Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Epigenetic, genetic and maternal effects enable stable centromere inheritance

Abstract

Centromeres are defined epigenetically by the histone H3 variant CENP-A. The propagation cycle by which pre-existing CENP-A nucleosomes serve as templates for nascent assembly predicts the epigenetic memory of weakened centromeres. Using a mouse model with reduced levels of CENP-A nucleosomes, we find that an embryonic plastic phase precedes epigenetic memory through development. During this phase, nascent CENP-A nucleosome assembly depends on the maternal Cenpa genotype rather than the pre-existing template. Weakened centromeres are thus limited to a single generation, and parental epigenetic differences are eliminated by equal assembly on maternal and paternal centromeres. These differences persist, however, when the underlying DNA of parental centromeres differs in repeat abundance, as assembly during the plastic phase also depends on sufficient repetitive centromere DNA. With contributions of centromere DNA and the Cenpa maternal effect, we propose that centromere inheritance naturally minimizes fitness costs associated with weakened centromeres or epigenetic differences between parents.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Evidence for epigenetic centromere memory through mouse reproduction.
Fig. 2: Male and female soma show reduced CENP-A compared to oocytes.
Fig. 3: Epigenetic differences between parental centromeres are not maintained.
Fig. 4: Centromere strength depends on maternally inherited CENP-A.
Fig. 5: CENP-A chromatin recovers in adult male F2 progeny from Cenpa+/+ WT F1 parents.
Fig. 6: Genetic contributions to centromere equalization in early embryogenesis.

Similar content being viewed by others

Data availability

Previously published microarray data for long poly-(A) tailed Cenpa mRNA in pre-implantation development is available freely on the NCBI Gene Expression Omnibus (GEO) database (accession no. GDS813 from reference series GSE1749). The 12 mouse genomes used for Cenpa 3′ UTR analysis are available from the NCBI BioProject database (https://www.ncbi.nlm.nih.gov/bioproject) under accession no. PRJNA669840. Data supporting the findings of this study are available from the corresponding author on reasonable request. Source data are provided with this paper.

Code availability

All codes used for statistical and distribution analysis are freely available as part of the R package ‘multimode’, described in ref. 58.

References

  1. Kixmoeller, K., Allu, P. K. & Black, B. E. The centromere comes into focus: from CENP-A nucleosomes to kinetochore connections with the spindle. Open Biol. 10, 200051 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Dumont, M. & Fachinetti, D. DNA sequences in centromere formation and function. Prog. Mol. Subcell. Biol. 56, 305–336 (2017).

    Article  CAS  PubMed  Google Scholar 

  3. Chmátal, L., Schultz, R. M., Black, B. E. & Lampson, M. A. Cell biology of cheating-transmission of centromeres and other selfish elements through asymmetric meiosis. Prog. Mol. Subcell. Biol. 56, 377–396 (2017).

    Article  PubMed  CAS  Google Scholar 

  4. Iwata-Otsubo, A. et al. Expanded satellite repeats amplify a discrete CENP-A nucleosome assembly site on chromosomes that drive in female meiosis. Curr. Biol. 27, 2365–2373.e8 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Akera, T., Trimm, E. & Lampson, M. A. Molecular strategies of meiotic cheating by selfish centromeres. Cell 178, 1132–1144 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Fishman, L. & Saunders, A. Centromere-associated female meiotic drive entails male fitness costs in monkeyflowers. Science 322, 1559–1562 (2008).

    Article  CAS  PubMed  Google Scholar 

  7. Voullaire, L. E., Slater, H. R., Petrovic, V. & Choo, K. H. A. A functional marker centromere with no detectable α-satellite, satellite III or CENP-B protein: activation of a latent centromere? Am. J. Hum. Genet. 52, 1153–1163 (1993).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Logsdon, G. A. et al. Human artificial chromosomes that bypass centromeric DNA. Cell 178, 624–639 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Depinet, T. W. et al. Characterization of neo-centromeres in marker chromosomes lacking detectable α-satellite DNA. Hum. Mol. Genet. 6, 1195–2204 (1997).

