Skip to main content
Log in

Pattern formation in auxin flux

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

The plant hormone auxin is fundamental for plant growth, and its spatial distribution in plant tissues is critical for plant morphogenesis. We consider a leading model of the polar auxin flux, and study in full detail the stability of the possible equilibrium configurations. We show that the critical states of the auxin transport process are composed of basic building blocks, which are isolated in a background of auxin depleted cells, and are not geometrically regular in general. The same model was considered recently through a continuous limit and a coupling to the von Karman equations, to model the interplay of biochemistry and mechanics during plant growth. Our conclusions might be of interest in this setting, since, for example, we establish the existence of Lyapunov functions for the auxin flux, proving in this way the convergence of pure transport processes toward the set of equilibrium points.

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

Similar content being viewed by others

References

  • Adler I (1974) A model of contact pressure in phyllotaxis. J Theor Biol 1:1–79

    Article  Google Scholar 

  • Atela P, Golé C, Hotton C (2002) A dynamical system for plant pattern formation. J Nonlinear Sci 12:641–676

    Article  MATH  MathSciNet  Google Scholar 

  • Barbier de Reuille P, Bohn-Courseau I, Ljung K, Morin H, Carraro N, Godin C, Traas J (2006) Computer simulations reveal properties of the cell-cell signaling network at the shoot apex in Arabidopsis. Proc Natl Acad Sci USA 103:1627–1632

    Article  Google Scholar 

  • Benaïm M (1997) Vertex-reinforced random walks and a conjecture of pemantle. Ann Prob 25:361–392

    Article  MATH  Google Scholar 

  • Benaïm M, Tarrès P (2008) Dynamics of vertex-reinforced random walks. ArXiv e-prints 0809.2739v3

  • Boudaoud A (2010) An introduction to the mechanics of morphogenesis for plant biologists. Trends Plant Sci 15:353–360

    Article  Google Scholar 

  • Diaconis P, Stroock D (1991) Geometric bounds for Eigenvalues of Markov chains. Ann Appl Probab 1:36–61

    Article  MATH  MathSciNet  Google Scholar 

  • Douady S, Couder Y (1996) Phyllotaxis as a dynamical self organizing process (Part I, II, III). J Theor Biol 178:255–312

    Article  Google Scholar 

  • Dumais J (2007) Can mechanics control pattern in plants? Curr Opin Plant Biol 10:58–62

    Article  Google Scholar 

  • Dumais J, Steele C (2000) New evidence for the role of mechanical forces in the shoot apex meristem. J Plant Growth Regul 19:7–18

    Google Scholar 

  • Ethier SN, Kurtz TG (1986) Markov processes: characterization and convergence. Wiley Ser Probab Math Stat

  • Gabriel JP, Hanisch H, Hirsch W (1988–1989) Prepatency and sexuality of parasitic worms: the hermaphroditic case, Anno 3, vol 4. Atti del colloquio di matematica, Edizione Cerfim Locarno

  • Green P (1980) Organogenesis—a biophysical view. Annu Rev Plant Physiol 31:51–82

    Article  Google Scholar 

  • Hamant O, Heisler MG, Jönsson H, Krupinski P, Uytterwaal M, Bokov P, Corson F, Sahlin P, Boudaoud A, Meyerowitz E, Couder Y, Traas J (2008) Developmental patterning by mechanical signals in Arabidopsis. Science 322:1650–1655

    Article  Google Scholar 

  • Hamant O, Traas J (2009) The mechanics behind plant development. New Phytol 185:369–385

    Article  Google Scholar 

  • Heisler MG, Jönsson H (2006) Modeling auxin transport and plant development. J Plant Growth Regul 25:302–312

    Article  Google Scholar 

  • Hofmeister W (1868) Handbuch der Physiologischen Botanik: Allgemeine Morphologie der Gewächse. Engelmann, Leipzig

    Google Scholar 

  • Jönsson H, Heisler MG, Shapiro BE, Mjolsness E, Meyerowitz EM (2006) An auxin-driven polarized transport model for phyllotaxis. Proc Natl Acad Sci USA 103:1633–1638

    Article  Google Scholar 

  • Kunz M (1995) Some analytical results about two physical models of phyllotaxis. Commun Math Phys 169:261–295

    Article  MATH  MathSciNet  Google Scholar 

  • LaSalle JP (1960) Some extensions of Liapunov’s second method. IRE Trans CT 7:520–527

