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.

  • Letter
  • Published:

Maternal age generates phenotypic variation in Caenorhabditis elegans

Abstract

Genetically identical individuals that grow in the same environment often show substantial phenotypic variation within populations of organisms as diverse as bacteria1, nematodes2, rodents3 and humans4. With some exceptions5,6,7, the causes are poorly understood. Here we show that isogenic Caenorhabditis elegans nematodes vary in their size at hatching, speed of development, growth rate, starvation resistance, fecundity, and also in the rate of development of their germline relative to that of somatic tissues. We show that the primary cause of this variation is the age of an individual’s mother, with the progeny of young mothers exhibiting several phenotypic impairments. We identify age-dependent changes in the maternal provisioning of the lipoprotein complex vitellogenin to embryos as the molecular mechanism that underlies the variation in multiple traits throughout the life of an animal. The production of sub-optimal progeny by young mothers may reflect a trade-off between the competing fitness traits of a short generation time and the survival and fecundity of the progeny.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Gene expression profiling reveals inter-individual variation in soma–germline phasing.
Figure 2: Slow development, early starvation and young mothers are associated with relative germline acceleration and lower brood size, with early progeny also short and sensitive to starvation.
Figure 3: Vitellogenin expression and provisioning is variable between individuals and increases with age.
Figure 4: Increased yolk provisioning explains phenotypic differences between day 1 and day 2 progeny.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

  1. Spudich, J. L. & Koshland, D. E. Jr. Non-genetic individuality: chance in the single cell. Nature 262, 467–471 (1976)

    Article  ADS  CAS  PubMed  Google Scholar 

  2. Kirkwood, T. B. et al. What accounts for the wide variation in life span of genetically identical organisms reared in a constant environment? Mech. Ageing Dev. 126, 439–443 (2005)

    Article  PubMed  Google Scholar 

  3. Gärtner, K. A third component causing random variability beside environment and genotype. A reason for the limited success of a 30 year long effort to standardize laboratory animals? Lab. Anim. 24, 71–77 (1990)

    Article  PubMed  Google Scholar 

  4. Wong, A. H., Gottesman, I. I. & Petronis, A. Phenotypic differences in genetically identical organisms: the epigenetic perspective. Hum. Mol. Genet. 14, R11–R18 (2005)

    Article  CAS  PubMed  Google Scholar 

  5. Raj, A., Rifkin, S. A., Andersen, E. & Van Oudenaarden, A. Variability in gene expression underlies incomplete penetrance. Nature 463, 913–918 (2010)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  6. Burga, A., Casanueva, M. O. & Lehner, B. Predicting mutation outcome from early stochastic variation in genetic interaction partners. Nature 480, 250–253 (2011)

    Article  ADS  CAS  PubMed  Google Scholar 

  7. Casanueva, M. O., Burga, A. & Lehner, B. Fitness trade-offs and environmentally induced mutation buffering in isogenic C. elegans. Science 335, 82–85 (2012)

    Article  ADS  CAS  PubMed  Google Scholar 

  8. Francesconi, M. & Lehner, B. The effects of genetic variation on gene expression dynamics during development. Nature 505, 208–211 (2014)

    Article  ADS  CAS  PubMed  Google Scholar 

  9. Ambros, V. & Horvitz, H. R. Heterochronic mutants of the nematode Caenorhabditis elegans. Science 226, 409–416 (1984)

    Article  ADS  CAS  PubMed  Google Scholar 

  10. Poullet, N. et al. Complex heterochrony underlies the evolution of Caenorhabditis elegans hermaphrodite sex allocation. Evolution 70, 2357–2369 (2016)

    Article  PubMed  Google Scholar 

  11. Beguet, B. & Brun, J. L. Influence of parental aging on the reproduction of the F1 in a hermaphrodite nematode Caenorhabditis elegans. Exp. Gerontol. 7, 195–206 (1972)

