b mutant lines, with these additional COs arising through the class II CO pathway[13,49,50]. Furthermore, FIDGTIN-L1 (FIGL1) represents an additional pathway that inhibits CO formation in a FANCM-independent manner[13]. Taken together, these findings suggest the existence of a delicate balance between NCOs and class II COs in plants.

    • As mentioned above, meiosis includes two rounds of segregation of chromosomes. At the end of meiosis I, the interlink between homologous chromosomes is disrupted, and the sister chromatids of homologous chromosomes that are separately connected by kinetochores are faithfully oriented and steered to opposite poles[13]. To date, meiosis-specific Recombination-deficient 8 (REC8) is found to be the exclusive meiosis-specific cohesion subunit in plants[51]. The plants with lesion in REC8 exhibit abnormal meiosis, including defects in DSB repairment and bipolar orientation of sister chromatids during meiosis I[51]. In rice, REC8 is also reported to be necessary for chromosome axis building and bouquet formation, homologous chromosome pairing and recombination[52]. Other cohesion subunits, including sister-chromatid cohesion protein 3 (SCC3), Structural Maintenance of Chromosome 1 (SMC1) and SMC3 are shared by both meiosis and mitosis to be responsible for separation of sister chromatids[51,53]. Additionally, it has been observed that SHUGOSHIN1 (SGO1) and Bub1-related kinase 1 (BRK1) are implicated in the protection of centromeric cohesion[54,55]. Notably, the mutants of SCC3, SGO1, and BRK1 have been found to induce untimely separation of sister chromatids, as documented in previous studies[51,54,55].

      Cyclin, cyclin-dependent kinases (CDKs) and the negative regulator anaphase-promoting complex (APC/C) constitute the conserved core regulatory module that ensures the faithful progression of the cell cycle for both meiosis and mitosis in eukaryotes[5658]. During cell division, cyclin associates with CDKs to activate its kinase activity, which is indispensable for the entry and progression of metaphase[59]. The progression of meiosis is closely correlated with the nuanced regulation of the activity of the cyclin-CDK module. For instance, the onset of anaphase requires rapid degradation of cyclin by APC/C to suppress CDK activity[58]. In contrast, the initiation of meiosis II through interkinesis requires moderate inhibition of the activity of cyclin-CDK. Intriguingly, the dramatic suppression of cyclin-CDK leads to the termination of meiosis. This cell cycle machinery is also highly conserved in plants[13].

      In Arabidopsis genome, several genes are associated with meiosis progression, including Tardy Asynchronous Meiosis (TAM), CDKs (CDKA;1, CDKB;1, CDKB1;2, CDKB2;1 and CDKB2;2), Three Division Mutation 1 (TDM1) and Omission of Second Division 1(OSD1)[6062]. During the first meiotic division, TAM, as an A-type cyclin associates with CDKA;1 to phosphorylate TDM1 to suppress its activity[62]. TDM1 is an APC/C component that is necessary for exit from meiosis II[63]. Additionally, TDM1 is supposed to be functional during the whole meiosis since it is always highly induced along with meiosis, while TAM is expressed only at meiosis I[60]. OSD1 is responsible for the onset of meiosis II by regulating cyclin-CDK activity in both rice and Arabidopsis[12,16]. Mutations in any of these genes induce the formation of polyploid gametes, suggesting that they are of great importance in maintaining normal meiosis progression.

    • After meiosis, the male and female gametophytes are formed after mitosis and cell differentiation[64,65]. Sperm cells encapsulated in the pollen are delivered to the ovule through the stigma and transmitting tissue to encounter egg cells for zygote formation[15]. A series of biological events, including pollen-stigma interaction, pollen tube germination and growth, pollen tube guidance, and pollen tube reception are carried out sequentially to ensure the successful selection and delivery of desired conspecific male gametes to egg cells[14]. Finally, double fertilization takes place in embryo sac when the two sperm cells unit with the egg cell and central cell to initiate the embryogenesis and endosperm development, respectively. To date, numerous advances have been made in deciphering the molecular underpinnings behind these processes (for review see[15,66,67]).