    MathSciNet  Google Scholar 

  • Levitov LS (1991) Energetics approach to phyllotaxis. Europhys Lett 14:533–539

    Article  Google Scholar 

  • Meinhardt H (1982) Models of biological pattern formation. Academic Press, London

    Google Scholar 

  • Mjolsness E (2006) The growth and development of some recent plant models: a viewpoint. J Plant Growth Regul 25:270–277

    Article  Google Scholar 

  • Motzkin T, Straus G (1965) Maxima for graphs a new proof of a theorem of Turán. Can J Math 17:533–540

    Article  MATH  MathSciNet  Google Scholar 

  • Newell A, Shipman P (2005) Plant and fibonacci. J Stat Phys 121:937–968

    Article  MATH  MathSciNet  Google Scholar 

  • Newell AC, Shipman PD, Sun Z (2008) Phyllotaxis: cooperation and competition between mechanical and biochemical processes. J Theor Biol 251:421–439

    Article  MathSciNet  Google Scholar 

  • Norris JR (1997) Markov chains. Camb Ser Stat Probab Math

  • Pemantle R (1992) Vertex-reinforced random walk. Probab Theory Relat Fields 92:117–136

    Article  MATH  MathSciNet  Google Scholar 

  • Reinhardt D (2005) Phyllotaxis—a new chapter in an old tale about beauty and magic numbers. Curr Opin Plant Biol 8:487–493

    Google Scholar 

  • Reinhardt D, Mandel T, Kuhlemeier C (2000) Auxin regulates the initiation and radial position of lateral organs. Plant Cell 12:501–518

    Google Scholar 

  • Reinhardt D, Pesce E, Stieger P, Mandel T, Baltensperger K, Bennett M, Traas J, Friml J, Kuhlemeier C (2003) Regulation of phyllotaxis by polar auxin transport. Nature 426:255–260

    Article  Google Scholar 

  • Sahlin P, Söderberg B, Jönsson H (2009) Regulated transport as a mechanism for pattern generation: capabilities for phyllotaxis and beyond. J Theor Biol 258:60–70

    Article  Google Scholar 

  • Shipman PD, Newell AC (2005) Polygonal plantform and phyllotaxis on plants. J Theor Biol 236:154–197

    Article  MathSciNet  Google Scholar 

  • Smith RS, Guyomarch’s S, Mandel T, Reinhardt D, Kuhlemeier C et al (2006) A plausible model of phyllotaxis. Proc Natl Acad Sci USA 103:1301–1306

    Article  Google Scholar 

  • Thornley J (1975) Phyllotaxis I. A mechanistic model. Ann Botany 39:491–507

    Google Scholar 

  • Turing A (1952) The chemical basis of morphogenesis. Philos Trans R Soc Lond 237:37–72

    Article  Google Scholar 

  • Waller D (1977) Optimisation of quadratic forms associated with graphs. Glasgow Math J 18:79–85

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

This work was supported by the University of Fribourg, and by the SystemsX “Plant growth in changing environments” project funding. Many thanks to D. Kierzkowski and C. Kuhlemeier for providing us the picture given in Fig. 1 and to Aleš Janka for its help in Matlab programming. We are very grateful to Didier Reinhardt for giving us the opportunity to learn parts of the actual knowledge on the role of the auxin flux in plant patterning.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to C. Mazza.

Appendix: A

Appendix: A

1.1 A.1 Proof of the existence of the solution (Proposition 1)

In the following, we use the notation \(\dot{{\varvec{a}}}\) instead of \(\frac{\mathrm{d}{\varvec{a}}}{\mathrm{d}t}\).

System (2) can be written as

$$\begin{aligned} \dot{a}_i=D\sum _{k\sim i}a_k+T\sum _{k\sim i}\left(\frac{a_k}{\kappa +\sum _{j\sim k}a_j}-\frac{a_k}{\kappa +\sum _{j\sim i}a_j}-\frac{D}{T}\right)a_i. \end{aligned}$$
(30)

Let \({\varvec{a}}\) be a solution of (30) with \({\varvec{a}}(0)\in \mathbb R ^L_{\ge 0}\).

