This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

11institutetext: S. D. Ryan 22institutetext: Department of Mathematics and Statistics, Cleveland State University, Cleveland, OH 44115, USA
Center for Applied Data Analysis and Modeling, Cleveland State University, Cleveland, OH 44115, USA
22email: [email protected]
33institutetext: Z. McCarthy 44institutetext: Department of Mathematics and Statistics, York University, Toronto, ON, Canada
Laboratory for Industrial and Applied Mathematics, Toronto, ON, Canada
Centre for Disease Modelling, York University, Toronto, ON, Canada
Fields-CQAM Mathematics for Public Health Laboratory, Toronto, ON, Canada
55institutetext: M. Potomkin 66institutetext: Department of Mathematics, University of California, Riverside, CA, 92521, USA
Corresponding Author 66email: [email protected]

Motor Protein Transport Along Inhomogeneous Microtubules

S. D. Ryan    Z. McCarthy    M. Potomkin
Abstract

Many cellular processes rely on the cell’s ability to transport material to and from the nucleus. Networks consisting of many microtubules and actin filaments are key to this transport. Recently, the inhibition of intracellular transport has been implicated in neurodegenerative diseases such as Alzheimer’s disease and Amyotrophic Lateral Sclerosis (ALS). Furthermore, microtubules may contain so-called defective regions where motor protein velocity is reduced due to accumulation of other motors and microtubule associated proteins. In this work, we propose a new mathematical model describing the motion of motor proteins on microtubules which incorporate a defective region. We take a mean-field approach derived from a first principle lattice model to study motor protein dynamics and density profiles. In particular, given a set of model parameters we obtain a closed-form expression for the equilibrium density profile along a given microtubule. We then verify the analytic results using mathematical analysis on the discrete model and Monte Carlo simulations. This work will contribute to the fundamental understanding of inhomogeneous microtubules providing insight into microscopic interactions that may result in the onset of neurodegenerative diseases. Our results for inhomogeneous microtubules are consistent with prior work studying the homogeneous case.

Keywords:
Mathematical Biology, Motor Proteins, Microtubules, Phase Transitions, Defective Transport
MSC:
34F05, 35Q92, 92B05

1 Introduction

Cells strongly rely on the ability to efficiently transport cargo via motor proteins. Fast active transport (FAT) is required to deliver materials such as proteins, mRNA, mitochrondria, vescicles, and organelles for use in a variety of cellular processes Ros08 ; Lakadamyali2014 . One of the three major means of intracellular transport is within networks of microtubules (MTs) Lakadamyali2014 . To accommodate FAT, motor proteins “walk” along MTs and actin filaments (AFs) with their cargo forming a “superhighway” for cellular transportation Klu05b ; Ros14 ; Str20 . Motor-protein families kinesin and dynein bind to the cellular material or cargo they carry as they move up and down the microtubule Gro07 ; Lakadamyali2014 . For a more general overview of molecular motor protein motion see review Mal04 .

Defects in microtubules are known to exist, but the current literature has yet to clarify their impact on molecular motor-based transport Lia16 . Defects in active transport, particularly axonal, have been implicated in the progression of various diseases Lakadamyali2014 . For instance, the defining characteristics of many neurodegenerative diseases such as Alzheimer’s disease and Amyotrophic Lateral Sclerosis (ALS) may be related to deficiencies in active transport within neurons Lakadamyali2014 . One such area of need for immediate study is the scenario where the microtubule paths used by motor proteins become congested, obstructed, or defective. Hallmarks and early indicators of neurodegenerative diseases are an accumulation of organelles and proteins in the cell body or axon, which inhibits active transport Lakadamyali2014 . Hence, understanding the nature of motor protein dynamics will provide insight in understanding the onset of these diseases and developing control strategies.

Advances in biophysical tools and imaging technology have allowed for many recent insightful in vitro experiments of motor protein behavior on microtubules Ros08 ; Lakadamyali2014 ; Lia16 . Motor-proteins may change directions, stop or pause briefly, increase their velocity, and also attach/detach from the microtubule Ros08 . Experimentally it is observed that this behavior may be attributed to the presence of MT-associated proteins Ros08 . Therefore, the cell’s ability to regulate active transport may be studied through tau MAP regions where motion of motor proteins is inhibited. We refer to these patches of high tau MAP concentration as defective regions due to their effects on the reduction of motor motility. We focus our study to modeling collective motor protein motility on MTs, first homogeneous and then defective.

Modeling motor protein motility on MTs has received recent theoretical attention. A one-dimensional discrete lattice model was studied in detail using a mean-field approach, including a full phase diagram for the stationary states ParFraFre2003 ; ParFraFre2004 ; Fre11 ; Klu08 ; Klu05 . The model was capable of predicting the emergence of interior layers by splitting the equilibrium density of motor proteins along the MT into two phases: low and high density. Such a co-existence corresponds to a traffic jam consisting of motor proteins translating along the MT in one direction, from the region with low density to the one with high density. In addition, a generalization of this lattice gas model has also been studied to account for local interactions between motors Fre18 . In particular, the effects of adjacent motors enhancing the detachment rate as well as motor crowding enhancing the frequency which motors become inactive or paused Fre18 . In a similar light, models featuring multiple “lanes” on a MT and motor proteins may switch lanes have also been studied Fre07 or between a filament and a tube in Mul05 . The common feature among these lattice gas models is they follow the totally asymmetric simple exclusion process (TASEP) framework ParFraFre2003 . These models also feature attachment and detachment processes for motors whose stationary states are described by the theory of Langmuir dynamics ParFraFre2003 .

In this paper, we seek to advance the current understanding of subcellular transport along microtubules with a defective region. In Section 2, we introduce a new discrete model for motor protein motility on a microtubule with a one-dimensional lattice following the TASEP paradigm and attachment/detachment dynamics. The distinctive feature of this model is that the MT consists of a defective region with decreased motor protein motility rate. From a discrete lattice formulation, we take the limit to recover the mean-field form of these equations. Next, in Section 2.3, we verify that our mean-field approximations produce results consistent with existing studies (e.g., ParFraFre2003 ; ParFraFre2004 ; Fre11 ) before moving to the defective region case where motility will be hindered. The main analytical results are then presented in Section 2.4, where we consider examples with the domain split into two regions, fast and slow, whereas the local and boundary attachment/detachment rates are varied as parameters. From these studies we find a closed-form expression of the solution given a set of the model parameters for attachment, detachment, and boundary conditions. We then verify that the analytical solution of the mean-field model is consistent with corresponding Monte Carlo simulations of the original discrete lattice model in Section 3. We note that if one needs to consider a wide range of problem parameters, for example, in model calibration or a control problem (e.g., finding the location and width of a defective region for a desired motor distribution), Monte Carlo simulations are prohibitively time consuming as compared to a closed-form expression when available. In the Appendix, we provide a new analytical approach to the solution of the homogeneous problem, based on the analysis of the phase portrait of the corresponding system of ODEs; then we give examples of applications of the approach for specific problem parameters. Overall, this work provides a critical result for inhomogeneous MTs consisting of multiple parts with different motility properties; namely, it can be modeled as segments of homeogeneous MTs linked by a matching flux condition. This greatly expands the utility of past studies that developed the theory of homogeneous MTs.

2 Mathematical Model

2.1 Discrete problem

Following the general TASEP paradigm, we construct a discrete model from first principles which generalizes previous models for homogeneous MTs to incorporate a defective region. Briefly, the TASEP paradigm states: (i) each binding site may be occupied by a maximum of one motor; (ii) motors move unidirectionally on the lattice and (iii) motors enter the lattice on the left side and exit the lattice on the right side ParFraFre2003 . Here we also account for the attachment and detachment of motor proteins on the MT interior as in works similar in scope focused on modeling ParFraFre2003 ; ParFraFre2004 ; Fre11 and experiment Var09 . Another recent work has focused on stochastic modeling with the goal of revealing how motor protein and filament properties affect transport Dal19 .

Specifically, consider a one-dimensional lattice {0=x0<x1<<xM=}\left\{0=x_{0}<x_{1}<...<x_{M}=\ell\right\}, representing sites which a motor may occupy on a microtubule of length \ell. Introduce ρin\rho_{i}^{n} which is the probability of finding a motor at site xix_{i} at time step nn. Probabilities at each lattice site 0ρin10\leq\rho_{i}^{n}\leq 1 for 1i<M1\leq i<M change during one time step via

ρin+1ρin=vi12ρi1n(1ρin)vi+12ρin(1ρi+1n)+ωA(1ρin)ωDρin.\rho_{i}^{n+1}-\rho_{i}^{n}=v_{i-\frac{1}{2}}\rho_{i-1}^{n}(1-\rho_{i}^{n})-v_{i+\frac{1}{2}}\rho_{i}^{n}(1-\rho_{i+1}^{n})+\omega_{A}(1-\rho_{i}^{n})-\omega_{D}\rho_{i}^{n}. (1)

The first term on the right-hand side of (1) says that the probability of finding a motor at site xix_{i} increases due to a possible jump of a motor from site xi1x_{i-1} to xix_{i} provided that the following jump condition is satisfied: there is a motor at site xi1x_{i-1} and site xix_{i} is vacant. As it is done in previous works on one-dimensional transport of active motors along a microtubule ParFraFre2003 ; ParFraFre2004 ; Fre11 , consider the problem in the mean-field approximation, that is, correlations are negligibly small and the probability of the jump condition is simply ρi1n(1ρin)\rho_{i-1}^{n}(1-\rho_{i}^{n}). Additionally, the coefficient vi12v_{i-\frac{1}{2}} accounts for inhomogeneity of the microtubule: if the jump condition is satisfied on the interval [xi1,xi][x_{i-1},x_{i}], then the jump occurs with probability (motility rate) vi12v_{i-\frac{1}{2}}, and these coefficient may change from site to site. The second term on the right-hand side of (1) is similar to the first term, but accounts for the decrease in probability ρin\rho_{i}^{n} due to a possible jump from site xix_{i} to xi+1x_{i+1}. The third term on the right-hand side of (1) describes the interaction of the microtubule with the exterior environment: a motor from outside can attach to the microtubule at site xix_{i} as well as a motor already occupying site xix_{i} can detach from the microtubule. Parameters ωA\omega_{A} and ωD\omega_{D} are the corresponding attachment and detachment rates.

