Search
2022 Volume 1
Article Contents
REVIEW   Open Access    

Parental regulation of seed development

More Information
  • Received Date: 23 October 2022
    Accepted Date: 17 November 2022
    Published Online: 07 December 2022
    Seed Biology  1 Article number: 7 (2022)  |  Cite this article
  • Angiosperms produce seeds with two zygotic tissues, namely the embryo and endosperm, from a unique double fertilization process. Seed development occurs within the maternal tissue and relies on maternal resources. Paternal tissue is not directly involved in seed development, and paternal regulation is usually based on the paternal genome of zygotic tissues in the filial generation. The complicated maternal-paternal communications and maternal-zygotic interactions result in distinguishable genetic effects on seed development. Here we review the conceptual framework of parental regulations on seed development. We summarize the common seed development process and look into the regulations pertaining to maternal and zygotic effects. Examples with more complicated interactions at the inter-tissue level are also discussed in the context of interwoven parental regulations.
  • 加载中
  • [1]

    Christenhusz MJM, Byng JW. 2016. The number of known plants species in the world and its annual increase. Phytotaxa 261:201−17

    doi: 10.11646/phytotaxa.261.3.1

    CrossRef   Google Scholar

    [2]

    Sliwinska E, Bewley JD, Gallagher R. 2014. Overview of Seed Development, Anatomy and Morphology. In Seeds: The Ecology of Regeneration in Plant Communities. Oxfordshire: CAB International. pp. 1−17. https://doi.org/10.1079/9781780641836.00

    [3]

    Bhatnagar SP, Sawhney V. 1981. Endosperm—Its morphology, ultrastructure, and histochemistry. International Review of Cytology 73:55−102

    doi: 10.1016/s0074-7696(08)61286-3

    CrossRef   Google Scholar

    [4]

    Olsen OA. 2007. Endosperm: developmental and molecular biology. Plant Cell Monographs. Heidelberg: Springer Berlin. pp. 26−43. https://doi.org/10.1007/978-3-540-71235-0

    [5]

    Garcia D, Saingery V, Chambrier P, Mayer U, Jürgens G, et al. 2003. Arabidopsis haiku mutants reveal new controls of seed size by endosperm. Plant Physiology 131:1661−70

    doi: 10.1104/pp.102.018762

    CrossRef   Google Scholar

    [6]

    Luo M, Dennis ES, Berger F, Peacock WJ, Chaudhury A. 2005. MINISEED3 (MINI3), a WRKY family gene, and HAIKU2 (IKU2), a leucine-rich repeat (LRR) KINASE gene, are regulators of seed size in Arabidopsis. PNAS 102:17531−36

    doi: 10.1073/pnas.0508418102

    CrossRef   Google Scholar

    [7]

    Zhou Y, Zhang X, Kang X, Zhao X, Zhang X, et al. 2009. SHORT HYPOCOTYL UNDER BLUE1 associates with MINISEED3 and HAIKU2 promoters in vivo to regulate Arabidopsis seed development. The Plant Cell 21:106−17

    doi: 10.1105/tpc.108.064972

    CrossRef   Google Scholar

    [8]

    Wang A, Garcia D, Zhang H, Feng K, Chaudhury A, et al. 2010. The VQ motif protein IKU1 regulates endosperm growth and seed size in Arabidopsis. The Plant Journal 63:670−79

    doi: 10.1111/j.1365-313X.2010.04271.x

    CrossRef   Google Scholar

    [9]

    Kang X, Li W, Zhou Y, Ni M. 2013. A WRKY transcription factor recruits the SYG1-like protein SHB1 to activate gene expression and seed cavity enlargement. PLoS Genetics 9:e1003347

    doi: 10.1371/journal.pgen.1003347

    CrossRef   Google Scholar

    [10]

    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

    [11]

    Xiong H, Wang W, Sun M-X. 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

    [12]

    Denney JO. 1992. Xenia includes metaxenia. HortScience 27:722−28

    doi: 10.21273/HORTSCI.27.7.722

    CrossRef   Google Scholar

    [13]

    Wang Z, Chen M, Chen T, Xuan L, Li Z, et al. 2014. TRANSPARENT TESTA2 regulates embryonic fatty acid biosynthesis by targeting FUSCA3 during the early developmental stage of Arabidopsis seeds. Plant Journal 77:757−69

    doi: 10.1111/tpj.12426

    CrossRef   Google Scholar

    [14]

    Chen M, Xuan L, Wang Z, Zhou L, Li Z, et al. 2014. TRANSPARENT TESTA8 inhibits seed fatty acid accumulation by targeting several seed development regulators in Arabidopsis. Plant Physiology 165:905−16

    doi: 10.1104/pp.114.235507

    CrossRef   Google Scholar

    [15]

    Chen M, Zhang B, Li C, Kulaveerasingam H, Chew FT, et al. 2015. TRANSPARENT TESTA GLABRA1 regulates the accumulation of seed storage reserves in Arabidopsis. Plant Physiology 169:391−402

    doi: 10.1104/pp.15.00943

    CrossRef   Google Scholar

    [16]

    Li C, Zhang B, Chen B, Ji L, Yu H. 2018. Site-specific phosphorylation of TRANSPARENT TESTA GLABRA1 mediates carbon partitioning in Arabidopsis seeds. Nature Communications 9:571

    doi: 10.1038/s41467-018-03013-5

    CrossRef   Google Scholar

    [17]

    Shi L, Katavic V, Yu Y, Kunst L, Haughn G. 2012. Arabidopsis glabra2 mutant seeds deficient in mucilage biosynthesis produce more oil. The Plant Journal 69:37−46

    doi: 10.1111/j.1365-313X.2011.04768.x

    CrossRef   Google Scholar

    [18]

    Li C, Chen B, Yu H. 2022. Splicing-mediated activation of SHAGGY-like kinases underpinning carbon partitioning in Arabidopsis seeds. The Plant Cell 34:2730−46

    doi: 10.1093/plcell/koac110

    CrossRef   Google Scholar

    [19]

    Vogiatzaki E, Baroux C, Jung JY, Poirier Y. 2017. PHO1 exports phosphate from the chalazal seed coat to the embryo in developing Arabidopsis seeds. Current Biology 27:2893−900

    doi: 10.1016/j.cub.2017.08.026

    CrossRef   Google Scholar

    [20]

    Sosso D, Luo D, Li Q, Sasse J, Yang J, et al. 2015. Seed filling in domesticated maize and rice depends on SWEET-mediated hexose transport. Nature Genetics 47:1489−93

    doi: 10.1038/ng.3422

    CrossRef   Google Scholar

    [21]

    Appelhagen I, Thiedig K, Nordholt N, Schmidt N, Huep G, et al. 2014. Update on transparent testa mutants from Arabidopsis thaliana: characterisation of new alleles from an isogenic collection. Planta 240:955−70

    doi: 10.1007/s00425-014-2088-0

    CrossRef   Google Scholar

    [22]

    Elliott RC, Betzner AS, Huttner E, Oakes MP, Tucker WQ, et al. 1996. AINTEGUMENTA, an APETALA2-like gene of Arabidopsis with pleiotropic roles in ovule development and floral organ growth. The Plant Cell 8:155−68

    doi: 10.1105/tpc.8.2.155

    CrossRef   Google Scholar

    [23]

    Klucher KM, Chow H, Reiser L, Fischer RL. 1996. The AINTEGUMENTA gene of Arabidopsis required for ovule and female gametophyte development is related to the floral homeotic gene APETALA2. The Plant Cell 8:137−53

    doi: 10.1105/tpc.8.2.137

    CrossRef   Google Scholar

    [24]

    Villanueva JM, Broadhvest J, Hauser BA, Meister RJ, Schneitz K, et al. 1999. INNER NO OUTER regulates abaxial–adaxial patterning in Arabidopsis ovules. Genes & Development 13:3160−69

    doi: 10.1101/gad.13.23.3160

    CrossRef   Google Scholar

    [25]

    Li N, Li Y. 2015. Maternal control of seed size in plants. Journal of Experimental Botany 66:1087−97

    doi: 10.1093/jxb/eru549

    CrossRef   Google Scholar

    [26]

    Li N, Li Y. 2014. Ubiquitin-mediated control of seed size in plants. Frontiers in Plant Science 5:332

    doi: 10.3389/fpls.2014.00332

    CrossRef   Google Scholar

    [27]

    Vanhaeren H, Chen Y, Vermeersch M, De Milde L, De Vleeschhauwer V, et al. 2020. UBP12 and UBP13 negatively regulate the activity of the ubiquitin-dependent peptidases DA1, DAR1 and DAR2. eLife 9:e52276

    doi: 10.7554/elife.52276

    CrossRef   Google Scholar

    [28]

    Garcia D, Fitz Gerald JN, Berger F. 2005. Maternal control of integument cell elongation and zygotic control of endosperm growth are coordinated to determine seed size in Arabidopsis. The Plant Cell 17:52−60

    doi: 10.1105/tpc.104.027136

    CrossRef   Google Scholar

    [29]

    Jofuku KD, Omidyar PK, Gee Z, Okamuro JK. 2005. Control of seed mass and seed yield by the floral homeotic gene APETALA2. PNAS 102:3117−22

    doi: 10.1073/pnas.0409893102

    CrossRef   Google Scholar

    [30]

    Leon-Kloosterziel KM, Keijzer CJ, Koornneef M. 1994. A seed shape mutant of Arabidopsis that is affected in integument development. The Plant Cell 6:385−92

    doi: 10.2307/3869758

    CrossRef   Google Scholar

    [31]

    McAbee JM, Hill TA, Skinner DJ, Izhaki A, Hauser BA, et al. 2006. ABERRANT TESTA SHAPE encodes a KANADI family member, linking polarity determination to separation and growth of Arabidopsis ovule integuments. The Plant Journal 46:522−31

    doi: 10.1111/j.1365-313X.2006.02717.x

    CrossRef   Google Scholar

    [32]