    Article  CAS  PubMed  Google Scholar 

  12. Seidel, H. S. et al. A novel sperm-delivered toxin causes late-stage embryo lethality and transmission ratio distortion in C. elegans. PLoS Biol. 9, e1001115 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Benton, T. G., St Clair, J. J. & Plaistow, S. J. Maternal effects mediated by maternal age: from life histories to population dynamics. J. Anim. Ecol. 77, 1038–1046 (2008)

    Article  CAS  PubMed  Google Scholar 

  14. Coburn, C. & Gems, D. The mysterious case of the C. elegans gut granule: death fluorescence, anthranilic acid and the kynurenine pathway. Front. Genet. 4, 151 (2013)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Fouad, A. D. et al. Quantitative assessment of fat levels in Caenorhabditis elegans using dark field microscopy. G3 (Bethesda) 7, 1811–1818 (2017)

    Article  CAS  Google Scholar 

  16. Klass, M. R. Aging in the nematode Caenorhabditis elegans: major biological and environmental factors influencing life span. Mech. Ageing Dev. 6, 413–429 (1977)

    Article  CAS  PubMed  Google Scholar 

  17. Johnson, T. E., Mitchell, D. H., Kline, S., Kemal, R. & Foy, J. Arresting development arrests aging in the nematode Caenorhabditis elegans. Mech. Ageing Dev. 28, 23–40 (1984)

    Article  CAS  PubMed  Google Scholar 

  18. Sterken, M. G., Snoek, L. B., Kammenga, J. E. & Andersen, E. C. The laboratory domestication of Caenorhabditis elegans. Trends Genet. 31, 224–231 (2015)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Kimble, J. & Sharrock, W. J. Tissue-specific synthesis of yolk proteins in Caenorhabditis elegans. Dev. Biol. 96, 189–196 (1983)

    Article  CAS  PubMed  Google Scholar 

  20. Grant, B. & Hirsh, D. Receptor-mediated endocytosis in the Caenorhabditis elegans oocyte. Mol. Biol. Cell 10, 4311–4326 (1999)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Chotard, L., Skorobogata, O., Sylvain, M. A., Shrivastava, S. & Rocheleau, C. E. TBC-2 is required for embryonic yolk protein storage and larval survival during L1 diapause in Caenorhabditis elegans. PLoS ONE 5, e15662 (2010)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  22. Van Rompay, L., Borghgraef, C., Beets, I., Caers, J. & Temmerman, L. New genetic regulators question relevance of abundant yolk protein production in C. elegans. Sci. Rep. 5, 16381 (2015)

    Article  ADS  PubMed  CAS  Google Scholar 

  23. Sharrock, W. J., Sutherlin, M. E., Leske, K., Cheng, T. K. & Kim, T. Y. Two distinct yolk lipoprotein complexes from Caenorhabditis elegans. J. Biol. Chem. 265, 14422–14431 (1990)

    CAS  PubMed  Google Scholar 

  24. Luo, S., Kleemann, G. A., Ashraf, J. M., Shaw, W. M. & Murphy, C. T. TGF-β and insulin signaling regulate reproductive aging via oocyte and germline quality maintenance. Cell 143, 299–312 (2010)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Rechavi, O. et al. Starvation-induced transgenerational inheritance of small RNAs in C. elegans. Cell 158, 277–287 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Jobson, M. A. et al. Transgenerational effects of early life starvation on growth, reproduction, and stress resistance in Caenorhabditis elegans. Genetics 201, 201–212 (2015)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Harvey, S. C. & Orbidans, H. E. All eggs are not equal: the maternal environment affects progeny reproduction and developmental fate in Caenorhabditis elegans. PLoS ONE 6, e25840 (2011)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Klosin, A., Casas, E., Hidalgo-Carcedo, C., Vavouri, T. & Lehner, B. Transgenerational transmission of environmental information in C. elegans. Science 356, 320–323 (2017)

    Article  ADS  CAS  PubMed  Google Scholar 

  29. Klosin, A. et al. Impaired DNA replication derepresses chromatin and generates a transgenerationally inherited epigenetic memory. Sci. Adv. 3, e1701143 (2017)