      In angiosperms, double fertilization is a highly ordered and sophisticated process that involves multiple steps, including gamete attachment, sperm activation, membrane fusion, plasmogamy, and karyogamy[66,68]. After release into the embryo sac, immobile sperm cells are steered towards the site of gamete fusion and attach to the female gametes[69]. The GAMETE EXPRESSED 2 (GEX2) gene, which encodes a sperm-specific transmembrane protein in Arabidopsis, has been reported to mediate the gamete attachment[70]. The conserved filamin-repeat domain of GEX2 forms an immunoglobulin (Ig)-like fold that is essential for gamete attachment and is also shared by FUS1 in Chlamydomonas reinhardtii (Chlamydomonas)[71]. Sperm cells with a lesion in GEX2 show defects in attaching to the egg and central cells, and the resultant double-fertilization failure and single fertilization typically cause aborted seeds[70,71].

      In the context of gamete fusion in flowering plants, little is known about the molecular mechanisms underlying the activation of sperm cells. In contrast, the signaling cascades initiated by gamete association prior to fusion have been more extensively studied in unicellular green algae and multicellular animals[72,73]. In Arabidopsis, upon the arrival of sperm cell, the membrane contact between the sperm and egg cells triggers the release of cysteine-rich EGG CELL 1 (EC1) by the egg cell through exocytosis to activate the sperm[74]. The secreted EC1 then triggers the translocation of the fusogen HAPLESS 2/GENERATIVE CELL SPECIFIC 1 (HAP2/GCS1) from endomembrane-associated sites to the surface of the sperm cell[74,75]. Furthermore, EC1 is reported to mediate the separation of the two physically linked sperm cells[76]. The positioning of sperm cells is also critical for the success of double fertilization. When the two sperms simultaneously attach to the egg cell, repositioning occurs with one sperm cell released and re-oriented towards the central cell and the other remains stagnant[77]. Proper adhesion of each sperm cell to one female gamete is then achieved[67,77].

      After adhesion, HAP2/GCS1, as the essential fusogen, acts as the plasma membrane merger for the male and female gametes[75,78]. In Chlamydomonas, hap2 mutant gametes can interact with female gametes and stopped at a 10 nm distance between the opposing membranes, failing in the further plasma membrane fusion[79]. The presence of orthologs of HAP2 in all eukaryotic clades, except fungi, indicates that HAP2 is a highly conserved gamete fusion protein in eukaryotes[80]. Interestingly, the structure of Chlamydomonas HAP2 exhibits striking similarities to class II viral fusion proteins, with three extracellular domains, a single transmembrane domain, and a smaller intracellular domain[81]. These class II viral fusion proteins homotrimerize upon entry into host endosomes, and a conformational change facilitates the fusion of the host membrane and viral envelope[81]. Similarly, recent structural analyses of Arabidopsis HAP2 have shown that an apical amphipathic helix is required for membrane insertion in vitro and fusion with the egg and central cell in vivo[80]. These discoveries suggest a model for double fertilization in Arabidopsis, where HAP2 functions as a unilateral gamete fusion protein, mediating the fusion of one sperm cell with the egg plasma membrane and the other with the central cell, without requiring a binding protein on the surface of the female gametes. However, this model raises intriguing questions, such as how the activity of HAP2 is regulated in preventing the premature sperm cell fusion and what is the underlying mechanism that guide the two distinct fusion events between sperm and female gametes.

      The absence of necessary genes for either gamete fusion or preventing polyspermy can result in haploid offspring in certain mutants, attributing to a double fertilization defect. Recent studies have identified two sperm cell-specific DOMAIN OF UNKOWN FUNCTION 679 membrane proteins (DMP8 and DMP9) that also facilitate gamete fusion, especially sperm-egg cell fusion[76]. The dmp8 dmp9 sperm cells are also defective in the secretion of HAP2 upon gamete activation by EC1[82]. Loss of DMP8/9 in Arabidopsis or the orthologs in maize, potato, tomato, Brassica napus, tobacco, watermelon, cabbage and Medicago truncatula causes maternal haploid offspring[8391]. In a recent study, the identification of two aspartic endopeptidases, EGG CELL-SECRETED 1 (ECS1) and ECS2, has shed light on the intricate spatial and temporal control of gamete interactions in Arabidopsis[92]. These peptidases are secreted into the extracellular space from the egg cells to regulate the nuclear and plasma membrane fusion between the male and female gametes[9396]. Interestingly, the generation of partially semigamous zygotes and then maternal haploid offspring containing only the genome of the egg has been observed in ecs1ecs2 double mutants in Arabidopsis and the rice esc1 mutant, through either selfing or hybridization[93].