    Article  CAS  PubMed  Google Scholar 

  10. Du Sart, D. et al. A functional neo-centromere formed through activation of a latent human centromere and consisting of non-α-satellite DNA. Nat. Genet. 16, 144–153 (1997).

    Article  PubMed  Google Scholar 

  11. Black, B. E. & Cleveland, D. W. Epigenetic centromere propagation and the nature of CENP-A nucleosomes. Cell 144, 471–479 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Foltz, D. R. et al. Centromere-specific assembly of CENP-A nucleosomes is mediated by HJURP. Cell 137, 472–484 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Jansen, L. E. T., Black, B. E., Foltz, D. R. & Cleveland, D. W. Propagation of centromeric chromatin requires exit from mitosis. J. Cell Biol. 176, 795–805 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Dunleavy, E. M. et al. HJURP is a cell-cycle-dependent maintenance and deposition factor of CENP-A at centromeres. Cell 137, 485–497 (2009).

    Article  CAS  PubMed  Google Scholar 

  15. Schuh, M., Lehner, C. F. & Heidmann, S. Incorporation of Drosophila CID/CENP-A and CENP-C into centromeres during early embryonic anaphase. Curr. Biol. 17, 237–243 (2007).

    Article  CAS  PubMed  Google Scholar 

  16. Moree, B., Meyer, C. B., Fuller, C. J. & Straight, A. F. CENP-C recruits M18BP1 to centromeres to promote CENP-A chromatin assembly. J. Cell Biol. 194, 855–871 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Ravi, M. & Chan, S. W. L. Haploid plants produced by centromere-mediated genome elimination. Nature 464, 615–618 (2010).

    Article  CAS  PubMed  Google Scholar 

  18. Comai, L. & Tan, E. H. Haploid induction and genome instability. Trends Genet. 35, 791–803 (2019).

    Article  CAS  PubMed  Google Scholar 

  19. Raychaudhuri, N. et al. Transgenerational propagation and quantitative maintenance of paternal centromeres depends on Cid/Cenp-A presence in Drosophila sperm. PLoS Biol. 10, e1001434 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Gassmann, R. et al. An inverse relationship to germline transcription defines centromeric chromatin in C. elegans. Nature 484, 534–537 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Liu, H., Kim, J. M. & Aoki, F. Regulation of histone H3 lysine 9 methylation in oocytes and early pre-implantation embryos. Development 131, 2269–2280 (2004).

    Article  CAS  PubMed  Google Scholar 

  22. Burton, A. et al. Heterochromatin establishment during early mammalian development is regulated by pericentromeric RNA and characterized by non-repressive h3k9me3. Nat. Cell Biol. 22, 767–778 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Puschendorf, M. et al. PRC1 and Suv39h specify parental asymmetry at constitutive heterochromatin in early mouse embryos. Nat. Genet. 40, 411–420 (2008).

    Article  CAS  PubMed  Google Scholar 

  24. Palmer, D. K., O’Day, K. & Margolis, R. L. The centromere specific histone CENP-A is selectively retained in discrete foci in mammalian sperm nuclei. Chromosoma 100, 32–36 (1990).

    Article  CAS  PubMed  Google Scholar 

  25. Palmer, D. K., Day, K. O., Trongt Le, H., Charbonneau, H. & Margolis, R. L. Purification of the centromere-specific protein CENP-A and demonstration that it is a distinctive histone. Proc. Natl Acad. Sci. USA 88, 3734–3738 (1991).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Brinkley, B. R. et al. Arrangements of kinetochores in mouse cells during meiosis and spermiogenesis. Chromosoma 94, 309–317 (1986).

    Article  CAS  PubMed  Google Scholar 

  27. Smoak, E. M., Stein, P., Schultz, R. M., Lampson, M. A. & Black, B. E. Long-term retention of CENP-A nucleosomes in mammalian oocytes underpins transgenerational inheritance of centromere identity. Curr. Biol. 26, 1110–1116 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Richter, J. D. Cytoplasmic polyadenylation in development and beyond. Microbiol. Mol. Biol. Rev. 63, 446–456 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Zeng, F., Baldwin, D. A. & Schultz, R. M. Transcript profiling during preimplantation mouse development. Dev. Biol. 272, 483–496 (2004).

