Skip to main content
Log in

Spatiotemporal dynamics of a reaction-diffusion model of pollen tube tip growth

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

Abstract

A reaction-diffusion model is proposed to describe the mechanisms underlying the spatial distributions of ROP1 and calcium on the pollen tube tip. The model assumes that the plasma membrane ROP1 activates itself through positive feedback loop, while the cytosolic calcium ions inhibit ROP1 via a negative feedback loop. Furthermore it is proposed that lateral movement of molecules on the plasma membrane are depicted by diffusion. It is shown that bistable or oscillatory dynamics could exist even in the non-spatial model, and stationary and oscillatory spatiotemporal patterns are found in the full spatial model which resemble the experimental data of pollen tube tip growth.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19

Similar content being viewed by others

References

  • Altschuler SJ, Angenent SB, Wang Y, Wu LF (2008) On the spontaneous emergence of cell polarity. Nature 454(7206):886–889

    Google Scholar 

  • Busenberg S, Huang W-Z (1996) Stability and Hopf bifurcation for a population delay model with diffusion effects. J Differ Equ 124(1):80–107

    MathSciNet  MATH  Google Scholar 

  • Chen S-S, Lou Y, Wei J-J (2018) Hopf bifurcation in a delayed reaction-diffusion-advection population model. J Differ Equ 264(8):5333–5359

    MathSciNet  MATH  Google Scholar 

  • Chen S-S, Shi J-P (2012) Stability and Hopf bifurcation in a diffusive logistic population model with nonlocal delay effect. J Differ Equ 253(12):3440–3470

    MathSciNet  MATH  Google Scholar 

  • Chen S-S, Shi J-P, Wei J-J (2013) Time delay-induced instabilities and Hopf bifurcations in general reaction-diffusion systems. J Nonlinear Sci 23(1):1–38

    MathSciNet  MATH  Google Scholar 

  • Chen S-S, Shi J-P, Wei J-J (2014) Bifurcation analysis of the Gierer-Meinhardt system with a saturation in the activator production. Appl Anal 93(6):1115–1134

    MathSciNet  MATH  Google Scholar 

  • Chen S-S, Yu J-S (2016a) Stability analysis of a reaction-diffusion equation with spatiotemporal delay and Dirichlet boundary condition. J Dyn Differ Equ 28(3–4):857–866

    MathSciNet  MATH  Google Scholar 

  • Chen S-S, Yu J-S (2016b) Stability and bifurcations in a nonlocal delayed reaction-diffusion population model. J Differ Equ 260(1):218–240

    MathSciNet  MATH  Google Scholar 

  • Chou C-S, Nie Q, Yi T-M (2008) Modeling robustness tradeoffs in yeast cell polarization induced by spatial gradients. PloS One 3(9):e3103

    Google Scholar 

  • Ding D-Q, Shi J-P, Wang Y (2017) Bistability in a model of grassland and forest transition. J Math Anal Appl 451(2):1165–1178

    MathSciNet  MATH  Google Scholar 

  • Edelstein-Keshet L, Holmes WR, Zajac M, Dutot M (2013) From simple to detailed models for cell polarization. Philos Trans R Soc Lond B Biol Sci 368(1629):20130003

    Google Scholar 

  • Feijó JA, Sainhas J, Holdaway-Clarke T, Cordeiro MS, Kunkel JG, Hepler PK (2001) Cellular oscillations and the regulation of growth: the pollen tube paradigm. Bioessays 23(1):86–94

    Google Scholar 

  • Gierer A, Meinhardt H (1972) A theory of biological pattern formation. Biol Cybern 12(1):30–39

    MATH  Google Scholar 

  • Goryachev AB, Pokhilko AV (2008) Dynamics of Cdc42 network embodies a Turing-type mechanism of yeast cell polarity. FEBS Lett 582(10):1437–1443

    Google Scholar 

  • Gu Y, Fu Y, Dowd P, Li S-D, Vernoud V, Gilroy S, Yang Z-B (2005) A rho family gtpase controls actin dynamics and tip growth via two counteracting downstream pathways in pollen tubes. J Cell Biol 169(1):127–138

    Google Scholar 

  • Guo S-J (2015) Stability and bifurcation in a reaction-diffusion model with nonlocal delay effect. J Differ Equ 259(4):1409–1448

    MathSciNet  MATH  Google Scholar 

  • Holmes WR, Edelstein-Keshet L (2016) Analysis of a minimal Rho-GTPase circuit regulating cell shape. Phys Biol 13(4):046001

    Google Scholar 

  • Hwang J-U, Gu Y, Lee Y-J, Yang Z-B (2005) Oscillatory ROP GTPase activation leads the oscillatory polarized growth of pollen tubes. Mol Biol Cell 16(11):5385–5399

    Google Scholar 

  • Jilkine A, Marée AFM, Edelstein-Keshet L (2007) Mathematical model for spatial segregation of the Rho-family GTPases based on inhibitory crosstalk. Bull Math Biol 69(6):1943–1978

    MathSciNet  MATH  Google Scholar 

  • Jin J-Y, Shi J-P, Wei J-J, Yi F-Q (2013) Bifurcations of patterned solutions in the diffusive Lengyel-Epstein system of CIMA chemical reactions. Rocky Mt J Math 43(5):1637–1674

    MathSciNet  MATH  Google Scholar 

  • Kondo S, Miura T (2010) Reaction-diffusion model as a framework for understanding biological pattern formation. Science 329(5999):1616–1620

    MathSciNet  MATH  Google Scholar 

  • Li H, Lin Y-K, Heath RM, Zhu M X, Yang Z-B (1999) Control of pollen tube tip growth by a rop gtpase–dependent pathway that leads to tip-localized calcium influx. Plant Cell 11(9):1731–1742