      In summary, the genetic alteration of some of the genes involved in double fertilization causes haploid production. Critical roles in gamete attachment and fusion are played by GEX2, HAP2/GCS1, DMP8/9, EC1 and ECS1/2 proteins, while only knockout of DMP8/9 and ECS1/2 have been tested and verified to induce haploid generation. This raises the possibility that the entry or physical adhesion of the sperm triggers certain zygotic programs, although the haploid generation occurs at a low frequency (less than 10%). Another commonality between DMP8/9 and ECS1/2 is that they both control preferential fertilization of the egg cell[76,94]. A more profound comprehension of the molecular mechanisms underlying double fertilization in angiosperms would facilitate the development of parthenogenesis.

    • Scientists have long sought to introduce apomixis into non-apomictic staple crops, as this could offer significant agronomic benefits. Two feasible routes for achieving this goal have been proposed: introgression and engineering synthetic apomixis. To date, researchers have achieved great success in engineering synthetic apomixis in Arabidopsis and rice using hybridization or genome-editing technology[17,18,97]. Synthetic apomixis involves three principles. First, meiosis, which normally produces haploid female gametophytes, must be circumvented (Fig. 2b). Second, embryogenesis must take place in the absence of nuclear fusion between male and female gametes or with the occurrence of genome elimination after gametes fusion (Fig. 3). Lastly, viable endosperm must develop via an autonomous or pseudogamous mechanism, or normal central cell fertilization. To bypass meiosis, Mitosis instead of Meiosis (MiMe) strategy can be employed to induce unreduced and nonrecombinant maternal gametes[12,16]. Parthenogenesis and elimination of the paternal genome are two approaches that can be used to create progenies with uniparental genome.

      Figure 3. 

      Schematic representation of genome elimination and parthenogenesis. The haploid offspring can be induced through crossing plant carrying modified CENH3 gene with (a) wild type pollen, (b) mutating MTL or (c) DMP gene, or (d) nectopic expression of BBMs in the egg cell. In (a), genome elimination occurs in crosses between wild type pollen and CEHN3-based haploid inducer. The chromosome of haploid inducer is unstable during mitotic divisions of the early embryo, and will be lost to produce haploid plants, which are genetically identical to the genome of wild type parental gametophyte. (b) Depicts the haploid induction by chromosome fragmentation event. In mtl mutants, chromosomes undergo fragmentation with varied frequencies from microsporocyte to embryogenesis stage. During double fertilization, DMP is supposed to bridge the gamete fusion, however, mutating DMP results in gamete fusion failure and haploid production (c). Additionally, parthenogenesis can be induced in sexual plants by ectopic expression of BBMs in the egg cell (d).

    • The MiMe strategy involves mutating several selected core meiosis-related genes to disrupt chromosome recombination, induce separation of sister chromatids during meiosis I, and totally abolish the meiosis II[16]. Among these genes, SPO11-1, PRD1, PRD2, PRD3/PAIR1, MTOPVIB, DFO and P31comet, which play key roles in DSB formation during meiotic recombination, can be mutated to prevent recombination events in plants[98]. Premature sister chromatid segregation during the first meiotic division can be achieved by mutating the REC8 gene, which encodes the exclusive meiosis-specific cohesion subunit[98]. OSD1, TAM, and TDM1 are pivotal genes controlling meiotic cell cycle transition and lesions in any one of these three genes abolish the meiosis II[98]. Combination of mutations in single gene of each group have been proven successful in establishing MiMe in Arabidopsis and rice (for review see[98]).