    Kirkbride RC, Lu J, Zhang C, Mosher RA, Baulcombe DC, et al. 2019. Maternal small RNAs mediate spatial-temporal regulation of gene expression, imprinting, and seed development in Arabidopsis. PNAS 116:2761−66

    doi: 10.1073/pnas.1807621116

    CrossRef   Google Scholar

    [33]

    Grover JW, Burgess D, Kendall T, Baten A, Pokhrel S, et al. 2020. Abundant expression of maternal siRNAs is a conserved feature of seed development. PNAS 117:15305−15

    doi: 10.1073/pnas.2001332117

    CrossRef   Google Scholar

    [34]

    Lu J, Zhang C, Baulcombe DC, Chen ZJ. 2012. Maternal siRNAs as regulators of parental genome imbalance and gene expression in endosperm of Arabidopsis seeds. PNAS 109:5529−34

    doi: 10.1073/pnas.1203094109

    CrossRef   Google Scholar

    [35]

    Adamski NM, Anastasiou E, Eriksson S, O'Neill CM, Lenhard M. 2009. Local maternal control of seed size by KLUH/CYP78A5-dependent growth signaling. PNAS 106:20115−20

    doi: 10.1073/pnas.0907024106

    CrossRef   Google Scholar

    [36]

    Fang W, Wang Z, Cui R, Li J, Li Y. 2012. Maternal control of seed size by EOD3/CYP78A6 in Arabidopsis thaliana. The Plant Journal 70:929−39

    doi: 10.1111/j.1365-313x.2012.04907.x

    CrossRef   Google Scholar

    [37]

    Wang J, Schwab R, Czech B, Mica E, Weigel D. 2008. Dual effects of miR156-targeted SPL genes and CYP78A5/KLUH on plastochron length and organ size in Arabidopsis thaliana. The Plant Cell 20:1231−43

    doi: 10.1105/tpc.108.058180

    CrossRef   Google Scholar

    [38]

    Li Y, Yu Y, Liu X, Zhang X, Su Y. 2021. The Arabidopsis MATERNAL EFFECT EMBRYO ARREST45 protein modulates maternal auxin biosynthesis and controls seed size by inducing AINTEGUMENTA. The Plant Cell 33:1907−26

    doi: 10.1093/plcell/koab084

    CrossRef   Google Scholar

    [39]

    Robert HS, Park C, Gutièrrez CL, Wójcikowska B, Pěnčík A, et al. 2018. Maternal auxin supply contributes to early embryo patterning in Arabidopsis. Nature Plants 4:548−53

    doi: 10.1038/s41477-018-0204-z

    CrossRef   Google Scholar

    [40]

    Shi C, Luo P, Du Y, Chen H, Huang X, et al. 2019. Maternal control of suspensor programmed cell death via gibberellin signaling. Nature Communications 10:3484

    doi: 10.1038/s41467-019-11476-3

    CrossRef   Google Scholar

    [41]

    Song X, Huang W, Shi M, Zhu M, Lin H. 2007. A QTL for rice grain width and weight encodes a previously unknown RING-type E3 ubiquitin ligase. Nature Genetics 39:623−30

    doi: 10.1038/ng2014

    CrossRef   Google Scholar

    [42]

    Bednarek J, Boulaflous A, Girousse C, Ravel C, Tassy C, et al. 2012. Down-regulation of the TaGW2 gene by RNA interference results in decreased grain size and weight in wheat. Journal of Experimental Botany 63:5945−55

    doi: 10.1093/jxb/ers249

    CrossRef   Google Scholar

    [43]

    Li Q, Li L, Yang X, Warburton ML, Bai G, et al. 2010. Relationship, evolutionary fate and function of two maize co-orthologs of rice GW2 associated with kernel size and weight. BMC Plant Biology 10:143

    doi: 10.1186/1471-2229-10-143

    CrossRef   Google Scholar

    [44]

    Xiong M, Feng G, Gao Q, Zhang C, Li Q, et al. 2022. Brassinosteroid regulation in rice seed biology. Seed Biology 1:2

    doi: 10.48130/SeedBio-2022-0002

    CrossRef   Google Scholar

    [45]

    Nagasawa N, Hibara KI, Heppard EP, Vander Velden KA, Luck S, et al. 2013. GIANT EMBRYO encodes CYP78A13, required for proper size balance between embryo and endosperm in rice. Plant Journal 75:592−605

    doi: 10.1111/tpj.12223

    CrossRef   Google Scholar

    [46]

    Yang W, Gao M, Yin X, Liu J, Xu Y, et al. 2013. Control of rice embryo development, shoot apical meristem maintenance, and grain yield by a novel cytochrome p450. Molecular Plant 6:1945−60

    doi: 10.1093/mp/sst107

    CrossRef   Google Scholar

    [47]

    Bellido AM, Distéfano AM, Setzes N, Cascallares MM, Oklestkova J, et al. 2022. A mitochondrial ADXR–ADX–P450 electron transport chain is essential for maternal gametophytic control of embryogenesis in Arabidopsis. PNAS 119:e2000482119

    doi: 10.1073/pnas.2000482119

    CrossRef   Google Scholar

    [48]

    Lotan T, Ohto Ma, Yee KM, West MA, Lo R, et al. 1998. Arabidopsis LEAFY COTYLEDON1 is sufficient to induce embryo development in vegetative cells. Cell 93:1195−205

    doi: 10.1016/S0092-8674(00)81463-4

    CrossRef   Google Scholar

    [49]

    Song J, Xie X, Chen C, Shu J, Thapa RK, et al. 2021. LEAFY COTYLEDON1 expression in the endosperm enables embryo maturation in Arabidopsis. Nature Communications 12:3963

    doi: 10.1038/s41467-021-24234-1

    CrossRef   Google Scholar

    [50]

    Parcy F, Valon C, Kohara A, Miséra S, Giraudat J. 1997. The ABSCISIC ACID-INSENSITIVE3, FUSCA3, and LEAFY COTYLEDON1 loci act in concert to control multiple aspects of Arabidopsis seed development. The Plant Cell 9:1265−77

    doi: 10.1105/tpc.9.8.1265

    CrossRef   Google Scholar

    [51]

    Horstman A, Li M, Heidmann I, Weemen M, Chen B, et al. 2017. The BABY BOOM transcription factor activates the LEC1-ABI3-FUS3-LEC2 network to induce somatic embryogenesis. Plant Physiology 175:848−57

    doi: 10.1104/pp.17.00232

    CrossRef   Google Scholar

    [52]

    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

    [53]

    Méndez-Hernández HA, Ledezma-Rodríguez M, Avilez-Montalvo RN, Juárez-Gómez YL, Skeete A, et al. 2019. Signaling overview of plant somatic embryogenesis. Frontiers in Plant Science 10:77

    doi: 10.3389/fpls.2019.00077

    CrossRef   Google Scholar

    [54]

    Zhao P, Shi C, Wang L, Sun M. 2022. The parental contributions to early plant embryogenesis and the concept of maternal-to-zygotic transition in plants. Current Opinion in Plant Biology 65:102144

    doi: 10.1016/j.pbi.2021.102144

    CrossRef   Google Scholar

    [55]

    Li L, Weigel D. 2021. One hundred years of hybrid necrosis: hybrid autoimmunity as a window into the mechanisms and evolution of plant–pathogen interactions. Annual Review of Phytopathology 59:213−37

    doi: 10.1146/annurev-phyto-020620-114826

    CrossRef   Google Scholar

    [56]

    Kawanabe T, Ishikura S, Miyaji N, Sasaki T, Wu LM, et al. 2016. Role of DNA methylation in hybrid vigor in Arabidopsis thaliana. PNAS 113:E6704−E6711

    doi: 10.1073/pnas.1613372113

    CrossRef   Google Scholar

    [57]

    Wang L, Greaves IK, Groszmann M, Wu LM, Dennis ES, Peacock WJ. 2015. Hybrid mimics and hybrid vigor in Arabidopsis. PNAS 112:E4959−E4967

    doi: 10.1073/pnas.1514190112

    CrossRef   Google Scholar

    [58]

    Jahnke S, Sarholz B, Thiemann A, Kühr V, Gutiérrez-Marcos JF, et al. 2010. Heterosis in early seed development: a comparative study of F1 embryo and endosperm tissues 6 days after fertilization. Theoretical and Applied Genetics 120:389−400

    doi: 10.1007/s00122-009-1207-y

    CrossRef   Google Scholar

    [59]

    Labroo MR, Studer AJ, Rutkoski JE. 2021. Heterosis and hybrid crop breeding: a multidisciplinary review. Frontiers in Genetics 12:643761

    doi: 10.3389/fgene.2021.643761

    CrossRef   Google Scholar

    [60]

    Liu W, Zhang Y, He H, He G, Deng X. 2022. From hybrid genomes to heterotic trait output: Challenges and opportunities. Current Opinion in Plant Biology 66:102193

    doi: 10.1016/j.pbi.2022.102193

    CrossRef   Google Scholar

    [61]

    Wang K. 2020. Fixation of hybrid vigor in rice: synthetic apomixis generated by genome editing. aBIOTECH 1:15−20

    doi: 10.1007/s42994-019-00001-1

    CrossRef   Google Scholar

    [62]

    Wang Q, Wang M, Chen J, Qi W, Lai J, et al. 2022. ENB1 encodes a cellulose synthase 5 that directs synthesis of cell wall ingrowths in maize basal endosperm transfer cells. The Plant Cell 34:1054−74

    doi: 10.1093/plcell/koab312

    CrossRef   Google Scholar

    [63]

    Rodrigues JA, Hsieh PH, Ruan D, Nishimura T, Sharma MK, et al. 2021. Divergence among rice cultivars reveals roles for transposition and epimutation in ongoing evolution of genomic imprinting. PNAS 118:e2104445118

    doi: 10.1073/pnas.2104445118

    CrossRef   Google Scholar

    [64]

    Luo M, Taylor JM, Spriggs A, Zhang H, Wu X, et al. 2011. A genome-wide survey of imprinted genes in rice seeds reveals imprinting primarily occurs in the endosperm. PLoS Genetics 7:e1002125

    doi: 10.1371/journal.pgen.1002125

    CrossRef   Google Scholar

    [65]