    Google Scholar 

  • Li X, Wang H, Zhang Z, Hastings A (2014) Mathematical analysis of coral reef models. J Math Anal Appl 416(1):352–373

    MathSciNet  MATH  Google Scholar 

  • Lo W-C, Park H-O, Chou C-S (2014) Mathematical analysis of spontaneous emergence of cell polarity. Bull Math Biol 76(8):1835–1865

    MathSciNet  MATH  Google Scholar 

  • Ludwig D, Jones DD, Holling CS (1978) Qualitative analysis of insect outbreak systems: the spruce budworm and forest. J Anim Ecol 47(1):315–332

    Google Scholar 

  • Luo N, Yan A et al (2017) Exocytosis-coordinated mechanisms for tip growth underlie pollen tube growth guidance. Nat Commun 8(1):1687

    Google Scholar 

  • Maini P, Painter K, Chau H (1997) Spatial pattern formation in chemical and biological systems. J Chem Soc Faraday Trans 93(20):3601–3610

    Google Scholar 

  • Mogilner A, Allard J, Wollman R (2012) Cell polarity: quantitative modeling as a tool in cell biology. Science 336(6078):175–179

    MathSciNet  MATH  Google Scholar 

  • Moore TI, Chou C-S, Nie Q, Jeon NL, Yi T-M (2008) Robust spatial sensing of mating pheromone gradients by yeast cells. PloS One 3(12):e3865

    Google Scholar 

  • Mori Y, Jilkine A, Edelstein-Keshet L (2011) Asymptotic and bifurcation analysis of wave-pinning in a reaction-diffusion model for cell polarization. SIAM J Appl Math 71(4):1401–1427

    MathSciNet  MATH  Google Scholar 

  • Mumby PJ, Hastings A, Edwards HJ (2007) Thresholds and the resilience of Caribbean coral reefs. Nature 450(7166):98–101

    Google Scholar 

  • Perko L (2001) Differential equations and dynamical systems, texts in applied mathematics, vol 7, 3rd edn. Springer, New York

    MATH  Google Scholar 

  • Rätz A, Röger M (2012) Turing instabilities in a mathematical model for signaling networks. J Math Biol 65(6–7):1215–1244

    MathSciNet  MATH  Google Scholar 

  • Scheffer M, Carpenter S, Foley JA, Folke C, Walker B (2001) Catastrophic shifts in ecosystems. Nature 413(6856):591–596

    Google Scholar 

  • Scheffer M, Hosper SH, Meijer ML, Moss B, Jeppesen E (1993) Alternative equilibria in shallow lakes. Trends Ecol Evol 8(8):275–279

    Google Scholar 

  • Seirin Lee S, Gaffney EA, Baker RE (2011) The dynamics of Turing patterns for morphogen-regulated growing domains with cellular response delays. Bull Math Biol 73(11):2527–2551

    MathSciNet  MATH  Google Scholar 

  • Shi Q-Y, Shi J-P, Song Y-L (2017) Hopf bifurcation in a reaction-diffusion equation with distributed delay and Dirichlet boundary condition. J Differen Equ 263(10):6537–6575

    MathSciNet  MATH  Google Scholar 

  • Shi Q-Y, Shi J-P, Song Y-L (2019a) , Effect of spatial average on the spatiotemporal pattern formation of reaction-diffusion systems, Preprint

  • Shi Q-Y, Shi J-P, Song Y-L (2019b) Hopf bifurcation and pattern formation in a diffusive delayed logistic model with spatial heterogeneity. Discrete Contin Dyn Syst Ser B 24(2):467–486

    MathSciNet  MATH  Google Scholar 

  • Simonett G (1995) Center manifolds for quasilinear reaction-diffusion systems. Differ Integral Equ 8(4):753–796

    MathSciNet  MATH  Google Scholar 

  • Staver AC, Archibald S, Levin SA (2011a) The global extent and determinants of savanna and forest as alternative biome states. Science 334(6053):230–232

    MATH  Google Scholar 

  • Staver AC, Archibald S, Levin SA (2011b) Tree cover in sub-Saharan Africa: rainfall and fire constrain forest and savanna as alternative stable states. Ecology 92(5):1063–1072

    Google Scholar 

  • Su Y, Wei J-J, Shi J-P (2009) Hopf bifurcations in a reaction-diffusion population model with delay effect. J Differ Equ 247(4):1156–1184

    MathSciNet  MATH  Google Scholar 

  • Turing AM (1952) The chemical basis of morphogenesis. Philos Trans R Soc Lond Ser B 237(641):37–72

    MathSciNet  MATH  Google Scholar 

  • Wang J-F, Shi J-P, Wei J-J (2011) Dynamics and pattern formation in a diffusive predator-prey system with strong Allee effect in prey. J Differ Equ 251(4–5):1276–1304

    MathSciNet  MATH  Google Scholar 

  • Wang J-F, Wei J-J, Shi J-P (2016) Global bifurcation analysis and pattern formation in homogeneous diffusive predator-prey systems. J Differ Equ 260(4):3495–3523

    MathSciNet  MATH  Google Scholar 

  • Xiao Z, Brunel N, Yang Z-B. Cui X.-P (2016) Constrained nonlinear and mixed effects of differential equation models for dynamic cell polarity signaling, arXiv:1605.00185

  • Yan A, Xu G-S, Yang Z-B (2009) Calcium participates in feedback regulation of the oscillating ROP1 Rho GTPase in pollen tubes. Proc Natl Acad Sci U.S.A. 106(51):22002–22007

    Google Scholar 

  • Yan X-P, Li W-T (2010) Stability of bifurcating periodic solutions in a delayed reaction-diffusion population model. Nonlinearity 23(6):1413–1431