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  30. Hodgkin, J. & Barnes, T. M. More is not better: brood size and population growth in a self-fertilizing nematode. Proc. R. Soc. Lond. B 246, 19–24 (1991)

    Article  ADS  CAS  Google Scholar 

  31. Dickinson, D. J., Pani, A. M., Heppert, J. K., Higgins, C. D. & Goldstein, B. Streamlined genome engineering with a self-excising drug selection cassette. Genetics 200, 1035–1049 (2015)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Mok, D. Z. L., Sternberg, P. W. & Inoue, T. Morphologically defined sub-stages of C. elegans vulval development in the fourth larval stage. BMC Dev. Biol. 15, 26 (2015)

    Article  PubMed  PubMed Central  Google Scholar 

  33. Wheeler, M. W., Park, R. M. & Bailer, A. J. Comparing median lethal concentration values using confidence interval overlap or ratio tests. Environ. Toxicol. Chem. 25, 1441–1444 (2006)

    Article  CAS  PubMed  Google Scholar 

  34. Carpenter, A. E. et al. CellProfiler: image analysis software for identifying and quantifying cell phenotypes. Genome Biol. 7, R100 (2006)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Wählby, C. et al. An image analysis toolbox for high-throughput C. elegans assays. Nat. Methods 9, 714–716 (2012)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Klapper, M. et al. Fluorescence-based fixative and vital staining of lipid droplets in Caenorhabditis elegans reveal fat stores using microscopy and flow cytometry approaches. J. Lipid Res. 52, 1281–1293 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. DePina, A. S. et al. Regulation of Caenorhabditis elegans vitellogenesis by DAF-2/IIS through separable transcriptional and posttranscriptional mechanisms. BMC Physiol. 11, 11 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Artyukhin, A. B., Schroeder, F. C. & Avery, L. Density dependence in Caenorhabditis larval starvation. Sci. Rep. 3, 2777 (2013)

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  39. Lee, I., Hendrix, A., Kim, J., Yoshimoto, J. & You, Y. J. Metabolic rate regulates L1 longevity in C. elegans. PLoS ONE 7, e44720 (2012); erratum 8, https://doi.org/10.1371/annotation/c69de5f4-dd02-4f92-9fc7-9a6a660a075e (2013)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  40. Lionaki, E. & Tavernarakis, N. Assessing aging and senescent decline in Caenorhabditis elegans: cohort survival analysis. Methods Mol. Biol. 965, 473–484 (2013)

    Article  CAS  PubMed  Google Scholar 

  41. Xiao, R. et al. RNAi interrogation of dietary modulation of development, metabolism, behavior, and aging in C. elegans. Cell Reports 11, 1123–1133 (2015)

    Article  CAS  PubMed  Google Scholar 

  42. Kamath, R. S. & Ahringer, J. Genome-wide RNAi screening in Caenorhabditis elegans. Methods 30, 313–321 (2003)

    Article  CAS  PubMed  Google Scholar 

  43. Rea, S. L., Ventura, N. & Johnson, T. E. Relationship between mitochondrial electron transport chain dysfunction, development, and life extension in Caenorhabditis elegans. PLoS Biol. 5, e259 (2007)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Vandesompele, J. et al. Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal control genes. Genome Biol. 3, research0034.1 (2002)

    Article  Google Scholar 

  45. Gubelmann, C. et al. GETPrime: a gene- or transcript-specific primer database for quantitative real-time PCR. Database (Oxford) 2011, bar040 (2011)

    Article  CAS  Google Scholar 

  46. Reinke, V., Gil, I. S., Ward, S. & Kazmer, K. Genome-wide germline-enriched and sex-biased expression profiles in Caenorhabditis elegans. Development 131, 311–323 (2004)