Stationary states of (1) solve the following system:

0=vi12ρi1(1ρi)vi+12ρi(1ρi+1)+ωA(1ρi)ωDρi.0=v_{i-\frac{1}{2}}\rho_{i-1}(1-\rho_{i})-v_{i+\frac{1}{2}}\rho_{i}(1-\rho_{i+1})+\omega_{A}(1-\rho_{i})-\omega_{D}\rho_{i}. (2)

This system is supplemented with boundary conditions corresponding to the attachment rate α\alpha at the left end and detachment rate β\beta at the right end of the microtubule:

ρ0=αandρM=1β.\rho_{0}=\alpha~{}~{}\text{and}~{}~{}\rho_{M}=1-\beta. (3)

We note here that it is assumed that boundary attachment rates have a corresponding relationship to those inside the microtubule, that is,

α=ωAωA+ωD and β=ωDωA+ωD,\alpha=\dfrac{\omega_{A}}{\omega_{A}+\omega_{D}}\text{ and }\beta=\dfrac{\omega_{D}}{\omega_{A}+\omega_{D}}, (4)

and vivv_{i}\equiv v, then the solution of (2)-(3) is simply a constant: ρωA/(ωA+ωD)\rho\equiv\omega_{A}/(\omega_{A}+\omega_{D}). However, in practice, rates α\alpha and β\beta are different from (4) which leads to non-trivial stationary solutions possessing interior jumps, even in the homogeneous case viconstv_{i}\equiv\text{const} ParFraFre2003 ; ParFraFre2004 .

We end this subsection with another form of (2) which is helpful for understanding the continuous limit presented in Section 2.2. Introduce the following notation for the flux between sites xi1x_{i-1} and xix_{i}:

Ji12=vi12ρi1(1ρi).J_{i-\frac{1}{2}}=v_{i-\frac{1}{2}}\rho_{i-1}(1-\rho_{i}). (5)

Note that if one interpolates ρi\rho_{i} by a smooth function ρ(x)\rho(x) such that ρ(xi)=ρi\rho(x_{i})=\rho_{i}, then performing Taylor expansions one can verify that

Δx2ρi12=vi121Ji12+ρi12(1ρi12)+o(Δx),\dfrac{\Delta x}{2}\rho^{\prime}_{i-\frac{1}{2}}=-v_{i-\frac{1}{2}}^{-1}J_{i-\frac{1}{2}}+\rho_{i-\frac{1}{2}}(1-\rho_{i-\frac{1}{2}})+o(\Delta x), (6)

where ρi12=ρ(xi12)\rho_{i-\frac{1}{2}}=\rho(x_{i-\frac{1}{2}}), Δx=xixi1\Delta x=x_{i}-x_{i-1}, and xi12=xi12Δxx_{i-\frac{1}{2}}=x_{i}-\frac{1}{2}\Delta x.

Going back to definition (5), we point out that, in terms of Ji±12J_{i\pm\frac{1}{2}}, equation (2) has the following form:

Ji+12Ji12=ωA(1ρi)ωDρiJ_{i+\frac{1}{2}}-J_{i-\frac{1}{2}}=\omega_{A}(1-\rho_{i})-\omega_{D}\rho_{i} (7)

2.2 Limiting continuous problem

We focus on the asymptotic behavior of solutions of (2)-(3) as MM\to\infty in the framework of the mean-field limit. Specifically, we introduce parameter ε:=M11\varepsilon:=\ell M^{-1}\ll 1 (here \ell represents the total length of microtubule) and lattice {xi=iε,i=0,..,M}\left\{x_{i}=i\varepsilon,\,i=0,..,M\right\} with the distance between lattice points ε\varepsilon. For small ε\varepsilon, we approximate the solution with the continuous function ρε(x)\rho_{\varepsilon}(x) defined on 0<x<0<x<\ell and derived from a discrete set of unknowns ρi\rho_{i} associated with the lattice points xix_{i}, i.e., ρε(xi)=ρi\rho_{\varepsilon}(x_{i})=\rho_{i}. Then for ε1\varepsilon\ll 1, the system of algebraic equations (2) for the unknown ρi\rho_{i}’s becomes a second order ordinary differential equation for unknown function ρε(x)\rho_{\varepsilon}(x):

x(v(x)(ε2xρε+ρε(1ρε)))=ΩA(ΩA+ΩD)ρε,\partial_{x}\left(v(x)\left(-\dfrac{\varepsilon}{2}\partial_{x}\rho_{\varepsilon}+\rho_{\varepsilon}(1-\rho_{\varepsilon})\right)\right)=\Omega_{A}-(\Omega_{A}+\Omega_{D})\rho_{\varepsilon}, (8)

where v(x)v(x) is the velocity the motor proteins move with at location xx, and ΩA/D=MωA/D\Omega_{A/D}=M\omega_{A/D} are properly rescaled attachment/detachment rates. The equalities in (3) become boundary conditions for ρε(x)\rho_{\varepsilon}(x):

ρε(0)=αandρε()=1β.\rho_{\varepsilon}(0)=\alpha~{}~{}\text{and}~{}~{}\rho_{\varepsilon}(\ell)=1-\beta. (9)
Remark 1

In order to obtain (8) from (2) we take the discrete-to-continuous limit ε0\varepsilon\to 0. Specifically, we write both (6) and (7), which considered together are equivalent to (2), with Δx=ε1\Delta x=\varepsilon\ll 1:

{ε2ρi12=vi121Ji12+ρi12(1ρi12)+o(ε),Ji=ΩA(1ρi)ΩDρi+o(ε).\left\{\begin{array}[]{l}\dfrac{\varepsilon}{2}\rho^{\prime}_{i-\frac{1}{2}}=-v_{i-\frac{1}{2}}^{-1}J_{i-\frac{1}{2}}+\rho_{i-\frac{1}{2}}(1-\rho_{i-\frac{1}{2}})+o(\varepsilon),\\ J_{i}^{\prime}=\Omega_{A}(1-\rho_{i})-\Omega_{D}\rho_{i}+o(\varepsilon).\end{array}\right. (10)

We then disregard the o(ε)o(\varepsilon) terms and write resulting equations for all x(0,)x\in(0,\ell):

{ε2ρ=v1J+ρ(1ρ),J=ΩA(1ρ)ΩDρ(=ΩA(ΩA+ΩD)ρ).\left\{\begin{array}[]{l}\dfrac{\varepsilon}{2}\rho^{\prime}=-v^{-1}J+\rho(1-\rho),\\ J^{\prime}=\Omega_{A}(1-\rho)-\Omega_{D}\rho\left(=\Omega_{A}-(\Omega_{A}+\Omega_{D})\rho\right).\end{array}\right. (11)

Finally, we find JJ from the first equation of the system above and substitute it into the second equation to derive (8) for ρ=ρε\rho=\rho_{\varepsilon}. Here both notations ρ\rho^{\prime} and xρ\partial_{x}\rho denote the derivative in xx. It turns out that the phase portrait of system (11) of two coupled first order differential equations is the key to constructing the solutions of (8)-(9), see Appendix A.

Remark 2

Alternatively, in the case of the continuous velocity v(x)v(x), this equation can be formally derived from (2) by using a Taylor expansion of a smooth function ρε(x)\rho_{\varepsilon}(x) which is again obtained by interpolation of the values ρi\rho_{i} on the mesh points xix_{i}. If v(x)v(x) is piecewise continuous with a finite number of jumps, then one needs to supplement equation (8), which holds inside the intervals of continuity of v(x)v(x), with a continuity condition for the flux

Jε(x):=v(x)(ε2ρε+ρε(1ρε)).J_{\varepsilon}(x):=v(x)\left(-\dfrac{\varepsilon}{2}\rho_{\varepsilon}^{\prime}+\rho_{\varepsilon}(1-\rho_{\varepsilon})\right).