    MathSciNet  MATH  Google Scholar 

  • Yang Z-B (2008) Cell polarity signaling in arabidopsis. Annu Rev Cell Deve Biol 24:551–575

    Google Scholar 

  • Yi F-Q, Gaffney E, Seirin-Lee S (2017) The bifurcation analysis of Turing pattern formation induced by delay and diffusion in the Schnakenberg system. Discrete Contin Dyn Syst Ser B 22(2):647–668

    MathSciNet  MATH  Google Scholar 

  • Yi F-Q, Wei J-J, Shi J-P (2009) Bifurcation and spatiotemporal patterns in a homogeneous diffusive predator-prey system. J Differ Equ 246(5):1944–1977

    MathSciNet  MATH  Google Scholar 

  • Yi T-M, Chen S-Q, Chou C-S, Nie Q (2007) Modeling yeast cell polarization induced by pheromone gradients. J Stat Phys 128(1–2):193–207

    MATH  Google Scholar 

  • Zheng Z-Z, Chou C-S, Yi T-M, Nie Q (2011) Mathematical analysis of steady-state solutions in compartment and continuum models of cell polarization. Math Biosci Eng 8(4):1135–1168

    MathSciNet  MATH  Google Scholar 

  • Zhou J, Shi J-P (2015) Pattern formation in a general glycolysis reaction-diffusion system. IMA J Appl Math 80(6):1703–1738

    MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Junping Shi.

Additional information

Publisher's Note

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

XP Cui is partially supported by National Science Foundation Grant ATD-1222718 and the University of California, Riverside AES-CE RSAP A01869; ZB Yang is partially supported by National Institute of General Medical Sciences Grant GM100130; JP Shi is partially supported by National Science Foundation Grant DMS-1715651; QY Shi is partially supported by China Scholarship Council.

Appendix

Appendix

Proof of Proposition 3.1

In this proposition, we study the number of roots of equation (3.4) with \(1<\alpha <2\). The function f(R) has the properties that

$$\begin{aligned} \lim _{R \rightarrow 0^+} f(R) = -\infty \quad \text {and} \quad \lim _{R \rightarrow \infty } f(R) = \infty . \end{aligned}$$
(6.1)

Also, we have the first derivative of f(R) as

$$\begin{aligned} f'(R)=\alpha R^{\alpha -1}-(\alpha -1)R^{\alpha -2}+k_3^2(\alpha -2)R^{\alpha -3}-k_3^2(\alpha -3)R^{\alpha -4}. \end{aligned}$$
(6.2)

Let

$$\begin{aligned} h(R)=\alpha R^3-(\alpha -1)R^2+k_3^2(\alpha -2)R-k_3^2(\alpha -3). \end{aligned}$$
(6.3)

So we have \(f'(R)=R^{\alpha -4}h(R)\).

Step 1. There exists a unique \(x_2>0\) such that \(h'(R)<0\) for \(0<R<x_2\), \(h'(R)>0\) for \(R>x_2\), and h(R) reaches the global minimum in \((0,\infty )\) at \(R=x_2\).

Note that the function h(R) has the properties that

$$\begin{aligned} h(0)=-k_3^2(\alpha -3)>0 \quad \text {and} \quad \lim _{R \rightarrow \infty } h(R) = \infty , \end{aligned}$$
(6.4)

and we have the first derivative of h(R) as

$$\begin{aligned} h'(R)=3\alpha R^2-2(\alpha -1)R+k_3^2(\alpha -2). \end{aligned}$$
(6.5)

Since \(\alpha \in (1,2)\), for (6.5), we have the discriminant \(\bigtriangleup _1=4(\alpha -1)^2-12k_3^2\alpha (\alpha -2)>0\). In this case, \(h'(R)=0\) must have two roots in \((-\infty ,\infty )\). Notice that \(\dfrac{2(\alpha -1)}{3\alpha }>0\) and \(\dfrac{k_3^2(\alpha -2)}{3\alpha }<0\) so \(h'(R)\) must have one negative root and one positive root. Let \(x_2\) be the positive root of \(h'(R)=0\). Since \(h'(0)=k_3^2(\alpha -2)<0\), we have \(h'(R)<0\) for \(0<R<x_2\), \(h'(R)>0\) for \(R>x_2\). Therefore, h(R) decreases for \(0<R<x_2\), and increases for \(R>x_2\). That is to say, if \(h(x_2)\ge 0\), then \(h(R)\ge 0\) for any \(R>0\), while \(h(R)=0\) has two positive solutions if \(h(x_2)<0\).

Step 2. There exist \(k_{31},k_{32}>0\) such that

$$\begin{aligned} h(x_2) \left\{ \begin{array}{cc} \ge 0, \quad &{} \text {if } k_{31}<k_3<k_{32}. \quad \quad \quad \quad \quad \\<0, \quad &{} \text {if } 0<k_3<k_{31} \text { or } k_3>k_{32}. \end{array} \right. \end{aligned}$$
(6.6)

Since \(h'(R)\) is an quadratic function and the relationship between h(R) and \(h'(R)\), we can have following two facts:

$$\begin{aligned} x_2&=\dfrac{(\alpha -1)+\sqrt{(\alpha -1)^2-3k_3^2\alpha (\alpha -2)}}{3\alpha }, \end{aligned}$$
(6.7)
$$\begin{aligned} h(x_2)=&\left( \dfrac{x_2}{3}-\dfrac{\alpha -1}{9\alpha }\right) h'(x_2)+\dfrac{6k_3^2\alpha (\alpha -2)-2(\alpha -1)^2}{9\alpha }x_2 \nonumber \\&+\dfrac{k_3^2(\alpha -1)(\alpha -2)-9k_3^2\alpha (\alpha -3)}{9\alpha }. \end{aligned}$$
(6.8)

