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:

Real-time conversion of tissue-scale mechanical forces into an interdigitated growth pattern

An Author Correction to this article was published on 23 June 2021

This article has been updated

Abstract

The leaf epidermis is a dynamic biomechanical shell that integrates growth across spatial scales to influence organ morphology. Pavement cells, the fundamental unit of this tissue, morph irreversibly into highly lobed cells that drive planar leaf expansion. Here, we define how tissue-scale cell wall tensile forces and the microtubule–cellulose synthase systems dictate the patterns of interdigitated growth in real time. A morphologically potent subset of cortical microtubules span the periclinal and anticlinal cell faces to pattern cellulose fibres that generate a patch of anisotropic wall. The subsequent local polarized growth is mechanically coupled to the adjacent cell via a pectin-rich middle lamella, and this drives lobe formation. Finite element pavement cell models revealed cell wall tensile stress as an upstream patterning element that links cell- and tissue-scale biomechanical parameters to interdigitated growth. Cell lobing in leaves is evolutionarily conserved, occurs in multiple cell types and is associated with important agronomic traits. Our general mechanistic models of lobe formation provide a foundation to analyse the cellular basis of leaf morphology and function.

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: Heterogeneous growth of the cell boundary in developing pavement cells.
Fig. 2: Tensile force orientation correlates with subcellular growth rates.
Fig. 3: Transfacial microtubules predict lobe initiation sites.
Fig. 4: Microtubule signals are concentrated at regions of elevated cell wall tensile stress.
Fig. 5: A system for inducible cell-autonomous lobing.
Fig. 6: Cell wall stress dictates the location and direction of furrow formation.
Fig. 7: A general model for lobe morphogenesis and interdigitated growth.

Similar content being viewed by others

Data availability

The Fiji macros, Python scripts and PDLP growth rate code in R used for all image analyses are available at https://github.com/yamsissamy/naturePlant_lobeInitiation. Abaqus input files for the FE models shown in the figures are available at http://tulips.unl.edu.

Change history

References

  1. Chitwood, D. H. & Sinha, N. R. Evolutionary and environmental forces sculpting leaf development. Curr. Biol. 26, R297–R306 (2016).

    Article  CAS  PubMed  Google Scholar 

  2. Swarup, R. et al. Localization of the auxin permease AUX1 suggests two functionally distinct hormone transport pathways operate in the Arabidopsis root apex. Genes Dev. 15, 2648–2653 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Savaldi-Goldstein, S., Peto, C. & Chory, J. The epidermis both drives and restricts plant shoot growth. Nature 446, 199–202 (2007).

    Article  CAS  PubMed  Google Scholar 

  4. Vaseva, I. I. et al. The plant hormone ethylene restricts Arabidopsis growth via the epidermis. Proc. Natl Acad. Sci. USA 115, E4130–E4139 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Andriankaja, M. et al. Exit from proliferation during leaf development in Arabidopsis thaliana: a not-so-gradual process. Dev. Cell 22, 64–78 (2012).

    Article  CAS  PubMed  Google Scholar 

  6. Das Gupta, M. & Nath, U. Divergence in patterns of leaf growth polarity is associated with the expression divergence of miR396. Plant Cell 27, 2785–2799 (2015).

    PubMed  PubMed Central  Google Scholar 

  7. Vofely, R. V., Gallagher, J., Pisano, G. D., Bartlett, M. & Braybrook, S. A. Of puzzles and pavements: a quantitative exploration of leaf epidermal cell shape. New Phytol. 221, 540–552 (2019).

    Article  CAS  PubMed  Google Scholar 

  8. Haberlandt, G. Physiological Plant Anatomy 4th edn (MacMillan, 1914).

  9. Panteris, E. & Galatis, B. The morphogenesis of lobed plant cells in the mesophyll and epidermis: organization and distinct roles of cortical microtubules and actin filaments. New Phytol. 167, 721–732 (2005).

    Article  CAS  PubMed  Google Scholar 

  10. Fu, Y., Gu, Y., Zheng, Z., Wasteneys, G. & Yang, Z. Arabidopsis interdigitating cell growth requires two antagonistic pathways with opposing action on cell morphogenesis. Cell 120, 687–700 (2005).