One can also consider (8) in the distributional sense. Then using the definition of Jε(x)J_{\varepsilon}(x), the integration of (8) in xx, and the absolute continuity of integrals together imply the continuity of flux Jε(x)J_{\varepsilon}(x):

Jε(x)\displaystyle J_{\varepsilon}(x) =\displaystyle= Jε(0)+0x[ΩA(ΩA+ΩD)ρε(s)]ds\displaystyle J_{\varepsilon}(0)+\int\limits_{0}^{x}\left[\Omega_{A}-(\Omega_{A}+\Omega_{D})\rho_{\varepsilon}(s)\right]\,\text{d}s (12)
=\displaystyle= Jε(0)+ΩAx(ΩA+ΩD)0xρε(s)ds.\displaystyle J_{\varepsilon}(0)+\Omega_{A}x-(\Omega_{A}+\Omega_{D})\int\limits_{0}^{x}\rho_{\varepsilon}(s)\,\text{d}s.

In what follows below, we assume that for all ε>0\varepsilon>0, there exists unique smooth solution 0ρε(x)<10\leq\rho_{\varepsilon}(x)<1 which solves (8) and satisfies the boundary conditions (9). Moreover, there exists a piecewise smooth function ρ0(x)\rho_{0}(x), referred to as the “outer solution”, such that

limε0ρε(x)=ρ0(x),for all 0x.\lim\limits_{\varepsilon\to 0}\rho_{\varepsilon}(x)=\rho_{0}(x),~{}~{}\text{for all }0\leq x\leq\ell. (13)

By calling ρ0(x)\rho_{0}(x) the outer solution we stick to the standard terminology of singularly perturbed ordinary differential equations Hol2013 . While the outer solution ρ0(x)\rho_{0}(x) is the pointwise limit of ρε(x)\rho_{\varepsilon}(x) for 0x0\leq x\leq\ell, it does not approximate the exact solution ρε(x)\rho_{\varepsilon}(x) uniformly on 0x0\leq x\leq\ell. Specifically, the outer solution ρ0(x)\rho_{0}(x) approximates ρε(x)\rho_{\varepsilon}(x) poorly in the vicinity of the jumps {xJ}\left\{x_{J}\right\}. To make the approximation uniform, one takes into account boundary layer terms of the form ρcorrector(x)=Y((xxJ)/ε)\rho_{\text{corrector}}(x)=Y((x-x_{J})/\varepsilon) whose distinguishing feature is that its slope, derivative in xx, is of the order of ε1\varepsilon^{-1}. Observe also that, even though ρ0(0)=α\rho_{0}(0)=\alpha and ρ0()=1β\rho_{0}(\ell)=1-\beta, it is possible that the outer solution ρ0(x)\rho_{0}(x) does not satisfy the boundary condition in the following sense: either limx0+ρ0(x)α\lim_{x\to 0^{+}}\rho_{0}(x)\neq\alpha or limxρ0(x)1β\lim_{x\to\ell^{-}}\rho_{0}(x)\neq 1-\beta.

Remark 3

We note that (1) possesses a unique constant solution ρiΩA/(ΩA+ΩD)\rho_{i}\equiv\Omega_{A}/(\Omega_{A}+\Omega_{D}). In what follows, we are interested in regimes where jamming can occur and, thus, for the sake of simplicity we restrict ourselves to the case when the attachment rate exceeds the detachment rate, that is, ΩA>ΩD\Omega_{A}>\Omega_{D}. This implies that the constant solution of (1) is greater than 1/21/2.

2.3 Homogeneous microtubule

Consider a constant motor protein motility rate v(x)v(x), representing a homogeneous microtubule, that is, viv0v_{i}\equiv v_{0}. The solution in the homogeneous case has been studied previously (e.g., ParFraFre2003 ; PopRakWilKolSch2003 ; ParFraFre2004 ). In this case, (8) reduces to

v0x(ε2xρε+ρε(1ρε))=ΩA(ΩA+ΩD)ρε,0<x<,\displaystyle v_{0}\partial_{x}\left(-\dfrac{\varepsilon}{2}\partial_{x}\rho_{\varepsilon}+\rho_{\varepsilon}(1-\rho_{\varepsilon})\right)=\Omega_{A}-(\Omega_{A}+\Omega_{D})\rho_{\varepsilon},\quad 0<x<\ell, (14)
ρε(0)=α,ρε()=1β.\displaystyle\rho_{\varepsilon}(0)=\alpha,\quad\rho_{\varepsilon}(\ell)=1-\beta. (15)

Recall that \ell is the length of a microtubule.

Equation (14) is a second order nonlinear ODE. If one formally passes to the limit ε0\varepsilon\to 0 in (14), then the term with the second derivative of ρε\rho_{\varepsilon} vanishes and this equation becomes first order where the solution cannot, in general, satisfy both boundary conditions in (15) as the boundary value problem is overdetermined. To describe the limiting solution, limε0ρε(x)\lim\limits_{\varepsilon\to 0}\rho_{\varepsilon}(x), we introduce an auxiliary function g(x;s,a)g(x;s,a), which is the solution of the initial value problem of the first order obtained from the formal limit as ε0\varepsilon\to 0 in (14):

{v0(12g)xg=ΩA(ΩA+ΩD)g,g(x=s;s,a)=a.\left\{\begin{array}[]{l}v_{0}(1-2g)\partial_{x}g=\Omega_{A}-(\Omega_{A}+\Omega_{D})g,\\ g(x=s;s,a)=a.\end{array}\right. (16)

In other words, function ρ0(x):=g(x;s,a)\rho_{0}(x):=g(x;s,a) is the smooth outer solution subject to a single boundary condition (or, equivalently, initial condition): ρ0|x=s=a\rho_{0}|_{x=s}=a. Equation (16) can be solved in an explicit form in terms of special functions (see ParFraFre2004 ).

Note that initial value problem (16) is not well-posed for a=0.5a=0.5. If a=0.5a=0.5, then there is no solution for x>sx>s, and for x<sx<s we define gg as the function solving the differential equation in (16) subject to the following condition:

limxsg(x;s,0.5)=0.5 and g(x;s,0.5)>0.5for all x<s.\lim\limits_{x\to s}g(x;s,0.5)=0.5\text{ and }g(x;s,0.5)>0.5\quad\text{for all }x<s. (17)

In other words, for the initial condition with a=0.5a=0.5, equation (16) admits two solutions for x<sx<s: one solution is less than 0.50.5, another one is above 0.50.5, and to describe the outer solution ρ0(x)\rho_{0}(x) we will need restrict our consideration to the upper one (the upper solution is stable in a certain sense, see Appendix A).

Function g(x;s,a)g(x;s,a) is not necessarily defined globally, for all xx. For example, consider xsx\geq s, then the solution exists on the interval (s,s+xa)(s,s+x_{a}) for some xa>0x_{a}>0 and at x=s+xax=s+x_{a} the slope of gg becomes unbounded. For example, if a<0.5a<0.5, then the value of xax_{a} can be found from the condition g(s+xa;s,a)=0.5g(s+x_{a};s,a)=0.5 which can be written as

xa=a0.5dg𝒗(g)=a0.5v0(12g)dgΩA(ΩA+ΩD)g,x_{a}=\int_{a}^{0.5}\dfrac{\text{d}g}{\boldsymbol{v}(g)}=\int_{a}^{0.5}\dfrac{v_{0}(1-2g)\,\text{d}g}{\Omega_{A}-(\Omega_{A}+\Omega_{D})g}, (18)

where function 𝒗(g)\boldsymbol{v}(g) is introduced in such a way that the differential equation from (16) is equivalent to xg=𝒗(g)\partial_{x}g=\boldsymbol{v}(g). One can compute the integral on the right-hand side of (18) to obtain an analytic formula for xax_{a}:

xa=v0ΩAΩD(ΩA+ΩD)2logΩAΩD2(ΩAa(ΩA+ΩD))+v012aΩA+ΩD.x_{a}=v_{0}\dfrac{\Omega_{A}-\Omega_{D}}{(\Omega_{A}+\Omega_{D})^{2}}\log\dfrac{\Omega_{A}-\Omega_{D}}{2(\Omega_{A}-a(\Omega_{A}+\Omega_{D}))}+v_{0}\dfrac{1-2a}{\Omega_{A}+\Omega_{D}}. (19)

The following theorem gives an explicit formula for the limiting solution of the (14)-(15) as ε0\varepsilon\to 0.

Theorem 2.1

Define ρ0(x):=limε0ρε(x)\rho_{0}(x):=\lim\limits_{\varepsilon\to 0}\rho_{\varepsilon}(x) for 0x0\leq x\leq\ell and ρε\rho_{\varepsilon} solving (14)-(15). Then

ρ0(x)={α,x=0,g(x;0,α),0<x<max{0,xJ},g(x;,max{0.5,1β}),max{0,xJ}<x<,1β,x=.\rho_{0}(x)=\left\{\begin{array}[]{ll}\alpha,&x=0,\\ g(x;0,\alpha),&0<x<\max\{0,x_{J}\},\\ g(x;\ell,\max\{0.5,1-\beta\}),&\max\{0,x_{J}\}<x<\ell,\\ 1-\beta,&x=\ell.\end{array}\right. (20)

If α1/2\alpha\geq 1/2, then xJ=0x_{J}=0. If α<1/2\alpha<1/2, then xJx_{J} is determined by

xJ:=min{x0|g(x;0,α)+g(x;,max{0.5,1β})1}.x_{J}:=\mathrm{min}\left\{x\geq 0\,|\,\,g(x;0,\alpha)+g(x;\ell,\max\left\{0.5,1-\beta\right\})\leq 1\right\}. (21)

The result of this theorem is consistent with previous works where the system (14)-(15) was studied (e.g., see Fre11 ; Fre07 ; Fre18 ). We relegate the proof of this theorem and examples of the application of the representation formula (20) to Appendix A.