We say that the function \(f:\mathbb R _+\rightarrow \mathbb R \) is instantaneously positive (i.p.) if there exists \(\delta >0\) so that \(f\) is strictly positive over \((0,\delta )\). If \(f(0)>0\) and \(f\) is continuous to the right at \(0\), then \(f\) is i.p.. It is also clear that if \(f\) admits a strictly positive right-hand derivative at \(0\), then it is i.p..

Let \(U\) be the open set \(U=\{{\varvec{x}}=(x_1,x_2,\ldots ,x_L)\in \mathbb R ^L;-\frac{\kappa }{2L}<x_i\}\). Since the right-hand member of (30) is continuous over \(U\), the general theory of o.d.e.’s provides the existence of a solution defined over a maximal interval \(0\in J^+\subset \mathbb R _+\) for any initial condition \({\varvec{a}}(0)\in U\). Moreover, the solution is unique because the right-hand member of (30) locally lipschitzian. Set for convenience

$$\begin{aligned} h_i(t)&= D\sum _{k\sim i}a_k(t)\quad \text{ and} \\ g_i(t)&= T\sum _{k\sim i}\left(\frac{a_k(t)}{\kappa +\sum _{j\sim k}a_j(t)}-\frac{a_k(t)}{\kappa +\sum _{j\sim i}a_j(t)}-\frac{D}{T}\right). \end{aligned}$$

The variation of constants formula allows us to write , \(\forall t\in J^+\),

$$\begin{aligned} a_i(t)=a_i(0)e^{\int _0^tg_i(s)ds}+\int \limits _0^t h_i(u)e^{-\int _u^tg_i(v)dv}du. \end{aligned}$$
(31)

Since \(a_i(0)\ge 0\), the first term in (31) is non-negative. Moreover if \(a_k(t)\) is i.p. for some \(k\sim i\), then according to (31), the same property holds for \(a_i(t)\). In particular, if \(a_k(0)>0\) for some \(k\sim i\), then by continuity \(a_k(t)\) is i.p. and thus also \(a_i(t)\).

The case \(D>0\):

Clearly, if \({\varvec{a}}(0)= {\varvec{0}}\), then the unique solution is identically \(0\). Otherwise, there exists \(1\le i_0\le L\) with \(a_{i_0}(0)>0\) and \(\forall j\sim i_0, a_j(t)\) is i.p.. Since our graph is supposed to be connected, every \(i\) admits a neighbour \(k\sim i\) with \(a_k(t)\) i.p.. Hence, \(a_i(t)\) is i.p. \(\forall i,1\le i\le L\).

The preceding arguments show that for any initial condition \({\varvec{a}}(0)\in \mathbb R _{\ge 0}^L\subset U\), all components of the solution of (30) are i.p.. Let us suppose that one of them admits the value \(0\) in \(J^+\). Since all components are continuous and their number is finite, there exists a first time \(t_0>0\) for which at least one component \(a_{i_0}(t_0)=0\) and all of them are strictly positive over \((0,t_0)\). According to (31), we have

$$\begin{aligned} a_{i_0}(t_0)=0=a_i(0)e^{\int _0^{t_0}g_i(s)ds}+\int \limits _0^{t_0}h_i(u)e^{-\int _u^{t_0}g_i(v)dv}du. \end{aligned}$$

Clearly \(h_i(t)>0\) over \(J^+\backslash \{0\}\) and since the first term is non-negative, we conclude to \(a_{i_0}(t_0)>0\), a contradiction. Therefore all \(a_i(t)\) are strictly positive over \(J^+\backslash \{0\}\).

The case \(D=0\): If \(a_i(0)>0\) , \(a_i(t)\) is i.p. by continuity. In that case \(a_i(t)=a_i(0)e^{\int _0^tg_i(s)ds}>0\) over \(J^+\). If \(a_i(0)=0\), the homogeneous equation for \(a_i(t)\) admits only the zero solution, and we remove the related \(i\)th component from (30).

The preceding modification implies that, in both cases, solution of (30) have strictly positive components over \(J^+\backslash \{0\}\). We also proved that \(\forall t \in J^+\) we have:

$$\begin{aligned} \sum _{1\le i\le L}a_i(t)=\sum _{1\le i\le L}a_i(0). \end{aligned}$$

As a consequence the solution of (30) is bounded and thus the unique solution of our problem is defined over \(J^+=[0,+\infty )\).