    • After successful induction of apomeiosis in sexual plants through MiMe strategy, the next hurdle towards clonal seeds is to circumvent the formation of zygote. The two ways to achieve this include genome elimination and parthenogenesis. The major difference between gametophyte apomixis and genome elimination in producing clonal seeds is whether fertilization is required. In genome elimination, either maternal or paternal chromosome set will be lost during early embryogenesis. To date, several strategies have been found to cause genome elimination, including CENH3-based crossing (Fig. 3a), interspecific outcrossing (Fig. 3b), and intraspecific hybridization.

      CENH3 is a highly conserved centromeric histone H3 present in all eukaryotes[99]. It consists of a conserved C-terminal histone fold domain and a highly variable N-terminal tail. During cell cycle, CENH3 recruits proteins to assemble kinetochores at the centromere, where spindle fibers attach to guide accurate chromosome segregation in mitosis and meiosis[99]. Studies have shown that homozygous null mutants of Atcenh3-1 exhibit mid-global stage embryo development arrest[99]. Intriguingly, the embryo-lethal phenotype can be rescued by fusing CENH3 with green fluorescent protein (GFP), or by using GFP-tailswap, a variant of GFP-CENH3 with the N-terminal tail of CENH3 substituted with the N-terminal of histone H3.3[99]. When these rescued plants are used as male or female parents in crosses with wild type plants, haploid offspring with uniparental chromosome sets can be observed, transforming the plant into a flexible haploid induction system that can efficiently induce both maternal and paternal haploids, with the latter having an induction rate of over 20%[99]. The mechanism underlying this phenomenon is thought to be postzygotic incompatibility, where the parental chromosome set carrying the structurally altered CENH3 at its centromeres is mitotically unstable and is therefore left behind in early embryonic divisions[99]. Moreover, it has been proposed that mutations in CENH3 may impair chromatin loading, resulting in smaller centromeres that cannot compete with the larger centromeres of the crossing parent[100]. This size dimorphism of parental centromeres can cause early loss of chromosomes, resulting in haploid and aneuploid progenies[100,101]. These serendipitous discoveries have stimulated researches on the relationship between CENH3 and genome elimination.

      Further studies have shown that orthologs of CENH3 from different species of Brassicaceae or even Zea mays can complement the embryo-lethal phenotype of Atcenh3-1 and generate haploid offspring when crossed with wild type plants[102]. In addition, genetic research has revealed that a single-amino-acid missense mutation in CENH3 is sufficient to mimic the effects of chimeric CENH3 or CENH3 from diverged species in inducing genome elimination[103]. However, despite the efforts of numerous researchers, the anticipated success has not been achieved in most crops, with only a few crops, such as wheat and maize, having successfully generated haploid induction systems based on CENH3, but with lower induction rates than those in Arabidopsis[101,104]. Excitingly, crosses between mutant lines expressing CENH3 variants and MiMe or dyad mutants produce clonal seeds in Arabidopsis[97].

      In a recent study by Li et al., it was discovered that the GFP-tailswap haploid induction system is highly sensitive to environmental temperature in terms of both pollen viability and haploid induction ability, compared to wild-type plants[105]. Specifically, even slight increases in environmental temperature, from 22 to 25 °C, led to a near-total loss of GFP-tailswap pollen viability, while significantly enhancing haploid induction ability. Conversely, decreasing the temperature had the opposite effect[105]. These findings have important implications for the extension of CENH3-based haploid induction systems to other crops and provides valuable clues for future studies, emphasizing the need to consider the influence of environments. Moreover, the research sheds light on the fundamental biological issues of CENH3 in maintaining the precise separation control of chromosomes during cell division and the mechanism of haploid formation.