Remark 4

The point xJx_{J}, if 0<xJ<0<x_{J}<\ell, is the location of the interior jump, that is, the outer solution ρ0(x)\rho_{0}(x) is a smooth solution of the differential equation from (16) on intervals (0,xJ)(0,x_{J}) and (xJ,)(x_{J},\ell) and it has one jump inside (0,)(0,\ell) at x=xJx=x_{J}. The value of xJx_{J} can also be found via numerical simulations of the following equation

g(xJ;0,α)+g(xJ;,max{0.5,1β})=1.g(x_{J};0,\alpha)+g(x_{J};\ell,\max\{0.5,1-\beta\})=1. (22)

This equation is equivalent to the continuity of fluxes J0=ρ0(1ρ0)J_{0}=\rho_{0}(1-\rho_{0}) at the point of the jump of the outer solution ρ0=ρ0(x)\rho_{0}=\rho_{0}(x):

v0ρα(1ρα)|xxJ=v0ρβ(1ρβ)|xxJ+,v_{0}\rho_{\alpha}(1-\rho_{\alpha})|_{x\to x_{J}-}=v_{0}\rho_{\beta}(1-\rho_{\beta})|_{x\to x_{J}+}, (23)

where ρα(x)=g(x;0,α)\rho_{\alpha}(x)=g(x;0,\alpha) and ρβ(x)=g(x;,max{0.5,1β})\rho_{\beta}(x)=g(x;\ell,\max\{0.5,1-\beta\}).

Remark 5

If one varies boundary conditions (15), then the outer solution ρ0(x)\rho_{0}(x) may stay unchanged (except values at x=0x=0 and x=x=\ell) for wide range of parameters α\alpha and β\beta. For example, denote ρβ(x):=g(x,,max{0.5,1β})\rho_{\beta}(x):=g(x,\ell,\max\left\{0.5,1-\beta\right\}). Theorem 2.1 implies that outer solution ρ0(x)\rho_{0}(x) coincides with ρβ(x)\rho_{\beta}(x) on the interval 0<x<10<x<1 for the following range of α\alpha:

1ρβ(0)α1.1-\rho_{\beta}(0)\leq\alpha\leq 1. (24)

Once α\alpha becomes smaller than the lower limit, 1ρβ(0)1-\rho_{\beta}(0), an interior jump appears in the outer solution ρ0(x)\rho_{0}(x).

The following corollary is important for the study of inhomogeneous microtubules in Section 2.4.

Corollary 1

Assume α1/2\alpha\geq 1/2 or

00.5v0(12ρ)ΩA(ΩA+ΩD)ρdρ.\int\limits_{0}^{0.5}\dfrac{v_{0}(1-2\rho)}{\Omega_{A}-(\Omega_{A}+\Omega_{D})\rho}\,\mathrm{d}\rho\leq\ell. (25)

Then

  • (i)

    limxρ0(x)=max{0.5,1β}\lim\limits_{x\to\ell-}\rho_{0}(x)=\max\left\{0.5,1-\beta\right\}.

  • (ii)

    If β<1/2\beta<1/2, then ρ0(x)\rho_{0}(x) is continuous at x=x=\ell, that is, limxρ0(x)=ρ0(0)=1β\lim\limits_{x\to\ell-}\rho_{0}(x)=\rho_{0}(0)=1-\beta.

  • (iii)

    If α1/2\alpha\geq 1/2, then there is no interior jump, that is, xJ0x_{J}\leq 0.

  • (iv)

    If 0<xJ<0<x_{J}<\ell, then α<1/2\alpha<1/2 and limx0+ρ0(x)=α\lim\limits_{x\to 0+}\rho_{0}(x)=\alpha (that is, ρ0(x)\rho_{0}(x) is continuous at x=0x=0).

Condition (25) excludes the case of low density solutions, that is, we exclude outer solutions of the form: ρ0(x)<0.5\rho_{0}(x)<0.5 for all x[0,)x\in[0,\ell). This regime is not consistent with jamming, which is the focus of this work. In other words, condition (25) imposes that the “left” part of solution, g(x;0,α)g(x;0,\alpha), cannot be extended to entire interval [0,)[0,\ell). The reason we would like to impose this condition below in the inhomogeneous case is because we then focus on cases when regions with high densities emerge, and thus traffic jams in motor transport are possible. By direct integration, condition (25) can be written as

v0ΩA+ΩD[1ΩAΩDΩA+ΩDlog(2ΩAΩD)].\dfrac{v_{0}}{\Omega_{A}+\Omega_{D}}\left[1-\dfrac{\Omega_{A}-\Omega_{D}}{\Omega_{A}+\Omega_{D}}\log\left(\dfrac{2}{\Omega_{A}-\Omega_{D}}\right)\right]\leq\ell. (26)

2.4 Inhomogeneous microtubule

In this section, we consider a non-constant motility rate v(x)v(x). Biologically, it corresponds to a inhomogeneous microtubule with different motor protein mobilities in different regions of the microtubule. Without loss of generality, we take =1\ell=1. We restrict ourselves to the case

v(x)={vL,0xx0,vR,x0<x1.v(x)=\left\{\begin{array}[]{ll}v_{L},&0\leq x\leq x_{0},\\ v_{R},&x_{0}<x\leq 1.\end{array}\right. (27)

This case may be considered as two coupled homogeneous microtubules meeting at interface x=x0x=x_{0} with coupling through the continuity of densities and fluxes. Specifically, we have the following system of equations:

vLx[(ε2ρε+ρε(1ρε))]=ΩA(ΩA+ΩD)ρε,0<x<x0,\displaystyle v_{L}\partial_{x}\left[\left(-\dfrac{\varepsilon}{2}\rho_{\varepsilon}^{\prime}+\rho_{\varepsilon}(1-\rho_{\varepsilon})\right)\right]=\Omega_{A}-(\Omega_{A}+\Omega_{D})\rho_{\varepsilon},~{}~{}~{}0<x<x_{0}, (28)
vRx[(ε2ρε+ρε(1ρε))]=ΩA(ΩA+ΩD)ρε,x0<x<1,\displaystyle v_{R}\partial_{x}\left[\left(-\dfrac{\varepsilon}{2}\rho_{\varepsilon}^{\prime}+\rho_{\varepsilon}(1-\rho_{\varepsilon})\right)\right]=\Omega_{A}-(\Omega_{A}+\Omega_{D})\rho_{\varepsilon},~{}~{}~{}x_{0}<x<1, (29)

and two coupling conditions:

  1. (1)

    continuity of ρε(x)\rho_{\varepsilon}(x) at x0x_{0}:

    ρε(x0)=ρε(x0+)=:ρε.\rho_{\varepsilon}(x_{0}^{-})=\rho_{\varepsilon}(x_{0}^{+})=:\rho_{\varepsilon}. (30)
  2. (2)

    continuity of flux Jε(x)J_{\varepsilon}(x) at x0x_{0}:

    vL(ε2ρε(x0)+ρε(1ρε))=vR(ε2ρε(x0+)+ρε(1ρε)).v_{L}\left(-\dfrac{\varepsilon}{2}\rho_{\varepsilon}^{\prime}(x_{0}^{-})+\rho_{\varepsilon}(1-\rho_{\varepsilon})\right)=v_{R}\left(-\dfrac{\varepsilon}{2}\rho_{\varepsilon}^{\prime}(x_{0}^{+})+\rho_{\varepsilon}(1-\rho_{\varepsilon})\right). (31)

The outer solution ρ0(x)=limε0ρε(x)\rho_{0}(x)=\lim\limits_{\varepsilon\to 0}\rho_{\varepsilon}(x) is not necessarily continuous, nevertheless due to (12) it satisfies the flux continuity condition:

vLρ0(1ρ0)|xx0=vRρ0(1ρ0)|xx0+.\displaystyle v_{L}\rho_{0}(1-\rho_{0})\biggl{|}_{x\to x_{0}^{-}}=v_{R}\rho_{0}(1-\rho_{0})\biggr{|}_{x\to x_{0}^{+}}. (32)
Remark 6

By looking at the system (28)-(29) one may think that the case of inhomogeneous v(x)v(x) is equivalent to the case of inhomogeneous attachment/detachment rates but with constant motility rates v(x)1v(x)\equiv 1:

x(ε2ρε+ρε(1ρε))=ΩA(x)(ΩA(x)+ΩD(x))ρε\partial_{x}\left(-\dfrac{\varepsilon}{2}\rho_{\varepsilon}^{\prime}+\rho_{\varepsilon}(1-\rho_{\varepsilon})\right)=\Omega_{A}(x)-(\Omega_{A}(x)+\Omega_{D}(x))\rho_{\varepsilon} (33)

with attachment/detachment rates ΩA(x)\Omega_{A}(x) and ΩD(x)\Omega_{D}(x) are ΩA/vL\Omega_{A}/v_{L} and ΩD/vL\Omega_{D}/v_{L} inside the left half of the microtubules, x<x0x<x_{0}, and ΩA/vR\Omega_{A}/v_{R} and ΩD/vR\Omega_{D}/v_{R} inside the right half, x>x0x>x_{0}. For differential equation (33), from the continuity of ρε(x)\rho_{\varepsilon}(x) and Jε(x)J_{\varepsilon}(x), one concludes that ρε\rho_{\varepsilon} is necessarily continuously differentiable, whereas the solution of (28)-(29) possesses a discontinuous derivative at x0x_{0} for vLvRv_{L}\neq v_{R} which follows from (31) and is written as

ρε(x0+)ρε(x0)=2εvRvL(vRvL)Jε(x0).\rho^{\prime}_{\varepsilon}(x_{0}^{+})-\rho^{\prime}_{\varepsilon}(x_{0}^{-})=\dfrac{2}{\varepsilon v_{R}v_{L}(v_{R}-v_{L})}J_{\varepsilon}(x_{0}). (34)

Therefore, problems for inhomogeneous v(x)v(x) and inhomogeneous ΩA,D(x)\Omega_{A,D}(x) are not equivalent.

