Skip to main content

Advertisement

Log in

Stochastic models for plant microtubule self-organization and structure

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

Abstract

One of the key enablers of shape and growth in plant cells is the cortical microtubule (CMT) system, which is a polymer array that forms an appropriately-structured scaffolding in each cell. Plant biologists have shown that stochastic dynamics and simple rules of interactions between CMTs can lead to a coaligned CMT array structure. However, the mechanisms and conditions that cause CMT arrays to become organized are not well understood. It is prohibitively time-consuming to use actual plants to study the effect of various genetic mutations and environmental conditions on CMT self-organization. In fact, even computer simulations with multiple replications are not fast enough due to the spatio-temporal complexity of the system. To redress this shortcoming, we develop analytical models and methods for expeditiously computing CMT system metrics that are related to self-organization and array structure. In particular, we formulate a mean-field model to derive sufficient conditions for the organization to occur. We show that growth-prone dynamics itself is sufficient to lead to organization in presence of interactions in the system. In addition, for such systems, we develop predictive methods for estimation of system metrics such as expected average length and number of CMTs over time, using a stochastic fluid-flow model, transient analysis, and approximation algorithms tailored to our problem. We illustrate the effectiveness of our approach through numerical test instances and discuss biological insights.

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
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

References

  • Allard JF, Wasteneys GO, Cytrynbaum EN (2010) Mechanisms of self-organization of cortical microtubules in plants revealed by computational simulations. Mol Biol Cell 21(2):278–286

    Article  Google Scholar 

  • Barton DA, Vantard M, Overall RL (2008) Analysis of cortical arrays from tradescantia virginiana at high resolution reveals discrete microtubule subpopulations and demonstrates that confocal images of arrays can be misleading. Plant Cell 20:982–994

    Article  Google Scholar 

  • Baulin VA, Marques CM, Thalmann F (2007) Collision induced spatial organization of microtubules. Biophys Chem 128(2–3):231–244

    Article  Google Scholar 

  • Burk DH, Ye ZH (2002) Alteration of oriented deposition of cellulose microfibrils by mutation of a Katanin-like microtubule-severing protein. Plant Cell 14(9):2145–2160

    Article  Google Scholar 

  • Burk DH, Liu B, Zhong R, Morrison WH, Ye ZH (2001) A Katanin-like protein regulates normal cell wall biosynthesis and cell elongation. Plant Cell 13(4):807–827

    Article  Google Scholar 

  • Deinum EE, Tindemans SH, Mulder BM (2011) Taking directions: the role of microtubule-bound nucleation in the self-organization of the plant cortical array. Phys Biol 8(5):056,002

    Article  Google Scholar 

  • Dixit R, Cyr R (2004a) Encounters between dynamic cortical microtubules promote ordering of the cortical array through angle-dependent modifications of microtubule behavior. Plant Cell 16(12):3274–3284

    Article  Google Scholar 

  • Dixit R, Cyr R (2004b) The cortical microtubule array: from dynamics to organization. Plant Cell 16(10):2546–2552

    Article  Google Scholar 

  • Dixit R, Chang E, Cyr R (2006) Establishment of polarity during organization of the acentrosomal plant cortical microtubule array. Mol Biol Cell 17(3):1298–1305

    Article  Google Scholar 

  • Dogterom M, Leibler S (1993) Physical aspects of the growth and regulation of microtubule structures. Phys Rev Lett 70(9):1347–1350

    Article  Google Scholar 

  • Ehrhardt DW, Shaw SL (2006) Microtubule dynamics and organization in the plant cortical array. Annu Rev Plant Biol 57:859–875

    Article  Google Scholar 

  • Eren EC (2012) Stochastic modeling and analysis of plant microtubule system characteristics. PhD thesis, Texas A&M University, College Station, TX

  • Eren EC, Dixit R, Gautam N (2010) A three-dimensional computer simulation model reveals the mechanisms for self-organization of plant cortical microtubules into oblique arrays. Mol Biol Cell 21(15):2674–2684

    Article  Google Scholar 

  • Eren EC, Gautam N, Dixit R (2012) Computer simulation and mathematical models of the noncentrosomal plant cortical microtubule cytoskeleton. Cytoskeleton. doi:10.1002/cm.21009

  • Goel A, Vogel V (2008) Harnessing biological motors to engineer systems for nanoscale transport and assembly. Nat Nanotechnol 3(8):465–475

    Article  Google Scholar 

  • Gray RM (1990) Entropy and Information Theory. Springer, New York

    Book  MATH  Google Scholar 

  • Hamant O, Heisler MG, Jonsson H, Krupinski P, Uyttewaal M, Bokov P, Corson F, Sahlin P, Boudaoud A, Meyerowitz EM et al (2008) Developmental patterning by mechanical signals in Arabidopsis. Science 322(5908):1650–1655

    Article  Google Scholar 

  • Hardham AR, Gunning BE (1978) Structure of cortical microtubule arrays in plant cells. J Cell Biol 77:14–34

    Article  Google Scholar 

  • Hawkins RJ, Tindemans SH, Mulder BM (2010) Model for the orientational ordering of the plant microtubule cortical array. Phys Rev E 82(1 Pt 1):011,911

    Article  Google Scholar 

  • Ishida T, Kaneko Y, Iwano M, Hashimoto T (2007) Helical microtubule arrays in a collection of twisting tubulin mutants of Arabidopsis thaliana. Proc Natl Acad Sci USA 104(20):8544–8549

    Article  Google Scholar 

  • Kawamura E, Wasteneys GO (2008) MOR1, the Arabidopsis thaliana homologue of xenopus MAP215, promotes rapid growth and shrinkage, and suppresses the pausing of microtubules in vivo. J Cell Sci 121(24):4114–4123

    Article  Google Scholar 

  • Khalil HK (2002) Nonlinear Systems. Prentice Hall, Upper Saddle River

    MATH  Google Scholar 

  • Kharoufeh JP, Gautam N (2004) Deriving link travel-time distributions via stochastic speed processes. Transp Sci 38(1):97–106

    Article  Google Scholar 

  • Long EF, Vaidya NK, Brandeau ML (2008) Controlling co-epidemics: analysis of HIV and tuberclosis infection dynamics. Oper Res 56(6):1366–1381

    Article  MATH  MathSciNet  Google Scholar 

  • Lu JL, Valois F, Dohler M, Barthel D (2008) Quantifying organization by means of entropy. IEEE Commun Lett 12(3):185–187

    Article  Google Scholar 

  • Martin MT, Plastinoa A, Rossob OA (2006) Generalized statistical complexity measures: geometrical and analytical properties. Physica A 369(2):439–462

    Article  Google Scholar 

  • Narayanan A, Kulkarni VG (1996) First passage times in fluid models with an application to two priority fluid systems. In: Proceedings of the 2nd International Computer Performance and Dependability Symposium. IEEE, Washington, DC, pp 166–175

  • Shannon CE (1948) A mathematical theory of communication. Bell Syst Tech J 27:379–423

    Article  MATH  MathSciNet  Google Scholar 

  • Shaw SL, Kamyar R, Ehrhardt DW (2003) Sustained microtubule treadmilling in Arabidopsis cortical arrays. Science 300(5626):1715–1718

    Article  Google Scholar 

  • Shi X, Ma Y (2010) Understanding phase behavior of plant cell cortex microtubule organization. Proc Natl Acad Sci 107(26):11,709–11,714

    Article  Google Scholar 

  • Tindemans SH, Hawkins RJ, Mulder BM (2010) Survival of the aligned: ordering of the plant cortical microtubule array. Phys Rev Lett 104(5):058,103

    Article  Google Scholar 

  • Vos JW, Dogterom M, Emons AMC (2004) Microtubules become more dynamic but not shorter during preprophase band formation: a possible, “search-and-capture” mechanism for microtubule translocation. Cell Motil Cytoskelet 57:246–258

    Article  Google Scholar 

  • Wasteneys GO, Ambrose JC (2009) Spatial organization of plant cortical microtubules: close encounters of the 2D kind. Trends Cell Biol 19(2):62–71

    Article  Google Scholar 

  • Wolff RW (1989) Stochastic Modeling and the Theory of Queues. Prentice Hall, New Jersey

    MATH  Google Scholar 

  • Zhang Q, Fishel EA, Bertroche T, Dixit R (2013) Microtubule severing at crossover sites by Katanin generates ordered cortical microtubule arrays in Arabidopsis. Curr Biol 23:2191–2195

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ezgi C. Eren.

Appendices

Appendix 1: Input parameter sets (dynamics)

$$\begin{aligned} Q&= [q_{m,n}], m,n \in \{GS,GP,SS,SP,PS,PP\} \\ V&= diag(v^m), m \in \{GS,GP,SS,SP,PS,PP\} \end{aligned}$$

Parameter Set I:

$$\begin{aligned}&Q = \begin{bmatrix} -8.925&\quad 6.72&\quad 1.485&\quad 0&\quad 0.72&\quad 0 \\ 2.427&\quad -4.632&\quad 0&\quad 1.485&\quad 0&\quad 0.72 \\ 3.537&\quad 0&\quad -11.192&\quad 6.72&\quad 0.935&\quad 0 \\ 0&\quad 3.537&\quad 2.427&\quad -6.898&\quad 0&\quad 0.935 \\ 5.05&\quad 0&\quad 2.376&\quad 0&\quad -14.1&\quad 6.72 \\ 0&\quad 5.05&\quad 0&\quad 2.376&\quad 2.427&\quad -9.85 \end{bmatrix}. \\&V = diag(0.91,3.69,-8.66,-5.88,-2.78,0) \end{aligned}$$

Parameter Set II:

$$\begin{aligned}&Q=\begin{bmatrix} -7.234&\quad 6.72&\quad 0.236&\quad 0&\quad 0.278&\quad 0 \\ 2.427&\quad -2.941&\quad 0&\quad 0.236&\quad 0&\quad 0.278 \\ 5&\quad 0&\quad -13.27&\quad 6.72&\quad 1.55&\quad 0 \\ 0&\quad 5&\quad 2.427&\quad -8.977&\quad 0&\quad 1.55 \\ 25.125&\quad 0&\quad 12.75&\quad 0&\quad -44.595&\quad 6.72 \\ 0&\quad 25.125&\quad 0&\quad 12.75&\quad 2.427&\quad -40.302 \end{bmatrix} \\&V=diag(0.72,3.5,-11.78,-9,-2.78,0) \end{aligned}$$

Parameter Set III:

$$\begin{aligned}&Q=\begin{bmatrix} -9.617&\quad 6.72&\quad 2.338&\quad 0&\quad 0.559&\quad 0 \\ 2.427&\quad -5.324&\quad 0&\quad 2.338&\quad 0&\quad 0.559 \\ 12.438&\quad 0&\quad -21.908&\quad 6.72&\quad 2.75&\quad 0 \\ 0&\quad 12.438&\quad 2.427&\quad -17.614&\quad 0&\quad 2.75 \\ 8.75&\quad 0&\quad 4.375&\quad 0&\quad -19.845&\quad 6.72 \\ 0&\quad 8.75&\quad 0&\quad 4.375&\quad 2.427&\quad -15.552 \end{bmatrix} \\&V=diag(3.72,6.5,-15.18,-12.4,-2.78,0) \end{aligned}$$

Parameter Set IV:

$$\begin{aligned}&Q=\begin{bmatrix} -7.537&\quad 6.72&\quad 0.535&\quad 0&\quad 0.282&\quad 0\\ 2.427&\quad -3.244&\quad 0&\quad 0.53521&\quad 0&\quad 0.282\\ 6.211&\quad 0&\quad -16.036&\quad 6.72&\quad 3.105&\quad 0\\ 0&\quad 6.211&\quad 2.427&\quad -11.742&\quad 0&\quad 3.105\\ 15.6&\quad 0&\quad 5.6&\quad 0&\quad -27.92&\quad 6.72\\ 0&\quad 15.6&\quad 0&\quad 5.6&\quad 2.427&\quad -23.627 \end{bmatrix} \\&V=diag(-0.28,2.5,-8.98,-6.2,-2.78,0) \end{aligned}$$

Parameter Set V:

$$\begin{aligned}&Q=\begin{bmatrix} -11.806&\quad 6.72&\quad 2.343&\quad 0&\quad 2.743&\quad 0\\ 2.427&\quad -7.512&\quad 0&\quad 2.343&\quad 0&\quad 2.743\\ 3.05&\quad 0&\quad -15.82&\quad 6.72&\quad 6.05&\quad 0\\ 0&\quad 3.05&\quad 2.427&\quad -11.527&\quad 0&\quad 6.05\\ 1.556&\quad 0&\quad 1.378&\quad 0&\quad -9.653&\quad 6.72\\ 0&\quad 1.556&\quad 0&\quad 1.378&\quad 2.427&\quad -5.36 \end{bmatrix} \\&V=diag(-0.78,2,-6.58,-3.8,-2.78,0) \end{aligned}$$

Parameter Set VI:

$$\begin{aligned}&Q=\begin{bmatrix} -8.925&\quad 6.72&\quad 1.485&\quad 0&\quad 0.72&\quad 0 \\ 2.427&\quad -4.632&\quad 0&\quad 1.485&\quad 0&\quad 0.72 \\ 3.537&\quad 0&\quad -11.192&\quad 6.72&\quad 0.935&\quad 0 \\ 0&\quad 3.537&\quad 2.427&\quad -6.898&\quad 0&\quad 0.935 \\ 5.05&\quad 0&\quad 2.376&\quad 0&\quad -14.1&\quad 6.72 \\ 0&\quad 5.05&\quad 0&\quad 2.376&\quad 2.427&\quad -9.85 \end{bmatrix}. \\&V=diag(3.69,3.69,-5.88,-5.88,0,0) \end{aligned}$$

Appendix 2: Proofs

Lemma 1

Let us define \(z_i=(y_i-x_i)\) \(i=1,\ldots ,N\). We group \(z_i\) values in three sets as follows:

$$\begin{aligned} \begin{aligned} I_+&=\{i \in \{1,N\}: z_i>0\}\\ I_0&=\{i \in \{1,N\}: z_i=0\}\\ I_-&=\{i \in \{1,N\}: z_i<0\}. \end{aligned} \end{aligned}$$

As \(\sum _{i=1}^N y_i-\sum _{i=1}^N x_i=0\), it follows that

$$\begin{aligned} \sum _{i=1}^N z_i=\sum _{i\in I_+} z_i + \sum _{i\in I_-} z_i=0. \end{aligned}$$
(23)

Let us divide \(z_i\) values into infinitesimal pieces of the same size, denoted by \(\varDelta _z>0\), such that for each \(i\in I_+\), \(z_i=w_i \varDelta _z\) and for each \(i\in I_-\), \(z_i=-w_i \varDelta _z\), \(i=1,\ldots ,N\), where \(\varDelta _z>0\) and \(w_i\) are positive real numbers. Equation (23) can be rewritten as

$$\begin{aligned} \sum _{i\in I_+} w_i - \sum _{i\in I_-} w_i=0. \end{aligned}$$
(24)

Hence, we have an equal number (\(\sum _{i\in I_+} w_i =\sum _{i\in I_-} w_i :=W\)) of \(\varDelta _z\) pieces that belong to sets \(I_+\) and \(I_-\). Note that \(\forall \) \(z_i\) with \(i \in I_+\) and \(z_j\) with \(j \in I_-\), it follows from the definition of \(z_i\) values that \(z_i>z_j\), and consequently \(x_i>x_j\) from the given ordering relation between the sequences \(x_i\), and \(z_i\); and finally \(f(x_i)<f(x_j)\) due to the decreasing property of \(f(\cdot )\). Let us redefine the sequence of \(x_i\), \(i=1,\ldots ,N\) such that its values are copied \(w_i\) times for each \(x_i\) value so that every \(\varDelta _z\) has its corresponding \(x'_{i'}\) and \(f(x'_{i'})\) values, where \(i'=1,\ldots ,W\). Adjusting \(I_+\) and \(I_-\) accordingly as \(I'_+\), \(I'_-\), for each \(\varDelta _z\) value that belongs to set \(I_+\), there is a \(\varDelta _z\) which is multiplied with a larger value in \(I_-\) in the following equation:

$$\begin{aligned} \sum _{i'\in I'_+} f(x'_{i'}) \varDelta _z - \sum _{i'\in I'_-} f(x'_{i'}) \varDelta _z<0, \end{aligned}$$

which gives the desired result.

Proposition 1

According to Lyapunov’s stability theory, a sufficient condition for the global asymptotic stability of an equilibrium point \(P^*\) is existence of a Lyapunov function \(\mathcal {L}(\cdot )\) such that

  • \(\mathcal {L}(t)\mid _P>0\), \(\forall P\ne P^*\) and \(\mathcal {L}(t)\mid _P=0\) only for \(P=P^*\),

  • \(\frac{\partial \mathcal {L}(t)}{\partial t}\mid _P<0\), \(\forall P\ne P^*\) and \(\frac{\partial \mathcal {L}(t)}{\partial t}\mid _{P=P^*}=0\).

We set our Lyapunov function as the entropy metric, \(H(\cdot )\) in Eq. (2). We already know that the first condition holds for \(H(\cdot )\), as \(H(t)\mid _{P=P^*}=0\) and \(H(t)\mid _P>0\) \(\forall P\ne P^*\). What is left to check is the sign of the derivative of the Lyapunov function with respect to \(t\), which is given by

$$\begin{aligned} \frac{\partial H(t)}{\partial t}=-\sum _{\theta =0}^{179} k'(\theta ,t) ln (k(\theta ,t))+k'(\theta ,t), \end{aligned}$$
(25)

where

$$\begin{aligned} k'(\theta ,t)=\frac{\partial \,\, k(\theta ,t)}{\partial t}. \end{aligned}$$
(26)

Let us define the total density of CMTs with length \(l\) and angle \(\theta \) at time \(t\) by

$$\begin{aligned} \bar{p}(l,\theta ,t):=\bar{p}_G(l,\theta ,t)+\bar{p}_S(l,\theta ,t)+\bar{p}_P(l,\theta ,t). \end{aligned}$$

Using Eq. (10), we can rewrite Eq. (26) as

$$\begin{aligned} k'(\theta ,t)=\frac{\left( \int _0^\infty l \frac{d \bar{p}(l,\theta ,t)}{dt}dl\right) \sum _{\theta =0}^{179}\int _0^\infty l \bar{p}(l,\theta ,t)dl\!-\!\left( \int _0^\infty l \bar{p}(l,\theta ,t)dl\right) \sum _{\theta =0}^{179}\int _0^\infty l \frac{d \bar{p}(l,\theta ,t)}{dt}dl}{ \left( \sum _{\theta =0}^{179}\int _0^\infty l \bar{p}(l,\theta ,t)dl\right) ^2}. \end{aligned}$$

Summing equations in (5) side by side, multiplying both sides with \(l\) and integrating with respect to \(l\) over \((0,\infty )\), we obtain the derivative of total length of all CMTs with angle \(\theta \) at time \(t\) as

$$\begin{aligned} L'_\theta (t) :&= \int _0^\infty l \frac{d \bar{p}(l,\theta ,t)}{dt}dl\\&= (v_G^+-v_S^-) \tilde{p}_{GS}(\theta ,t)+v_G^+ \tilde{p}_{GP}(\theta ,t)\\&\quad -(v_S^++v_S^-) \tilde{p}_{SS}(\theta ,t)-v_S^+\tilde{p}_{SP}(\theta ,t)-v_S^- \tilde{p}_{PS}(\theta ,t)\\&\quad -v_G^+\int _0^\infty \bar{p}_G(l,\theta ,t) l dl\sum _{\theta ' \in \varTheta ^*} p_b|sin(\theta -\theta ')|\int _{l'} dl' l' \bar{p}(l',\theta ',t)\\&\quad +v_G^+ \int _{l'} dl' l' \bar{p}(l',\theta ,t)\sum _{\theta ' \in \varTheta ^*} p_b|sin(\theta -\theta ')\int _0^\infty \bar{p}_G(l,\theta ',t) l dl, \end{aligned}$$

where \(\tilde{p}_{m}(\theta ,t):= \int _0^{\infty } p_m(l,\theta ,t) dl\), \(m\) \(\in \) \(\{GS,GP,SS,SP,PS,PP\}\) stands for the total density of CMTs with angle \(\theta \) at state \(m\) at time \(t\). We denote the sum of \(L'_\theta (t)\) over all \(\theta \) by \( \sum L'(t):= \sum _{\theta =0}^{179} L'_\theta (t)\), which gives the derivative of total length of all CMTs in the system at time \(t\). Defining total length of CMTs with angle \(\theta \) at time \(t\) as \(L_\theta (t):=\int _0^\infty l p(l,\theta ,t)dl\), and the total length of all CMTs at time \(t\) as \(\sum L(t):=\sum _{\theta =1}^{360} L_\theta (t)\), and plugging these into Eq. (25) results in

$$\begin{aligned} \frac{dH(t)}{dt}=-\sum _{\theta =0}^{179} \frac{L'_\theta (t)\ln \left( \frac{L_\theta (t)}{\sum L(t)}\right) \sum L(t)-L_\theta (t)\ln \left( \frac{L_\theta (t)}{\sum L(t)}\right) \sum L'(t)}{(\sum L(t))^2}. \end{aligned}$$

Rearranging terms, we obtain

$$\begin{aligned} \frac{dH(t)}{dt}=-\sum _{\theta =0}^{179} \frac{\frac{L'_\theta (t)}{\sum L'(t)}\ln \left( \frac{L_\theta (t)}{\sum L(t)}\right) -\frac{L_\theta (t)}{\sum L(t)}\ln \left( \frac{L_\theta (t)}{\sum L(t)}\right) }{\sum L(t) \sum L'(t)}. \end{aligned}$$
(27)

Let us denote \(a_{\theta }:=\frac{L'_\theta (t)}{\sum L'(t)}\) and \(b_{\theta }:=\frac{L_\theta (t)}{\sum L(t)}\). By definition, it follows that \(\sum _{\theta =0}^{179} a_{\theta }=1\), \(\sum _{\theta =0}^{179} b_{\theta }=1\), and \(b_i>0\). Assuming \(\sum L'(t) >0\), i.e. the net total CMT length change in time is positive, we are interested in the sign of

$$\begin{aligned} \sum -\ln (b_{\theta }) (a_{\theta } -b_{\theta }). \end{aligned}$$
(28)

If the sign for Expression (28) is negative, then Eq. (27) is negative, i.e. \(\frac{dH}{dt} <0\), and the stability condition is satisfied. A sufficient condition to ensure this follows from Lemma 1 as \(-ln(b)\) is a decreasing function of \(b\) for \(0<b<1\). Accordingly, we require the two sequences \(a_\theta \) and \(b_\theta \), and their difference \(a_\theta -b_\theta \) to increase and decrease in the same order. This roughly means that if CMTs with an angle \(\theta \) have a larger total length compared to the total length of CMTs with angle \(\theta '\ne \theta \), they also grow larger in ratio in total length on average, and vice versa. This property follows by careful observation of model equations and the property that \(p_b(l)\) is decreasing in \(l\).

Finally, in order to fulfill \(\sum L'(t) >0\), it is required that the problem parameters satisfy

$$\begin{aligned}&\sum _\theta (v_G^+-v_S^-) p_{GS}(\theta ,t)+v_G^+ p_{GP}(\theta ,t)-(v_S^++v_S^-) p_{SS}(\theta ,t)-v_S^+p_{SP}(\theta ,t)\nonumber \\&\qquad -v_S^- p_{PS}(\theta ,t)>0, \end{aligned}$$
(29)

\(\forall t\), which can be stated as

$$\begin{aligned} \sum _\theta p^+_G(\theta ,t) v_{G^+}-p^+_S(\theta ,t) v_{S^+}- p^-_S(\theta ,t) v_{S^-}>0, \end{aligned}$$
(30)

by rearranging terms.

Theorem 1

Consider \( F_{ab}(l,t+h)\), where \(h\) is a small positive real number. It can be written as

$$\begin{aligned} F_{ab}(l,t+h)=P\{\tau \le t+h,M(\tau )=b|L(0)=l,M(0)=a\}. \end{aligned}$$

Conditioning on the first transition from the initial state, we obtain

$$\begin{aligned} F_{ab}(l,t+h)=&P\{\tau \le t+h,M(\tau )=b|L(0)=l,M(0)=a\}\\ =&\sum _{c \ne a} P\{\tau \le t+h,M(\tau )=b|L(0)=l,M(0)=a,M(h)=c\}\\&\qquad P\{M(h)=c|M(0)=a,L(0)=l\}\\&+P\{\tau \le t+h,M(\tau )=b|L(0)=l,M(0)=a,M(h)=a\}\\&\qquad P\{M(h)=a|M(0)=a,L(0)=l\}. \end{aligned}$$