We easily check that the system \(\dot{{\varvec{a}}}=f({\varvec{a}})\) is conservative, as

$$\begin{aligned} \sum _i \dot{a}_i (t)&= D \sum _i\sum _{k \sim i} (a_k -a_i)+T \sum _i \sum _{k\sim i}a_k a_i\left(\frac{N_i-N_k}{N_k N_i}\right)\\&= D \sum _i(d_i a_i-d_i a_i)+\frac{1}{2} T \sum _i\sum _{k\sim i}\frac{a_k a_i}{N_k N_i}((N_i-N_k)+(N_k-N_i))=0, \end{aligned}$$

where \(N_k=\kappa +\sum _{j\sim k}a_k\), and where \(d_i\) is the degree of \(i\) (that is the number of neighbours of \(i\)).\(\square \)

1.2 A.2 Proof of the Jacobian matrix in the irreducible case (Proposition 3)

We first give the Jacobian, for an arbitrary \({\varvec{a}}\). We have

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_j} =\frac{a_i}{N_j}-\frac{a_i}{N_i}+\sum _{k\sim i}\frac{a_i a_k}{N_i^2}-\sum _{k\sim i, k\sim j}a_k \frac{a_i}{N_k^2}, \end{aligned}$$
(32)

(where the last term is due to the triangles in the graph) when \(j\sim i\), that is, \(i\) and \(j\) are nearest neighbours. When \(i=j\), one gets

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_i} =\sum _{k\sim i}\frac{a_k}{N_k}-a_i\sum _{k\sim i}\frac{a_k}{N_k^2} -\frac{\sum _{k\sim i}a_k}{N_i}. \end{aligned}$$
(33)

The remaining non-vanishing partial derivatives correspond to nodes \(j\) located at distance 2 of \(i\) in the graph, that is, to nodes \(j\) such that \(j\sim k\) for some \(k\sim i\), \(j\ne i\) but \(i\not \sim j\). Then

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_j} =-\sum _{j\sim k,\ k\sim i}\frac{a_i a_k}{N_k^2}. \end{aligned}$$
(34)

When \(N_i = N\), \(\forall i\), these expressions simplify to

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_j} =\sum _{k\sim i}\frac{a_i a_k}{N_i^2}=\frac{N-\kappa }{N^2}a_i-\frac{a_i}{N^2}\sum _{k\sim i, k\sim j}a_k. \end{aligned}$$

If \(j\sim i\),

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_i}=-\frac{N-\kappa }{N^2}a_i, \end{aligned}$$

and

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_j} =-\sum _{k\sim i, k\sim j}\frac{a_i a_k}{N_k^2}=-\frac{a_i}{N^2}\sum _{k\sim i, k\sim j}a_k, \end{aligned}$$

if \(j\sim k\) for some \(k\sim i\), \(j\ne i\) but \(i\not \sim j\).

Consider the sub-matrix \({\varvec{M}}\) given by \({\varvec{M}}=(\partial f_i({\varvec{a}})/\partial a_j)_{j\sim i}\). Let \(\text{ diag}({\varvec{a}})\) be the diagonal matrix whose diagonal elements are given by \({\varvec{a}}\). The perturbation associated with the triangles contained in the graph is represented by the term \(-\frac{a_i}{N^2}\sum _{k\sim i, k\sim j}a_k\) in \(\frac{\partial f_i({\varvec{a}})}{\partial a_j}\) for \(j\sim i\), and the related matrix is given by

$$\begin{aligned} \!\!\!\!\!\! \left(-\frac{a_i}{N^2}\sum _{k\sim i, k\sim j}a_k\right)\gamma _{ij}&= \left(-\frac{a_i}{N^2}\sum _{k}\gamma _{ik}a_k\gamma _{kj}\right)\gamma _{ij}\\ \quad&= \left( -\frac{1}{N^2}\left(\text{ diag}({\varvec{a}}){\varvec{\varGamma }} \text{ diag}({\varvec{a}}){\varvec{\varGamma }}-\mathrm{diag}( \text{ diag}({\varvec{a}}){\varvec{\varGamma }} \text{ diag}({\varvec{a}}){\varvec{\varGamma }})\right)_{ij}\right)\gamma _{ij}\\ \quad&= \left( -\frac{1}{N^2}(\text{ diag}({\varvec{a}}){\varvec{\varGamma }} \text{ diag}({\varvec{a}}){\varvec{\varGamma }})_{ij}+\frac{N-\kappa }{N^2}\text{ diag}({\varvec{a}})_{ij}\right)\gamma _{ij}.\\ \end{aligned}$$