By analogy with function gg from (16) in the case of homogeneous microtubules, we introduce gL(x;sL,aL)g_{L}(x;s_{L},a_{L}) and gR(x;sR,aR)g_{R}(x;s_{R},a_{R}) as solutions of the following initial value problems:

{vL(12gL)xgL=ΩA(ΩA+ΩD)gL,gL(sL;sL,aL)=aL,vR(12gR)xgR=ΩA(ΩA+ΩD)gR,gR(sR;sR,aR)=aR.\left\{\begin{array}[]{l}v_{L}(1-2g_{L})\partial_{x}g_{L}=\Omega_{A}-(\Omega_{A}+\Omega_{D})g_{L},~{}~{}~{}g_{L}(s_{L};s_{L},a_{L})=a_{L},\\ v_{R}(1-2g_{R})\partial_{x}g_{R}=\Omega_{A}-(\Omega_{A}+\Omega_{D})g_{R},~{}~{}~{}g_{R}(s_{R};s_{R},a_{R})=a_{R}.\end{array}\right. (35)

In what follows we consider separately two cases:

  • (i)

    fast - slow microtubule: vL>vRv_{L}>v_{R},

  • (ii)

    slow - fast microtubule: vR>vLv_{R}>v_{L}.

Before we formulate our main result for these two cases, we note that the difficulty in the determination of the outer solution ρ0(x)\rho_{0}(x) is finding the value of ρ0\rho_{0} at the interface, A:=ρ0(x0)A:=\rho_{0}(x_{0}). Once AA is found, one can use Theorem 2.1 to restore ρ0(x)\rho_{0}(x) in both intervals [0,x0][0,x_{0}] and [x0,1][x_{0},1].

Theorem 2.2

Consider vL>vRv_{L}>v_{R} and assume that condition (25) holds with v0=vLv_{0}=v_{L} and =x0\ell=x_{0}. Let ρ0(x)\rho_{0}(x) be the outer solution of system (28)-(29) equipped with coupling conditions (30)-(31). Then function ρ0(x)\rho_{0}(x) has a jump at x=x0x=x_{0} and is given by

ρ0(x)={α,x=0,gL(x;0,α),0<x<max{0,xJ},gL(x;x0,A),max{0,xJ}<xx0,gR(x;1,max{0.5,1β}),x0<x<1,1β,x=1,\rho_{0}(x)=\left\{\begin{array}[]{ll}\alpha,&x=0,\\ g_{L}(x;0,\alpha),&0<x<\max\{0,x_{J}\},\\ g_{L}(x;x_{0},A),&\max\{0,x_{J}\}<x\leq x_{0},\\ g_{R}(x;1,\max\{0.5,1-\beta\}),&x_{0}<x<1,\\ 1-\beta,&x=1,\end{array}\right. (36)

where xJx_{J} is determined from the continuity of fluxes:

gL(xJ;0,α)+gL(xJ;x0,A)=1,\displaystyle g_{L}(x_{J};0,\alpha)+g_{L}(x_{J};x_{0},A)=1, (37)

and A=(vL+vL24JR2)/(2vL)A=\left(v_{L}+\sqrt{v_{L}^{2}-4J^{2}_{R}}\right)/(2v_{L}) where JR=vRgR(1gR)J_{R}=v_{R}g_{R}(1-g_{R}) with

gR:=gR(x0;1,max{0.5,1β}).g_{R}:=g_{R}(x_{0};1,\max\{0.5,1-\beta\}).
Proof

First, we show that the outer solution ρ0(x)\rho_{0}(x) has a jump at x=x0x=x_{0}. Indeed, from continuity of fluxes (31) with ρε(x0)=A+o(1)\rho_{\varepsilon}(x_{0})=A+o(1) we get:

vLρε(x0)=vRρε(x0+)+vLvR2εA(1A)+o(1ε).v_{L}\rho_{\varepsilon}^{\prime}(x_{0}^{-})=v_{R}\rho_{\varepsilon}^{\prime}(x_{0}^{+})+\dfrac{v_{L}-v_{R}}{2\varepsilon}A(1-A)+o\left(\dfrac{1}{\varepsilon}\right). (38)

Since AA is strictly between 0 and 1, we find that one of the derivatives (either the left or right one) is of the order ε1\varepsilon^{-1} corresponding to a jump.

Next, denote the limits of the outer solution from the left and right at x0x_{0} by

AL:=ρ0|xx0 and AR:=ρ0|xx0+.A_{L}:=\rho_{0}|_{x\to x_{0}^{-}}\text{ and }A_{R}:=\rho_{0}|_{x\to x_{0}^{+}}.

Then the continuity of fluxes (32) is written as vLAL(1AL)=vRAR(1AR)v_{L}A_{L}(1-A_{L})=v_{R}A_{R}(1-A_{R}), and since vL>vRv_{L}>v_{R} we have AL(1AL)<1/4A_{L}(1-A_{L})<1/4. By applying Corollary 1 (ii) we find A=AL>0.5A=A_{L}>0.5 and thus, there is no jump from the left at x0x_{0}. Then, following Corollary 1 (iii), there is no interior jump, and ρ0(x)\rho_{0}(x) in interval [x0,1][x_{0},1] is determined by Corollary 1 (i), which is a boundary condition for ρ0(x)\rho_{0}(x) at x0=1x_{0}=1.

Specifically, by Theorem 2.1 we have ρ0(x)=gR(x;1,max{0.5,1β})\rho_{0}(x)=g_{R}(x;1,\max\{0.5,1-\beta\}) for x0<x<1x_{0}<x<1. Then AR=gR:=gR(x0;1,max{0.5,1β})A_{R}=g_{R}:=g_{R}(x_{0};1,\max\{0.5,1-\beta\}) and AA is the solution of the quadratic equation vLA(1A)=vRAR(1AR)v_{L}A(1-A)=v_{R}A_{R}(1-A_{R}), which is strictly greater than 0.50.5. Thus, we found AA, and the expression for ρ0(x)\rho_{0}(x) in [0,x0][0,x_{0}] is found by using the representation formula (20) from Theorem 2.1. ∎

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 1: Examples 1 & 2 for inhomogeneous microtubulus. The upper row depicts plots of outer solutions ρ0(x)\rho_{0}(x). The lower row depicts solutions as trajectories in the (ρ,J)(\rho,J)-plane; note that these trajectories are continuous in JJ and have discontinuities (jumps) in ρ\rho. Red (lower) and blue (upper) dashed lines are arcs J=vRρ(1ρ)J=v_{R}\rho(1-\rho) and J=vLρ(1ρ)J=v_{L}\rho(1-\rho), respectively.

Next, we illustrate formula (36) of Theorem 2.2 by the following two examples with x0=0.5x_{0}=0.5, vL=1.0v_{L}=1.0, vR=0.5v_{R}=0.5, ΩA=0.8\Omega_{A}=0.8 and ΩD=0.2\Omega_{D}=0.2.

Example 1. α=0.2\alpha=0.2 and β=0.3\beta=0.3.

ρ0(x)={0.2,x=0,gL(x;0.5,0.907),0<x0.5,(A=0.5(1±14JR)0.907)gR(x;1.0,0.7),0.5<x1.0.\rho_{0}(x)=\left\{\begin{array}[]{ll}0.2,&x=0,\\ g_{L}(x;0.5,0.907),&0<x\leq 0.5,\,(A=0.5(1\pm\sqrt{1-4J_{R}})\approx 0.907)\\ g_{R}(x;1.0,0.7),&0.5<x\leq 1.0.\end{array}\right.

Example 2. α=0.05\alpha=0.05 and β=0.7\beta=0.7.