    Article  CAS  PubMed  Google Scholar 

  30. Ishiuchi, T. et al. Reprogramming of the histone H3.3 landscape in the early mouse embryo. Nat. Struct. Mol. Biol. 28, 38–49 (2021).

    Article  CAS  PubMed  Google Scholar 

  31. Barnhart, M. C. et al. HJURP is a CENP-A chromatin assembly factor sufficient to form a functional de novo kinetochore. J. Cell Biol. 194, 229–243 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Fujita, Y. et al. Priming of centromere for CENP-A recruitment by human hmis18α, hmis18β and M18BP1. Dev. Cell 12, 17–30 (2007).

    Article  CAS  PubMed  Google Scholar 

  33. Nardi, I. K., Zasadzińska, E., Stellfox, M. E., Knippler, C. M. & Foltz, D. R. Licensing of centromeric chromatin assembly through the Mis18α-Mis18β heterotetramer. Mol. Cell 61, 774–787 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Stellfox, M. E., Nardi, I. K., Knippler, C. M. & Foltz, D. R. Differential binding partners of the Mis18α/β YIPPEE domains regulate Mis18 complex recruitment to centromeres. Cell Rep. 15, 2127–2135 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Bodor, D. L. et al. The quantitative architecture of centromeric chromatin. eLife 3, e02137 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Masumoto, H., Masukata, H., Muro, Y., Nozaki, N. & Okazaki, T. A human centromere antigen (CENP-B) interacts with a short specific sequence in alphoid DNA, a human centromeric satellite. J. Cell Biol. 109, 1963–1973 (1989).

    Article  CAS  PubMed  Google Scholar 

  37. Fachinetti, D. et al. DNA sequence-specific binding of CENP-B enhances the fidelity of human centromere function. Dev. Cell 33, 314–327 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Kumon, T. et al. Parallel pathways for recruiting effector proteins determine centromere drive and suppression. Cell 184, 4904–4918 (2021).

    Article  CAS  PubMed  Google Scholar 

  39. Rossant, J. Postimplantation development of blastomeres isolated from 4 and 8 cell mouse eggs. J. Embryol. Exp. Morphol. 36, 283–290 (1976).

    CAS  PubMed  Google Scholar 

  40. Lepikhov, K. & Walter, J. Differential dynamics of histone H3 methylation at positions K4 and K9 in the mouse zygote. BMC Dev. Biol. 4, 12 (2004).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Lepikhov, K. et al. Evidence for conserved DNA and histone H3 methylation reprogramming in mouse, bovine and rabbit zygotes. Epigenetics Chromatin 1, 8 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Xu, Q. & Xie, W. Epigenome in early mammalian development: inheritance, reprogramming and establishment. Trends Cell Biol. 28, 237–253 (2018).

    Article  CAS  PubMed  Google Scholar 

  43. Xie, B. et al. Histone H3 lysine 27 trimethylation acts as an epigenetic barrier in porcine nuclear reprogramming. Reproduction 151, 9–16 (2016).

    Article  CAS  PubMed  Google Scholar 

  44. Van Der Heijden, G. W. et al. Asymmetry in histone H3 variants and lysine methylation between paternal and maternal chromatin of the early mouse zygote. Mech. Dev. 122, 1008–1022 (2005).

    Article  PubMed  CAS  Google Scholar 

  45. Hou, H. et al. Centromeres are dismantled by foundational meiotic proteins Spo11 and Rec8. Nature 591, 671–676 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Monen, J., Maddox, P. S., Hyndman, F., Oegema, K. & Desai, A. Differential role of CENP-A in the segregation of holocentric C. Elegans chromosomes during meiosis and mitosis. Nat. Cell Biol. 7, 1248–1255 (2005).