The matrix \({\varvec{M}}\) is now given by

$$\begin{aligned} {\varvec{M}}=\frac{\text{ diag}({\varvec{a}})}{N^2}(N-\kappa )({\varvec{\varGamma }}-\mathbf{id})- \frac{1}{N^2}(\text{ diag}({\varvec{a}}){\varvec{\varGamma }} \text{ diag}({\varvec{a}}) {\varvec{\varGamma }}-(N-\kappa )\text{ diag}({\varvec{a}}))\circ {\varvec{\varGamma }}, \end{aligned}$$

where \(\circ \) represents the matrix Hadamard product, i.e. the multiplication component by component.

Similarly, the perturbation of \({\varvec{M}}\) by \(\left(\frac{\partial f_i({\varvec{a}})}{\partial a_j}\right)_{i\sim k, k\sim j, i\not \sim j, i\ne j}\) can be written as

$$\begin{aligned}&\left(-\frac{a_i}{N^2}\sum _{\begin{array}{c} k\sim i, k\sim j,\\ \ i \not \sim j, i\ne j \end{array}} a_k \right)\gamma _{ij}= \left(-\frac{a_i}{N^2}\sum _{k}\gamma _{ik}a_k\gamma _{kj}\right)(1-\gamma _{ij}-\mathbf{id}_{ij})\\&\qquad = \left( -\frac{1}{N^2}(\text{ diag}({\varvec{a}}){\varvec{\varGamma }} \text{ diag}({\varvec{a}}){\varvec{\varGamma }})_{ij}+\frac{N-\kappa }{N^2}\text{ diag}({\varvec{a}})_{ij}\right)(1-\gamma _{ij}-\mathbf{id}_{ij}).\\ \end{aligned}$$

The related Jacobian is thus given by \({\varvec{M}}+\left(\frac{\partial f_i({\varvec{a}})}{\partial a_j}\right)_{i\sim k, k\sim j, i\not \sim j, i\ne j}\), that is

$$\begin{aligned} \mathrm{d}f({\varvec{a}})&= \frac{\text{ diag}({\varvec{a}})}{N^2}(N{-}\kappa )({\varvec{\varGamma }}{-}\mathbf{id}){-} \frac{1}{N^2}\left(\text{ diag}({\varvec{a}}){\varvec{\varGamma }} \text{ diag}({\varvec{a}}) {\varvec{\varGamma }}-(N-\kappa )\text{ diag}({\varvec{a}})\right) \circ {\varvec{\varGamma }}\\&- \frac{1}{N^2}\left(\text{ diag}({\varvec{a}}){\varvec{\varGamma }} \text{ diag}({\varvec{a}}) {\varvec{\varGamma }}-(N-\kappa )\text{ diag}({\varvec{a}})\right) \circ ({\varvec{1}}-{\varvec{\varGamma }}-\mathbf{id})\\&= \frac{\text{ diag}({\varvec{a}})}{N^2}(N{-}\kappa )({\varvec{\varGamma }}-\mathbf{id})- \frac{1}{N^2}\left(\text{ diag}({\varvec{a}}){\varvec{\varGamma }} \text{ diag}({\varvec{a}}) {\varvec{\varGamma }} -(N-\kappa )\text{ diag}({\varvec{a}})\right) \circ ({\varvec{1}}-\mathbf{id})\\&= \frac{\text{ diag}({\varvec{a}})}{N^2}(N-\kappa )({\varvec{\varGamma }}-\mathbf{id})- \frac{1}{N^2}(\text{ diag}({\varvec{a}}){\varvec{\varGamma }} \text{ diag}({\varvec{a}}) {\varvec{\varGamma }}-(N-\kappa )\text{ diag}({\varvec{a}})),\\ \end{aligned}$$

where \({\varvec{1}}\) is the matrix composed of ones only. The last equality is a consequence of the fact that the diagonal of \(\text{ diag}({\varvec{a}}){\varvec{\varGamma }} \text{ diag}({\varvec{a}}) {\varvec{\varGamma }}-(N-\kappa )\text{ diag}({\varvec{a}})\) vanishes. Hence,

$$\begin{aligned} \mathrm{d}f({\varvec{a}})=\frac{\text{ diag}({\varvec{a}}){\varvec{\varGamma }}}{N^2}((N-\kappa )\mathbf{id}-\text{ diag}({\varvec{a}}){\varvec{\varGamma }} ) =\frac{\text{ diag}({\varvec{a}}){\varvec{\varGamma }}}{N^2}( c\ \mathbf{id}-\text{ diag}({\varvec{a}}){\varvec{\varGamma }} ), \end{aligned}$$

proving the result.