    • The phenomenon of uniparental genome elimination in interspecific outcrosses between monocots was first observed in 1970. Although the underlying mechanism is not yet fully understood, this phenomenon has been applied in agriculture to produce maternal haploid lines, such as triploid wheat, which is generated by crossing hexaploid wheat with maize, sorghum, or pearl millet[106,107]. Selective elimination of parental genomes can occur in interspecific hybridization, such as in the haploid offspring induced by the interspecific hybridization of barley (Hordeum vulgare L.)[108]. In crosses between wheat and pearl millet, the paternal genome of pearl millet is selectively eliminated[109], and a series of biological events, including paternal genome separation, chromosome structure rearrangement, micronucleus formation, and micronucleus degradation, are induced in this process[110]. In outcrosses between wheat and maize, a haploid wheat embryo forms, possibly due to the small genome size of maize and defective spindle attachment to the centromere[111], or defective gamete interactions. The formed wheat zygotic embryo gradually degrades after several divisions, and normal viable endosperm cannot form after successful fertilization of central cell due to post-zygotic cross barrier[111]. However, given the inadequate understanding of the underlying molecular mechanism, the use of outcross in producing clonal seeds of crops is still constrained.

    • The use of Stock6-derived haploid inducer lines for intraspecific hybridization is a commonly employed technique in maize breeding, as established by Coe[112]. The haploid induction capability of Stock-6 is due to the mutations in ZmPLA1 and ZmDMP [87,113]. ZmPLA1, which encodes a maize phospholipase that localized on the endo-plasma membrane (endo-PM) that wraps the sperm cells within the cytoplasm of the pollen tube[114], was initially identified as qhir1, a quantitative trait locus[115], and is also known as MATRILINEAL (MTL)[116] and NOT LIKE DAD (NLD)[114]. Single nucleus sequencing of the zmpla1 haploid inducer lines revealed an unexpectedly high frequency of sperm DNA fragmentation (Fig. 3b), which has been suggested to be the cause of paternal genome loss following fertilization and the subsequent production of maternal haploids[117]. However, the mechanism triggering DNA fragmentation remains unknown. Recent comprehensive omics analysis has revealed a possible mechanism involving an ROS burst for haploid induction in zmpla1[7], suggesting that a ROS burst causes an imbalance in the redox state of the metabolome, leading to sperm DNA breakage. Once the constraints of the DNA-repair machinery are overcome, the breakage may extend from the centromeric regions to the entirety of the sperm genome, ultimately leading to fragmentation. Subsequently, continuous DNA fragmentation leads to the loss of the male genome and/or chromosomes with defective centromeres, ultimately culminating in haploid induction following fertilization (Fig. 3b). Furthermore, through the use of omics analysis, a unique peroxidase, ZmPOD65, has been identified as being specific to pollen and serving as a novel gene that regulates haploid induction[7]. This discovery serves to reinforce the relationship between reactive oxygen species and haploid induction in zmpla1.

      MTL-based chromosome fragmentation has been utilized in generating haploid inducing lines in other staple crops. For instance, by mutating TaMTL in wheat using gene editing technique can create a haploid inducer line with a haploid induction rate (HIR) of up to 18.9%[118]. Haploid induction is also achieved in foxtail millet (Setaria italica) by gene editing of SiMTL[119]. In rice, the selfing progeny of mtl mutant includes haploids and doubled haploids with HIR of 4.44%. Moreover, simultaneously mutating OsMTL and three MiMe inducing genes REC8, PAIR1, and OSD1 can result in maternal clonal seeds, suggesting that the mechanism is likely to be conserved in plants[17]. However, expanding this method to Arabidopsis is hampered by the fact that the most relevant gene of MTL in Arabidopsis, AtPLP2, is only expressed in vegetative organs[120]. MTL-based genome elimination also relates to genome instability but occurs before fertilization. Therefore, searching for an endo-PM-specific phospholipase in dicots may be effective in haploid induction in dicot species. Intriguingly, recent studies indicate that the phospholipases expressed in pollen or gynoecium are also correlated to haploid induction. In Arabidopsis, null allele mutation of a gynoecium-expressed phospholipase All (pPLAIIγ) produces maternal haploid seeds, and the haploid induction in pplaIIγ is accompanied by the internalization of PIN1 at the plasma membrane of the basal funiculus[121]. In Zea mays, mutation line impaired in pollen-specific PHOSPHOLIPASE D3 (ZmPLD3) generates maternal haploids, which shows comparable haploid induction rate with mtl[122]. Additionally, mutating PLD3 in mtl background triples the haploid induction rate from 1.19% to 4.13%, indicating synergistic effect of PLD3 and MTL in haploid induction[122]. ZmPLD3 is highly conserved in other cereals, indicating the potential in developing PLD3-based haploid-inducer in other staple crops.