    Article  PubMed  CAS  Google Scholar 

  47. Prosée, R. F. et al. Trans-generational inheritance of centromere identity requires the CENP-A N-terminal tail in the C. elegans maternal germ line. PLoS Biol. 19, e3000968 (2021).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Malik, H. S. & Henikoff, S. Major evolutionary transitions in centromere complexity. Cell 138, 1067–1082 (2009).

    Article  CAS  PubMed  Google Scholar 

  49. Maheshwari, S. et al. Naturally occurring differences in CENH3 affect chromosome segregation in zygotic mitosis of hybrids. PLoS Genet. 11, e1004970 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Bao, J. & Bedford, M. T. Epigenetic regulation of the histone-protamine transiiton during spermiogenesis. Reproduction 151, R55–R70 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Rathke, C., Baarends, W. M., Awe, S. & Renkawitz-Pohl, R. Chromatin dynamics during spermiogenesis. Biochim. Biophys. Acta 1839, 155–168 (2014).

    Article  CAS  PubMed  Google Scholar 

  52. Ameijeiras-Alonso, J., Crujeiras, R. M. & Rodríguez-Casal, A. Mode testing, critical bandwidth and excess mass. Test 28, 900–919 (2019).

    Article  Google Scholar 

  53. Stein, P. & Schindler, K. Mouse oocyte microinjection, maturation and ploidy assessment. J. Vis. Exp. 53, 2851 (2011).

    Google Scholar 

  54. Chatot, C. L., Ziomek, C. A., Bavister, B. D., Lewis, J. L. & Torres, I. An improved culture medium supports development of random-bred 1-cell mouse embryos in vitro. J. Reprod. Fertil. 86, 679–688 (1989).

    Article  CAS  PubMed  Google Scholar 

  55. Dia, F., Strange, T., Liang, J., Hamilton, J. & Berkowitz, K. M. Preparation of meiotic chromosome spreads from mouse spermatocytes. J. Vis. Exp. 129, 55378 (2017).

    Google Scholar 

  56. Taft, R. In vitro fertilization in mice. Cold Spring Harb. Protoc. https://doi.org/10.1101/pdb.prot094508 (2017).

  57. R: a Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2017).

  58. Ameijeiras-Alonso, J., Crujeiras, R. M. & Rodriguez-Casal, A. multimode: An R Package for Mode Assessment. J. Stat. Soft. 97, 1–32 (2021).

    Article  Google Scholar 

Download references

Acknowledgements

We thank D. P. Dudka, V. Fu and M. Barmada for assistance with genotyping, G. Logsdon for cloning a protein expression vector, M. Gerace for antigen preparation, D. P. Dudka for help with multiple sequence alignments and R. M. Schultz, M. S. Bartolomei and M. T. Levine for comments and discussion. This work was supported by the NIH (HD058730 to B.E.B. and M.A.L).

Author information

Authors and Affiliations

Authors

Contributions

A. Das contributed to experiments, quantifications, data analysis and statistical analysis, animal husbandry and genotyping. A.I.-O. carried out experiments and quantification for some of Fig. 3g. J.D.-M. prepared and characterized new reagents and assisted with statistical analysis. A. Destouni performed the initial experimentation in zygotes and early embryos. K.G.B. carried out animal husbandry and genotyping. A. Das, B.E.B. and M.A.L. contributed to experimental design, data interpretation and writing. B.E.B. and M.A.L. provided supervision and sourced funding.

Corresponding authors

Correspondence to Ben E. Black or Michael A. Lampson.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks Hiroshi Kimura and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 CENP-A chromatin is reduced in the soma of Cenpa+/− heterozygous animals in the P0 generation.

a, Bone marrow metaphase spreads: each pair of CENP-A foci represents sister centromeres in mitosis. Scale bars: 5 μm (main panel), 1μm (inset). b, Quantification of CENP-A foci intensities in control (grey) and P0 (yellow) generation in soma. N = 166, 170 centromeres (top to bottom). ** P < 0.0001, Mann–Whitney U test (two-tailed). Error bars: median ± 95% CI. Source numerical data are available in source data.

Source data

Extended Data Fig. 2 Weakened centromeres in the male germline are independent of meiotic stage.