As \(M(t)\) process is independent of \(L(0)\), and the length would change by \(v^a h\) by time \(h\), when the CMT is in state \(a\) at time 0,

$$\begin{aligned}&F_{ab}(l,t+h)\\&\quad = \sum _{c \ne a} P\{\tau \le t+h,M(\tau )=b|L(h)=l+v^a h,M(h)\!=\!c\}P\{M(h)\!=\!c|M(0)\!=\!a\}\\&\qquad +P\{\tau \le t+h,M(\tau )=b|L(h)=l+v^a h, M(h)=a\}P\{M(h)=a|M(0)=a\}\\&\quad = \sum _{c \ne a} P\{\tau \le t,M(\tau )=b|L(0)=l+v^a h,M(0)=c\}P\{M(h)=c|M(0)=a\}\\&\qquad +P\{\tau \le t,M(\tau )=b|L(0)=l+v^a h, M(0)=a\}P\{M(h)=a|M(0)=a\}. \end{aligned}$$

As the transition probability from state \(a\) to \(c\) in time \(h\) is given by \(q_{ac} h+o(h)\) if \(c \ne a\) and \(1+q_{aa} h+o(h)\) if \(c = a\), where \(o(h)\) is a collection of terms of higher order than \(h\) such that \(o(h)/h \rightarrow 0\) as \(h \rightarrow 0\), it follows

$$\begin{aligned} F_{ab}(l,t+h)=\sum _{c \ne a} F_{cb}(l+v^a h,t) q_{ac} h +F_{ab}(l+v^a h,t) (q_{aa} h+1)+o(h). \end{aligned}$$

Subtracting \(F_{ab}(l,t)\) from each side of the equation, dividing by \(h\) and rearranging terms results in

$$\begin{aligned}&\frac{F_{ab}(l,t+h)-F_{ab}(l,t)}{h}\\&\quad =\frac{F_{ab}(l+v^a h,t)-F_{ab}(l,t)}{h}+\sum _c q_{ac}F_{cb}(l+v^a h,t)+o(h)/h. \end{aligned}$$

Letting \(h\rightarrow 0\) yields Eq. (15), and rewriting in the matrix form gives Eq. (16). Next, we describe the boundary conditions for all \(a,b\) and \(t\). As the lifetime would be zero if CMT appeared with zero length at state \(b\) such that \(v^b<0\), it follows

$$\begin{aligned} F_{bb}(0,t) = 1 \qquad \hbox {for} \qquad v^b<0. \end{aligned}$$

The second boundary condition,

$$\begin{aligned} F_{ab}(0,t) = 0 \qquad \hbox {for} \qquad a\ne b, v^a<0, \end{aligned}$$

follows from the fact that although the lifetime is zero, the probability that the state is \(b\) when the lifetime is reached is zero (since at time \(t=0\) the state is \(a\) with \(v^a<0\)). Finally, the last two conditions follow from the fact that lifetime cannot be reached at state \(b\) at time \(t=0\) if the initial state is \(a \ne b\) for any initial length; or if the initial state is \(b\) for a positive initial length.

Theorem 2

Taking the LT of Eq. (16) with respect to \(t\) gives

$$\begin{aligned} (w I-Q) F_b^*(l,w)= V \frac{\delta F_b^*(l,w)}{\delta l}. \end{aligned}$$
(31)

Taking the LT of Eq. (31) with respect to \(l\) results in

$$\begin{aligned} (w I-Q) F_b^{**}(s,w)= V [sF_b^{**}(s,w)-F^*_b(0,w)]. \end{aligned}$$
(32)

Define \(e_j\) as the \(j\mathrm{th}\) unit vector. Plugging in the boundary condition

$$\begin{aligned} F_b^*(0,w)=w^{-1}e_j \qquad if \qquad v^j<0, \end{aligned}$$

we get

$$\begin{aligned} (V s- w I+Q) \tilde{F}_b^*(s,w)= w^{-1}(V e_j). \end{aligned}$$

Rearranging terms yields Eq. (17).

Appendix 3: Algorithms for estimation of system metrics

figure a
figure b

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Eren, E.C., Dixit, R. & Gautam, N. Stochastic models for plant microtubule self-organization and structure. J. Math. Biol. 71, 1353–1385 (2015). https://doi.org/10.1007/s00285-015-0860-9

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-015-0860-9

Keywords

Navigation