ρ0(x)={gL(x;0.0,0.05),0x<xJ,(xJ0.101)gL(x;0.5,0.904),xJ<x0.5,gR(x;1.0,0.5+),0.5<x<1.0,(A0.904)0.3,x=1.0.\rho_{0}(x)=\left\{\begin{array}[]{ll}g_{L}(x;0.0,0.05),&0\leq x<x_{J},\,(x_{J}\approx 0.101)\\ g_{L}(x;0.5,0.904),&x_{J}<x\leq 0.5,\\ g_{R}(x;1.0,0.5+),&0.5<x<1.0,\,(A\approx 0.904)\\ 0.3,&x=1.0.\end{array}\right.

Solutions from Examples 1 and 2 are depicted in Fig. 1. The main difference between Examples 1 and 2 is that in Example 2 there is an interior jump in the left sub-interval (0,0.5)(0,0.5), while the solution in Example 1 does not possesses a jump in neither of the two sub-intervals, left (0,0.5)(0,0.5) and right (0.5,1)(0.5,1) (or, equivalently, xJ=0x_{J}=0 in (36) for Example 1).

Theorem 2.3

Consider vR>vLv_{R}>v_{L} and assume condition (25) holds with both v0=vLv_{0}=v_{L}, =x0\ell=x_{0} and v0=vRv_{0}=v_{R}, =1x0\ell=1-x_{0}. Let ρ0(x)\rho_{0}(x) be the outer solution of system (28)-(29) equipped with coupling conditions (30)-(31). Denote also

A^R:=gR(x0;1,max{0.5,1β}).\hat{A}_{R}:=g_{R}(x_{0};1,\max\{0.5,1-\beta\}). (39)

Then

A={1214vL4vR,vRA^R(1A^R)>0.25vL,12+14vRvLA^R(1A^R),vRA^R(1A^R)0.25vL,A=\left\{\begin{array}[]{ll}\dfrac{1}{2}-\sqrt{\dfrac{1}{4}-\dfrac{v_{L}}{4v_{R}}},&\qquad v_{R}{\hat{A}_{R}(1-\hat{A}_{R})>0.25v_{L}},\\ &\\ \dfrac{1}{2}+\sqrt{\dfrac{1}{4}-\dfrac{v_{R}}{v_{L}}\hat{A}_{R}(1-\hat{A}_{R})},&\qquad v_{R}{\hat{A}_{R}(1-\hat{A}_{R})\leq 0.25v_{L}},\end{array}\right. (40)

and function ρ0(x)\rho_{0}(x) is given by

ρ0(x)={α,x=0,gL(x;0,α),0<x<max{0,xJ(L)},gL(x;x0,max{A,0.5}),max{0,xJ(L)}x<x0,gR(x;x0,min{A,0.5}),x0xxJ(R),gR(x;1,max{0.5,1β}),xJ(R)<x<1,1β,x=1.\rho_{0}(x)=\left\{\begin{array}[]{ll}\alpha,&\qquad x=0,\\ g_{L}(x;0,\alpha),&\qquad 0<x<\max\{0,x_{J}^{(L)}\},\\ g_{L}(x;x_{0},\max\{A,0.5\}),&\qquad\max\{0,x_{J}^{(L)}\}\leq x<x_{0},\\ g_{R}(x;x_{0},\min\{A,0.5\}),&\qquad x_{0}\leq x\leq x_{J}^{(R)},\\ g_{R}(x;1,\max\{0.5,1-\beta\}),&\qquad x_{J}^{(R)}<x<1,\\ 1-\beta,&\qquad x=1.\end{array}\right. (41)

Here xJ(L)x_{J}^{(L)} and xJ(R)x_{J}^{(R)} are determined from the continuity of fluxes:

gL(xJ(L);0,α)+gL(xJ(L);x0,max{A,0.5})=1,\displaystyle g_{L}(x_{J}^{(L)};0,\alpha)+g_{L}(x_{J}^{(L)};x_{0},\max\{A,0.5\})=1,
gR(xJ(R);x0,min{A,0.5})+gR(xJ(R);1,max{0.5,1β})=1.\displaystyle g_{R}(x_{J}^{(R)};x_{0},\min\{A,0.5\})+g_{R}(x_{J}^{(R)};1,\max\{0.5,1-\beta\})=1.
Proof

As in the proof of Theorem 2.2, introduce AL:=ρ0|xx0A_{L}:=\rho_{0}|_{x\to x_{0}^{-}} and AR:=ρ0|xx0+A_{R}~{}:=~{}\rho_{0}|_{x\to x_{0}^{+}}. Then continuity of fluxes at x0x_{0} reads

vLAL(1AL)=vRAR(1AR).v_{L}A_{L}(1-A_{L})=v_{R}A_{R}(1-A_{R}). (42)

By Corollary 1 (i) we get AL0.5A_{L}\geq 0.5. Next, we consider two cases. If

vRvLA^R(1A^R)1/4,\dfrac{v_{R}}{v_{L}}\hat{A}_{R}(1-\hat{A}_{R})\leq 1/4,

then AR=A^RA_{R}=\hat{A}_{R} and AL0.5A_{L}\geq 0.5 solves (42), that is,

A=AL=12+14vRvLA^R(1A^R).A=A_{L}=\dfrac{1}{2}+\sqrt{\dfrac{1}{4}-\dfrac{v_{R}}{v_{L}}\hat{A}_{R}(1-\hat{A}_{R})}.

If vRvLA^R(1A^R)>1/4,\dfrac{v_{R}}{v_{L}}\hat{A}_{R}(1-\hat{A}_{R})>1/4, then ρ0(x)\rho_{0}(x) does not coincide with gR(x;1,max{0.5,1β})g_{R}(x;1,\max\{0.5,1-\beta\}) on the entire interval (x0,1)(x_{0},1), there is an interior jump at x0<xJ(R)<1x_{0}<x_{J}^{(R)}<1 and by Corollary 1 (iii) and (iv), AR<1/2A_{R}<1/2 and A=ARA=A_{R}. Then AL=1/2A_{L}=1/2 and ρ0(x)\rho_{0}(x) on intervals (0,x0)(0,x_{0}) and (x0,1)(x_{0},1) is found by (20). ∎

Next we illustrate formula (41) by two examples with vR>vLv_{R}>v_{L}. Namely, set vL=0.5v_{L}=0.5 and vR=1.0v_{R}=1.0 and consider x0=0.5x_{0}=0.5, ΩA=0.8\Omega_{A}=0.8 and ΩD=0.2\Omega_{D}=0.2.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2: Examples 3 & 4 for a inhomogeneous microtubule. The upper row depicts plots of outer solutions ρ0(x)\rho_{0}(x). The lower row depicts solutions as trajectories in (ρ,J)(\rho,J) plane, red (lower) and blue (upper) dashed lines are arcs J=vLρ(1ρ)J=v_{L}\rho(1-\rho) and J=vRρ(1ρ)J=v_{R}\rho(1-\rho), respectively.

Example 3. α=0.1\alpha=0.1 and β=0.7\beta=0.7.

ρ0(x)={gL(x;0.0,0.1),0x<xJ(L),(xJ(L)0.097)gL(x;0.5,0.5+),xJ(L)<x<0.5,gR(x;0.5,A),0.5x<xJ(R),(A=12(112),xJ(R)0.624)gR(x;1.0,0.5+),xJ(R)<x<1.0,0.3,x=1.0.\rho_{0}(x)=\left\{\begin{array}[]{ll}g_{L}(x;0.0,0.1),&\qquad 0\leq x<x_{J}^{(L)},\,(x_{J}^{(L)}\approx 0.097)\\ g_{L}(x;0.5,0.5+),&\qquad x_{J}^{(L)}<x<0.5,\\ g_{R}(x;0.5,A),&\qquad 0.5\leq x<x_{J}^{(R)},\,(A=\frac{1}{2}(1-\frac{1}{\sqrt{2}}),\,x_{J}^{(R)}\approx 0.624)\\ g_{R}(x;1.0,0.5+),&\qquad x_{J}^{(R)}<x<1.0,\\ 0.3,&\qquad x=1.0.\end{array}\right.

Example 4. α=0.1\alpha=0.1 and β=0.05\beta=0.05.

ρ0(x)={gL(x;0.0,0.1),0x<xJ(L),(xJ(L)0.09)gL(x;0.5,A),xJ(L)x0.5,(A0.703)gR(x;1.0,0.95),0.5<x1.\rho_{0}(x)=\left\{\begin{array}[]{ll}g_{L}(x;0.0,0.1),&\qquad 0\leq x<x_{J}^{(L)},\,(x_{J}^{(L)}\approx 0.09)\\ g_{L}(x;0.5,A),&\qquad x_{J}^{(L)}\leq x\leq 0.5,\,(A\approx 0.703)\\ g_{R}(x;1.0,0.95),&\qquad 0.5<x\leq 1.\end{array}\right.

Solutions from Examples 3 and 4 are depicted in Fig. 2. These two examples illustrate two possibilities for slow-fast microtubules: when ρ0(x)\rho_{0}(x) is continuous from the right at x0x_{0} (Example 3) and when it is continuous from the left at x0x_{0} (Example 4).

3 Monte Carlo Simulations

We now show that the results from Section 2.4 involving the mean-field continuous model in the inhomogeneous case are consistent with those from Monte Carlo simulations. To this end, we return to the discrete problem (2)-(3) with M=500M=500 lattice sites. We perform R=104R=10^{4} realizations and compare the resulting discrete densities with the continuum equation (8) as ε0\varepsilon\to 0, computed by representation formulas from Theorem 2.2 and 2.3.