Therefore, we have

$$\begin{aligned}&h(x_2)\ge 0 \end{aligned}$$
(6.9)
$$\begin{aligned} \Leftrightarrow \quad&x_2\le \dfrac{(\alpha -1)(\alpha -2)-9\alpha (\alpha -3)}{2(\alpha -1)^2-6k_3^2\alpha (\alpha -2)}k_3^2 \end{aligned}$$
(6.10)
$$\begin{aligned} \Leftrightarrow \quad&\dfrac{(\alpha -1)+\sqrt{(\alpha -1)^2-3k_3^2\alpha (\alpha -2)}}{3\alpha }\le \dfrac{(\alpha -1)(\alpha -2)-9\alpha (\alpha -3)}{2(\alpha -1)^2-6k_3^2\alpha (\alpha -2)}k_3^2 \end{aligned}$$
(6.11)
$$\begin{aligned} \Leftrightarrow \quad&Ak_3^4+Bk_3^2+C\ge 0, \end{aligned}$$
(6.12)

where

$$\begin{aligned} A=&\alpha (\alpha -2)^3<0, \end{aligned}$$
(6.13)
$$\begin{aligned} B=&2(\alpha -1)^2(\alpha -2)^2-18(\alpha -1)(\alpha -2)+27>0, \end{aligned}$$
(6.14)
$$\begin{aligned} C=&(\alpha -1)^3(\alpha -3)<0 . \end{aligned}$$
(6.15)

Notice that the discriminant of the quadratic function \(Ay^2+By+C\) is

$$\begin{aligned} \bigtriangleup _2&=[2(\alpha -1)^2(\alpha -2)^2-18(\alpha -1)(\alpha -2){+}27]^2-4\alpha (\alpha -1)^3(\alpha -2)^3(\alpha -3) \end{aligned}$$
(6.16)
$$\begin{aligned}&=[2(\alpha -1)^2(\alpha -2)^2-18(\alpha -1)(\alpha -2)+27]^2 \nonumber \\&\qquad -4(\alpha -1)^4(\alpha -2)^4 +8(\alpha -1)^3(\alpha -2)^3 \end{aligned}$$
(6.17)
$$\begin{aligned}&>-18(\alpha -1)(\alpha -2)[4(\alpha -1)^2(\alpha -2)^2 \nonumber \\&\qquad -18(\alpha -1)(\alpha -2)+27]+8(\alpha -1)^3(\alpha -2)^3 \end{aligned}$$
(6.18)
$$\begin{aligned}&=-2(\alpha -1)(\alpha -2)[32(\alpha -1)^2(\alpha -2)^2-162(\alpha -1)(\alpha -2)+243]. \end{aligned}$$
(6.19)

Since the discriminant of the quadratic function \(32y^2-162y+243\) is \(\bigtriangleup _3=162^2-4\times 32\times 243=-4860<0\), then \(32(\alpha -1)^2(\alpha -2)^2-162(\alpha -1)(\alpha -2)+243>0\) for any \(\alpha \). Therefore

$$\begin{aligned} \bigtriangleup _2>-2(\alpha -1)(\alpha -2)[32(\alpha -1)^2(\alpha -2)^2-162( \alpha -1)(\alpha -2)+243]>0. \end{aligned}$$
(6.20)

So the quadratic equation \(Ay^2+By+C=0\) has two real-valued solutions \(k_{31}^*<k_{32}^*\). Because \(k_{31}^*+k_{32}^*=-\dfrac{B}{A}>0\) and \(k_{31}^*k_{32}^*=\dfrac{C}{A}>0\), then \(k_{31}^*\) and \(k_{32}^*\) are both positive. Moreover we can have that \(Ak_3^4+Bk_3^2+C\ge 0\) if and only if \(k_{31}^*\le k_3^2\le k_{32}^*\). Now let \(k_{31}=\sqrt{k_{31}^*}\) and \(k_{32}=\sqrt{k_{32}^*}\), we reach the conclusion in (6.6).

Step 3. We consider the number of roots of equation \(f(R)=0\) in (3.4) for each case in (6.6). In the case where \(h(x_2)\ge 0\), we would have \(h(R)\ge 0\) for any \(R>0\) because h(R) decreases for \(0<R<x_2\), and increases for \(R>x_2\). That is to say, \(f'(R)=R^{\alpha -4}h(R)>0\) for any \(R>0\). So f(R) increases for all \(R>0\). According to property (6.1), \(f(R)=0\) has one unique positive root.

On the other hand, when \(h(x_2)<0\), \(h(R)=0\) has two positive solutions. Let \(0<r_1<x_2<r_2\) be the solutions of \(h(R)=0\). Then \(h(R)>0\) if \(R\in [0,r_1)\cup (r_2,+\infty )\) and \(h(R)<0\) if \(R\in (r_1,r_2)\). That is to say,

$$\begin{aligned} f'(R)=R^{\alpha -4}h(R) \left\{ \begin{array}{c@{\qquad }c}>0, &{} \text {if } R \in (0,r_1), \\ <0, &{} \text {if } R \in (r_1,r_2), \\ >0, &{} \quad \text {if } R \in (r_2,+\infty ) . \end{array} \right. \end{aligned}$$
(6.21)

Therefore, f(R) increases for \(0<R<r_1\), decreases for \(r_1<R<r_2\), and increases when \(R>r_2\). Then we know that

  1. 1.

    If \(f(r_1)f(r_2)>0\), then \(f(R)=0\) has one unique positive solution.

  2. 2.

    If \(f(r_1)f(r_2)=0\), then \(f(R)=0\) has two positive solutions.

  3. 3.

    If \(f(r_1)f(r_2)<0\), then \(f(R)=0\) has three positive solutions.

Define

$$\begin{aligned} l(R)=R^{\alpha -3}(1-R)(R^2+k_3^2). \end{aligned}$$
(6.22)

Then \(f(r_1)f(r_2)=[k_2-l(r_1)][k_2-l(r_2)]\). Since \(r_1\) and \(r_2\) are solutions of \(h(R)=0\), \(r_1\) and \(r_2\) only depends on \(\alpha \) and \(k_3\). So there exists \(k_{21}\), \(k_{22}\) which only depends on \(\alpha \), \(k_3\) and are defined as

$$\begin{aligned} k_{21}=r_1^{\alpha -3}(1-r_1)(r_1^2+k_3^2), \;\; k_{22}=r_2^{\alpha -3}(1-r_2)(r_2^2+k_3^2), \end{aligned}$$
(6.23)

such that

  1. 1.

    If \(k_2<k_{21}\) or \(k_2>k_{22}\), then \(f(R)=0\) has one unique positive solution.

  2. 2.

    If \(k_2=k_{21}\) or \(k_2=k_{22}\), then \(f(R)=0\) has two positive solutions.

  3. 3.

    If \(k_{21}<k_2<k_{22}\), then \(f(R)=0\) has three positive solutions.

We claim that \(0<k_{21}<k_{22}\) for \(0<k_3<k_{31}\), while \(k_{21}<k_{22}<0\) for \(k_3>k_{32}\). This is equivalent to \(r_1<r_2<1\) for \(0<k_3<k_{31}\), while \(1<r_1<r_2\) for \(k_3<k_{32}\). Notice that \(h(1)=1+k_3^2>0\), and that h(R) decreases for \(0<R<x_2\) and increases for \(R>x_2\). So we only need to prove that \(h'(1)=\alpha +2+k_3^2(\alpha -2)<0\) for \(0<k_3<k_{31}\), while \(h'(1)>0\) for \(k_3>k_{32}\). In fact, \(k_{31}^2\) and \(k_{32}^2\) are two positive roots of equation \(Ay^2+By+C=0\), where ABC are defined as (6.13), (6.14) and (6.15) respectively. Notice that \(A<0\) and

$$\begin{aligned} A\left( \dfrac{2+\alpha }{2-\alpha }\right) ^2+B\left( \dfrac{2+\alpha }{2-\alpha }\right) +C =32\left( \alpha -\dfrac{1}{4}\right) ^2+54\dfrac{\alpha }{2-\alpha }>0, \quad \text {for } 1<\alpha <2. \end{aligned}$$
(6.24)

So \(k_{31}^2<\dfrac{2+\alpha }{2-\alpha }<k_{32}^2\). Therefore, we have

$$\begin{aligned} h'(1)&=\alpha +2+k_3^2(\alpha -2)>\alpha +2+k_{31}^2(\alpha -2)>0, \quad \text {if }0<k_3<k_{31}, \end{aligned}$$
(6.25)
$$\begin{aligned} h'(1)&=\alpha +2+k_3^2(\alpha -2)<\alpha +2+k_{32}^2(\alpha -2)<0, \quad \text {if }k_3>k_{32}. \end{aligned}$$
(6.26)

So we have proved that \(0<k_{21}<k_{22}\) for \(0<k_3<k_{31}\), while \(k_{21}<k_{22}<0\) for \(k_3>k_{32}\). Therefore we reach the conclusion about number of solution of \(f(R)=0\). \(\square \)

Proof of Proposition 3.2

By the Center Manifold Theorem (Page 116 in Perko (2001)), we can compute the center manifold near the equilibrium (0, 0):

$$\begin{aligned} C=\vartheta (R)=\frac{1}{k_1(2-\alpha )}R^{2-\alpha }+o(R^{2-\alpha }). \end{aligned}$$
(6.27)

Then, by substituting (6.27) into the first equation of the kinetic system (3.1), we obtain the following scalar system which gives the flow of Eq. (3.1) on the center manifold:

$$\begin{aligned} R_t=k_{1}R^{\alpha }(1-R)-k_1k_{2}\dfrac{R\vartheta ^2(R)}{\vartheta ^2(R)+k_3^2}>0, \;\; \text {for}~0<R<\delta . \end{aligned}$$
(6.28)

Thus, we know that the flow on the center manifold is moving away from the origin and it is an unstable orbit.

Next we show that there is an invariant region near \(R=0\), \(C>0\) for Eq. (3.1). Define

$$\begin{aligned} O=\left\{ (R,C): 0\le R\le \left( \frac{k_2}{2(k_3^2+\delta ^2)}\right) ^{\frac{1}{\alpha -1}}C^{\frac{2}{\alpha -1}},~0\le C\le \delta \right\} . \end{aligned}$$

It is obvious that \(R=0\) is invariant for (3.1). Then, if \(C=\delta ,~0\le R\le \left( \dfrac{k_2}{2(k_3^2+\delta ^2)}\right) ^{\frac{1}{\alpha -1}}\delta ^{\frac{2}{\alpha -1}}\), since \(1<\alpha <2\) then \(\frac{2}{\alpha -1}>1\), so one can choose \(\delta >0\) small enough so that \(\left( \dfrac{k_2}{2(k_3^2+\delta ^2)}\right) ^{\frac{1}{\alpha -1}}\delta ^{\frac{2}{\alpha -1}}\le \delta \). By using \(C=\delta \), we have \(C^{\prime }<0\).