    • After generating a diploid clonal gamete, autonomous transition to embryonic development can be achieved through ectopic expression of the transcription factor BABY BOOM (BBM) or its orthologs in the egg cell of monocots or dicots[18,123,124]. BBMs belong to the eudicot AINTEGUMENTA (euANT) family and play a crucial role in various processes, including embryogenesis, root development, and somatic embryogenesis[125]. In Pennisetum squamulatum, several BBM-like (BBML) genes located in the apospory-specific genomic region (ASGR) control the parthenogenesis of Pennisetum squamulatum[126,127]. PsASGR-BBMLs are orthologs of BnBBM, and their ectopic expression in the egg cell can induce parthenogenesis in tobacco, pearl millet, rice and maize[18,123,128130]. Intriguingly, the expression of PsASGR-BBM1 driven by the Arabidopsis egg cell-specific AtDD45/EC1.2 promoter resulted in more autonomously developed egg cells in rice and maize than when driven by its native promoter[130]. However, this strategy was unsuccessful in Arabidopsis, and the native promoter of PsASGR-BBM1 has failed to be transcribed in Arabidopsis[130]. Recently, ectopic expression of BnBBM in the egg cell of Arabidopsis and the dicot crops Brassica napus and Solanum lycopersicon was suggested to initiate early embryogenesis in the absence of fertilization[124]. Moreover, haploid induction rates of the egg-cell-expressed BBM lines are even increased in crossing with dmp in the case of Arabidopsis and tomato, implying a synergistic effect between dmp and ectopic BBM expression on parthenogenesis[124].

      Rice has four OsBBM homologous genes, among them, BBM1, BBM3, and BBM4, are exclusively expressed in sperm cells before fertilization[18]. Ectopic expression of OsBBMs in rice egg cell can initiate parthenogenesis and result in viable seeds when the endosperm develops after the normal fertilization of the central cell[18]. As shown in other species (for review see[131]), the ability of OsASGR-BBM1/OsBBM1 in initiating somatic embryogenesis before fertilization was also confirmed[132]. The successful engineering of synthetic apomixis in rice involves the mutation of three MiMe-causing genes, namely PAIR1, REC8, and OSD1 using the CRISPR/Cas9 technique, and the ectopic expression of OsBBM1 in the egg cell[18]. Nonetheless, this strategy is beset with low seed setting rates and low clonal seed ratio, which considerably limit its implementation in crops. Subsequent research aimed at improving this approach reported the enhancement of cloning seed proportion to as high as 90% in the hybrid rice variety BRS-CIRAD 302[133]. This was achieved through the combination of ectopic OsBBM1 expression in the egg cell and the mutation of PAIR1, REC8, and OSD1 genes through a single vector[133]. A new synthetic apomixis strategy for rice that combines the ectopic expression of BBM4 with the MiMe strategy has been developed for hybrid rice, resulting in the generation of Fixation of hybrids 2 (Fix2) plants that exhibit normal growth during the vegetative growth stage[134]. Additionally, the Fix2 plants displayed a high seed setting rate of 80.9%−86.1%, similar to that of normal hybrid rice at 82.1%−86.6%[134]. Although the seed setting rate of Fix2 strategy was not affected, the cloning seed proportion was quite low (1.7%)[134]. Thus, further research efforts are needed to develop a synthetic apomixis strategy for hybrid rice with high seed-setting rates and high cloning seed induction rates.