Refer to caption
Refer to caption
Figure 3: Examples 3 & 4 are simulated with R=104R=10^{4} realizations of the lattice model, described in Section 3. Simulations use the same parameters as Examples 3 & 4 for inhomogeneous microtubules above (analytical solution is solid black line; see also Figure 2). We see a strong quantitative agreement between the Monte Carlo (red) and analytical (black) solutions.

Specifically, we adapt a similar algorithm to the one in ParFraFre2003 . In each realization r=1,,Rr=1,...,R we consider tuple {ν(n)(i)}\{\nu^{(n)}(i)\} where nn stands for the iteration step number and ν(n)(i)=1\nu^{(n)}(i)=1 if the iith site is occupied at the nnth iteration step and ν(n)(i)=0\nu^{(n)}(i)=0, if otherwise. Initially, microtubule is empty, i.e., ν(0)(i)=0\nu^{(0)}(i)=0 for i=1,,Mi=1,...,M. For each time step, n=1,,Nn=1,...,N discrete dynamics of {ν(n)(i)}i=1M\left\{\nu^{(n)}(i)\right\}_{i=1}^{M} is described by the following procedure:

  • 1.

    Choose randomly site ii (all sites are equiprobable).

  • 2.

    If i=1i=1 and ν(n)(1)=0\nu^{(n)}(1)=0, then ν(n+1)(1)=1\nu^{(n+1)}(1)=1 with probability α\alpha.

  • 3.

    If i=Mi=M and ν(n)(M)=1\nu^{(n)}(M)=1, then ν(n+1)(M)=0\nu^{(n+1)}(M)=0 with probability β\beta.

  • 4.

    If 1<i<M1<i<M, then

    • if ν(n)(i)=0\nu^{(n)}(i)=0, then ν(n+1)(i)=1\nu^{(n+1)}(i)=1 and ν(n+1)(i1)=0\nu^{(n+1)}(i-1)=0 with probability vi12v_{i-\frac{1}{2}} provided that ν(n)(i1)=1\nu^{(n)}(i-1)=1;

    • if ν(n)(i)=1\nu^{(n)}(i)=1, then ν(n+1)(i)=0\nu^{(n+1)}(i)=0 and ν(n+1)(i+1)=1\nu^{(n+1)}(i+1)=1 with probability vi+12v_{i+\frac{1}{2}} provided that ν(n)(i+1)=0\nu^{(n)}(i+1)=0.

  • 5.

    If 1<i<M1<i<M, then

    • if ν(n+1)(i)=1\nu^{(n+1)}(i)=1 after step 4, then let ν(n+1)(i)=0\nu^{(n+1)}(i)=0 with probability ωd\omega_{d};

    • if ν(n+1)(i)=0\nu^{(n+1)}(i)=0 after step 4, then let ν(n+1)(i)=1\nu^{(n+1)}(i)=1 with probability ωa\omega_{a}.

Finally, after running NN steps and RR realizations, we assign ρiMCS:=νi(N)r\rho_{i}^{\text{MCS}}:=\langle\nu_{i}^{(N)}\rangle_{r}.

Monte Carlo simulations are in very good agreement with the outer solutions derived in Section 2.4. Specifically, results of Monte Carlo simulations corresponding to Example 3 and 4 from Section 2.4 are depicted in Fig. 3; one can see agreement between histograms obtained from Monte Carlo simulations (red) and analytical solutions (black). Observe that number of realizations RR and number of iteration steps NN needed to reach equilibrium are critical for recovering the sharp transitions observed near the interface of the inhomogeneous microtubule. For Monte Carlo histograms in Fig. 3, we were to take N=2.5106N=2.5\cdot 10^{6} time steps in order to guarantee that a transient solution has reached equilibrium for each of R=104R=10^{4} realizations. The large numbers for both realizations and iteration steps lead to the observation that the reproduction of equilibrium profiles (solutions) in Examples 1-4 by Monte Carlo simulations to be very time consuming, whereas Theorems 2.2 and 2.3 give explicit formulas which require only numerical integration of at most four ODEs for the auxiliary functions g(x;s,a)g(x;s,a).

4 Discussion

In this work, we present a mathematical model to describe dynamics of motor proteins on microtubules. Using methods from asymptotic analysis, we provide closed-form expressions for motor protein density solutions. We also provide verification of the results of mathematical analysis by Monte Carlo simulation with the discrete MT model. The mathematical model may serve as a convenient framework for studying experimental data. Even more, the modeling and analysis may assist in inferring in vivo dynamics where biophysical imaging is limited in the crowded cellular environments. It is also important to note that the model presented herein is consistent with prior theoretical results for the homogeneous case (e.g., ParFraFre2003 ; ParFraFre2004 ; Fre11 ).

The model approach developed herein provides additional advantages over the prior approach of Frey et al. Fre18 and others Klu08 ; Klu05 while remaining faithful in the homogeneous case. Most notably, the model is developed to study inhomogeneous regimes where large density profiles can result in the emergence of internal boundary layers. Beyond the obvious application to motor protein dynamics along a microtubule, this also provides insight into traffic flow problems. The PDE governing the density of cars has a similar form to the equation governing the density of motor proteins here. This work also provides an additional example of the power of analyzing discrete ODE model systems by passing to the limit and obtaining a mean-field PDE.

What made this work challenging is that a priori initial data cannot predict regions of low or high density. Even within the Monte Carlo simulations we observe that they must be run for a significant length of time to capture all the feature of the solutions (e.g, interior boundary layers, sharp transitions etc.). An additional challenge lies in experimental verification given the current state of technology. Once imaging technology improves combined with advancements in biophysical knowledge, the theory developed in this manuscript can be rigorously tested experimentally both in vitro and in vivo. This will be crucial in verifying model parameter regimes corresponding to biologically realistic results.

This work lays the foundation for future work in understanding inhibited transport along microtubules. The model we present may be augmented to account for more biological realism in describing motor protein dynamics and intracellular transport. Realistically there are several “lanes” on these MTs which motor proteins move laterally and they may switch lanes. Similarly, motor proteins may also change directions when encountering patches Ros08 . Furthermore, transport takes place on highly complex 3-dimensional networks of many MTs and AFs. Hence modeling the intersections between MTs would be of interest as well as analyzing the composite density profiles using the analysis presented in this work. In addition, the cargo transported to and from the cell nucleus and cell wall is carried by motor proteins Lakadamyali2014 ; Gro07 and the given model may be augmented to account for this cargo. We also note that motor proteins transfer from MT to MT within the cell, and the model as well as analysis developed here may serve as a foundation for this study.

Overall, the model for an inhomogeneous microtubule presented here can inform motor protein dynamics in rough regimes where transport properties are not consistent along given trajectories. This will ultimately lay the groundwork for fundamental understanding of the onset of neurodegenerative diseases. The inhomogeneous microtubule model may be used to investigate how one can control transport properties of motor proteins in high density regimes along microtubules. Given the structure of a microtubule, can one devise conditions so that the equilibrium solution contains no high density regimes (jams) by understanding or imposing defects along its surface? Also, given a distribution of inhomogeneous regions (N>2N>2) on a microtubule can we predict the equilibrium solution? The answers to these questions may be the source of further investigation in a future work.

Acknowledgements.
The work of SR was supported by the Cleveland State University Office of Research through a Faculty Research Development Grant.