On the boundary \(R=\left( \dfrac{k_2}{2(k_3^2+\delta ^2)}\right) ^{\frac{1}{\alpha -1}}C^{\frac{2}{\alpha -1}}\), we have

$$\begin{aligned} \begin{aligned} \frac{d}{d t}\left( \dfrac{R^{\frac{\alpha -1}{2}}}{C}\right)&=\dfrac{\frac{\alpha -1}{2}C R^{\frac{\alpha -3}{2}}R^{\prime }-R^{\frac{\alpha -1}{2}}C^{\prime }}{C^2}=\dfrac{R^{\frac{\alpha -3}{2}}}{C^2}\left( \frac{\alpha -1}{2}C R^{\prime }-R C^{\prime }\right) \\&=\dfrac{R^{\frac{\alpha -3}{2}}}{C^2}\left[ \frac{\alpha -1}{2} Ck_1\left( R^{\alpha }-R^{\alpha +1}-k_2\frac{R C^2}{C^2+k_3^2}\right) -R^2+C R\right] \\&\le \dfrac{R^{\frac{\alpha -3}{2}}}{C}\left[ \frac{\alpha -1}{2}k_1\left( R^{\alpha }-k_2\frac{R C^2}{C^2+k_3^2}\right) +R\right] \\&\le \dfrac{R^{\frac{\alpha -3}{2}}}{C}\left[ (\alpha -1)k_1R^{\alpha }-\frac{(\alpha -1)k_1k_2R C^2}{2(C^2+k_3^2)}\right] \\&\le \dfrac{R^{\frac{\alpha -3}{2}}}{C}\left[ (\alpha -1)k_1\left( R^{\alpha }-\frac{(\alpha -1)k_1k_2R C^2}{2(\delta ^2+k_3^2)}\right) \right] \\&=(\alpha -1)R^{\frac{\alpha -1}{2}}k_1 C^2\left[ \frac{R^{\alpha -1}}{C^2}-\frac{k_2}{2(\delta ^2+k_3^2)}\right] =0, \end{aligned} \end{aligned}$$

and the first inequality holds for R small enough: \(R<\frac{\alpha -1}{2}k_1R^{\alpha }\). The above calculation implies that the dynamics of (3.1) is inward on \(R=\left( \dfrac{k_2}{2(k_3^2+\delta ^2)}\right) ^{\frac{1}{\alpha -1}}C^{\frac{2}{\alpha -1}}\). This shows that O is an invariant region for Eq. (3.1), and any orbit in O converges to the origin. It is also clear that when \(R=0,~C>0\) (the positive C-axis), we have \((R_t,C_t)=(0,-C)\). So we know that all the solutions starting from \(R=0,~C>0\) will always stay on this curve and eventually converge to the origin. One can choose a maximum orbit \(R=h_s(C)\) so that all orbits such that \(0\le R\le h_s(C)\) converge to the origin. Then other trajectories exhibits saddle behavior near the origin. \(\square \)

Proof of Theorem 3.3

First, we look at the determinant of the Jacobian matrix \(J(R_j,R_j)\): \(\text {Det}(J(R_j,R_j))=k_1R_j(f_2'(R_j)-f_1'(R_j))\), where \(f_1,f_2\) are defined in (3.6). From (3.6), we have

$$\begin{aligned}&f_2(R)-f_1(R)=\dfrac{k_2R^2}{R^2+k_3^2}-R^{\alpha -1}(1-R) \end{aligned}$$
(6.29)
$$\begin{aligned} \Leftrightarrow&f(R)=k_2-R^{\alpha -3}(1-R)(R^2+k_3^2)=\dfrac{R^2+k_3^2}{R^2}(f_2(R)-f_1(R)) \end{aligned}$$
(6.30)
$$\begin{aligned} \Leftrightarrow&f'(R)=-\dfrac{2k_3^2}{R^3}(f_2(R)-f_1(R))+\dfrac{R^2+k_3^2}{R^2}(f_2'(R)-f_1'(R)). \end{aligned}$$
(6.31)

From (6.30), we know that \(f_2(R)-f_1(R)=0\) when \(f(R)=0\). Since positive steady states \((R_j,R_j)\) (\(j=1,2,3\)) satisfy \(f(R_j)=0\), we have \(f_2(R_j)-f_1(R_j)=0\). Therefore, from (6.31), we know

$$\begin{aligned}&f'(R_j)>0 \Leftrightarrow f_2'(R_j)-f_1'(R_j)>0 \Leftrightarrow \text {Det}(J(R_j,R_j))>0; \\&f'(R_j)=0 \Leftrightarrow f_2'(R_j)-f_1'(R_j)=0 \Leftrightarrow \text {Det}(J(R_j,R_j))=0; \\&f'(R_j)<0 \Leftrightarrow f_2'(R_j)-f_1'(R_j)<0 \Leftrightarrow \text {Det}(J(R_j,R_j))<0. \end{aligned}$$

According to the proof of Proposition 3.1, we have the following result of \(\text {Det}(J(R_j,R_j))\): there exists a constant \(k_{31}>0\) such that

  1. 1.

    If \(0<k_3<k_{31}\), then there exists \(r_1\) and \(r_2\), which are two positive solutions of \(h(R)=0\), such that

    1. (a)

      \(f'(R)>0\Rightarrow \text {Det}(J(R,R))>0\) for \(R \in (0,r_1) \cup (r_2,1)\);

    2. (b)

      \(f'(R)<0\Rightarrow \text {Det}(J(R,R))<0\) for \(R \in (r_1,r_2)\).

  2. 2.

    If \(k_{31}<k_3\), then for any \(0<R<1\), we always have \(f'(R)>0\Rightarrow \text {Det}(J(R,R))>0\).

Here we want to point out that from (6.23), we have \(k_2(r_1)=k_{21}\) and \(k_2(r_2)=k_{22}\).