      In addition to BBMs, a PARTHENOGENESIS (PAR) gene in apomictic dandelion has been cloned recently, which encodes a K2-2 zinc finger, ethylene-responsive element binding factor-associated amphiphilic repression (EAR)-domain protein[135]. Ectopic expression of PAR, under the control of AtEC1 promoter in sexual lettuce, induces haploid embryo formation in unfertilized embryo sac[135]. However, the haploid embryo undergoes abortion at later developmental stages due to the inadequate support from abnormal and nonsexual endosperm[135]. The recessive par gene are expressed in the pollen of sexual dandelion, and the two PAR homologs in Arabidopsis[135], DUO1 activated zinc finer 3 (DAZ3) and transcriptional repressor of EIN3-dependent ethylene-response 1 (TREE1), are among the highest expressed genes in Arabidopsis sperm cells[136]. Therefore, it is supposed that the function of PAR is to lift the autonomous inhibition on embryogenesis of unfertilized egg cell[135]. Overall, this study highlights the potential application of PAR homologs in inducing parthenogenesis in other dicots.

    • As the last step, the formation of viable endosperm is critical for the success of synthetic apomixis. In most natural apomicts, endosperm develops from the fertilized central cell[137], however, endosperm can also form autonomously in some apomictic species[138140]. It is worth noting that the bypassing of central cell fertilization is not essential for the induction of clonal seeds, and all reported instances of synthetic apomixis have relied on the normal fertilization of the central cell[17,18,97,133,134]. Nevertheless, the autonomous formation of endosperm represents the ultimate culmination towards the realization of fertilization-independent synthetic apomixis.

      In Hieracium, the occurrence of autonomous endosperm development represents a qualitative trait governed by one single dominant locus[141]. Additionally, this trait is regulated by multiple other unknown gene(s)[141], implying a complex underlying mechanism for autonomous endosperm formation in natural apomictic species. While progress in elucidating the mechanisms of autonomous endosperm formation in natural apomicts has been sluggish, considerable knowledge have been amassed regarding the induction of autonomous endosperm formation in sexual plants. Notably, the pivotal role is played by Fertilization-Independent Seed (FIS) class genes, which encode the proteins involved in the Fertilization-Independent Seed-Polycomb Repressive Complex2 (FIS-PRC2) in embryo sac[142147]. The FIS-PRC2 members include MEDEA (MEA), FIS2, FERTILIZATION INDEPENDENT ENDOSPERM (FIE), MULTICOPY SUPRESSOR OF IRA (MSI1), and BORGIA (BGA)[142146]. Disruptions in these FIS genes induce autonomous endosperm formation in sexual Arabidopsis[142145]. Moreover, the penetrance rate of autonomous endosperm formation, which estimates the percentage of ovules carrying these fis mutant, ranges from 41.2% to 92.4% in Arabidopsis[142]. However, fis mutants consistently exhibit abnormal endosperm development or embryo lethality[142146,148]. To date, the orthologs of FIS genes have been cloned in many cereal plants, such as rice, maize and barley[149154]. Encouragingly, mutations in many of these orthologs have been demonstrated to induce autonomous endosperm formation[149,150,154]. Further investigation reveals that the chromatin modifying complex FIS-PRC2 mediates the trimethylation of the histone H3 Lysine-27, thereby suppressing gene expression[155,156]. The causes underlying autonomous endosperm development in mutants of the PRC2 genes are likely attributed to the activation of the auxin biosynthesis pathway in the central cell of Arabidopsis and rice[155,157160].

      Interestingly, a recent study has shed light on the impact of mutations in two rice orthologs of FIE. These mutations have been found to abolish the suppression of asexual embryo development and autonomous endosperm formation prior to fertilization[161]. Moreover, it has been observed that the asexual embryo development is prompted by the autonomous endosperm, which also leads to the transcription of male-genome expressed genes such as OsBBM1, OsBBM2, and OsWUSHCEL-related homeobox 8/9 (OsWOX8/9) in the asexual embryo[161]. These findings suggest that PRC genes are involved in the silencing of male-genome expressed genes. Additionally, it is noteworthy that the syncytial phase and cellularization of endosperm development occur independently of embryo presence[162,163]. Considering these insights, it is tempting to speculate that combining MiMe with fis mutants could be a feasible approach towards achieving fertilization-independent synthetic apomixis.