    Article  CAS  PubMed  Google Scholar 

  47. Rockman, M. V., Skrovanek, S. S. & Kruglyak, L. Selection at linked sites shapes heritable phenotypic variation in C. elegans. Science 330, 372–376 (2010)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  48. Spencer, W. C. et al. A spatial and temporal map of C. elegans gene expression. Genome Res. 21, 325–341 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Chikina, M. D., Huttenhower, C., Murphy, C. T. & Troyanskaya, O. G. Global prediction of tissue-specific gene expression and context-dependent gene networks in Caenorhabditis elegans. PLOS Comput. Biol. 5, e1000417 (2009)

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  50. Hastie, T. & Stuetzle, W. Principal curves. J. Am. Stat. Assoc. 84, 502–516 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  51. Iwasaki, K., McCarter, J., Francis, R. & Schedl, T. emo-1, a Caenorhabditis elegans Sec61p gamma homologue, is required for oocyte development and ovulation. J. Cell Biol. 134, 699–714 (1996)

    Article  CAS  PubMed  Google Scholar 

  52. Kennedy, B. P. et al. The gut esterase gene (ges-1) from the nematodes Caenorhabditis elegans and Caenorhabditis briggsae. J. Mol. Biol. 229, 890–908 (1993)

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by a European Research Council Consolidator grant (616434), the Spanish Ministry of Economy and Competitiveness (BFU2011-26206 and SEV-2012-0208), the AXA Research Fund, the Bettencourt Schueller Foundation, Agència de Gestió d’Ajuts Universitaris i de Recerca (AGAUR, 2014 SGR 831), the EMBL-CRG Systems Biology Program, and the CERCA Program/Generalitat de Catalunya. M.F.P. was partially supported by a FPI-Severo Ochoa fellowship. Expression profiling was performed in the CRG Genomics core facility and microscopy in the CRG Advanced Light Microscopy Facility. Some strains were provided by the CGC, which is funded by National Institutes of Health Office of Research Infrastructure Programs (P40 OD010440).

Author information

Authors and Affiliations

Authors

Contributions

M.F.P., M.F. and B.L. conceived the model, designed experiments and wrote the manuscript. M.F.P., M.F. and C.H.-C. performed experiments and analysed the data.

Corresponding author

Correspondence to Ben Lehner.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Additional information

Reviewer Information Nature thanks D. Gems, O. Rechavi and the other anonymous reviewer(s) for their contribution to the peer review of this work.

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

Extended data figures and tables

Extended Data Figure 1 Spermatogenesis and oogenesis gene expression signatures extracted by independent component analysis.

a, Scatterplot showing gene loadings on the oogenesis component (x axis) and spermatogenesis component (y axis) extracted by independent component analysis. Spermatogenesis and oogenesis46 genes are highlighted. b, The oogenesis and spermatogenesis components in the reference time series46 ranked with principal curves. The three replicates per time point in the time series are considered separately. c, Early- and late-moulting signatures in the reference time series ranked according to germline rank. d, Enrichment (measured by odds) of oogenesis-46, spermatogenesis-46 and hypodermis-specific48 genes in the top 10% most variable genes among single worms judged by expression residuals after correcting for germline developmental stage. e, Enrichment (measured by log odds) for genes for which variance is better explained by germline (positive log odds) or hypodermal (negative log odds) rank in tissue-specific gene sets. Variation in other somatic tissues is best explained by hypodermal gene expression signatures. Error bars represent 95% confidence intervals. Gene sets are listed in Methods.

Extended Data Figure 2 Slow development, early starvation and young mothers are associated with altered temporal distribution of progeny production, with early progeny displaying enlarged gut granules, slow development and reduced resistance to larval starvation.