    Article  CAS  PubMed  Google Scholar 

  11. Szymanski, D. B. The kinematics and mechanics of leaf expansion: new pieces to the Arabidopsis puzzle. Curr. Opin. Plant Biol. 22C, 141–148 (2014).

    Article  Google Scholar 

  12. Gao, Y. et al. Auxin binding protein 1 (ABP1) is not required for either auxin signaling or Arabidopsis development. Proc. Natl Acad. Sci. USA 112, 2275–2280 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Belteton, S. A., Sawchuk, M. G., Donohoe, B. S., Scarpella, E. & Szymanski, D. B. Reassessing the roles of PIN proteins and anticlinal microtubules during pavement cell morphogenesis. Plant Physiol 176, 432–449 (2018).

    Article  CAS  PubMed  Google Scholar 

  14. Xu, T. et al. Cell surface- and Rho GTPase-based auxin signaling controls cellular interdigitation in Arabidopsis. Cell 143, 99–110 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Armour, W. J., Barton, D. A., Law, A. M. & Overall, R. L. Differential growth in periclinal and anticlinal walls during lobe formation in Arabidopsis cotyledon pavement cells. Plant Cell 27, 2484–2500 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Sapala, A. et al. Why plants make puzzle cells, and how their shape emerges. eLife https://doi.org/10.7554/eLife.32794 (2018).

  17. Altartouri, B. et al. Pectin chemistry and cellulose crystallinity govern pavement cell morphogenesis in a multi-step mechanism. Plant Physiol. 181, 127–141 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Haas, K. T., Wightman, R., Meyerowitz, E. M. & Peaucelle, A. Pectin homogalacturonan nanofilament expansion drives morphogenesis in plant epidermal cells. Science 367, 1003–1007 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Elsner, J., Lipowczan, M. & Kwiatkowska, D. Differential growth of pavement cells of Arabidopsis thaliana leaf epidermis as revealed by microbead labeling. Am. J. Bot. 105, 257–265 (2018).

    Article  CAS  PubMed  Google Scholar 

  20. Sampathkumar, A. et al. Subcellular and supracellular mechanical stress prescribes cytoskeleton behavior in Arabidopsis cotyledon pavement cells. eLife 3, e01967 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Zhang, C., Halsey, L. E. & Szymanski, D. B. The development and geometry of shape change in Arabidopsis thaliana cotyledon pavement cells. BMC Plant Biol. 11, 27 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  22. Elsner, J., Michalski, M. & Kwiatkowska, D. Spatiotemporal variation of leaf epidermal cell growth: a quantitative analysis of Arabidopsis thaliana wild-type and triple cyclinD3 mutant plants. Ann. Bot. 109, 897–910 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Paredez, A. R., Somerville, C. R. & Ehrhardt, D. W. Visualization of cellulose synthase demonstrates functional association with microtubules. Science 312, 1491–1495 (2006).

    Article  CAS  PubMed  Google Scholar 

  24. Baskin, T. I. Anisotropic expansion of the plant cell wall. Annu. Rev. Cell Dev. Biol. 21, 203–222 (2005).

    Article  CAS  PubMed  Google Scholar 

  25. Green, P. B. The spiral growth pattern of the cell wall in Nitella axillaris. Am. J. Bot. 41, 403–409 (1954).

    Article  Google Scholar 

  26. Yanagisawa, M. et al. Patterning mechanisms of cytoskeletal and cell wall systems during leaf trichome morphogenesis. Nat. Plants 1, 15014 (2015).

    Article  CAS  PubMed  Google Scholar 

  27. Wong, J. H. et al. Basic proline-rich protein-mediated microtubules are essential for lobe growth and flattened cell geometry. Plant Physiol. 181, 1535–1551 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Majda, M. et al. Mechanochemical polarization of contiguous cell walls shapes plant pavement cells. Dev. Cell 43, 290–304 (2017).

    Article  CAS  PubMed  Google Scholar 

  29. Bidhendi, A. J., Altartouri, B., Gosselin, F. P. & Geitmann, A. Mechanical stress initiates and sustains the morphogenesis of wavy leaf epidermal cells. Cell Rep. 28, 1237–1250 (2019).

    Article  CAS  PubMed  Google Scholar 

  30. Fayant, P. et al. Finite element model of polar growth in pollen tubes. Plant Cell 22, 2579–2593 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Fujita, M. et al. The anisotropy1 D604N mutation in the Arabidopsis cellulose synthase1 catalytic domain reduces cell wall crystallinity and the velocity of cellulose synthase complexes. Plant Physiol. 162, 74–85 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Uyttewaal, M. et al. Mechanical stress acts via katanin to amplify differences in growth rate between adjacent cells in Arabidopsis. Cell 149, 439–451 (2012).

    Article  CAS  PubMed  Google Scholar 

  33. Ambrose, J. C., Shoji, T., Kotzer, A. M., Pighin, J. A. & Wasteneys, G. O. The Arabidopsis CLASP gene encodes a microtubule-associated protein involved in cell expansion and division. Plant Cell 19, 2763–2775 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Kotzer, A. M. & Wasteneys, G. O. Mechanisms behind the puzzle: microtubule–microfilament cross-talk in pavement cell formation. Can. J. Bot. 84, 594–603 (2006).

    Article  CAS  Google Scholar 

  35. Whittington, A. T. et al. MOR1 is essential for organizing cortical microtubules in plants. Nature 411, 610–613 (2001).

    Article  CAS  PubMed  Google Scholar 

  36. Mouille, G. et al. Homogalacturonan synthesis in Arabidopsis thaliana requires a Golgi-localized protein with a putative methyltransferase domain. Plant J. 50, 605–614 (2007).

    Article  CAS  PubMed  Google Scholar 

  37. Wu, T.-C., Belteton, S. A., Pack, J., Szymanski, D. B. & Umulis, D. M. LobeFinder: a convex-hull based method for quantitative boundary analysis of lobed plant cells. Plant Physiol. 171, 2331–2342 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Paredez, A. R., Persson, S., Ehrhardt, D. W. & Somerville, C. Genetic evidence that cellulose synthase activity influences microtubule cortical array organization. Plant Physiol. 147, 1723–1734 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Li, S., Lei, L., Somerville, C. R. & Gu, Y. Cellulose synthase interactive protein 1 (CSI1) links microtubules and cellulose synthase complexes. Proc. Natl Acad. Sci. USA 109, 185–190 (2012).

    Article  CAS  PubMed  Google Scholar 

  40. Staff, L. et al. The hidden geometries of the Arabidopsis thaliana epidermis. PLoS ONE 7, e43546 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. El-Assal, S. E., Le, J., Basu, D., Mallery, E. L. & Szymanski, D. B. DISTORTED2 encodes an ARPC2 subunit of the putative Arabidopsis ARP2/3 complex. Plant J. 38, 526–538 (2004).

    Article  CAS  Google Scholar 

  42. Harris, W. M. Ultrastructural observations on the mesophyll cells of pine leaves. Can. J. Bot. 49, 1107–1109 (1971).

    Article  Google Scholar 

  43. Verger, S., Long, Y., Boudaoud, A. & Hamant, O. A tension–adhesion feedback loop in plant epidermis. eLife https://doi.org/10.7554/eLife.34460 (2018)

  44. Apostolakos, P., Galatis, B. & Panteris, E. Microtubules in cell morphogenesis and intercellular space formation in Zea mays leaf mesophyll and Pilea cadierei epithem. J. Plant Physiol. 137, 591–601 (1991).

    Article  Google Scholar 

  45. Panteris, E., Apostolakos, P. & Galatis, B. Microtubule organization and cell morphogenesis in two semi-lobed cell types of Adiantum capillus-veneris L. leaflets. New Phytol. 125, 509–520 (1993).

    Article  CAS  PubMed  Google Scholar 

  46. Panteris, E., Apostolakos, P. & Galatis, B. Sinuous ordinary epidermal cells: behind several patterns of waviness, a common morphogenetic mechanism. New Phytol. 127, 771–780 (1994).

    Article  CAS  PubMed  Google Scholar 

  47. Fu, Y., Li, H. & Yang, Z. The ROP2 GTPase controls the formation of cortical fine F-actin and the early phase of directional cell expansion during Arabidopsis organogenesis. Plant Cell 14, 777–794 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Fu, Y., Gu, Y., Zheng, Z., Wasteneys, G. & Yang, Z. Arabidopsis interdigitating cell growth requires two antagonistic pathways with opposing action on cell morphogenesis. Cell 120, 687–700 (2005).

    Article  CAS  PubMed  Google Scholar 

  49. Kutschera, U., Bergfeld, R. & Schopfer, P. Cooperation of epidermis and inner tissues in auxin-mediated growth of maize coeloptiles. Planta 170, 168–180 (1987).

    Article  CAS  PubMed  Google Scholar 

  50. Chan, J. & Coen, E. Interaction between autonomous and microtubule guidance systems controls cellulose synthase trajectories. Curr. Biol. 30, 941–947 (2020).

    Article  CAS  PubMed  Google Scholar 

  51. Elliott, A. & Shaw, S. L. Microtubule array patterns have a common underlying architecture in hypocotyl cells. Plant Physiol. 176, 307–325 (2018).

    Article  CAS  PubMed  Google Scholar 

  52. Xu, T. et al. Cell surface ABP1–TMK auxin-sensing complex activates ROP GTPase signaling. Science 343, 1025–1028 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Zhao, F. et al. Microtubule-mediated wall anisotropy contributes to leaf blade flattening. Curr. Biol. 30, 3972–3985 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Ren, T., Weraduwage, S. M. & Sharkey, T. D. Prospects for enhancing leaf photosynthetic capacity by manipulating mesophyll cell morphology. J. Exp. Bot. 70, 1153–1165 (2019).

    Article  CAS  PubMed  Google Scholar 

  55. He, W., Adachi, S., Sage, R. F., Ookawa, T. & Hirasawa, T. Leaf photosynthetic rate and mesophyll cell anatomy changes during ontogenesis in backcrossed indica × japonica rice inbred lines. Photosynth. Res. 134, 27–38 (2017).

    Article  CAS  PubMed  Google Scholar 

  56. Higaki, T. et al. A theoretical model of jigsaw-puzzle pattern formation by plant leaf epidermal cells. PLoS Comput. Biol. 12, e1004833 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  57. Jarvis, M. C., Briggs, S. P. H. & Knox, J. P. Intercellular adhesion and cell separation in plants. Plant Cell Environ. 26, 977–989 (2003).

    Article  Google Scholar 

  58. Rolland-Lagan, A.-G., Bangham, J. A. & Coen, E. Growth dynamics underlying petal shape and asymmetry. Nature 422, 161–163 (2003).

    Article  CAS  PubMed  Google Scholar 

  59. Kuchen, E. E. et al. Generation of leaf shape through early patterns of growth and tissue polarity. Science 335, 1092–1096 (2012).

    Article  CAS  PubMed  Google Scholar 

  60. Hervieux, N. et al. A mechanical feedback restricts sepal growth and shape in Arabidopsis. Curr. Biol. https://doi.org/10.1016/j.cub.2016.03.004 (2016).

  61. Kierzkowski, D. et al. A growth-based framework for leaf shape development and diversity. Cell 177, 1405–1418 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Fendrych, M. et al. The Arabidopsis exocyst complex is involved in cytokinesis and cell plate maturation. Plant Cell 22, 3053–3065 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Ambrose, C., Allard, J. F., Cytrynbaum, E. N. & Wasteneys, G. O. A CLASP-modulated cell edge barrier mechanism drives cell-wide cortical microtubule organization in Arabidopsis. Nat. Commun. 2, 430 (2011).

    Article  PubMed  Google Scholar 

  64. Nakamura, M., Ehrhardt, D. W. & Hashimoto, T. Microtubule and katanin-dependent dynamics of microtubule nucleation complexes in the acentrosomal Arabidopsis cortical array. Nat. Cell Biol. 12, 1064–1070 (2010).

    Article  CAS  PubMed  Google Scholar 

  65. Oppenheimer, D. G. et al. Essential role of a kinesin-like protein in Arabidopsis trichome morphogenesis. Proc. Natl Acad. Sci. USA 94, 6261–6266 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Thomas, C. L., Bayer, E. M., Ritzenthaler, C., Fernandez-Calvino, L. & Maule, A. J. Specific targeting of a plasmodesmal protein affecting cell-to-cell communication. PLoS Biol. 6, e7 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Nelson, B. K., Cai, X. & Nebenfuhr, A. A multicolored set of in vivo organelle markers for co-localization studies in Arabidopsis and other plants. Plant J. 51, 1126–1136 (2007).

    Article  CAS  PubMed  Google Scholar 

  68. Belteton, S. A., Sawchuk, M. G., Donohoe, B. S., Scarpella, E. & Szymanski, D. B. Reassessing the roles of PIN proteins and anticlinal microtubules during pavement cell morphogenesis. Plant Physiol. 176, 432–449 (2018).

    Article  CAS  PubMed  Google Scholar 

  69. Gutierrez, R., Lindeboom, J. J., Paredez, A. R., Emons, A. M. & Ehrhardt, D. W. Arabidopsis cortical microtubules position cellulose synthase delivery to the plasma membrane and interact with cellulose synthase trafficking compartments. Nat. Cell Biol. 11, 797–806 (2009).

    Article  CAS  PubMed  Google Scholar 

  70. Sbalzarini, I. F. & Koumoutsakos, P. Feature point tracking and trajectory analysis for video imaging in cell biology. J. Struct. Biol. 151, 182–195 (2005).

    Article  CAS  PubMed  Google Scholar 

  71. Forouzesh, E., Goel, A., Mackenzie, S. A. & Turner, J. A. In vivo extraction of Arabidopsis cell turgor pressure using nanoindentation in conjunction with finite element modeling. Plant J. 73, 509–520 (2013).

    Article  CAS  PubMed  Google Scholar 

  72. Zsivanovits, G., MacDougall, A. J., Smith, A. C. & Ring, S. G. Material properties of concentrated pectin networks. Carbohydr. Res. 339, 1317–1322 (2004).

    Article  CAS  PubMed  Google Scholar 

  73. Puspoki, Z., Storath, M., Sage, D. & Unser, M. Transforms and operators for directional bioimage analysis: a survey. Adv. Anat. Embryol. Cell Biol. 219, 69–93 (2016).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We thank D. Jackson (Cold Spring Harbor Laboratory) for his generous gift of the PDLP3–GFP transgenic line and Y. Gu (Pennsylvania State University) for her generous gift of the csi1-3 mutant and csi1-3; YFP–CesA6; RFP–TUA5 transgenic line. This material is based on work supported by the National Science Foundation under grant no. 1715544 to D.B.S. and J.A.T.

Author information

Authors and Affiliations

Authors

Contributions

D.B.S. and S.A.B. conceptualized the project. S.A.B. and W.L. developed the methodology. S.A.B., W.L., M.Y., M.I.Q., M.K.S., M.W.M. and F.A.H. conducted the investigation. S.A.B. and D.B.S. wrote the original draft. S.A.B., W.L., M.Y., M.I.Q., M.K.S., M.W.M., F.A.H., J.A.T. and D.B.S. reviewed and edited the draft. D.B.S. administered the project. D.B.S. and J.A.T. acquired the funding.

Corresponding author

Correspondence to Daniel B. Szymanski.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Plants thanks Olivier Hamant 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

Extended Data Fig. 1 Methodology for PDLP growth analysis, their reliability as fiducial marks, and their usage in lobing segments.

Growth dependent separation of PDLP puncta pairs along an example segment. a, Using the plasma-membrane marker segments were traced at the anticlinal/outer-periclinal junction. b-c, Segment straightening followed by re-slicing produces a face-view of the anticlinal wall. d, Non-overlapping PDLPs including manually labeled 3-way-junctions were tracked over time. e, PDLP-pair distance is used to obtain sub-segmental growth rates from linear fit models over the time course. f, Sub-segmental elemental growth rates mapped onto the initial segment shape. PDLPs are stable features in the anticlinal wall and their displacement reflects localized diffuse growth. g-i, Cold-treated PDLP3:GFP; PIP2:mCherry seedlings were imaged at 4-hour intervals. j-k, Normalized signal intensity plots and overlapping peaks from PDLP3 particles of the two segments (g-i) at all 3 time-points. l, Boxplots of measured sub-segmental growth rates of cold-treated (CT) and room temperature (RT) grown seedlings. Center lines show the medians; box limits indicate the 25th and 75th percentiles as determined by R software; whiskers extend 1.5 times the interquartile range from the 25th and 75th percentiles, outliers are represented by dots. n = 14, 38 sub-segments from 3 and 9 independent segments respectively. Sub-segmental growth rates from lobing segments. m-o, Plots of normalized segment length at lobe detection (gray outline) and final (black outline) time-points color coded for normalized percent growth rate per hour. Cyan and green arrowheads mark the location of newly and established lobes respectively (Top panel). Perimeter to center of mass plots highlight warping regions between lobe detection and final timepoint. Scale bar = 5 µm (a-d), 10 µm (g-i).

Extended Data Fig. 2 FE model optimization and correlation of maximum principal stress orientation with subsegment growth rate.

a, Microscopic image of the cells which were used to study the effect of the partial geometry of cells for the stress response under turgor pressure. The star shows the location of the maximum deflection of the periclinal wall for Cell 2. b, 3D finite element (FE) model of the two complete cells and the distribution of maximum principal stress (Max PS) on both anticlinal walls (AW). c, Distribution of the mean of the Max PS from the upper 1/8th of the AW for both cells for three example results of partial cells in terms of the area used in the analysis. d, Maximum of the absolute value of the percentage difference between mean Max PS distributions in the upper 1/8th of the AW in each simulation relative to using the complete geometry of cells. It indicates that the FE analysis can use a partial cell. If the area includes the maximum deflection the distribution of the Max PS does not change substantially and the magnitude is reduced by < 10%. e, Comparison of the average of Max PS from the different upper areas of materials in the AW. It turns out that the mean of Max PS of the different upper areas in the walls has the same distribution pattern but varies in magnitude. f, Optical image of the cells were used to study the correlation of the Max PS and the growth rate of the cell-cell interfaces. g, Cell 2 from the measurement is defined partially, but the maximum deflection area can be seen in the image. h,i, Comparison of maximum deflection of periclinal wall for Cell 2 in (fg) from the simulations using different remaining areas of the cell. The results of (i) is very consistent with the optical measurement in (g). jl, Like the study described in (f), four other segments of cells were used for the comparison, and the results show that the horizontal component of Max PS along the segment length has a strong correlation with the growth rate of the associated wall. The scale bar for all = 10 µm.

Extended Data Fig. 3 The microtubule and cellulose systems are required for lobe formation.

Mature pavement cell shapes, magenta outline, were analyzed using LobeFinder. Mutants affecting the microtubule- (b, c, i-k), cellulose- (d, e), and pectinase- (f), and secretion (g) were compared to wild-type (a, i). The temperature-sensitive mutant mor1-1 was tested at permissive (h) and restrictive temperatures (j) and compared to its wild-type counterpart (i). Effects of 0.2% pectinase (m), 1 µM oryzalin (n), and 5 nM Isoxaben (o) on cell shape from 1 to 2 DAG compared to a buffer treated control (l). A combination of 0.2% Pectinase with a reduced concentration of Oryzalin [1 µM] (p) and the combination of 0.2% Pectinase and 5 nM Isoxaben (q) both completely halted the production on new lobes (Supplementary Table 1 – for results). Cellulose synthase complex marked with YFP:CesA6 strongly colocalizes with cortical microtubules in early stage pavement cells. s, 8-min projection of 15-sec intervals, single plane time course shows how CesA tracks resemble and co-localize (t – white arrows) with cortical microtubules, mCherry:TuA5 (r). Evidence of colocalization between cellulose synthase complex marked with CesA3:GFP and cortical microtubules at the anticlinal wall. u, Top-down view of segment resliced. Dotted line shows shape and direction of segmentation. White arrows on resliced images are CesA complexes that localized to anticlinal microtubules. The colocalization of cellulose and microtubules was not affected in 8-minute projection of 15-sec intervals, single plane time-lapses, in the csi1 mutant background where CesA tracks (YFP:CesA6 - w) resemble and co-localized (x – white arrows) with cortical microtubules, RFP:TUA5 (v). Resliced views of the anticlinal wall also show evidence of cellulose and microtubule co-localization in the csi1 mutant. CesA colocalization with anticlinal microtubules is evident in the csi1 mutant background. y, Top-down view of segment resliced with dotted line showing the shape and direction of segmentation. White arrows on resliced images on the right are CesA complexes that localize to anticlinal microtubules. The anticlinal wall at the apex remains straight during lobe formation. z, Snapshots of a lobing segment before, (a) during, (b′-c′) and after lobe formation. d′, Cross-sectional views of the anticlinal wall at the lobe apex from locations specified in (z-c′). e′, Histogram of anticlinal wall tilt showing how the vast majority of measure wall regions were completely perpendicular to the leaf surface. Scale bars = 50 µm (a-q), 25 µm (k-q), 5 µm (r-d′).

Extended Data Fig. 4 Transfacial microtubule localization and frequency analyses as a function of lobe initiation.

Analysis of microtubule turnover in time series data. Pairwise correlation results between the initial time-point and each subsequent time-point (blue line), a sliding window of paired consecutive frames over the time course (green dashed line), and each time-point to a randomly selected time-point (magenta dotted line). The random correlation results average (gray dotted line) is the background level for microtubule correlations. a-b, 15-sec interval time-lapses. c-d, 1-min interval time-lapses. e-f, 5-minute interval time-lapses. Each sampling regime had a similar pattern of correlation decay from the first time-point to background levels after ~10 min. Microtubule orientation in ROIs in opposing cells that undergo lobe initiation does not show detectable differences in alignment and coherence. g-i, Snapshots of plasma-membrane (PIP2:mCherry) of a lobing segment, yellow boxes are ROIs used to monitor microtubule behaviors in opposing cells lobe centered at the future lobe region. j-l, Snapshots of the microtubules (GFP:TUB6). Microtubules orientation and coherency results using red ovals whose tilt reflects the orientation and their aspect ratios reflect the coherency of the microtubules at the time point shown. m-n, Scatterplot of microtubules orientation (green) and microtubule coherency (purple) in cell 1 (g -top) and cell 2 (g - bottom) as a function of time before (< time-point 19) and after (≥ time-point 19) lobe detection. Distribution of manually identified transfacial microtubules in lobing segments. o-s, Top - segment shape before (blue) and after (orange) lobe detection. Yellow arrowheads are existing lobes. Cyan arrowheads are newly formed lobes. The top histogram (green) depicts transfacial microtubules at the future convex cell at the lobe initiation site. The bottom histogram (magenta) depicts transfacial microtubules of the future concave cell at the lobe initiation site. Time-lapsed analysis of the occurrence and location of transfacial microtubules in six different lobing segments. t-y, The number of transfacial microtubules in the lobing cortical domain in the future convex (green) and future concave (magenta) cells. Scale bar = 5 µm.

Extended Data Fig. 5 Cross-correlation analyses of microtubules, cell wall stress, and lobe initiation.