    Hsieh TF, Ibarra CA, Silva P, Zemach A, Eshed-Williams L, et al. 2009. Genome-wide demethylation of Arabidopsis endosperm. Science 324:1451−4

    doi: 10.1126/science.1172417

    CrossRef   Google Scholar

    [66]

    Gehring M, Bubb KL, Henikoff S. 2009. Extensive demethylation of repetitive elements during seed development underlies gene imprinting. Science 324:1447−51

    doi: 10.1126/science.1171609

    CrossRef   Google Scholar

    [67]

    Batista RA, Köhler CJG. 2020. Genomic imprinting in plants — revisiting existing models. Genes & Development 34:24−36

    doi: 10.1101/gad.332924.119

    CrossRef   Google Scholar

    [68]

    Choi Y, Gehring M, Johnson L, Hannon M, Harada JJ, et al. 2002. DEMETER, a DNA glycosylase domain protein, is required for endosperm gene imprinting and seed viability in Arabidopsis. Cell 110:33−42

    doi: 10.1016/S0092-8674(02)00807-3

    CrossRef   Google Scholar

    [69]

    Xiao W, Gehring M, Choi Y, Margossian L, Pu H, et al. 2003. Imprinting of the MEA Polycomb gene is controlled by antagonism between MET1 methyltransferase and DME glycosylase. Developmental Cell 5:891−901

    doi: 10.1016/S1534-5807(03)00361-7

    CrossRef   Google Scholar

    [70]

    Xiao W, Brown RC, Lemmon BE, Harada JJ, Goldberg RB, Fischer RL. 2006. Regulation of seed size by hypomethylation of maternal and paternal genomes. Plant Physiology 142:1160−68

    doi: 10.1104/pp.106.088849

    CrossRef   Google Scholar

    [71]

    Tiwari S, Schulz R, Ikeda Y, Dytham L, Bravo J, et al. 2008. MATERNALLY EXPRESSED PAB C-TERMINAL, a novel imprinted gene in Arabidopsis, encodes the conserved C-terminal domain of polyadenylate binding proteins. The Plant Cell 20:2387−98

    doi: 10.1105/tpc.108.061929

    CrossRef   Google Scholar

    [72]

    Jullien PE, Kinoshita T, Ohad N, Berger F. 2006. Maintenance of DNA methylation during the Arabidopsis life cycle is essential for parental imprinting. The Plant Cell 18:1360−72

    doi: 10.1105/tpc.106.041178

    CrossRef   Google Scholar

    [73]

    Kinoshita T, Miura A, Choi Y, Kinoshita Y, Cao X, et al. 2004. One-way control of FWA imprinting in Arabidopsis endosperm by DNA methylation. Science 303:521−23

    doi: 10.1126/science.1089835

    CrossRef   Google Scholar

    [74]

    Gehring M, Huh JH, Hsieh T-F, Penterman J, Choi Y, et al. 2006. DEMETER DNA glycosylase establishes MEDEA polycomb gene self-imprinting by allele-specific demethylation. Cell 124:495−506

    doi: 10.1016/j.cell.2005.12.034

    CrossRef   Google Scholar

    [75]

    Jullien PE, Katz A, Oliva M, Ohad N, Berger F. 2006. Polycomb group complexes self-regulate imprinting of the Polycomb group gene MEDEA in Arabidopsis. Current Biology 16:486−92

    doi: 10.1016/j.cub.2006.01.020

    CrossRef   Google Scholar

    [76]

    Martínez G, Panda K, Köhler C, Slotkin RK. 2016. Silencing in sperm cells is directed by RNA movement from the surrounding nurse cell. Nature Plants 2:16030

    doi: 10.1038/nplants.2016.30

    CrossRef   Google Scholar

    [77]

    Satyaki PRV, Gehring M. 2019. Paternally acting canonical RNA-directed DNA methylation pathway genes sensitize Arabidopsis endosperm to paternal genome dosage. The Plant Cell 31:1563−78

    doi: 10.1105/tpc.19.00047

    CrossRef   Google Scholar

    [78]

    Ibarra CA, Feng X, Schoft VK, Hsieh TF, Uzawa R, et al. 2012. Active DNA demethylation in plant companion cells reinforces transposon methylation in gametes. Science 337:1360−64

    doi: 10.1126/science.1224839

    CrossRef   Google Scholar

    [79]

    Calarco JP, Borges F, Donoghue MT, Van Ex F, Jullien PE, et al. 2012. Reprogramming of DNA methylation in pollen guides epigenetic inheritance via small RNA. Cell 151:194−205

    doi: 10.1016/j.cell.2012.09.001

    CrossRef   Google Scholar

    [80]

    Pignatta D, Erdmann RM, Scheer E, Picard CL, Bell GW, et al. 2014. Natural epigenetic polymorphisms lead to intraspecific variation in Arabidopsis gene imprinting. eLife 3:e03198

    doi: 10.7554/eLife.03198

    CrossRef   Google Scholar

    [81]

    Gent JI, Higgins KM, Swentowsky KW, Fu F, Zeng Y, et al. 2022. The maize gene maternal derepression of r1 encodes a DNA glycosylase that demethylates DNA and reduces siRNA expression in the endosperm. The Plant Cell 34:3685−701

    doi: 10.1093/plcell/koac199

    CrossRef   Google Scholar

    [82]

    Schon MA, Nodine MD. 2017. Widespread contamination of Arabidopsis embryo and endosperm transcriptome data sets. The Plant Cell 29:608−17

    doi: 10.1105/tpc.16.00845

    CrossRef   Google Scholar

    [83]

    Köhler C, Page DR, Gagliardini V, Grossniklaus U. 2005. The Arabidopsis thaliana MEDEA Polycomb group protein controls expression of PHERES1 by parental imprinting. Nature Genetics 37:28−30

    doi: 10.1038/ng1495

    CrossRef   Google Scholar

    [84]

    Köhler C, Hennig L, Spillane C, Pien S, Gruissem W, et al. 2003. The Polycomb-group protein MEDEA regulates seed development by controlling expression of the MADS-box gene PHERES1. Genes & Development 17:1540−53

    doi: 10.1101/gad.257403

    CrossRef   Google Scholar

    [85]

    Batista RA, Moreno-Romero J, Qiu Y, van Boven J, Santos-González J, et al. 2019. The MADS-box transcription factor PHERES1 controls imprinting in the endosperm by binding to domesticated transposons. eLife 8:e50541

    doi: 10.7554/elife.50541

    CrossRef   Google Scholar

    [86]

    Hornslien KS, Miller JR, Grini PE. 2019. Regulation of parent-of-origin allelic expression in the endosperm. Plant Physiology 180:1498−519

    doi: 10.1104/pp.19.00320

    CrossRef   Google Scholar

    [87]

    Moreno-Romero J, Jiang H, Santos-González J, Köhler C. 2016. Parental epigenetic asymmetry of PRC 2-mediated histone modifications in the Arabidopsis endosperm. EMBO Journal 35:1298−311

    doi: 10.15252/embj.201593534

    CrossRef   Google Scholar

    [88]

    Moreno-Romero J, Toro-De León D, Yadav VK, Santos-González J, Köhler C. 2019. Epigenetic signatures associated with imprinted paternally expressed genes in the Arabidopsis endosperm. Genome Biology 20:41

    doi: 10.1186/s13059-018-1612-0

    CrossRef   Google Scholar

    [89]

    Autran D, Baroux C, Raissig MT, Lenormand T, Wittig M, et al. 2011. Maternal epigenetic pathways control parental contributions to Arabidopsis early embryogenesis. Cell 145:707−19

    doi: 10.1016/j.cell.2011.04.014

    CrossRef   Google Scholar

    [90]

    Wolff P, Jiang H, Wang G, Santos-González J, Köhler C. 2015. Paternally expressed imprinted genes establish postzygotic hybridization barriers in Arabidopsis thaliana. eLife 4:e10074

    doi: 10.7554/eLife.10074

    CrossRef   Google Scholar

    [91]

    Raissig MT, Baroux C, Grossniklaus U. 2011. Regulation and flexibility of genomic imprinting during seed development. The Plant Cell 23:16−26

    doi: 10.1105/tpc.110.081018

    CrossRef   Google Scholar

    [92]

    Pignatta D, Novitzky K, Satyaki PRV, Gehring M. 2018. A variably imprinted epiallele impacts seed development. PLoS Genetics 14:e1007469

    doi: 10.1371/journal.pgen.1007469

    CrossRef   Google Scholar

    [93]

    Dai D, Mudunkothge JS, Galli M, Char SN, Davenport R, et al. 2022. Paternal imprinting of dosage-effect defective1 contributes to seed weight xenia in maize. Nature Communications 13:5366

    doi: 10.1038/s41467-022-33055-9

    CrossRef   Google Scholar

    [94]

    Adams S, Vinkenoog R, Spielman M, Dickinson HG, Scott RJ. 2000. Parent-of-origin effects on seed development in Arabidopsis thaliana require DNA methylation. Development 127:2493−502

    doi: 10.1242/dev.127.11.2493

    CrossRef   Google Scholar

    [95]

    Scott RJ, Spielman M, Bailey J, Dickinson HG. 1998. Parent-of-origin effects on seed development in Arabidopsis thaliana. Development 125:3329−41

    doi: 10.1242/dev.125.17.3329

    CrossRef   Google Scholar

    [96]

    Leblanc O, Pointe C, Hernandez M. 2002. Cell cycle progression during endosperm development in Zea mays depends on parental dosage effects. The Plant Journal 32:1057−66

    doi: 10.1046/j.1365-313X.2002.01491.x

    CrossRef   Google Scholar

    [97]

    Sekine D, Ohnishi T, Furuumi H, Ono A, Yamada T, et al. 2013. Dissection of two major components of the post-zygotic hybridization barrier in rice endosperm. The Plant Journal 76:792−99

    doi: 10.1111/tpj.12333

    CrossRef   Google Scholar

    [98]