. Because oocytes were analysed at metaphase I and spermatocytes at prophase I (Fig. 1), we confirmed that F1 spermatocytes also show weakened centromeres at metaphase I. Images (a) and quantification (b) of F1 spermatocytes show CENP-A reduced to a similar level at metaphase I (70.54 ± 7.1% of control) as prophase I. Each of the CENP-A foci represents four centromeres (a pair of homologous chromosomes, each with two sisters). N = 330 (control), 284 (F1 progeny). Scale bars: 5 μm (main panel), 1μm (inset). Quantification of SYCP3 foci from the same cells (c) shows no decrease (114.90 ± 5.6% of control). N = 235 (control), 259 (F1 progeny). ** P < 0.001, Mann–Whitney U Test (two-tailed). Error bars: median ± 95% CI. Source numerical data are available in source data.

Source data

Extended Data Fig. 3 Littermate analysis showing that weakened centromeres persist in the male but not female germline.

a, Data from Fig. 1c replotted as CENP-A levels per animal, averaged over all centromeres in each animal and normalized to controls (dashed line). N = 10,10,10, 9, 7 animals. The F1 male but not the female germline and the male and female soma are significantly lower than the controls **P < 0.001, *P < 0.05 n.s.: P > 0.05, Wilcoxon signed sum rank test (two-tailed). b, CENP-A quantifications in spermatocytes and oocytes from littermates from one set of parents. N = 121, 431, 60, 259, 246, 105 centromeres (top to bottom). Female germline levels are significantly elevated compared to littermate male germline levels. **P < 0.0001, Mann–Whitney U Test (two-tailed). Error bars: median ± 95% CI. Source numerical data are available in source data.

Source data

Extended Data Fig. 4 CENP-A nucleosomes are retained through the replacement of canonical nucleosomes with protamines during spermiogenesis.

a, Quantification and b, images showing CENP-A levels are reduced to 42.7 ± 1.5% in spermatids from Cenpa+/− males compared to WT males, similar to the reduction measured in prophase spermatocytes (Fig. 1c). N = 20 (control), 32 (Cenpa+/−) spermatids. Error bars: median ± 95% CI. Scale bars: 5 μm (main panel), 1 μm (inset). Source numerical data are available in source data.

Source data

Extended Data Fig. 5 Model to explain equalization of epigenetic differences and subsequent memory.

a, Assumptions used for the modelling. b, Epigenetic inheritance of CENP-A as determined in cycling somatic cells in culture by replication coupled dilution and G1 reloading. c, Example calculation and graph for CENP-A assembly in the first two embryonic cell cycles for progeny of a WT x WT cross. For simplicity, initial CENP-A levels are set to 100 and 50 on the maternal and paternal centromeres, respectively, based on our measurements in zygotes (Fig. 3c). At each S-phase, CENP-A levels are diluted by half on each centromere, and we assume equal assembly on maternal and paternal centromeres in the following G1. Assembly in the first cell cycle depends on the maternal pool, set to 100 for a zygote from a WT female, giving an increase of 50 on both maternal and paternal centromeres. Assembly in the second cell cycle depends on the zygotic pool, which is set to 100 for a WT zygotic genotype. d, Graphs from similar calculations as b, for the designated crosses. Initial CENP-A levels are set to 50 for maternal centromeres from Cenpa+/− mothers and 40 for paternal centromeres from Cenpa+/− fathers, based on our measurements (Fig. 1c and Fig. 3c). Arrows indicate equal assembly on maternal and paternal centromeres. In the first cell cycle, assembly is from a maternal pool of 100 (black arrows) or 50 (yellow arrows) for WT or Cenpa+/− mothers, respectively. In the second cell cycle, assembly is from a zygotic pool of 100 (purple arrows), reflecting a WT zygotic genotype. Calculations show equalization by the four-cell stage in all crosses. Furthermore, crosses with reduced maternal contribution (H♀) equalize to a lower level, which is then remembered through development. Source numerical data are available in source data.