a, Schematic indicating the establishment of a synchronized population by allowing eggs to hatch for 1 h and transferring newly hatched larvae to a new plate. b, Progeny production per day for early-, mid- or late-moulting worms (n = 9, 11, 10 worms). c, Schematic indicating comparison of worms continuously fed after hatching with worms that underwent a transient 5-h starvation immediately after hatching. d, Progeny production per day for worms either continuously fed or transiently starved for 5 h immediately after hatching (n = 12, 13 worms). e, Schematic showing establishment of parallel cohorts of progeny from day 1, day 2 or day 3 mothers and subsequent phenotypic assays. f, Progeny production per day of offspring of day 1, day 2 or day 3 mothers. The start of the first day of egg laying for each cohort is counted from the point where a majority of the population have at least 1 embryo in utero (n = 18, 18, 19 worms). g, Confocal microscopy images of live L1 larvae from day 1, 2 or 3 mothers showing blue autofluorescence from intestinal lysosome-related organelles (‘gut granules’). Insets show expanded section of main image indicated by white box (shown in Fig. 2m). All images are maximum projections. Scale bars (main image and inset), 5 μm. h, i, Gut granule mean diameter (h) and count (i) in individual newly hatched L1 larvae from day 1, 2 or 3 mothers (n = 3, 3, 3 larvae). j, Intensity of blue autofluorescence from gut granules in segments along the anterior–posterior axis of digitally straightened newly hatched L1 larvae from day 1, 2 or 3 mothers (n = 32, 29, 16 larvae). Representative straightened brightfield and fluorescence images are positioned correspondingly along the x axis. Scale bar, 20 μm. k, Worm developmental stage determined by scoring vulval morphology of L4 larvae from day 1, day 2 or day 3 mothers after 42 h of feeding (n = 70, 61, 49 larvae). l, Length of larvae 40 h after recovery from a 7-day L1 starvation from day 1, day 2 or day 3 mothers (n = 26, 33, 26 larvae). m, Brightfield and epifluorescence images showing sterile progeny from day 1 mothers recovered for 96 h from a 14-day L1 starvation, or a worm that was continuously fed, with nuclei stained by DAPI. Gonads of worms recovered from starvation often exhibit severe abnormalities, such as failure to migrate (*), hyper-proliferation (+) and endomitotic oocytes51 (arrowhead). DAPI images are max intensity projection of Z-stacks following rolling ball background subtraction. Scale bars, 100 μm. n, Survival curves of progeny of day 1 or day 2 mothers upon recovery from L1 starvation for 12 h (day 1, n = 39, 66 worms, day 2, n = 52, 56, 55 worms), 1 week (day 1, n = 35, 40, 45 worms, day 2, n = 39, 47, 52 worms) or 2 weeks (day 1, n = 56, 65 worms; day 2, n = 51, 61, 68 worms). Each line represents a biological replicate. Boxplots represent median values, interquartile ranges and Tukey whiskers with individual data points superimposed. Error bars show s.e.m. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001. Statistical tests used were Kruskal–Wallis test with Dunn’s multiple comparisons tests (k, l), generalized linear mixed model analysis (h, i), two-way ANOVA with Bonferroni multiple comparisons post-tests (j), log-rank test (n). Results were replicated at least 2 times independently. NS, not significant; YA, young adult.

Source data

Extended Data Figure 3 Maternal age affects progeny phenotypes in C. elegans wild isolates.

a, b, Boxplot showing length at hatching of L1 larvae from day 1 (grey), day 2 (blue) or day 3 (green) mothers of CB4856 (a, n = 51, 87, 53 larvae) and PB306 (b, n = 134, 237, 120 larvae). c, d, Boxplot showing length of larvae after 42 h of feeding from day 1, day 2 or day 3 mothers of CB4856 (c, n = 82, 66, 106 larvae) and PB306 (d, n = 168, 150, 93 larvae). e, f, Bar chart indicating worm developmental stage determined by scoring vulval morphology of L4 larvae after 42 h of feeding from day 1, day 2 or day 3 mothers of CB4856 (e, n = 81, 49, 101 larvae) and PB306 (f, n = 75, 85, 72 larvae). g, h, Boxplot showing length of larvae 48 h after recovery from a 7-day L1 starvation from day 1, day 2 or day 3 mothers of CB4856 (g, n = 174, 132, 185 larvae) and PB306 (h, n = 71, 62, 26 larvae). i, j, Confocal microscopy images of live L1 larvae from day 1, 2 or 3 mothers of CB4856 (i) and PB306 (j) showing blue autofluorescence from intestinal lysosome-related organelles (‘gut granules’). Insets show expanded section of main image indicated by white box. All images are maximum projections. Scale bars (main image and inset), 5 μm. Boxplots represent median values, interquartile ranges and Tukey whiskers with individual data points superimposed. *P < 0.05, **P < 0.01, ***P < 0.001. Statistical tests used were the Kruskal–Wallis test with Dunn’s multiple comparisons tests (ah). Results were replicated at least twice independently. NS, not significant; YA, young adult.

Source data

Extended Data Figure 4 Vitellogenin expression and loading into embryos increase with age.

a, Representative images of early embryos from day 1, day 2 and day 3 BCN9070 (vit-2::gfp) adults and wild-type N2 embryo for autofluorescence. Brightfield and epifluorescence images are shown. Scale bars, 20 μm. b, Average fluorescence intensity per embryo from immunostaining with an antibody against YP170, the protein encoded by vit-1 to vit-5, in 2–75-cell wild-type N2 embryos from day 1 or day 3 mothers. Duplicate stainings performed for each age group are shown (n = 41, 76; 58, 48 embryos). Fluorescence values are normalized to the mean of day 1 replicate A. c, Representative images of embryos extruded from day 1 or day 3 gravid hermaphrodites and stained with anti-YP170 antibody (green) and DAPI (blue). Scale bars, 20 μm. d, Total green autofluorescence of early embryos from wild-type N2 day 2 mothers treated with empty vector, RNAi against the constitutive intestinal gene ges-1 (ref. 52) as an additional control, combined vit-5 and vit-6 RNAi, and rme-2 RNAi (n = 58, 40, 22, 42 embryos). Fluorescence intensities are normalized to the mean of the empty vector control group. Representative images are shown above the corresponding group. e, Total green autofluorescence of early embryos from wild-type N2 day 1, day 2 or day 3 mothers (n = 92, 125, 70 embryos). Data shown are pooled from two trials analysed together. Fluorescence intensities are normalized to the mean of the day 1 group within each trial. Representative images from a single trial are shown above the corresponding group. f, mRNA transcripts quantified by qPCR of the yolk receptor rme-2 in day 1, day 2 and day 3 adults. Transcript levels were normalized to the geometric mean of two housekeeping genes (cdc-42 and rpl-27). Data shown are from three biological replicates. g, Representative images of day 1, day 2 and day 3 BCN9070 adults carrying an in-frame gfp knock-in at the C terminus of vit-2 and wild-type N2 worm for autofluorescence. Brightfield and epifluorescence images are shown. Worms have been digitally straightened. Scale bars, 100 μm. h, Representative lysates of day 1, 2 or 3 adult hermaphrodites run on an SDS–PAGE denaturing protein gel and stained with SYPRO Ruby fluorescent protein stain. Additionally, to confirm the identity of peaks corresponding to yolk proteins, samples are shown from hermaphrodites treated with empty vector control RNAi, rme-2 (yolk receptor) RNAi to induce yolk accumulation, and vit-5 and vit-6 RNAi to deplete yolk. All samples shown were run on the same gel. Myosin, YP170 (encoded by vit-1 to vit-5) and YP115 and YP88 (both encoded by vit-6) are indicated by arrows. For gel source data, see Supplementary Fig. 1. i, Estimate of the intestinal volume of adult hermaphrodites (n = 7, 9, 9 worms). Bar shows mean ± s.e.m. Boxplots represent median values, interquartile ranges and Tukey whiskers with individual data points superimposed. Error bars in barcharts and dot plots show s.e.m. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001, one-way ANOVA with Tukey’s multiple comparison tests (f, i). Statistical tests used were Kruskal–Wallis test with Dunn’s multiple comparisons test (d), generalized linear model analysis (b, e). NS, not significant; Ab, antibody.

Source data

Extended Data Figure 5 Maternal vitellogenin expression and embryonic lipid content increase with maternal age in C. elegans wild isolates.

a, b, Bar chart of total fluorescence of fixed embryos from day 1, 2 or 3 mothers of CB4856 (a, 719, 1,315, 249 embryos) and PB306 (b, n = 1,090, 839, 330 embryos) stained with BODIPY 493/503, a neutral lipid stain. Data shown are pooled from 4–5 biological replicates. c, d, Bar chart of mRNA transcripts quantified by qPCR of vit-2 and vit-6 in day 1, day 2 and day 3 adults of CB4856 (c) and PB306 (d). Transcript levels were normalized to the geometric mean of three housekeeping genes (cdc-42, rpl-27 and Y45F10D.4). Data shown are from three biological replicates. Error bars in bar charts show s.e.m. *P < 0.05, **P < 0.01, ***P < 0.001. Statistical tests used are generalized linear mixed model analysis (a, b), one-way ANOVA with Tukey’s multiple comparison tests (panel c, d). NS, not significant.

Source data

Extended Data Figure 6 Progeny depleted of embryonic yolk by maternal vitellogenin knockdown recapitulate phenotypes observed in progeny of young mothers.

a, Schematic showing experimental method used to generate yolk-depleted progeny by maternal vit-5 and vit-6 RNAi treatment. b, mRNA transcripts quantified by qPCR of the 6 vit genes in day 2 adults treated with or empty vector control (blue), vit-5 RNAi (magenta), vit-6 RNAi (yellow) or a mixture of vit-5 and vit-6 RNAi (turquoise). Transcript levels were normalized to the geometric mean of three housekeeping genes (cdc-42, rpl-27 and Y45F10D.4). Data shown are from three biological replicates. c, The constitutive intestinal gene ges-1 (ref. 52) was used as an additional control. mRNA transcripts quantified by qPCR for ges-1, vit-2 and vit-6 in day 2 adults treated with empty vector control (blue) or ges-1 RNAi (brown). Transcript levels were normalized to the geometric mean of three housekeeping genes (cdc-42, rpl-27 and Y45F10D.4). Data shown are from three biological replicates. d, Length at hatching of L1 larvae from day 2 mothers treated with empty vector control (blue), ges-1 RNAi (brown) or vit-5 and vit-6 RNAi (turquoise) (n = 103, 130, 159 larvae). e, Confocal microscopy images of live L1 larvae from day 2 mothers treated with vit-5 and vit-6 RNAi, ges-1 RNAi or empty vector control showing blue autofluorescence from intestinal lysosome-related organelles (‘gut granules’). Insets show expanded section of main image indicated by white box. All images are maximum projections. Scale bars (main image and inset), 5 μm. f, Worm developmental stage determined by scoring vulval morphology of L4 larvae after 46 h of feeding from day 2 mothers treated with empty vector control (blue), ges-1 RNAi (brown) or vit-5 and vit-6 RNAi (turquoise) (n = 48, 36, 47 larvae). g, Time after hatching by which half the progeny of mothers treated with empty vector control, ges-1 RNAi or vit-5 and vit-6 RNAi had undergone the L4-to-adult moult (blue) or carried a fertilized embryo in utero (red). Each time point for each transition is calculated by sampling the population twice. (Sample sizes: empty vector, n = 199 and 198 for the L4-to-adult moult, n = 171 and 168 for the first embryo; ges-1, n = 204 and 206 for the L4-to-adult moult, n = 198 and 168 for the first embryo; vit-5 and vit-6, n = 201 and 202 for the L4-to-adult moult, n = 179 and 200 for the first embryo). h, Time between the L4-to-adult moult and the first embryo for the data shown in g. i, Length of larvae from day 2 mothers treated with empty vector control (blue), ges-1 RNAi (brown) or vit-5 and vit-6 RNAi (turquoise) after 46 h of feeding (n = 67, 56, 57 larvae). j, Length of larvae at early L4 vulval substages in progeny of day 2 mothers treated with empty vector control (blue), ges-1 RNAi (brown) or vit-5 and vit-6 RNAi (turquoise) (L4.1, n = 15, 15, 9 larvae; L4.2, n = 7, 8, 5 larvae, L4.3, n = 18, 22, 13 larvae). k, Total brood size of progeny of mothers treated with empty vector control (blue), ges-1 RNAi (brown) or vit-5 and vit-6 RNAi (turquoise) (n = 13, 14, 11 worms). l, Length 44 h after recovery from a 7-day L1 starvation of larvae from day 2 mothers treated with empty vector control (blue), ges-1 RNAi (brown) or vit-5 and vit-6 RNAi (turquoise) (n = 30, 33, 85 larvae). m, Percentage of adult progeny of day 2 mothers treated with empty vector control (blue), ges-1 RNAi (brown) or vit-5 and vit-6 RNAi (turquoise) displaying normal phenotypic outcomes 96 h after recovery from a 10-day L1 starvation (n = 610, 567, 1,135 worms). n, Percentage of adult progeny of day 2 mothers treated with empty vector control (blue), ges-1 RNAi (brown) or vit-5 and vit-6 RNAi (turquoise) that successfully produced offspring 168 h after recovery from a 11-day L1 starvation (n = 103, 116, 124 worms). Boxplots represent median values, interquartile ranges and Tukey whiskers with individual data points superimposed. Error bars in b and c represent s.e.m.; error bars in g, h, j, m and n represent 95% confidence intervals. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001. Statistical tests used were two-tailed t-tests (c), Kruskal–Wallis test with Dunn’s multiple comparisons tests (d, f, i, l), ratio tests (g, h), one-way ANOVA with Tukey’s multiple comparisons tests (k), two-way ANOVA with Bonferroni multiple comparisons tests (j), pairwise Fisher’s exact test with Bonferroni multiple corrections (m, n). Data shown are not all from the same trial but represent typical outcomes. All results were replicated at least twice. NS, not significant.