    Wang L, Yuan J, Ma Y, Jiao W, Ye W, et al. 2018. Rice interploidy crosses disrupt epigenetic regulation, gene expression, and seed development. Molecular Plant 11:300−14

    doi: 10.1016/j.molp.2017.12.006

    CrossRef   Google Scholar

    [99]

    Kradolfer D, Wolff P, Jiang H, Siretskiy A, Köhler C. 2013. An imprinted gene underlies postzygotic reproductive isolation in Arabidopsis thaliana. Developmental Cell 26:525−35

    doi: 10.1016/j.devcel.2013.08.006

    CrossRef   Google Scholar

    [100]

    Wang G, Jiang H, de León GDT, Martinez G, Köhler C. 2018. Sequestration of a transposon-derived siRNA by a target mimic imprinted gene induces postzygotic reproductive isolation in Arabidopsis. Developmental Cell 46:696−705.E4

    doi: 10.1016/j.devcel.2018.07.014

    CrossRef   Google Scholar

    [101]

    Martinez G, Wolff P, Wang Z, Moreno-Romero J, Santos-González J, et al. 2018. Paternal easiRNAs regulate parental genome dosage in Arabidopsis. Nature Genetics 50:193−98

    doi: 10.1038/s41588-017-0033-4

    CrossRef   Google Scholar

    [102]

    Jiang H, Moreno-Romero J, Santos-González J, De Jaeger G, Gevaert K, et al. 2017. Ectopic application of the repressive histone modification H3K9me2 establishes post-zygotic reproductive isolation in Arabidopsis thaliana. Genes & Development 31:1272−87

    doi: 10.1101/gad.299347.117

    CrossRef   Google Scholar

    [103]

    Erdmann RM, Satyaki PR, Klosinska M, Gehring M. 2017. A small RNA pathway mediates allelic dosage in endosperm. Cell Reports 21:3364−72

    doi: 10.1016/j.celrep.2017.11.078

    CrossRef   Google Scholar

    [104]

    Huang F, Zhu Qh, Zhu A, Wu X, Xie L, et al. 2017. Mutants in the imprinted PICKLE RELATED 2 gene suppress seed abortion of fertilization independent seed class mutants and paternal excess interploidy crosses in Arabidopsis. The Plant Journal 90:383−95

    doi: 10.1111/tpj.13500

    CrossRef   Google Scholar

    [105]

    Lafon-Placette C, Hatorangan MR, Steige KA, Cornille A, Lascoux M, et al. 2018. Paternally expressed imprinted genes associate with hybridization barriers in Capsella. Nature Plants 4:352−57

    doi: 10.1038/s41477-018-0161-6

    CrossRef   Google Scholar

    [106]

    Schatlowski N, Wolff P, Santos-González J, Schoft V, Siretskiy A, et al. 2014. Hypomethylated pollen bypasses the interploidy hybridization barrier in Arabidopsis. The Plant Cell 26:3556−68

    doi: 10.1105/tpc.114.130120

    CrossRef   Google Scholar

    [107]

    Huc J, Dziasek K, Pachamuthu K, Woh T, Köhler C, et al. 2022. Bypassing reproductive barriers in hybrid seeds using chemically induced epimutagenesis. The Plant Cell 34:989−1001

    doi: 10.1093/plcell/koab284

    CrossRef   Google Scholar

    [108]

    Bayer M, Nawy T, Giglione C, Galli M, Meinnel T, et al. 2009. Paternal control of embryonic patterning in Arabidopsis thaliana. Science 323:1485−88

    doi: 10.1126/science.1167784

    CrossRef   Google Scholar

    [109]

    Ueda M, Aichinger E, Gong W, Groot E, Verstraeten I, et al. 2017. Transcriptional integration of paternal and maternal factors in the Arabidopsis zygote. Genes & Development 31:617−27

    doi: 10.1101/gad.292409.116

    CrossRef   Google Scholar

    [110]

    Anderson SN, Johnson CS, Chesnut J, Jones DS, Khanday I, et al. 2017. The zygotic transition is initiated in unicellular plant zygotes with asymmetric activation of parental genomes. Developmental Cell 43:349−58.E4

    doi: 10.1016/j.devcel.2017.10.005

    CrossRef   Google Scholar

    [111]

    Khanday I, Skinner D, Yang B, Mercier R, Sundaresan V. 2019. A male-expressed rice embryogenic trigger redirected for asexual propagation through seeds. Nature 565:91−5

    doi: 10.1038/s41586-018-0785-8

    CrossRef   Google Scholar

    [112]

    Zhao Y, Wang S, Wu W, Li L, Jiang T, Zheng B. 2018. Clearance of maternal barriers by paternal miR159 to initiate endosperm nuclear division in Arabidopsis. Nature Communications 9:5011

    doi: 10.1038/s41467-018-07429-x

    CrossRef   Google Scholar

    [113]

    Kasahara RD, Notaguchi M, Nagahara S, Suzuki T, Susaki D, et al. 2016. Pollen tube contents initiate ovule enlargement and enhance seed coat development without fertilization. Science Advances 2:e1600554

    doi: 10.1126/sciadv.1600554

    CrossRef   Google Scholar

    [114]

    Heydlauff J, Serbes IE, Vo D, Mao Y, Gieseking S, et al. 2022. Dual and opposing roles of EIN3 reveal a generation conflict during seed growth. Molecular Plant 15:363−71

    doi: 10.1016/j.molp.2021.11.015

    CrossRef   Google Scholar

    [115]

    Birchler JA. 2014. Interploidy hybridization barrier of endosperm as a dosage interaction. Frontiers in Plant Science 5:281

    doi: 10.3389/fpls.2014.00281

    CrossRef   Google Scholar

    [116]

    Kato A, Birchler JA. 2006. Induction of tetraploid derivatives of maize inbred lines by nitrous oxide gas treatment. Journal of Heredity 97:39−44

    doi: 10.1093/jhered/esj007

    CrossRef   Google Scholar

    [117]

    Li C, Gong X, Zhang B, Liang Z, Wong CE, et al. 2020. TOP1α,UPF1, and TTG2 regulate seed size in a parental dosage–dependent manner. PLoS Biology 18:e3000930

    doi: 10.1371/journal.pbio.3000930

    CrossRef   Google Scholar

    [118]

    Picard CL, Povilus RA, Williams BP, Gehring M. 2021. Transcriptional and imprinting complexity in Arabidopsis seeds at single-nucleus resolution. Nature Plants 7:730−38

    doi: 10.1038/s41477-021-00922-0

    CrossRef   Google Scholar

    [119]

    Povilus RA, Gehring M. 2022. Maternal-filial transfer structures in endosperm: A nexus of nutritional dynamics and seed development. Current Opinion in Plant Biology 65:102121

    doi: 10.1016/j.pbi.2021.102121

    CrossRef   Google Scholar

    [120]

    Haig D. 2013. Kin conflict in seed development: an interdepengdent but fractious collective. Annual Review of Cell and Developmental Biology 29:189−211

    doi: 10.1146/annurev-cellbio-101512-122324

    CrossRef   Google Scholar

    [121]

    Zhang B, Li C, Li Y, Yu H. 2020. Mobile TERMINAL FLOWER1 determines seed size in Arabidopsis. Nature Plants 6:1146−57

    doi: 10.1038/s41477-020-0749-5

    CrossRef   Google Scholar

    [122]

    Kang IH, Steffen JG, Portereiko MF, Lloyd A, Drews GN. 2008. The AGL62 MADS domain protein regulates cellularization during endosperm development in Arabidopsis. The Plant Cell 20:635−47

    doi: 10.1105/tpc.107.055137

    CrossRef   Google Scholar

    [123]

    Xu W, Fiume E, Coen O, Pechoux C, Lepiniec L, Magnani E. 2016. Endosperm and nucellus develop antagonistically in Arabidopsis seeds. The Plant Cell 28:1343−60

    doi: 10.1105/tpc.16.00041

    CrossRef   Google Scholar

    [124]

    Figueiredo DD, Batista RA, Roszak PJ, Hennig L, Köhler C. 2016. Auxin production in the endosperm drives seed coat development in Arabidopsis. eLife 5:e20542

    doi: 10.7554/eLife.20542

    CrossRef   Google Scholar

    [125]

    Pires ND, Bemer M, Müller LM, Baroux C, Spillane C, Grossniklaus U. 2016. Quantitative genetics identifies cryptic genetic variation involved in the paternal regulation of seed development. PLoS Genetics 12:e1005806

    doi: 10.1371/journal.pgen.1005806

    CrossRef   Google Scholar

    [126]

    Zhang H, Li X, Wang W, Li H, Cui Y, et al. 2022. SERKs regulate embryonic cuticle integrity through the TWS1-GSO1/2 signaling pathway in Arabidopsis. New Phytologist 233:313−28

    doi: 10.1111/nph.17775

    CrossRef   Google Scholar

    [127]

    Doll NM, Royek S, Fujita S, Okuda S, Chamot S, et al. 2020. A two-way molecular dialogue between embryo and endosperm is required for seed development. Science 367:431−35

    doi: 10.1126/science.aaz4131

    CrossRef   Google Scholar

    [128]

    Xing Q, Creff A, Waters A, Tanaka H, Goodrich J, et al. 2013. ZHOUPI controls embryonic cuticle formation via a signalling pathway involving the subtilisin protease ABNORMAL LEAF-SHAPE1 and the receptor kinases GASSHO1 and GASSHO2. Development 140:770−79

    doi: 10.1242/dev.088898

    CrossRef   Google Scholar

    [129]

    Yang S, Johnston N, Talideh E, Mitchell S, Jeffree C, et al. 2008. The endosperm-specific ZHOUPI gene of Arabidopsis thaliana regulates endosperm breakdown and embryonic epidermal development. Development 135:3501−9

    doi: 10.1242/dev.026708

    CrossRef   Google Scholar

    [130]

    Moussu S, Doll NM, Chamot S, Brocard L, Creff A, et al. 2017. ZHOUPI and KERBEROS mediate embryo/endosperm separation by promoting the formation of an extracuticular sheath at the embryo surface. The Plant Cell 29:1642−56

    doi: 10.1105/tpc.17.00016

    CrossRef   Google Scholar

    [131]

    Yuan J, Chen S, Jiao W, Wang L, Wang L, et al. 2017. Both maternally and paternally imprinted genes regulate seed development in rice. New Phytologist 216:373−87

    doi: 10.1111/nph.14510

    CrossRef   Google Scholar

    [132]

    Ma A, McDermaid A, Xu J, Chang Y, Ma Q. 2020. Integrative methods and practical challenges for single-cell multi-omics. Trends in Biotechnology 38:1007−22

    doi: 10.1016/j.tibtech.2020.02.013

    CrossRef   Google Scholar

  • Cite this article

    Li C, Yu H. 2022. Parental regulation of seed development. Seed Biology 1:7 doi: 10.48130/SeedBio-2022-0007
    Li C, Yu H. 2022. Parental regulation of seed development. Seed Biology 1:7 doi: 10.48130/SeedBio-2022-0007

Figures(4)  /  Tables(1)

Article Metrics

Article views(4330) PDF downloads(732)

Other Articles By Authors

REVIEW   Open Access    

Parental regulation of seed development

Seed Biology  1 Article number: 7  (2022)  |  Cite this article

Abstract: Angiosperms produce seeds with two zygotic tissues, namely the embryo and endosperm, from a unique double fertilization process. Seed development occurs within the maternal tissue and relies on maternal resources. Paternal tissue is not directly involved in seed development, and paternal regulation is usually based on the paternal genome of zygotic tissues in the filial generation. The complicated maternal-paternal communications and maternal-zygotic interactions result in distinguishable genetic effects on seed development. Here we review the conceptual framework of parental regulations on seed development. We summarize the common seed development process and look into the regulations pertaining to maternal and zygotic effects. Examples with more complicated interactions at the inter-tissue level are also discussed in the context of interwoven parental regulations.

    • Gymnosperms and angiosperms are called seed plants because they reproduce by seeds. In particular, the angiosperms, also known as flowering plants or higher plants, are the most diverse and widespread group of land plants on earth[1]. As indicated by their names, gymnosperm seeds are exposed without the protection layer, while angiosperm seeds are embedded in the maternal fruit. The seed origin of angiosperms is also different from that of gymnosperms. Angiosperm seeds result from a unique double fertilization process, in which one sperm nucleus (haploid; 1n) fertilizes the egg cell (haploid; 1n) and another sperm nucleus fertilizes the central cell (either 1n + 1n or 2n). The sperm-egg fusion produces the embryo (diploid; 2n), while the sperm-central cell fusion develops into the endosperm (triploid; 3n), which is an angiosperm-specific and terminally differentiated tissue that provides nutrition to the embryo or young seedling (Fig. 1a).

      Figure 1. 

      Seed development in angiosperms. (a) Double fertilization (leftmost panels) initiates embryo and endosperm formation (right panels) across successive stages of seed development in Arabidopsis (dicot model) and rice (monocot model). (b) Different types of angiosperm endosperms. Dots denote endosperm nuclei, while ellipses denote the embryo sac before fertilization or the endosperm after fertilization. (c) Various maternal and paternal effects on the regulation of seed development. (a) & (c) Maternal and paternal components are indicated in red and blue, respectively. The seed coat is indicated in light brown. For the zygotic tissues, the endosperm and embryo are indicated in pink and purple, respectively.

      Except for basal angiosperm species, the majority of angiosperms can be roughly divided into two groups, monocots and dicots, which exhibit various features of the seed structure. Some dicot seeds, including those of the model plant Arabidopsis (Arabidopsis thaliana), bear degraded endosperms as the nutrients are mainly stored in mature embryos (Fig. 1a). In the monocot Poaceae family, including popular crops such as rice (Oryza sativa), maize (Zea mays), barley (Hordeum vulgare), and wheat (Triticum aestivum), the commonly termed 'seed' is the caryopsis (a kind of fruit), in which the seed coat is fused to the pericarp (the fruit coat). The mature seeds of these crops have well-developed endosperms that store nutrients (Fig. 1a). Although seed development in these monocot crops coincides with fruit development in nature, their basic stages are comparable with those in dicots[2] (Fig. 1a).

      The embryogenesis process is geometrically different in dicots and monocots[2]. For example, the first zygotic division in Arabidopsis is asymmetric, resulting in a small apical cell and a large basal cell. The cell lineage from the basal cell generates the suspensor and part of the embryonic root apical meristem, while the other embryonic tissues generate from the apical cell. In contrast, the rice zygote undergoes random divisions to generate a cluster of cells before differentiation (Fig. 1a), indicating that establishment of embryonic patterning is much later in monocots than in dicots. Moreover, embryonic differentiation in Arabidopsis is along the apical-basal axis with bilateral symmetry, whereas embryonic differentiation in rice exhibits an evident dorsal-ventral axis with both shoot and root meristem cells occurring at the ventral side.

      As the featured structure of angiosperm seeds, endosperms are classified into three types: nuclear type, cellular type, and helobial type[3]. The nuclear-type endosperm is the most common type in which the primary endosperm undergoes karyokinesis repeatedly without cell wall formation to produce free nuclei at earlier stages. The cell wall only appears during endosperm cellularization to separate individual nuclei (Fig. 1a & b). In contrast, the cellular-type endosperm proliferates via complete cytokinesis with cell wall formation from the very beginning (Fig. 1b). The helobial-type endosperm is an intermediate type in which the chalazal endosperm undergoes complete cytokinesis once or twice, while the micropyle endosperm undergoes karyokinesis (Fig. 1b).

      Both Arabidopsis and rice develop the nuclear-type endosperm (Fig. 1a & b). In Arabidopsis, endosperm cellularization occurs during the embryo status at the heart stage to the early torpedo stage, except that its chalazal endosperm never undergoes cellularization[2]. After endosperm cellularization, endosperm cells undergo endoreplication in Arabidopsis, whereas in monocots, numerous additional rounds of mitoses occur between endosperm cellularization and endoreduplication[4]. Endosperm cellularization is crucial for seed development[59]. Generally, the over-proliferated endosperm is associated with delayed or failed cellularization, resulting in larger or aborted seeds, respectively. In contrast, less-proliferated and accelerated endosperm cellularization leads to smaller seeds. At the end of seed development, the Arabidopsis endosperm is almost consumed by the embryo except for a one-cell layer adjacent to the seed coat, whereas the mature rice endosperm takes the major volume of a seed and differentiates into several functional regions[2]. Notably, although endosperm and embryo development are closely correlated, their individual development can proceed autonomously albeit defectively when the accompanying part is completely lost[10,11], indicating that the endosperm and embryo develop both independently and dependently.

      In general, a mature angiosperm seed contains at least the diploid seed coat (parent generation; maternal sporophytic tissue), the diploid embryo (filial generation), and the triploid endosperm (filial generation) (Fig. 1a). Such heterogeneity implies that seed development is regulated by interwoven signaling networks. Before fertilization, maternal and paternal gametophytic effects influence the formation of gametophytic embryo sac and pollen prior to seed development. After fertilization, factors pertaining to filial tissues (zygotic tissues) could play a more specific role in embryonic or endospermic development. Because the 'paternal sporophyte' is not involved in seed development, a paternal effect is equivalent to the paternal gametophytic effect and zygotic paternal effect (sometimes known as xenia effects[12]). Notably, since seed development depends on the maternal support in the course of the whole seed developmental process, the maternal sporophytic effects play pivotal roles both before and after fertilization (Fig. 1c). In the following sections, we discuss parental effects, including paternal effects, in the context of maternal, zygotic, and inter-tissue regulation. The genes discussed in these sections are summarized in Table 1.

      Table 1.  Information on the genes discussed in this review.

      Gene nameAbbreviationGene IDFunction noteReference
      ABERRANT TESTA SHAPEATSAT5G42630KANADI family transcription factor[30,31]
      ABNORMAL LEAF-SHAPE 1ALE1AT1G62340Subtilisin-like serine protease[126129]
      ABSCISIC ACID INSENSITIVE 3ABI3AT3G24650B3 domain transcription factor[4850]
      ADMETOSADMAT4G11940J-domain chaperone[99]
      ADRENODOXIN 1ADX1AT4G05450Adrenodoxin[47]
      ADRENODOXIN 2ADX2AT4G21090Adrenodoxin[47]
      ADRENODOXIN REDUCTASEADXRAT4G32360Adrenodoxin reductase[47]
      AGAMOUS-LIKE 40AGL40AT4G36590MADS-box family transcription factor[32]
      AGAMOUS-LIKE 62AGL62AT5G60440MADS-box family transcription factor[122]
      AGAMOUS-LIKE 91AGL91AT3G66656MADS-box family transcription factor[32]
      AINTEGUMENTAANTAT4G37750AP2 family transcription factor[22, 23]
      APETALA2AP2AT4G36920AP2 family transcription factor[29]
      BABY BOOMBBMAT5G17430AP2 family transcription factor[5153]
      CYTOCHROME P450 FAMILY 78 A7CYP78A7AT5G09970Cytochrome p450 family[36,37]
      CYTOCHROME P450 FAMILY 78 A9CYP78A9AT3G61880Cytochrome p450 family[36,37]
      DA1DA1AT1G19270Ubiquitin-activated peptidase[26,27]
      DA2DA2AT1G78420RING-type E3 ubiquitin ligase[26]
      DEMETERDMEAT5G04560DNA glycosylase[65,66,68,
      69,78,80]
      DOSAGEEFFECT DEFECTIVE 1DED1Zm00001eb050770MYB family transcription factor[93]
      ENDOSPERM BREAKDOWN1ENB1Zm00001eb061800Cellulose synthase 5[62]
      ENHANCER OF da1-1 3EOD3 (CYP78A6)AT2G46660Cytochrome p450 family[36]
      ETHYLENE INSENSITIVE 3EIN3AT3G20770Transcription regulator[114]
      FERTILIZATION INDEPENDENT SEED 2FIS2AT2G35670PRC2 component[72,89]
      FLOWERING WAGENINGENFWAAT4G25530Homeodomain-containing transcription factor[73]
      FUSCA3FUS3AT3G26790B3 domains transcription factor[4850]
      GASSHO1GSO1AT4G20140Leucine rich repeat (LRR) receptor-like kinase[126,127]
      GASSHO2GSO2AT5G44700Leucine rich repeat (LRR) receptor-like kinase[126,127]
      GIANT EMBRYOGE (OsCYP78A13)LOC_Os07g41240Cytochrome p450 family[45,46]
      GLABRA2GL2AT1G79840Homeodomain-containing transcription factor[17]
      GRAIN WEIGHT 2GW2LOC_Os02g14720RING-type E3 ubiquitin ligase[41]
      HAIKU1IKU1AT1G55600Plant-specific VQ motif-containing protein[5,6,8]
      HAIKU2IKU2AT3G19700Leucine rich repeat (LRR) kinase[57]
      HOMEDOMAIN GLABROUS 3HDG3AT2G32370Homeodomain-containing transcription factor[92]
      INDUCER OF CBF EXPRESSIONICE 1ICE1AT3G26744bHLH family transcription factor[128, 129]
      INNER NO OUTERINOAT1G23420YABBY family transcription factor[24]
      KERBEROSKRSAT1G50650STIG1 family of peptide[130]
      KLUHKLU (CYP78A5)AT1G13710Cytochrome p450 family[35]
      LEAFY COTYLEDON 1LEC1AT1G21970Nuclear factor Y transcription factor[4850]
      LEAFY COTYLEDON 2LEC2AT1G28300B3 domains transcription factor[4850]
      MATERNAL DEREPRESSION OF r1MDR1 (DNG101)Zm00001eb202980DNA glycosylase[81]
      MATERNAL EFFECT EMBRYO ARREST45MEE45AT4G00260B3 domains transcription factor[38]
      MATERNALLY EXPRESSED PAB C-TERMINALMPCAT3G19350C-terminal domain of poly(A) binding protein[71]
      MEDEAMEAAT1G02580PRC2 component[69,72,74,75,
      83,84,89]
      METHYLTRANSFERASE 1MET1AT5G49160Methyltransferase 1[70, 83,84,
      86,106]
      MINISEED3MINI3AT1G55600WRKY family transcription factor, WRKY10[6,7]
      MIR159aMIR159aAT1G73687MicroRNA[112]
      MIR159bMIR159bAT1G18075MicroRNA[112]
      MIR159cMIR159cAT2G46255MicroRNA[112]
      MYB33MYB33AT5G06100MYB family transcription factor[112]
      MYB65MYB65AT3G11440MYB family transcription factor[112]
      PHERES 1PHE1(AGL37)AT1G65330MADS-box family transcription factor[8385]
      PHOSPHATE 1PHO1AT3G23430Phosphate transporter[19]
      PICKLE RELATED 2PKR2AT4G31900Chromatin remodeling factor[104]
      OsBBM1OsBBM1LOC_Os11g19060AP2 family transcription factor[110,111]
      SHAGGY-LIKE KINASE 11SK11AT5G26751GSK3 family/SHAGGY-like protein kinase[16, 18]
      SHAGGY-LIKE KINASE 12SK12AT3G05840GSK3 family/SHAGGY-like protein kinase[16,18]
      SHORT HYPOCOTYL UNDER BLUE1SHB1AT4G25350homologous with SYG1 protein family members, transcription regulator[9]
      SHORT SUSPENSORSSPAT2G17090Receptor-like cytoplasmic protein kinase[108,109]
      SmD1bSmD1bAT4G02840Smith protein[18]
      TERMINAL FLOWER1TFL1AT5G03840Phosphatidylethanolamine binding protein (PEBP) family member[121]
      TOPOISOMERASE IαTOP1αAT5G55300DNA topoisomerase[117]
      TRANSPARENT TESTA 16TT16 (AGL32)AT5G23260MADS-box family transcription factor[123]
      TRANSPARENT TESTA 2TT2AT5G35550MYB family transcription factor[13]
      TRANSPARENT TESTA 8TT8AT4G09820bHLH family transcription factor[14]
      TRANSPARENT TESTA GLABRA 1TTG1AT5G24520WD40-motif containing transcription regulator[15, 16]
      TRANSPARENT TESTA GLABRA 2TTG2AT2G37260WRKY family transcription factor, WRKY44[28,117]
      TWISTED SEED 1TWS1AT5G01075Signaling peptide precursor[126, 127]
      UBIQUITIN-SPECIFIC PROTEASE 12UBP12AT5G06600Deubiquitination enzyme[27]
      UBIQUITIN-SPECIFIC PROTEASE 13UBP13AT3G11910Deubiquitination enzyme[27]
      UP-FRAMESHIFT SUPPRESSOR 1UPF1AT5G47010RNA helicase[117]
      YODAYDAAT1G63700Member of MEKK subfamily, involved in MAPK cascade[108,109]
      ZHOUPIZOUAT1G49770bHLH family transcription factor[128, 129]
      ZmGW2-CHR4ZmGW2-CHR4Zm00001eb204560RING-type E3 ubiquitin ligase[43]
      ZmGW2-CHR5ZmGW2-CHR5Zm00001eb238650RING-type E3 ubiquitin ligase[43]
      ZmSWEET4cZmSWEET4cZm00001eb236820Sugar transporter[20]
    • A maternal effect is generally defined as a phenomenon in which the offspring phenotype is determined by the genotype of its mother. However, the angiosperm seeds bear mixed features of two generations, which makes the maternal effects of angiosperms more complicated than those of animals.

      Usually, maternal effects can be observed from reciprocal crosses, where the F1 progenies that have the same genetic background might show different phenotypes from the mother (Fig. 2a & b). For example, several TRANSPARENT TESTA (TT) genes in Arabidopsis, including TT2[13], TT8[14], and TRANSPARENT TESTA GLABRA1 (TTG1)[15,16], show maternal effects on the accumulation of seed oil in F1 progenies. The seeds from tt mutants pollinated with wild-type and its own pollen have similarly higher levels of seed oil than wild-type seeds, whereas the seeds from wild-type plants pollinated with tt pollen do not exhibit such a phenotype. Likewise, the regulators acting upstream and downstream of TTG1, including SmD1b, SHAGGY-LIKE KINASE 11/12 (SK11/12), and GLABRA2 (GL2), show maternal effects on regulating seed oil levels[1618]. In addition, phosphate (Pi) exporters localized in the chalazal seed coat are crucial for Pi flux between the chalazal seed coat and the embryo, and such a remote control is evidenced by grafting assays[19]. Sugar transporters expressed in the maternal seed coat are responsible for transferring hexoses across the basal endosperm transfer layer to the starch-storing endosperm in rice and maize[20]. These findings suggest that maternal effects play a central regulatory role in relocating resources into seeds.

      Figure 2. 

      Maternal control of seed development. (a) Symbols of maternal and filial tissues appearing in this figure. (b) Scheme of typical maternal effects. The phenotype of developing or mature seeds is determined by the maternal genotype. (c) Scheme of on-site effects of the maternal tissue. The phenotype is restricted to the tissue inherited from the mother, and thus determined by the maternal genotype. (d) Characterization of gametophytic maternal effects by test crosses. As the phenotype of developing or mature seeds is determined by the genotype of the female gametophyte, phenotypic segregation is observable in F1 progenies of test crosses. (e) Characterization of sporophytic maternal effects by test crosses. As the phenotype of developing or mature seeds is determined by the genotype of the female sporophyte, phenotypic segregation is unobservable in F1 progenies of test crosses.

      It's worth noting that the tt mutants display another common phenotype of the non-pigmented seed coat[21]. However, regulation of such a seed phenotype is not attributed to canonical maternal effects because the integument-derived seed coat and its pigmentation are directly inherited from the mother generation (Fig. 2c). Likewise, maternal defects of the integument and megaspore are also not a consequence of canonical maternal effects because the filial generation is not involved as exemplified by the regulation conferred by AINTEGUMENTA (ANT)[22,23] and INNER NO OUTER (INO)[24].

      In contrast, the integument influences on seed size are considered maternal effects[25]. The integument proliferation and expansion make the cavity for filial development. In this regard, the 'DA' pathway, where 'DA' means 'large' in Chinese, contains a group of genes in the ubiquitin pathway and regulates the seed size maternally and sporophytically[26,27]. TTG2[28], APETALA2 (AP2)[29], and ABERRANT TESTA SHAPE (ATS)[30,31] also act maternally to control integument characteristics. In addition, the signals from the maternal integument also control the outcome of filial development. Maternal siRNAs, which are produced in sporophytic tissues, such as the integument, or transcribed by the maternal alleles in the endosperm, repress the AGAMOUS-LIKE transcription factors (AGLs) in the endosperm to regulate endosperm development[3234]. Moreover, several cytochrome P450s (CYPs), including KLUH (KLU; CYP78A5)[35], ENHANCER OF da1-1 3 (EOD3; CYP78A6)[36], and possibly CYP78A7/A9[36,37], generate mobile maternal signals to regulate seed size. Besides, maternal auxin provided by the integument regulates embryonic cell proliferation and patterning[38,39], while maternal gibberellin is crucial for the programmed cell death of the embryonic suspensor[40].

      In monocots, there are several functionally conserved pathways that exert maternal effects. For example, the causal genes of the QTL GRAIN WEIGHT 2 (GW2) in rice[41], wheat[42], and maize[43] maternally regulate seed size. These genes encode E3 ligases homologous to DA2 in Arabidopsis. However, different seed structures of monocots and dicots (Fig. 1a) implicate partially distinct regulatory mechanisms. As the seeds of the grass family are usually merged with the pericarp and covered by husks, the maternal effects may be related to fruit or flower tissues. A lot of brassinosteroid-related mutants exhibit altered grain size and shape[44], which are at least attributed to the misregulation of cell division or expansion in lemma and palea. Besides, different seed structures also indicate different functional modes of homologous genes in monocots and dicots. For example, GIANT EMBRYO (GE) encodes CYP78A13 and regulates the balance of endosperm and embryo development in rice[45,46], while its Arabidopsis homologs do not exert such an effect.

      In contrast to the father, both maternal sporophyte and female gametophyte are involved in the control of seed development. Thus, the maternal effects may act sporophytically or gametophytically. Maternal gametophytic effects refer to the phenotype of offspring determined by the haplotype inherited from the mother. It can be distinguished from the sporophytic effects by test crosses, where heterozygotes are used as maternal plants to be pollinated with the pollen from a homozygous donor (Fig. 2d). A maternal gametophytic effect results in a 1:1 segregation ratio in the progenies from the test cross, whereas a maternal sporophytic effect does not cause phenotypic segregation of the progenies from the test cross (Fig. 2e). Although the phenotype of the maternal gametophytic effect is similar to that caused by maternally imprinted genes (MEGs; see Fig. 3), they are conceptually different. A maternal gametophytic effect could be explained by the biologically active components inherited from the maternal gametophyte, while MEGs are responsible for de novo synthesis of biologically active components in the filial tissue. For example, ADRENODOXIN REDUCTASE (ADXR), ADRENODOXINS (ADXs), and their targets, mitochondrial cytochrome P450s, are important regulators of the mitochondrial steroidogenic pathway in female gametophytes, and they influence early embryogenesis in a maternal gametophytic manner[47].

      Figure 3. 

      Zygotic control of seed development. (a) A typical zygotic effect causes phenotype segregation among F1 siblings. (b) Scheme of Mendelian and non-Mendelian inheritance patterns of F1 progenies with zygotic effects. Left panels: possible phenotypes of F1 progenies with a recessive, semi-dominant, or dominant mutation. Right panels: possible phenotypes of F1 progenies with a non-mendelian mutation. Such patterns of non-mendelian inheritance are likely related to parental interactions. Ellipse indicates the quantified range of a phenotype. (c) Scheme of endospermic factors regulating F1 phenotypes in a parental-dependent manner. (d) Typical mechanisms underlying gene imprinting. MEG, maternally imprinted gene; PEG, paternally imprinted gene. (e) Phenotypic assumptions based on unbalanced parental dosage. The first four panels from the left show phenotypic patterns of reciprocal crosses between wild-type plants and plants with loss of function or overexpression of imprinted genes. The fifth panel shows phenotypic patterns of interploidy crosses in Arabidopsis, while the sixth panel shows a smilar phenotype between the paternal-excess cross (2nd column) and the cross with loss of MEG (meg) (4th column). Such phenotypes are suppressed by loss of PEG (peg) (3rd and 5th columns). Ellipse indicates the quantified range of a phenotype, while half ellipse indicates a possible abortive phenotype.

    • In contrast to maternal effects, zygotic effects delineate a phenomenon where the offspring phenotype is determined by its own genotype. Theoretically, any autonomous regulation of embryogenesis should show a zygotic effect. For example, the LEAFY COTYLEDON (LEC) class genes in Arabidopsis, including LEC1/2, ABSCISIC ACID INSENSITIVE3 (ABI3), and FUSCA3 (FUS3), are major regulators of embryogenesis and endosperm development[4850]. The zygotic effects are manifested in these lec mutants as phenotypic segregation of individually developing seeds is observable in siliques of heterozygous plants (Fig. 3a). The filial LEC class genes are regulated by BABY BOOM (BBM)[51], which is one of the major inducers of early embryogenesis, and ectopic expression of BBM is sufficient to induce asexual and somatic embryo development[52,53]. As BBM is expressed in maternal sporophyte and gametophyte cells as well as filial zygotic cells[52], its effect on LEC genes implies a transition from maternal control to zygotic control. This transition is associated with the zygotic genome activation (ZGA)[54], in which both maternal and paternal genomes start to exert function in the filial cells.

      In some special scenarios, filial phenotypes are superior or inferior to those of both parents, which is known as hybrid vigor (heterosis) or hybrid necrosis, respectively (Fig. 3b). These effects are not explained by the genetic background of F1 progenies, regardless of whether the mutation is recessive, dominant, or semi-dominant/dosage-dependent. Such patterns of non-Mendelian inheritance are likely related to parental interactions, including at the epigenetic level, although the mechanisms are so far unclear at the molecular level (Fig. 3b). Investigations so far have shown that hybrid necrosis is physiologically similar to auto-immunity and depends on the interactions between pairs of parent-of-origin compounds[55], while epigenetic regulation provides a possible platform for hybrid vigor[56,57] because the interaction of parental epi-alleles confers new characteristics in the F1 progenies. Overall, hybrid vigor and necrosis that influence the F1 seed development are special cases of zygotic effects[58], which are of particular value for crop breeding, including phenotypic improvement of filial generations[5961].

      Moreover, the endosperm is also considered as filial (zygotic) tissue. For canonically recessive or dominant mutants with endosperm defects, the filial segregation of heterozygous plants is observed as those with embryonic defects. For example, the maize ears of endosperm breakdown1 (enb1) heterozygous mutants contain normal or endosperm-defective kernels according to the genotypes of individual kernels[62]. However, whilst the genome is equally inherited from the parents (maternal : paternal = 1:1) in the embryo, the parental contributions are unequal in the endosperm (maternal : paternal = 2:1). Therefore, the F1 progenies of reciprocal crosses display different endosperm genotypes and parent-of-origin effects are thus expected in reciprocal crosses if the mutation is semi-dominant, dosage-dependent, or parentally biased (Fig. 3c). A group of genes with allele-specific expression depending on their parental origin is called imprinted genes. These genes are known as MEGs or paternally expressed genes (PEGs), as their transcription is only activated in the allele inherited from the mother or father, respectively. Imprinted genes are mainly found in the endosperm of both monocots and dicots[6366] and regulate seed phenotypes in a parent-of-origin manner (Fig. 3c). The establishment of imprinting generally requires high activity of DNA demethylation and H3K27me3 deposition in the central cell as well as early endosperm, compared to the sperm[67].

      The establishment of MEGs can be achieved by relieving MEGs from repression compared to the paternal allele, which depends on the passive activation of DEMETER (DME; eraser for DNA methylation) in the central cell[68] (Fig. 3d). Consequently, loss of maternal DME leads to seed abortion because of impaired imprinting[69]. Meanwhile, mutation of METHYLTRANSFERASE 1 (MET1), which is a CG DNA methyltransferase, also leads to a parent-of-origin effect on seed size because of the imprinting disturbance in the endosperm[70]. The biased parental DNA methylation by DME and MET1 accounts for the maternal allele-specific expression of FLOWERING WAGENINGEN (FWA), FERTILIZATION INDEPENDENT SEED 2 (FIS2), and MATERNALLY EXPRESSED PAB C-TERMINAL (MPC)[7173], and at least partially for MEDEA (MEA)[69,72,74,75]. Besides, the paternal alleles of some MEGs can be actively silenced by the noncanonical RNA-directed DNA methylation (RdDM) pathway activated in nurse cells of gametes[7679], which leaves the remaining maternal allele active in the endosperm (Fig. 3d).

      However, except for the parental differential DNA methylation, a significant number of MEGs are established by unknown mechanisms beyond DNA methylation[80,81]. As the putative MEGs could have been contaminated by the genes expressed in maternal sporophytic tissues[82], it remains an open question if alternative pathways for MEG establishment exist or not. In addition, some imprinted genes are not conserved among different accessions of the same species[80], suggesting that the imprinting status could be dynamic or altered during evolution.

      The establishment of PEGs is achieved mainly by silencing of the maternal allele (Fig. 3d). For example, it is hypothesized that DNA methylation of the 3' flanking sequence of PHERES 1 (PHE1), a PEG, excludes H3K27me3 deposition in the endosperm. The maternal allele of PHE1 is demethylated in the central cell by DME, facilitating subsequent H3K27me3-mediated silencing in the endosperm, whereas the paternal allele of PHE1 keeps DNA methylation, which is maintained by MET1 and remains active in the endosperm[83,84]. As an AGAMOUS-LIKE transcription factor, PHE1 further controls the imprinting of other loci in the endosperm[85]. In a genomic view, some PEGs are downregulated when the paternal loss of MET1 is introduced[86], indicating that paternal DNA methylation is common for PEG activation. Although H3K27me3 is a core silencing mark for the maternal alleles of PEGs[87], these loci are not highly correlated with DME-mediated DNA demethylation[80]. It is possible that H3K27me3 itself functions to build the parental asymmetry independently of DNA methylation (Fig. 3d). Notably, silencing of the maternal allele of PEGs by H3K27me3 is frequently associated with non-CG DNA methylation and H3K9me2 histone modifications, which could be the subsequent mechanisms contributing to imprinting establishment[88]. Given that the maternal PRC2 components, such as MEA and FIS2, regulate PEGs as well, maternal regulation is generally more dominant over paternal regulation[89]. This is in agreement with the notion that seed development relies on maternal tissues.

      Although many imprinting genes have been identified, most mutants of imprinted genes (especially PEGs) do not show obvious phenotypes[90,91] except that there are increasing literature reporting the link between imprinted genes and potential seed phenotypes[92,93]. The most known imprinting-related phenotype is seed abortion caused by endosperm overproliferation and cellularization failure. Theoretically, a seed phenotype related to a given imprinted gene is either maternally or paternally determined, although imprinted genes fundamentally function zygotically (Fig. 3e). Interestingly, a dramatic parent-of-origin effect is observed in the interploidy reciprocal cross in Arabidopsis, where the F1 progenies from the reciprocal parental origins show opposite phenotypes[94,95]. Similar phenomenon in reciprocal interploidy crosses is also reported in monocots, such as maize and rice[9698]. Although tetraploid seeds are generally larger than diploid seeds, the F1 seeds from the maternal excess cross (♀tetraploid × diploid♂) are smaller than diploid seeds (precocious endosperm cellularization), while the F1 seeds from the paternal excess cross (♀diploid × tetraploid♂) are larger than tetraploid seeds or even aborted (delayed/failed endosperm cellularization) (Fig. 3e). Like the hybrid vigor or necrosis, such patterns of non-Mendelian inheritance in interploidy crosses indicate parental interactions.