Source data

Extended Data Fig. 6 3’ UTR of Cenpa message has hallmarks of dormant maternal mRNA.

a, Polyadenylation (addition of a poly (A) tail) of mRNA is a mechanism to control gene expression. Nuclear polyadenylation is an essential part of post-transcriptional processing of most mRNAs, dictated by the ubiquitous cis-element 3’ UTR hexameric motif AATAAA (nuclear polyadenylation element, NPE). Dormant maternal mRNAs are deposited in the oocyte with short poly(A) tails and are translationally inactive. After fertilization, these maternal mRNAs undergo translation by elongation of the poly(A) tail, controlled by a cytoplasmic polyadenylation element (CPE) usually present within 100 nt upstream of the NPE28. We find conserved CPEs in the mouse, human and frog Cenpa 3’ UTRs (CPE I = TTTTAT or CPE II = TTTTAA) upstream of the NPE as expected for dormant maternal mRNAs. b, Analysis of 12 sequenced rodent species38 reveals that CPEs (CPE I in bold boxes and CPE II in dashed boxes) are present upstream of the NPE in every species as expected for a maternal effect gene.

Extended Data Fig. 7 Symmetric distribution of H3K9me3 at the four-cell stage.

Representative cell from four-cell embryos for each of the two denoted crosses with H3K9me3 (red), CENP-A (green) and DNA (blue). H3K9me3 is present on both maternal and paternal chromatin at this stage, in contrast to zygotes (Fig. 3b and Fig. 6b–e). Scale bars: 5 μm.

Extended Data Fig. 8 CENP-A intensity distribution changes from bimodal to unimodal in early embryogenesis.

Graphs show locations of the modes in each distribution from Fig. 6a. a, The WT x WT and WT♀ x H♂ zygote distributions contain two modes (dashed lines) on either side of a central antimode (dip, pointed lines) characteristic of bimodal distributions52. The separation between the two modes is greater in the WT♀ x H♂ cross as expected. In addition, the ratios of the values of the two modes (x-axis) denoted under each cross agree well with the ratios of paternal to maternal centromere intensities calculated in Figs. 3c and 6f. b,c, Similar plots of four-cell embryos (b) from the same crosses show a single central mode characteristic of a unimodal population, like the F1 adult spermatocytes (c), which represents a uniform centromere population. The ratio of the modes in bimodal or the value of the mode in unimodal distribution is indicated below the graphs. Source numerical data are available in source data.

Source data

Extended Data Fig. 9 Genetic pathway for centromere equalization.

a, Quantifications of maternal (pink) and paternal (blue) CENP-A and CENP-C intensities in zygotes from a WT x WT control for the Cenpb−/− strain38, with average paternal/maternal CENP-A or CENP-C ratios above; N = 46, 42, 237, 231 centromeres (left to right). Error bars: median ± 95% CI. Although these animals are in a CF-1/C57BL/6J/DBA/2J background, CENP-A and CENP-C ratios in WT zygotes using mothers from this background are consistent with those of C57BL/6J alone (Fig. 6b,f). Source numerical data are available in source data.

Source data

Supplementary information

Reporting Summary

Supplementary Table 1

Replicate information for all relevant figures.

Source data

Source Data Fig. 1

Numerical source data.

Source Data Fig. 2

Numerical source data.

Source Data Fig. 3

Numerical source data.

Source Data Fig. 4

Numerical source data.

Source Data Fig. 5

Numerical source data.

Source Data Fig. 6

Numerical source data.

Source Data Extended Data Fig. 1

Numerical source data.

Source Data Extended Data Fig. 2

Numerical source data.

Source Data Extended Data Fig. 3

Numerical source data.

Source Data Extended Data Fig. 4

Numerical source data.

Source Data Extended Data Fig. 5

Numerical source data.

Source Data Extended Data Fig. 8

Numerical source data.

Source Data Extended Data Fig. 9

Numerical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Das, A., Iwata-Otsubo, A., Destouni, A. et al. Epigenetic, genetic and maternal effects enable stable centromere inheritance. Nat Cell Biol 24, 748–756 (2022). https://doi.org/10.1038/s41556-022-00897-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-022-00897-w

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing