Skip to main content
Log in

The Dynamics of Root Growth: A Geometric Model

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

A new model for macroscopic root growth based on a dynamical Riemannian geometry is presented. Assuming that the thickness of the root is much less than its length, the model is restricted to growth in one dimension (1D). We treat 1D tissues as continuous, deformable, growing geometries for sizes larger than 1 mm. The dynamics of the growing root are described by a set of coupled tensor equations for the metric of the tissue and velocity field of material transport in non-Euclidean space. These coupled equations represent a novel feedback mechanism between growth and geometry. We compare 1D numerical simulations of these tissue growth equations to two measures of root growth. First, sectional growth along the simulated root shows an elongation zone common to many species of plant roots. Second, the relative elemental growth rate calculated in silico exhibits spatio-temporal dynamics recently characterized in high-resolution root growth studies but which thus far lack a biological hypothesis to explain them. In our model, these dynamics are a direct consequence of considering growth as both a geometric reaction–diffusion process and expansion due to a distributed source of new materials.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  • Audoly B, Boudaoud A (2003) Self-similar structures near boundaries in strained systems. Phys Rev Lett. doi:10.1103/PhysRevLett.91.086105

  • Band LR, Úbeda-Tomás S, Dyson RJ, Middleton AM, Hodgman TC, Owen MR, Jensen OE, Bennett MJ, King JR (2012) Growth-induced hormone dilution can explain the dynamics of plant root cell elongation. Proc Natl Acad Sci 109(19):7577–7582

    Article  Google Scholar 

  • Basu P, Pal A (2012) A new tool for analysis of root growth in the spatio-temporal continuum. New Phytol 195(1):264–274

    Article  Google Scholar 

  • Beemster GT, Baskin TI (1998) Analysis of cell division and elongation underlying the developmental acceleration of root growth in Arabidopsis thaliana. Plant Physiol 116(4):1515–1526

    Article  Google Scholar 

  • Chavarría-Krauser A, Jäger W, Schurr U (2005) Primary root growth: a biophysical model of auxin-related control. Funct Plant Biol 32(9):849–862

    Article  Google Scholar 

  • Chavarría-Krauser A, Nagel KA, Palme K, Schurr U, Walter A, Scharr H (2008) Spatio-temporal quantification of differential growth processes in root growth zones based on a novel combination of image sequence processing and refined concepts describing curvature production. New Phytol 177(3):811–821

    Article  Google Scholar 

  • d’Inverno R (1992) Introducing Einstein’s relativity. Clarendon Press, Oxford

    MATH  Google Scholar 

  • Erickson RO, Silk WK (1980) The kinematics of plant growth. Sci Am 242(5):134–151

    Article  Google Scholar 

  • Hejnowicz Z, Romberger JA (1984) Growth tensor of plant organs. J Theor Biol 110(1):93–114

    Article  Google Scholar 

  • Kardar M (2007) Statistical physics of fields. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Kardar M, Parisi G, Zhang YC (1986) Dynamic scaling of growing interfaces. Phys Rev Lett 56(9):889

    Article  MATH  Google Scholar 

  • Klein Y, Efrati E, Sharon E (2007) Shaping of elastic sheets by prescription of non-Euclidean metrics. Science 315(5815):1116–1120

    Article  MathSciNet  MATH  Google Scholar 

  • Lockhart JA (1965) An analysis of irreversible plant cell elongation. J Theor Biol 8(2):264–275

    Article  Google Scholar 

  • Niklas KJ (1992) Plant biomechanics: an engineering approach to plant form and function. University of Chicago Press, Chicago

    Google Scholar 

  • Pletcher RH, Tannehill JC, Anderson D (2012) Computational fluid mechanics and heat transfer. CRC Press, Boca Raton

    MATH  Google Scholar 

  • Pulwicki J (2015) Dynamics of plant growth; a theory based on Riemannian geometry. Ph.D. thesis, University of Calgary

  • Saito T, Soga K, Hoson T, Terashima I (2006) The bulk elastic modulus and the reversible properties of cell walls in developing Quercus leaves. Plant Cell Physiol 47(6):715–725

    Article  Google Scholar 

  • Sharp RE, Silk WK, Hsiao TC (1988) Growth of the maize primary root at low water potentials. I. Spatial distribution of expansive growth. Plant Physiol 87(1):50–57

    Article  Google Scholar 

  • Silk WK (1989) Growth rate patterns which maintain a helical tissue tube. J Theor Biol 138(3):311–328

    Article  Google Scholar 

  • Silk WK, Erickson RO (1979) Kinematics of plant growth. J Theor Biol 76(4):481–501

    Article  Google Scholar 

  • Taiz L, Zeiger E (2010) Plant physiology. Sinauer Associates, Sunderland, MA

    Google Scholar 

  • Topping P (2006) Lectures on the Ricci flow, vol 325. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Villa E, Matarrese S, Maino D (2014) Cosmological dynamics: from the Eulerian to the Lagrangian frame—I. Newtonian approximation. J Cosmol Astropart Phys 06:41–62

    Article  MathSciNet  Google Scholar 

  • Walter A, Feil R, Schurr U (2003) Expansion dynamics, metabolite composition and substance transfer of the primary root growth zone of zea mays l. grown in different external nutrient availabilities. Plant Cell Environ 26(9):1451–1466

    Article  Google Scholar 

  • Walter A, Spies H, Terjung S, Küsters R, Kirchgeßner N, Schurr U (2002) Spatio-temporal dynamics of expansion growth in roots: automatic quantification of diurnal course and temperature response by digital image sequence processing. J Exp Bot 53(369):689–698

    Article  Google Scholar 

  • Wang L, Beyer ST, Cronk QC, Walus K (2011) Delivering high-resolution landmarks using inkjet micropatterning for spatial monitoring of leaf expansion. Plant Methods 7(1):1–10

    Article  Google Scholar 

  • van der Weele CM, Jiang HS, Palaniappan KK, Ivanov VB, Palaniappan K, Baskin TI (2003) A new algorithm for computational image analysis of deformable motion at high spatial and temporal resolution applied to root growth. roughly uniform elongation in the meristem and also, after an abrupt acceleration, in the elongation zone. Plant Physiol 132(3):1138–1148

    Article  Google Scholar 

Download references

Acknowledgements

This project was partially funded through an Alberta Innovates Graduate Student Scholarship to J.P. and a Natural Sciences and Engineering Research Council of Canada (NSERC) Discovery Grant to D.H.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to David Hobill.

Appendix: Riemannian Geometry: A Short Introduction

Appendix: Riemannian Geometry: A Short Introduction

1.1 1. The Metric Tensor

This appendix introduces some of the fundamental concepts of Riemannian geometry those who might be unfamiliar with the subject or require a re-familiarization with the material. Reviews of Riemannian geometry and tensor analysis can be presented in one of two ways: in (1) an intrinsic coordinate free form or (2) in an explicit coordinate dependent form using index notation. Since the latter formulation provides a means for developing computational schemes and making comparisons with observation, index notation will be used throughout this review.

Although it is assumed that the geometry of root growth is described by a one-dimensional space or manifold, everything presented here will assume that there exists a set of curvilinear coordinates \(\{x^i\}\) where \(i=1,2, \ldots N\) in an N-dimensional space. Therefore, a vector quantity \(\mathbf {A}\) will have N components represented by \(A^i\) in such a manifold.

The coordinate differences between points do not necessarily give the physical separations between the points that define the coordinates. Suppose two points in the manifold P and Q are infinitesimally separated. Point P will have coordinates \(x^i\), while point Q will have coordinates \(x^i + \hbox {d}x^i\). The physical separation d l between the points is determined by the metric tensor \(g_{ik}\) with \(N^2\) components and the quadratic differential form:

$$\begin{aligned} \hbox {d}l^2 = \sum ^N_{i = 1} \sum ^N_{k = 1} g_{ik} \hbox {d}x^i \hbox {d}x^k . \end{aligned}$$

The length \(\hbox {d}l\) is a scalar quantity and is invariant under arbitrary transformations of the coordinates \(x^i \rightarrow \tilde{x}^i(x)\).

For Cartesian coordinates (\(x^1=x\), \(x^2=y\), \(x^3=z\)) the physical distance between P and Q in a flat 3D space is given by the standard Pythagorean relation:

$$\begin{aligned} \hbox {d}l^2 = \hbox {d}x^2 + \hbox {d}y^2 + \hbox {d}z^2 \end{aligned}$$

and the metric tensor is just the unit tensor (with 1’s along the diagonal):

In the more general case the metric tensor will have components that depend on the coordinates \(x^i\) and perhaps time. In what follows, the Einstein summation notation will be used so as to simplify the expressions that appear in all subsequent computations. This convention drops the summation sign over repeated indices and simply assumes that any pair of the same index in an expression implicitly assumes a summation over that index through the range of that index from 1 to N. Therefore, the double sum in the expression for \(\hbox {d}l^2\) given above can be rewritten as:

$$\begin{aligned} {\hbox {d}}l^2 = g_{ik} \hbox {d}x^i \hbox {d}x^k \end{aligned}$$
(28)

where the sum over i and k is now indicated by the fact that both indices appear in pairs. Riemannian geometries require that the metric tensor is a symmetric tensor:

$$\begin{aligned} g_{ik} = g_{ki} \end{aligned}$$

and the components of the metric tensor lead to a positive definite values of \(\hbox {d}l^2\).

While scalars are mathematical quantities that remain invariant under coordinate transformations, vectors and tensors will have different components in different coordinate systems. If the different coordinate systems are curvilinear or non-orthogonal then displacement vectors will have different mathematical properties compared to vectors obtained from the gradient of scalar functions. The Jacobian matrix of the transformation \(x^i \rightarrow \tilde{x}^i(x)\)

$$\begin{aligned} \tilde{J}^i_j = \frac{\partial \tilde{x}^i}{\partial {x}^j} \end{aligned}$$

and its inverse:

$$\begin{aligned} {J}^j_i = \frac{\partial {x}^j}{\partial \tilde{x}^i} \end{aligned}$$

lead to the transformation rules for contravariant vector components \(A^i\) (i.e. those that transform like differentials)

$$\begin{aligned} \tilde{A}^k = \tilde{J}^k_i A^i = \frac{\partial \tilde{x}^k}{\partial {x}^i} A^i \end{aligned}$$
(29)

and covariant vector components \(A_i\) (i.e. those that transform like gradients)

$$\begin{aligned} \tilde{A}_k = J^i_k A_i = \frac{\partial x^i}{\partial \tilde{x}^k} A_i. \end{aligned}$$
(30)

In fact it is these transformation rules that can be used to define the difference between the contravariant and the covariant components of a vector \(\mathbf {A}\). The metric tensor \(g_{ik}\) and its inverse \(g^{ik}\) (which obeys \(g^{im}g_{mk} = \delta ^i_k\)—the Kronecker delta) can be considered as an operator that transforms contravariant components into covariant components and vice versa. That is:

$$\begin{aligned} A_i = g_{ik}A^k \qquad \mathrm{and} \qquad A^i = g^{ik} A_k. \end{aligned}$$

NB: The pair of indices appearing in the Einstein summation must have one contravariant component (superscript) and one covariant component (subscript); therefore, the square of the magnitude of a vector \(\mathbf {A}\) is:

$$\begin{aligned} |\mathbf {A}|^2 = A^2 = A_iA^i = g_{ik}A^kA^i. \end{aligned}$$

The rules for tensor transformations follow in a similar manner. One must associate a Jacobian matrix with each index that describes the tensors components. Since tensor requires multiple indices, one can have all contravariant components, all covariant components or a mixture of both. All the possible second-rank tensor components transform as:

$$\begin{aligned} \tilde{T}^{ik} = \frac{\partial \tilde{x}^i}{\partial {x}^m} \frac{\partial \tilde{x}^k}{\partial {x}^n} T^{mn} \end{aligned}$$

for completely contravariant components of a second-rank tensor.

$$\begin{aligned} \tilde{T}_{ik} = \frac{\partial {x}^m}{\partial \tilde{x}^i} \frac{\partial {x}^n}{\partial \tilde{x}^k} T_{mn} \end{aligned}$$

for completely covariant components of a second-rank tensor.

$$\begin{aligned} \tilde{T}^{i}_{k} = \frac{\partial \tilde{x}^i}{\partial {x}^m} \frac{\partial {x}^n}{\partial \tilde{x}^k} T^m_n \end{aligned}$$

for the mixed components of a second-rank tensor.

These transformation properties are required if the mathematical relationships between tensor quantities are to remain valid in all coordinate systems. This allows for the establishment of general laws and equations that remain invariant under the transformation of coordinates. Hence the approach taken in this article. While the components of different objects may change under coordinate transformations, the fundamental relationship between the scalars, vectors and tensors remains unchanged under a transformation of coordinates.

1.2 2. The Connection Coefficients

Higher-rank tensors can be created by taking tensor products of lower-rank tensors. For example a second-rank tensor can be constructed from two vectors \(A^i\) and \(B^i\) along with a scalar \(\kappa \) by:

$$\begin{aligned} T^{ik} = \kappa A^i B^k \end{aligned}$$

Since there is no summation here, \(T^{ik}\) consists of the ordered product of the two vectors. Alternatively higher-rank tensors can be computed by taking derivatives of lower-rank tensors. However, for a general coordinate transformation partial differentiation does not obey the tensorial transformation rules established above. Therefore, a new form of differentiation is required in order to maintain the tensorial characteristics of derivatives of tensors.

This requires the introduction of a geometric quantity that removes the non-tensorial terms from the partial derivative to yield a tensor (or covariant) derivative that has the proper transformation properties. This quantity \(\varGamma ^i_{jk}\) is called a “connection coefficient”, and it creates the covariant derivative for a contravariant vector to create a mixed second-rank tensor:

$$\begin{aligned} \nabla _k A^i = \frac{\partial A^i}{\partial x^k} + \varGamma ^i_{jk}A^j. \end{aligned}$$
(31)

The covariant derivative of a covariant vector is a completely covariant second-rank tensor:

$$\begin{aligned} \nabla _k A_i = \frac{\partial A_i}{\partial x^k} - \varGamma ^j_{ik}A_j. \end{aligned}$$
(32)

A special set of connection coefficients that can be written in terms of the metric and its derivatives called Christoffel symbols of the second kind leave the metric tensor invariant to the tensor derivative, and this fact leads to:

$$\begin{aligned} \varGamma ^i_{jk} = \frac{1}{2}g^{i\ell }\left( \frac{\partial g_{j\ell }}{\partial x^k} + \frac{\partial g_{\ell k}}{\partial x^j} - \frac{\partial g_{jk}}{\partial x^\ell } \right) . \end{aligned}$$
(33)

Due to the symmetry of the metric the connection coefficients are symmetric in the covariant indices:

$$\begin{aligned} \varGamma ^i_{jk} = \varGamma ^i_{kj}. \end{aligned}$$

The connection coefficients are also important in that they define how a particle with no external forces acting on it travel along the geodesics of a curved geometry. In flat space such a particle follows a straight line with constant velocity. In a curved manifold the velocity of a particle \(v^i\) obeys the “geodesic motion equation”:

$$\begin{aligned} \frac{\hbox {d}v^i}{\hbox {d}s} = - \varGamma ^{i}_{jk} v^j v^k \end{aligned}$$
(34)

where the parameter s is a scalar quantity that parameterizes the position of the particle along the geodesic.

1.3 3. Measures of Curvature

Unlike mixed partial derivatives, mixed tensor derivatives do not commute. Since the tensor derivatives are tensors it can be expected that the commutator of covariant derivatives is also a tensor. In fact it can be shown that:

$$\begin{aligned} \nabla _j \nabla _k A_i- \nabla _k \nabla _j A_i = R^{\ell }_{ijk}A_{\ell } \end{aligned}$$

where the new tensor \(R^a_{bcd}\) is a quantity computed from partial derivatives of the Christoffel symbols and products of the Christoffel symbols. It is the famous “Riemann tensor” that provides an unambiguous measure of local curvature of the manifold. Explicitly it is

$$\begin{aligned} R^i_{jk\ell } = \frac{\partial \varGamma ^i_{j\ell }}{\partial x^k} -\frac{\partial \varGamma ^i_{jk}}{\partial x^\ell } +\varGamma ^m_{j\ell } \varGamma ^i_{mk} -\varGamma ^m_{jk} \varGamma ^i_{m\ell } \end{aligned}$$
(35)

obtained after much algebra. This is a fourth-rank mixed tensor and is a nonlinear functional of the metric and its first and second partial derivatives.

With this tensor Riemann was able to complete the work begun by Gauss by finding an expression for the curvature of a space from knowing its intrinsic geometric properties without reference to any embeddings of that space in a higher-dimensional manifold. This is the motivation for introducing Riemannian curvature into a growth model. The local interactions that take place in the manifold and are independent of the external space in which the manifold exists.

From the symmetries of the metric and Christoffel symbols one can show that the number of independent components of the Riemann tensor is \(\frac{1}{12}N^2(N^2-1)\). Thus, a 3D manifold has six independent components to the Riemann tensor, and a 2D surface has only one independent component.

It is possible to create other tensors that measure some form of curvature. For example the Ricci curvature tensor \(R_{ab}\) = \(R^c_{acb}\) or:

$$\begin{aligned} R_{ik} = R^j_{ijk} = \frac{\partial \varGamma ^j_{ik}}{\partial x^j} -\frac{\partial \varGamma ^j_{ij}}{\partial x^k} +\varGamma ^j_{ik } \varGamma ^m_{jm} -\varGamma ^j_{im} \varGamma ^m_{jk} \end{aligned}$$
(36)

The Ricci tensor can be thought of as an averaged curvature measure since it does not contain all of the required local information regarding curvature. However, on a 2D surface the one component of the Riemann tensor is transferred to the Ricci tensor. For a 2D surface with coordinates \((x^1=u,x^2=v)\) the distance between points can be given by:

$$\begin{aligned} \hbox {d}\ell ^2 = f(u,v) [\hbox {d}u^2 + \hbox {d}v^2] \end{aligned}$$

where f(uv) is a position-dependent “scale factor”. The Ricci tensor for this metric is computed to be:

$$\begin{aligned} R_{11} = R_{22} = \frac{1}{2f^2} \left[ f\frac{\partial ^2 f}{\partial u^2} + f\frac{\partial ^2 f}{\partial v^2} - \left( \frac{\partial f}{\partial u} \right) ^2 - \left( \frac{\partial f}{\partial v} \right) ^2 \right] \end{aligned}$$
(37)

Although a 1D manifold has no Riemannian curvature, the geometric KPZ operator given by Eq. (10) has the same structure as the Ricci tensor. It is linear in the second spatial derivative and quadratic in the gradient of the scale factor. Finally forming the inner product of the Ricci curvature with the metric leads to a scalar curvature measure:

$$\begin{aligned} R = g^{ab}R_{ab} \end{aligned}$$

and this quantity is often called the Ricci scalar.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pulwicki, J., Hobill, D. The Dynamics of Root Growth: A Geometric Model. Bull Math Biol 79, 1820–1845 (2017). https://doi.org/10.1007/s11538-017-0308-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-017-0308-z

Keywords

Navigation