Source data

Extended Data Figure 7 Increased yolk provisioning accounts for the differences in development and growth between day 1 and day 2 progeny.

a, Worm developmental stage determined by scoring vulval morphology of L4 larvae after 42 h of feeding from day 1 mothers (grey), day 2 mothers treated with empty vector control (blue) and day 2 mothers treated with 10% rme-2 RNAi (orange) (n = 69, 66, 52 embryo). b, Length 48 h after recovery from a 7-day L1 starvation of larvae from day 1 mothers (grey), day 2 mothers treated with empty vector control (blue) and day 2 mothers treated with 10% rme-2 RNAi (orange) (n = 295, 206, 293 larvae). Boxplots represent median values, interquartile ranges and Tukey whiskers with individual data points superimposed. **P < 0.01, ***P < 0.001, ****P < 0.0001. A Kruskal–Wallis test with Dunn’s multiple comparisons tests was used. Data shown (here and in Fig. 4) are from a single trial. Results were replicated twice independently.

Source data

Extended Data Table 1 Sequences of qPCR primers used in this study

Related audio

Supplementary information

Life Sciences Reporting Summary (PDF 72 kb)

Supplementary Figure 1

This file contains he uncropped source data for the gel shown in Extended Data Figure 4h. (PDF 506 kb)

Supplementary Tables 1 and 2

This file contains table 1 (supplementary to Fig. 1). Table including projection of single worm expression profiles onto the germline and somatic expression spaces and their germline and soma ranking. It also contains table 2 (supplementary to Extended Data Fig. 1). Table including residuals of single worm gene expression after correcting for germline and soma developmental stage. The table also indicates for each gene whether germline or somatic ranking better explains the observed variation and includes gene annotation for the tissue-specific gene sets. Both tables are combined as separate sheets in a single .xlsx file. (XLSX 16107 kb)

PowerPoint slides

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Perez, M., Francesconi, M., Hidalgo-Carcedo, C. et al. Maternal age generates phenotypic variation in Caenorhabditis elegans. Nature 552, 106–109 (2017). https://doi.org/10.1038/nature25012

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature25012

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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