(a, d) Heatmap of σ-max values on the anticlinal wall for two lobing segments. (b, e) Density plots of stress alignment at the anticlinal wall for cell 1 (green) and cell 2 (magenta). (c, f) Face-view of cortical microtubules at the anticlinal wall for segment depicted on A and D respectively. g-i, Time-lapse images of the plasma-membrane, (PIP2:mCherry - magenta) and microtubules (GFP:TUB6 - green) of a lobing segment that was captured for > 4 hours before the lobe was detected. j-l, Outline of the segment shape for segment corresponding to the live-cell images above where the lobe initiation cortical domain is marked using cyan vertical lines and lobes peak locations are marked by a black vertical line. m, Density plot of manually scored transfacial microtubules where cell 1 has a peak at the future convex side of a lobe before its detection. n, The accumulated microtubule signal method can detect enriched microtubule signal (\(Enrichement = 1 - \frac{{Area\,under\,curve_{Cell1}}}{{Area\,under\,curve_{Cell2}}} \ast 100\)) in cell1 within the future convex domain. o, Summary of outer periclinal microtubule enrichment in future convex cells measured from 11 independent lobing events. Boxplots show the distribution of microtubule enrichment of future lobe convex domain prior to lobe detection. Center lines show the medians; box limits indicate the 25th and 75th percentiles as determined by R software; whiskers extend 1.5 times the interquartile range from the 25th and 75th percentiles, outliers are represented by dots. The number of time-points at 10 min intervals is listed above each boxplot. Example images of two additional lobing pairs that could be converted to FE models and tested for MT and tensile stress correlations at sites of microtubule enrichment and lobe formation. (p, t) Summed microtubule signal projected images of segments before lobe detection. (q, u) Enrichment of microtubules at the future convex side of a lobe as a function of time before lobe detection. (r, v) Line plots of FE σ Max predictions along the anticlinal wall of cell 1 (green) and cell 2 (magenta), microtubule accumulated signals along the shared anticlinal wall (black dashed line). (s, w) Line plots of FE σ Max predictions along the outer periclinal wall of cell 1 (green) and cell 2 (magenta), microtubule accumulated signals outer periclinal wall of cell 1 (green dashed line) and cell 2 (magenta dashed line). Symmetry breaking events are not defined by a pre-existing segment shape. x-z, Snapshots of eight-hour time-lapse with plasma-membrane (PIP2:mCherry - magenta) and microtubule (GFP:TUB6 - green) markers of a segment that has a symmetry-breaking event against the established bulging shape. (a′ - top) Accumulated microtubule signal plot and level of enrichment within the future convex cortical region (cell 1) (a′ - bottom) Accumulated microtubule signal plot (cell 2). Scale bars = 5 µm for c, d, gi; 10 µm for xz.

Extended Data Fig. 6 Localized microfibril alignment can drive lobe initiation: an FE analysis.

a, Overview of a lobing pair of pavement cells. A transfacial patch of anisotropic material is located in one cell while the rest of the material in both cells is isotropic. b, The material in the connected anticlinal wall in the same cell is also defined as anisotropic. Note that no transition zone near the edges of the patch is used. c, The anisotropic patch has a width of 2μm, and the simulated MF direction is aligned with the arrows. d, Maximum principal stress map after pressurization and relaxation for the completely isotropic case; this result serves as a reference for the deformation. e, Stress map for the case which includes the anisotropic patch, which shows a stress concentration within the patch area. f,g, Comparison of the cell-cell interface geometry for the isotropic (reference – green) and anisotropic (magenta) models.

Supplementary information

Supplementary Information

Supplementary Tables 1 and 2.

Reporting Summary

Supplementary Video 1

5.5-hour time-lapse of expanding pavement cells taken at 30-min intervals. Displacements of trackable bead pairs were used to evaluate the subsegmental growth rates using PDLPs (green) and plasma membrane (magenta).

Supplementary Video 2

Representative time-lapse where both the plasma membrane (magenta) and microtubules (green) are tracked as a function of lobe initiation.

Supplementary Video 3

Local feature detection method to identify and track segment protrusions that develop into lobes.

Supplementary Video 4

Time-lapse of a lobing segment in which views of the outer-periclinal (top panels) and anticlinal (bottom panels) cell surfaces are aligned and projected for both the initiating (left) and following (right) cells.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Belteton, S.A., Li, W., Yanagisawa, M. et al. Real-time conversion of tissue-scale mechanical forces into an interdigitated growth pattern. Nat. Plants 7, 826–841 (2021). https://doi.org/10.1038/s41477-021-00931-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41477-021-00931-z

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