    • Synthetic apomixis offers great opportunities for permanent fixation of hybrid vigor and genetic stability of staple crops. By bypassing meiosis and fertilization, synthetic apomixis can produce genetically identical offspring, eliminating the need for costly and time-consuming breeding efforts. Although much progresses have been achieved in the past two decades, the molecular understanding of apomixis and robust strategies for application of synthetic apomixis in major crops are still limited. Compared to the conserved MiMe strategy, the haploid induction appears to be the major hurdles to be resolved in this field. Moreover, MiMe mutations has only been studied in few species, whether it works in other crops still need investigation. In addition, achieving a basic understanding of the distinction between sexual and asexual reproduction remains a significant challenge.

      Advancements in our understanding of gamete interaction and the activation of the zygotic genome have the potential to lead to the identification of new genes that can be leveraged for the engineering of synthetic apomixis in crops. In addition, new tools have emerged that can be used to screen for potential gene targets and to dissect gene function and regulation networks. For example, single-cell CRISPR screening techniques, such as DAP-seq, CROP-seq, CRISP-seq, or Perturb-seq[164], and CRISPR-Cas-based genetic and epigenetic manipulation of gene expression systems[165], as well as comprehensive omics analyses including transcriptome, metabolome, quantitative proteome, and protein modification analysis[7], are rapidly developing. These technologies have the potential to accelerate the development of synthetic apomixis, as they can be used to identify the desired genes related to haploid induction and MiMe in plants. The knowledge of the transcriptional regulation of genes also depends on an understanding of the cis-elements and chromatin-based regulations in the genome sequence.

      There are also several challenges that must be addressed before synthetic apomixis can be successfully applied to staple crops. One major challenge is the complexity of the apomixis trait, which involves the coordination of multiple genetic pathways. Developing synthetic apomixis systems that can be efficiently transferred to target crops will require a thorough understanding of these pathways and how they interact with one another. Another challenge is the potential for unintended consequences, such as the loss of genetic diversity and the accumulation of deleterious mutations over time. To address these concerns, researchers must carefully monitor the performance of synthetic apomixis lines over multiple generations and develop strategies to maintain genetic diversity within breeding populations. In addition, there may be regulatory challenges associated with the adoption of synthetic apomixis in agriculture, for example, concerns about the safety and environmental impact of genetically modified crops, or doubt on the need for such technologies given the availability of alternative breeding strategies. Overall, while synthetic apomixis holds great promise for improving staple crops like maize and soybeans, there are a number of scientific and technical challenges that must be overcome before this technology can be widely adopted.

      In summary, the successful engineering of synthetic apomixis in rice signifies the commencement of a novel agricultural revolution. However, the translation of this breakthrough to other major crops, such as maize and soybean, presents various limitations. Despite the potential difficulties and challenges, the development of synthetic apomixis has enormous benefits in the field of crop breeding, and it is imperative that continued efforts should be made to overcome the existing obstacles.

      • This study was supported by the Strategic Priority Research Program of the Chinese Academy of Science (XDA24020306) and the National Key Research and Development Program of China (2022YFF1003500).

      • The authors declare that they have no conflict of interest. Hongju Li and KejianWang are the Editorial-Board members of Seed Biology who were blinded from reviewing or making decisions on the manuscript. The article was subject to the journal's standard procedures, with peer-review handled independently of these Editorial-Board members and their research groups.

      • Copyright: © 2023 by the author(s). Published by Maximum Academic Press on behalf of Hainan Yazhou Bay Seed Laboratory. This article is an open access article distributed under Creative Commons Attribution License (CC BY 4.0), visit https://creativecommons.org/licenses/by/4.0/.
    Figure (3)  References (165)
  • About this article
    Cite this article
    Li S, Wang J, Jia S, Wang K, Li H. 2023. Synthetic apomixis: from genetic basis to agricultural application. Seed Biology 2:10 doi: 10.48130/SeedBio-2023-0010
    Li S, Wang J, Jia S, Wang K, Li H. 2023. Synthetic apomixis: from genetic basis to agricultural application. Seed Biology 2:10 doi: 10.48130/SeedBio-2023-0010

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return