Next we look at the trace of Jacobian matrix (3.5): \(\text {Tr}(J(R_j,C_j))=k_1R_jf_1'(R_j)-1\). Define a new function

$$\begin{aligned} g(R)=Rf_1'(R)=R^{\alpha -1}[(\alpha -1)-\alpha R]. \end{aligned}$$
(6.32)

We observe that g(R) has the following properties:

$$\begin{aligned} g(0)=0, \quad g\left( \dfrac{\alpha -1}{\alpha }\right) =0, \quad \text {and} \quad \lim _{R \rightarrow \infty } g(R) = \infty . \end{aligned}$$
(6.33)

Also we have the first derivative of g(R) as

$$\begin{aligned} g'(R)=R^{\alpha -2}[(\alpha -1)^2-\alpha ^2R]. \end{aligned}$$
(6.34)

Hence the function g(R) increases for \(0<R<\displaystyle \left( \dfrac{\alpha -1}{\alpha }\right) ^2\) and decreases for \(R>\displaystyle \left( \dfrac{\alpha -1}{\alpha }\right) ^2\). So g(R) achieves its maximum at \(R=\displaystyle \left( \dfrac{\alpha -1}{\alpha }\right) ^2\) with \(g\left( \left( \dfrac{\alpha -1}{\alpha }\right) ^2\right) =\left( \dfrac{\alpha -1}{\alpha }\right) ^{2\alpha -1}\).

Therefore we conclude that

  1. 1.

    If \(k_1<\left( \dfrac{\alpha }{\alpha -1}\right) ^{2\alpha -1}\), then

    $$\begin{aligned} \text {Tr}(J(R,R))=k_1g(R)-1<\left( \dfrac{\alpha }{\alpha -1}\right) ^{2\alpha -1}\left( \dfrac{\alpha -1}{\alpha }\right) ^{2\alpha -1}-1=0. \end{aligned}$$
    (6.35)
  2. 2.

    If \(k_1>\left( \dfrac{\alpha }{\alpha -1}\right) ^{2\alpha -1}\), then there exists \(0<{\tilde{R}}_1<{\tilde{R}}_2\), such that \(g({\tilde{R}}_1)=g({\tilde{R}}_2)=\dfrac{1}{k_1}\). Therefore,

    1. (a)

      If \({\tilde{R}}_1<R<{\tilde{R}}_2\), then \(\text {Tr}(J(R,R))=k_1g(R)-1>k_1g({\tilde{R}}_1)=0.\)

    2. (b)

      If \(0<R<{\tilde{R}}_1\) or \({\tilde{R}}_2<R<1\), then \(\text {Tr}(J(R,R))=k_1g(R)-1<k_1g({\tilde{R}}_1)=0.\)

\(\square \)

Proof of Proposition 3.4

From the proof of Theorem 3.3, we can easily get Part 1 in Proposition 3.4. So here we only discuss Part 2: the case that \(0<k_3<k_{31}\). Since \(\text {Det}(J(R_2,C_2))<0\), the steady state \((R_2,C_2)\) is always a saddle point. So we focus on the positive steady states \((R_1,C_1)\) and \((R_3,C_3)\). To prove the results in Proposition 3.4, we need to determine the order of the possible bifurcation points: \(r_1\), \(r_2\), \({\tilde{R}}_1\) and \({\tilde{R}}_2\), where \(r_1\) and \(r_2\) are the steady state bifurcation points satisfying \(h(r_1)=h(r_2)=0\) with h(R) defined in (6.3), and \({\tilde{R}}_1\), \({\tilde{R}}_2\) are possible Hopf bifurcation points satisfying \(g({\tilde{R}}_1)=g({\tilde{R}}_2)=1/k_{1}\). Then, by the results of Theorem 3.3, we can obtain the stability of each steady state.

First, we prove that \(g(r_1)>g(r_2)\) always holds. From the definition of h(R), we know that

$$\begin{aligned} h(r_1)=\alpha r_1^3-(\alpha -1)r_1^2+k_3^2(\alpha -2)r_1-k_3^2(a\alpha -3)=0. \end{aligned}$$
(6.36)

Multiplying (6.36) by \(r_1^{\alpha -3}\), we have

$$\begin{aligned} \alpha r_1^{\alpha }-(\alpha -1)r_1^{\alpha -1}+k_3^2(\alpha -2)r_1^{\alpha -2}-k_3^2(a\alpha -3)r_1^{\alpha -3}=0, \end{aligned}$$
(6.37)

which together with \(g(r_1)=-\alpha r_1^{\alpha }+(\alpha -1)r_1^{\alpha -1}\) from (6.32) implies that

$$\begin{aligned} g(r_1)=k_3^2(\alpha -2)r_1^{\alpha -2}-k_3^2(\alpha -3)r_1^{\alpha -3}. \end{aligned}$$
(6.38)

Define

$$\begin{aligned} G(R)=k_3^2(\alpha -2)R^{\alpha -2}-k_3^2(\alpha -3)R^{\alpha -3},~R\in (0,1),~\alpha \in (1,2). \end{aligned}$$
(6.39)

By direct calculation, we have \(G^{\prime }(R)=k_3^2(\alpha -2)^{2}R^{\alpha -3}-k_3^2(\alpha -3)^{2}R^{\alpha -4}\) and \(G^{\prime }(R)<0\) for \(R\in \left( 0, \left( \frac{\alpha -3}{\alpha -2}\right) ^2\right) \supset (0,1)\). Therefore, G(R) is strictly decreasing for \(R\in (0,1)\). By the fact that \(0<r_1<r_2<1\), immediately we reach the conclusion that \(g(r_1)>g(r_2)\).

Now we consider the case that \(0<k_3<k_{31}\) which implies the existence of multiple steady states. For the convenience of discussion, we define

$$\begin{aligned} {\tilde{g}}(R)=g(R)-1/k_1, \end{aligned}$$
(6.40)

then we know that \({\tilde{g}}\) has two zeros \({\tilde{R}}_1\) and \({\tilde{R}}_2\). For the order of \(r_1\), \(r_2\), \({\tilde{R}}_1\) and \({\tilde{R}}_2\), we have the following six possible situations:

  1. (i)

    \(r_1<r_2<{\tilde{R}}_1<{\tilde{R}}_2\). We show that this case will not happen. By the property of h(R), it is not difficult to verify that \(h\left( \left( \frac{\alpha -1}{\alpha }\right) ^{2}\right) >0=h(r_2)\), so we know that \(\left( \frac{\alpha -1}{\alpha }\right) ^{2}<r_2\). Because \(\left( \frac{\alpha -1}{\alpha }\right) ^{2}\) is the maximum point of \({\tilde{g}}(R)\) and \({\tilde{R}}_1\) is the smallest root of \({\tilde{g}}(R)\), so we have \({\tilde{R}}_1<r_2\) which is a contradiction to the assumption.

  2. (ii)

    \(r_1<{\tilde{R}}_1<r_2<{\tilde{R}}_2\). By the fact that \({\tilde{g}}(R)>0\) for \(R\in ({\tilde{R}}_1,{\tilde{R}}_1)\) and \(g(R)<0\) for \(R\in (0,{\tilde{R}}_1)\cup ({\tilde{R}}_2,1)\), it is easy to obtain that \({\tilde{g}}(r_1)<0\) since \(r_1<{\tilde{R}}_1\), which is equivalent to \(k_1<1/g(r_1)\). Also, by \({\tilde{g}}(r_2)>0\), we have \(k_1>1/g(r_2)\). However, it has been proved that \(g(r_1)>g(r_2)\), so the set \((1/g(r_2),1/g(r_1))\) is empty, which means that this case cannot happen.

  3. (iii)

    \({\tilde{R}}_1<r_1<r_2<{\tilde{R}}_2\) (see Fig. 7c). Because that \(r_1,~r_2\in ({\tilde{R}}_1,{\tilde{R}}_2)\), so we have \({\tilde{g}}(r_1)>0\) and \({\tilde{g}}(r_2)>0\), which is equivalent to \(k_1>1/g(r_2)\). In this case, by Theorem 3.3, we know that Hopf bifurcations occur at both of \((R_1,R_1)\) and \((R_3,R_3)\).

  4. (iv)

    \({\tilde{R}}_1<r_1<{\tilde{R}}_2<r_2\) (see Fig. 7d). By similar argument, since \(r_1\in ({\tilde{R}}_1,{\tilde{R}}_2)\) and \(r_2>{\tilde{R}}_2\), we can obtain that \({\tilde{g}}(r_1)>0\) and \({\tilde{g}}(r_2)<0\) which imply that \(k_1\in (1/g(r_1),1/g(r_2))\). In this case, from \({\tilde{R}}_1<r_1<{\tilde{R}}_2<r_2\) and Theorem 3.3, a Hopf bifurcation only occurs at \((R_3,R_3)\) and does not occur at \((R_1,R_1)\).

  5. (v)

    \(r_1<{\tilde{R}}_1<{\tilde{R}}_2<r_2\) (see Fig. 7e). Similarly, we have \({\tilde{g}}(r_1)<0\) and \({\tilde{g}}(r_2)<0\), then it can be inferred that \(k_{11}<k_1<1/g(r_1)\). In this case, no Hopf bifurcation can occur. Also, we have \(\left( \frac{\alpha -1}{\alpha }\right) ^{2}>r_1\) in this case, which will be used later.

  6. (vi)

    \({\tilde{R}}_1<{\tilde{R}}_2<r_1<r_2\) (see Fig. 7f). In this case, we still have \(k_{11}<k_1<1/g(r_1)\), but the difference with case (v) is that \(\left( \frac{\alpha -1}{\alpha }\right) ^{2}<r_1\). In this case, two Hopf bifurcations occur at \((R_1,R_1)\).

So in order to distinguish the last two cases, we define \({\tilde{k}}_3\) to be the value of \(k_3\) such that \(r_1=\left( \frac{\alpha -1}{\alpha }\right) ^{2}\), and it is easy to calculate that \({\tilde{k}}_3\) is given by (3.16). So case (v) is for \(0<k_3<{\tilde{k}}_3\) which is equivalent to \(r_1<\left( \frac{\alpha -1}{\alpha }\right) ^{2}\) and case (vi) is for \({\tilde{k}}_3<k_3<k_{31}\) which implies \(r_1>\left( \frac{\alpha -1}{\alpha }\right) ^{2}\). Also we must have that \({\tilde{k}}_3<k_{31}\). Suppose not, then first we assume that \(k_3>k_{32}\), then \(h(R)=0\) has two positive solutions, but both of them should be bigger than 1 and here we have \(0<R<1\) which is a contradiction. If \(k_{31}<k_3<k_{32}\), then \(h(R)=0\) has no roots, so it contradicts with the fact that \(h(R)=0\) has one of positive roots at \((\dfrac{\alpha -1}{\alpha })^2<1\) when \(k_3={\tilde{k}}_3\). Therefore, we can conclude that \({\tilde{k}}_3<k_{31}\). Finally if \(0<k_1<k_{11}\), then \({\tilde{g}}(R)\) has no zeros, \((R_1,R_1)\) and \((R_3,R_3)\) are both always linearly stable and Hopf bifurcation will not occur, which is similar to (v) above.

In summary the case (c) is implied by (iii) above, case (d) is implied by (iv) above, case (e) is implied by (v) and the case of \(0<k_1<k_{11}\), and case (f) is implied by (vi) above. The proof is completed. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tian, C., Shi, Q., Cui, X. et al. Spatiotemporal dynamics of a reaction-diffusion model of pollen tube tip growth. J. Math. Biol. 79, 1319–1355 (2019). https://doi.org/10.1007/s00285-019-01396-7

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-019-01396-7

Keywords

Mathematics Subject Classification

Navigation