      Opposite phenotypes between maternal excess and paternal excess crosses imply that MEGs and PEGs tend to restrict and promote endosperm growth, respectively. This is also supported by the findings that seed abortion caused by defective MEGs can be partially rescued by additional loss of some PEGs in Arabidopsis[84,99]. In particular, PEGs are critical for seed abortion caused by paternal excess, which is known as the triploid block or interploidy barrier. Mutants of several PEGs and mutants with failed PEG establishment suppress the phenotype of paternal excess[77,90,99105], although they do not cause visible defects per se in diploids (Fig. 3e). Global paternal demethylation bypasses triploid block[106,107], suggesting that paternal epigenome is crucial. It is possible that imprinted genes function together as a genomic feature rather than acting as individual regulators to regulate seed phenotypes. Therefore, obvious seed phenotypes are only found in the genetic backgrounds with dramatic or global disturbance of parental balance, such as interploidy progenies or mutants of imprinted genes that are epigenetic mark builders and general transcription factors.

    • Different cell types with distinct genetic backgrounds are involved in seed development, and the communications among cell types make the underlying regulations more complicated and cannot be simply interpreted as maternal or zygotic effect. Signal communications among multiple cell types are expected, but the relevant mechanisms are largely obscure. For example, pollen-delivered SHORT SUSPENSOR (SSP) mRNA is only translated in the zygote soon after fertilization to regulate the zygotic YODA (YDA) pathway, thereby controlling the asymmetric division[108,109], while paternal OsBBM1 transcripts delivered by the pollen trigger early embryogenesis in rice[110,111]. In addition, paternal miR159 represses central cell-inherited MYB33/65 to allow endosperm nuclear division[112] (Fig. 4a). Such paternal triggers could be more widespread, as there is evidence showing that pollen tube contents can mimic fertilization and induce the growth of maternal sporophytic tissues[113].

      Figure 4. 

      Inter-tissue communication in seed development. (a) Inter-tissue communication at the beginning of seed development. Paternal SSP mRNA from the pollen affects the zygotic YDA pathway to determine zygote division. Paternal miR159 from the pollen quenches maternal MYB33/65 to initiate nascent endosperm division. Other contents in the pollen tube can also trigger ovule growth, which mimicks fertilization. (b) Inter-tissue communication in early seed development. Synergid nuclei affect the maternal-paternal genome ratio in the nascent endosperm, which is oppositely regulated by sporophytic and gametophytic EIN3. The communication between maternal antipodal cells and paternal cues in the nascent endosperm relies on the relative dosage of maternal and paternal TTG2, which is transcriptionally regulated by TOP1α and UPF1. Nascent endosperm-female gametophyte communication is also suggested, although the mechanisms are yet unknown. (c) Inter-tissue communication regulating endosperm and integument development. Chalazal-transcribed TFL1 functions in the peripheral endosperm to regulate endosperm cellularization. This module also infers a potential maternal-filial communication at the chalazal part. Endosperm regulators, AGLs, are regulated by maternal siRNAs from both the endosperm and maternal tissues. AGL62 in turn regulates maternal nucellus degradation via the maternal TT16 and integument growth via the maternal PcG complex. (d) Inter-tissue communication between endosperm and embryo. When the cuticle barrier between endosperm and embryo is not established, endosperm-expressed LEC1 relocates into the embryo to exert its function. The integrity of such a barrier is monitored by two-way communication, in which the precursor of the embryo-expressed TWS1 peptide (TWS1pre) is processed in the endosperm by ALE1 and the mature peptide signal moves back into the embryo to activate the GSO1/2-pathway. (e) Color legend shows different elements in this figure. Single- and double-headed arrows indicate one-way and reciprocal regulations, respectively. Dashed arrow indicates a putative regulation. Green single-headed arrow indicates protein movement, while green gradient single-headed arrow indicates protein movement along with the maturation process.

      Parental communications are crucial at the initial stage of seed development, especially with regard to the female gametophytic cues and nascent endosperm (Fig. 4b). It has been recently revealed that nascent endosperm growth is related to the disintegration of synergid-derived nuclei, which affects the overall maternal-paternal ratio in the nascent endosperm. The disintegration of synergid nuclei is inhibited by maternal sporophytic ETHYLENE INSENSITIVE 3 (EIN3), but promoted by gametophytic EIN3[114]. Besides, the nitrous oxide treatment in maize causes defective endosperms in the affected kernel, where the dosage balance between the newly synthesized compounds in nascent endosperm and compounds inherited from female gametophytes is changed without affecting the maternal-paternal balance in the endosperm, implying an as-yet-unknown female gametophyte-endosperm communication[115,116]. Such communication is also found in Arabidopsis. In addition to the well-known sporophytic functions of TTG2, the relative parental dosage of gametophytic TTG2 also affects the final seed size[117]. Maternal TTG2 is expressed in antipodal cells, while paternal TTG2 is expressed in the nascent endosperm inherited from the sperm. Such a parental module influences the outcome of interploidy crosses, although it is still unknown why the maternal and paternal TTG2 exert antagonistic functions[117]. Interestingly, antipodal cells, whose foci are later replaced by chalazal endosperm, are most extensively regulated by parental cues[118]. The chalazal part is also crucial for nutrient exchange between maternal and filial tissues[119], echoing the principles inferred by the parental conflict hypothesis[120]. These findings hint that it may be common for a gene to act oppositely in different parental contexts at the beginning of seed development.

      At the late seed development stage, TERMINAL FLOWER1 (TFL1) regulates seed size by affecting endosperm cellularization. TFL1 is transcribed in the chalazal endosperm, but its protein is relocated into the peripheral endosperm to take action (Fig. 4c). Genetic data reveals the maternal effect of TFL1 mutation, suggesting a potential signal exchange between maternal and zygotic tissues[121]. At this stage, endosperm-maternal sporophyte communications are common. The major endosperm regulator AGAMOUS-LIKE 62 (AGL62)[122], which is not imprinted but under the paternal control of PHE1[85], can guide maternal nucellus degradation by promoting maternal TT16 expression[123], and regulates auxin transport from the endosperm to the integument to repress maternal Polycomb Group (PcG) function[124] (Fig. 4c). Considering the previously mentioned maternal regulation of the endosperm AGLs by siRNAs[3234], reciprocal ways exist for the paternal-maternal antagonism. Although more details remain to be revealed regarding the maternal-paternal interaction, the paternal effect is generally weaker than the maternal one[125].

      After endosperm cellularization, the endosperm-embryo communications are crucial for the seed maturation process (Fig. 4d). The endosperm-produced LEC1 is transported into the embryo to participate in the transcriptional programming during embryo maturation[49]. The overall LEC1 expression level is more correlated to the maternal allele because of the 2:1 genome ratio in the endosperm. Meanwhile, an embryo-produced peptide TWISTED SEED1 (TWS1) is processed by endosperm-expressed subtilisin-like protease ABNORMAL LEAF-SHAPE 1 (ALE1) and in turn perceived by embryo-presented receptors GASSHO1/2 (GSO1/2) and their co-receptors[126,127] when the cuticle barrier between embryo and endosperm is not fully established. This two-way communication acts downstream of the regulatory module in which the endosperm-expressed transcription factors ZHOUPI (ZOU) and INDUCER OF CBF EXPRESSIONICE 1 (ICE1) regulate embryo development by controlling the expression of ALE1 in the embryo-surrounding region (ESR)[128,129] and another putative signal-function peptide KERBEROS (KRS) in the endosperm[130]. These components function together to build a molecular sensor for cuticle integrity between endosperm and embryo, which is a marker delineating seed development stages.

    • Apart from the extensive studies showing the nature of parental regulation on seed development in Arabidopsis, emerging studies have also shown that such regulations are valuable for crop engineering. In rice, a significant number of imprinted genes are associated with grain yield quantitative trait loci with the potential function of regulating nutrient metabolism and endosperm development[131]. These findings echo the parental conflict hypothesis: mothers restrict the resource allocation for seed development to feed all their descendants, while fathers help their offspring evade this maternal restriction[120]. Therefore, investigation of parental regulations on seed development is certainly important for improving seed yield and quality for various crops.

      Because of the tissue complexity and genetic diversity, histological and genetic analyses are essential for functional studies of potential parental interactions. However, such data only provide a rough framework to assess the nature of a potential regulation, while the detailed mechanisms must be revealed by other combined approaches. Traditional biochemical assays and in vitro tests have inherent disadvantages in revealing mechanisms during seed development because of missing of intercellular information, which is, however, critical for understanding seed development. Emerging single-cell technologies are likely good platforms to reveal cellular relationships during seed development. Using single-cell technologies, DNA methylation, chromatin accessibility, protein abundance, and gene perturbation can be investigated at the sub-tissue level[132]. For example, single-nucleus sequencing of Arabidopsis endosperm has revealed the functional partitioning among endosperm nuclei, with the chalazal endosperm showing the most parentally biased expression[118]. This is consistent with the fact that the chalazal part is the interface of maternal and filial tissues, which could be the frontline of maternal-filial signal communications. With a clear understanding of the parental interplay among various cell types involved in seed development, valuable molecular targets could be identified and precisely modified for crop improvement.

      • This work was supported by Singapore Food Story R&D Programme (SFS_RND_SUFP_001_04) and the Reimagine Research Grant from the National University of Singapore. We apologize to authors whose excellent work could not be cited owing to space limitations.

      • The authors declare that they have no conflict of interest.

      • Copyright: © 2022 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 (4)  Table (1) References (132)
  • About this article
    Cite this article
    Li C, Yu H. 2022. Parental regulation of seed development. Seed Biology 1:7 doi: 10.48130/SeedBio-2022-0007
    Li C, Yu H. 2022. Parental regulation of seed development. Seed Biology 1:7 doi: 10.48130/SeedBio-2022-0007

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return