1.3 A.3 Proof of the spectrum in the reducible case (Proposition 4)

Set \(I=\{i \in \varLambda : a_i=0\}\), and consider the subgraphs \(\gamma _p\) of \(G\) induced by the nodes of \(J=\varLambda \setminus I\), with \(\gamma _p=(\varLambda _p,E_p)\), \(1 \le p \le P\). For any related equilibrium point \({\varvec{a}}\), the restrictions \({\varvec{a}}\vert _{\gamma _p}\) satisfy the linear systems \({\varvec{\varGamma }}_{\gamma _p}{\varvec{a}}\vert _{\gamma _p}=c_{\gamma _p} {\varvec{1}}\vert _{\gamma _p}\). Set \(N_{\gamma _p}=c_{\gamma _p}+ \kappa \).

Equations (32)–(34) allow the computation of the entries of the Jacobian matrix. By first looking at the diagonal entries, for \(i\in \varLambda _p\), one has

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_i} = - a_i \frac{N_{\gamma _p}-\kappa }{N_{\gamma _p}^2}, \end{aligned}$$

providing the diagonal entry of the Jacobian of \(f\vert _{\gamma _p}({\varvec{a}}\vert _{\gamma _p})\). When \(i\not \in \varLambda _p\), a similar computation yields

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_i} = \sum _{k\sim i} \frac{a_k}{N_k} - \frac{N_i-\kappa }{N_i}. \end{aligned}$$

We then compute the entries \((i,j)\) for \(j \sim i\), \(i,j \in \varLambda _p\) and \(1\le p \le P\):

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_j} =a_i \frac{N_{\gamma _p} - \kappa }{N_{\gamma _p}^2}-\sum _{k\sim i, k\sim j} a_k \frac{a_i}{N_k^2} = a_i \frac{N_{\gamma _p} - \kappa }{N_{\gamma _p}^2}-\sum _{k\sim i, k\sim j, k \in \varLambda _p} \frac{a_k a_i}{N_{\gamma _p}^2} , \end{aligned}$$

This corresponds to the \((i,j)\) entry of the Jacobian of \(f\vert _{\gamma _p}({\varvec{a}}\vert _{\gamma _p})\). Similarly, when \(i \in \varLambda _p\) for some \(p\) and \(j \not \in \varLambda _p\)

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_j} = \frac{a_i}{N_j}- \frac{a_i}{N_{\gamma _p}} {+} \sum _{k\sim i} \frac{a_i a_k}{N_{\gamma _p}^2}{-}\sum _{\begin{array}{c} k\sim i, k\sim j, \\ k \in \varLambda _p \end{array}} \frac{a_k a_i}{N_{\gamma _p}^2} = \frac{a_i}{N_j} - a_i \frac{\kappa }{N_{\gamma _p}^2} - \frac{a_i}{N_{\gamma _p}^2} \sum _{\begin{array}{c} k\sim i, k\sim j,\\ k \in \varLambda _p \end{array}} a_ k . \end{aligned}$$

Finally, when \(i,j \not \in \cup _p\varLambda _p\), or equivalently when both \(i\) and \(j\) belongs to \(I\)

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_j}=0. \end{aligned}$$

We next consider \((i,j)\) entries where \(j\) is at a distance 2 of \(i\) in the graph \(G\), that is when \(j\) is such that \(j\sim k\) for some \(k\sim i\), \(j\ne i\) and \(j \not \sim i\). For \(i,j,k \in \varLambda _p\)

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_j}=-a_i \sum _{j \sim k, k \sim i} \frac{a_k}{N_{\gamma _p}^2}. \end{aligned}$$

This is the \((i,j)\) entry of the Jacobian of \(f\vert _{\gamma _p}({\varvec{a}}\vert _{\gamma _p})\).

Similarly, for \(i,k \in \varLambda _p, j\not \in \varLambda _p\) (\(\Rightarrow j \in I\))

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_j}=-a_i \sum _{j \sim k, k \sim i} \frac{a_k}{N_{\gamma _p}^2} . \end{aligned}$$

Next, for \(i \, \text{ or} \, k \not \in \varLambda _p, \forall j \in \varLambda \)

$$\begin{aligned} \frac{\partial f_i({\varvec{a}})}{\partial a_j}=0. \end{aligned}$$

Permuting conveniently the indices, the Jacobian \(\mathrm{d}f({\varvec{a}})\) can be written as

$$\begin{aligned} \mathrm{d}f({\varvec{a}}) = \begin{pmatrix} {\varvec{d}_n}&{\varvec{0}} \\ *&\mathrm{d}f^{\gamma } \end{pmatrix} \end{aligned}$$

where \({\varvec{d}_n}\) is a diagonal matrix \(n \times n\) with entries \(\lambda _i:= \sum _{k\sim i} \frac{a_k}{N_k} - \frac{N_i-\kappa }{N_i}\), for \(i \in I\), and \(\mathrm{d}f^{\gamma }\) is a block diagonal matrix, each block being equal to the Jacobian of \(f\) restricted on each subgraph \(\gamma _p\). The permutation allows grouping of all indices \(i \in I\) in the same block, and all indices related to the subgraphs \(\gamma _p\) are also arranged together. It follows that the spectrum of \(\mathrm{d}f({\varvec{a}})\) is :

$$\begin{aligned} \mathrm{spec}(\mathrm{d}f({\varvec{a}}))= \bigcup _{p=1}^P \mathrm{spec}\left(\mathrm{d}f\vert _{\gamma _p}({\varvec{a}}\vert _{\gamma _p})\right) \cup \left\{ \sum _{k\sim i} \frac{a_k}{N_k} - \frac{N_i-\kappa }{N_i}, i\in I \right\} . \end{aligned}$$

1.4 A.4 Proof of Proposition 5

To prove the Proposition 5, we will use the following Proposition, see Gabriel et al. (1989).

Proposition 9

Let \(f:\mathbb R _+\rightarrow \mathbb R \) be twice differentiable and bounded together with \(\ddot{f}\). If, as \(n\rightarrow +\infty \), \(t_n\uparrow +\infty \) and \(f(t_n)\rightarrow \liminf _{t\rightarrow +\infty } f(t)\) (or \(f(t_n)\rightarrow \limsup _{t\rightarrow +\infty } f(t)\)), then \(\dot{f}(t_n)\rightarrow 0\).

Proof

If \(a_k(0)=0\) for all \(1\le k\le L\), then the unique solution is identically zero. Otherwise \(\sum _{k=1}^L a_k(0)>0\). Let us suppose that for some \(i\in \{1,\ldots ,L\}\),

$$\begin{aligned} \liminf _{t\rightarrow +\infty }a_i(t)=0. \end{aligned}$$

Let us introduce the notation \(\underline{a}_i=\liminf _{t\rightarrow +\infty }a_i(t)\). Since \(a_i(t)\) is bounded together with its second derivative, the preceding proposition applies and for any sequence \(t_n\uparrow +\infty \) such that \(a_i(t_n)\rightarrow \underline{a}_i\), we have \(\dot{a}_i(t_n)\rightarrow 0\) as \(n\rightarrow +\infty \). Every \(a_k(t_n)\) being bounded in the right-hand member of the equation for \(\dot{a}_i(t_n)\), we conclude that \(\lim _{n\rightarrow +\infty }D\sum _{k\sim i}a_k(t_n)=0\). The non-negativity of each \(a_k(t_n)\) entails \(\lim _{n\rightarrow +\infty }a_k(t_n)=0=\underline{a}_k\) for every \({k\sim i}\). According to the above proposition, \(\lim _{n\rightarrow +\infty }\dot{a}_k(t_n)=0\) for every \({k\sim i}\) and since the graph is connected, repeating the same argument provides \(\lim _{n\rightarrow +\infty }\dot{a}_j(t_n)=0\) for every \(j\in \{1,\ldots ,L\}\). Thus \(0=\lim _{n\rightarrow +\infty }\sum _{1\le j\le L}a_j(t_n)=\sum _{k=1}^L a_k(0)>0\), a contradiction. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Feller, C., Gabriel, J.P., Mazza, C. et al. Pattern formation in auxin flux. J. Math. Biol. 68, 879–909 (2014). https://doi.org/10.1007/s00285-013-0655-9

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-013-0655-9

Keywords

Navigation