References

  • (1) Parmeggiani, A., Franosch, T. and Frey, E., 2003. Phase coexistence in driven one-dimensional transport. Physical review letters, 90, p.086601.
  • (2) Popkov, V., Rákos, A., Willmann, R.D., Kolomeisky, A.B. and Schütz, G.M., 2003. Localization of shocks in driven diffusive systems without particle number conservation. Physical Review E, 67, p.066117.
  • (3) Parmeggiani, A., Franosch, T. and Frey, E., 2004. Totally asymmetric simple exclusion process with Langmuir kinetics. Physical Review E, 70, p.046101.
  • (4) Gross, S. P., Vershinin, M. and Shubeita G. T., 2007. Cargo transport: Two motors are sometimes better than one. Current Biology, 17, R478-R486.
  • (5) Reese, L., Melbinger, A. and Frey, E., 2011. Crowding of molecular motors determines microtubule depolymerization. Biophysical Journal, 101 p. 2190-2200.
  • (6) M. H. Holmes, Introduction to Perturbation Methods, 2nd edition. Texts in Applied Mathematics, 20. Springer, New York, 2013. xviii+436 pp.
  • (7) Ross, J. L., Ali, M. Y. and Warshaw, D. M., 2008. Cargo transport: molecular motors navigate a complex cytoskeleton Current Opinion in Cell Biology, 20 p. 41-47.
  • (8) Lakadamyali, M., 2014. Navigating the cell: how motors overcome roadblocks and traffic jams to efficiently transport cargo Phys. Chem. Chem. Phys., 16 p. 5907.
  • (9) Reichenbach, T., Frey, E. and Franosch, T., 2007. Traffic jams induced by rare switching events in two-lane transport New Journal of Physics, 9 p. 159.
  • (10) Tan, R., Lam, A.J., Tan, T., Han, J., Nowakowski, D.W., Vershinin, M., Simo, S., Ori-McKenney, K.M. and McKenney, R.J., 2019. Microtubules gate tau condensation to spatially regulate microtubule functions. Nature cell biology, 21(9), pp.1078-1085.
  • (11) Oelz, D. and Mogilner, A., 2016. A drift-diffusion model for molecular motor transport in anisotropic filament bundles. Discrete and Continuous Dynamical Systems, 36:8 p. 4553-4567.
  • (12) Rank, M. and Frey, E., 2018. Crowding and pausing strongly affect dynamics of kinesin-1 motors along microtubules. Biophysical journal, 115(6), 1068-1081.
  • (13) Klumpp, S., Chai, Y., and Lipowsky, R., 2008. Effects of the chemomechanical stepping cycle on the traffic of molecular motors. Physical Review E, 78, 041909.
  • (14) Klumpp, S., Nieuwenhuizen, T. M., and Lipowsky, R., 2005. Self-organized density patterns of molecular motors in arrays of cytoskeletal filaments. Biophysical Journal, 88, pg. 3118-3132.
  • (15) Klumpp, S., and Lipowsky, R., 2005. Cooperative cargo transport by several molecular motors. PNAS, 102(48), pg. 17284-17289.
  • (16) Varga, V., Leduc, C., Bormuth, V., Diez, S., and Howard, J., 2009. Kinesin-8 motors act cooperatively to mediate length-dependent microtubule depolymerization. Cell, 138, pg. 1174-1183.
  • (17) Liang, W. H., Li, Q., Rifat Faysal, K. M., King, S. J., Gopinathan, A., and Xu, J., 2016. Microtubule defects influence kinesin-based transport in vitro. Biophysical Journal, 110, pg. 2229-2240.
  • (18) Mallik, R., and Gross, S. P., 2004. Molecular motors: Strategies to get along. Current Biology, 14, pg. R971-R982.
  • (19) Müller, M. J. I., Klumpp, S., and Lipowsky, R., 2005. Molecular motor traffic in a half-open tube. Journal of Physics: Condensed Matter, 17, pg. S3839-S3850.
  • (20) Rossi, L. W., Radtke, P. K., and Goldman, C., 2014. Long-range cargo transport on crowded microtubules: The motor jamming mechanism. Physica A, 401, pg. 319-329.
  • (21) Striebel, M., Graf, I. R., and Frey, E., 2020. A mechanistic view of collective filament motion in active nematic networks. Biophysical Journal, 118(2), pg. 313-324.
  • (22) Dallon, J. C., Leduc, C., Etienne-Manneville, S., Portet, S., 2019. Stochastic modeling reveals how motor protein and filament properties affect intermediate filament transport. Journal of Theoretical Biology, 464, pg. 132-148.

Appendix A Homogeneous microtubules: Proof of Theorem 2.1 and Corollary 1

Equation (14) may be rewritten in the form of a system of two first order ODEs for density ρε\rho_{\varepsilon} and flux JεJ_{\varepsilon} (see also (11)):

{ε2ρε=v01Jε+ρε(1ρε),Jε=ΩA(ΩA+ΩD)ρε.\left\{\begin{array}[]{rl}\dfrac{\varepsilon}{2}\rho_{\varepsilon}^{\prime}&=-v_{0}^{-1}J_{\varepsilon}+\rho_{\varepsilon}(1-\rho_{\varepsilon}),\\ J_{\varepsilon}^{\prime}&=\Omega_{A}-(\Omega_{A}+\Omega_{D})\rho_{\varepsilon}.\end{array}\right. (43)
Refer to caption
Refer to caption
Figure 4: Left: Phase portrait for (43) with ε=0.01\varepsilon=0.01, ΩA=0.7\Omega_{A}=0.7 and ΩD=0.3\Omega_{D}=0.3; Right: Sketch of the phase portrait for (43) with ε1\varepsilon\ll 1, the black circle represents the stationary point.

Next, we discuss the phase portrait for this system with ε1\varepsilon\ll 1, depicted in Fig. 4. Away from curve γ\gamma defined by

γ:={(ρ,J)|J=v0ρ(1ρ) and 0ρ1,0Jv0/4},\gamma:=\left\{(\rho,J)\left|J=v_{0}\rho(1-\rho)\text{ and }\begin{array}[]{l}0\leq\rho\leq 1,\\ 0\leq J\leq v_{0}/4\end{array}\right.\right\}, (44)

the trajectories of (43), parametrized by 0x0\leq x\leq\ell, have almost horizontal slope in (ρ,J)(\rho,J) plane. This is because the slope of ρε\rho_{\varepsilon} is of the order ε1\varepsilon^{-1}, that is ρε(x)ε1\rho_{\varepsilon}^{\prime}(x)\sim\varepsilon^{-1}, whenever the point (ρε(x),Jε(x))(\rho_{\varepsilon}(x),J_{\varepsilon}(x)) is away from γ\gamma (it follows from the first equation in (43)). It would be natural to expect that as ε\varepsilon vanishes, trajectory {(ρε(x),Jε(x)),0x}\left\{(\rho_{\varepsilon}(x),J_{\varepsilon}(x)),0\leq x\leq\ell\right\} approaches the arch γ\gamma and this trajectory is contained in a given thin neighborhood of γ\gamma for sufficiently small ε\varepsilon. In this subsection, it will be shown that the behavior of the solution is more complicated than simply evolving near γ\gamma.

To describe how the solution ρε(x)\rho_{\varepsilon}(x) behaves for ε1\varepsilon\ll 1, we introduce the following notation for parts of curve γ\gamma. Namely,

γl\displaystyle\gamma_{l} :=\displaystyle:= γ{0ρ<0.5},\displaystyle\gamma\cap\left\{0\leq\rho<0.5\right\},
γr,+\displaystyle\gamma_{r,+} :=\displaystyle:= γ{0.5ρρeq},\displaystyle\gamma\cap\left\{0.5\leq\rho\leq\rho_{\text{eq}}\right\},
γr,\displaystyle\gamma_{r,-} :=\displaystyle:= γ{ρeqρ1}.\displaystyle\gamma\cap\left\{\rho_{\text{eq}}\leq\rho\leq 1\right\}.

Here ρeq:=ΩA/(ΩA+ΩD)\rho_{\text{eq}}:=\Omega_{A}/(\Omega_{A}+\Omega_{D}). Let us also introduce the following horizontal segment

Γ:={(ρ,J):J=v0/4,0ρ0.5},\Gamma:=\left\{(\rho,J):J=v_{0}/4,~{}0\leq\rho\leq 0.5\right\},

and the solution g(x;s,a)g(x;s,a) to (16), i.e., the initial value problem of the first order obtained by the formal limit as ε0\varepsilon\to 0 in (14):

v0(12g)xg=ΩA(ΩA+ΩD)g,g(s;s,a)=a.v_{0}(1-2g)\partial_{x}g=\Omega_{A}-(\Omega_{A}+\Omega_{D})g,~{}~{}~{}g(s;s,a)=a. (45)

First, note that γl\gamma_{l}, which is the left part of the curve γ\gamma, is unstable, that is all trajectories, excluding γl\gamma_{l}, are directed away from γl\gamma_{l} in the vicinity of γl\gamma_{l}. The right part of the curve γ\gamma, consisting of curve segments γr,+\gamma_{r,+} and γr,\gamma_{r,-}, is stable, attracting all trajectories in its vicinity, except those that follow Γ\Gamma. We note that this exception, when γr,+\gamma_{r,+} loses its stability, occurs at the interface point (ρ=1/2,J=v0/4)(\rho=1/2,J=v_{0}/4) where γr,+\gamma_{r,+} meets γl\gamma_{l}. All trajectories reaching this point near (not necessarily intersecting) the curves γr,+\gamma_{r,+} and γl\gamma_{l} continue along Γ\Gamma.

Given specific values of α,β(0,1)\alpha,\beta\in(0,1) in boundary conditions (15), the statement of Theorem 2.1 as well as representation formula (20) can be simply verified by careful inspection of the phase portrait depicted in Fig. 4. Specifically, for all 0<α,β<10<\alpha,\beta<1, one can draw a path {(ρ(x),J(x)):0x}\left\{(\rho(x),J(x)):0\leq x\leq\ell\right\} along arrows in Fig. 4 (right), which starts at vertical line ρ=α\rho=\alpha and ends at vertical line ρ=1β\rho=1-\beta, and such a path will be unique for given α\alpha and β\beta (see also left column of Fig. 5 for specific examples). Instead of checking each couple (α,β)(\alpha,\beta), one would split ranges of (α,β)(\alpha,\beta) into sub-domains within which the outer solution has constant or smoothly varying shape, as it is done in proof below.

Proof of Theorem 2.1. Consider the following functions:

ρα(x)=g(x;0,α) and ρβ(x)=g(x;,max{0.5,1β}).\rho_{\alpha}(x)=g(x;0,\alpha)\text{ and }\rho_{\beta}(x)=g(x;\ell,\max\left\{0.5,1-\beta\right\}).

These functions can be thought of as one-sided solutions (i.e., satisfying one of the boundary conditions, either ρ(0)=α\rho(0)=\alpha or ρ()=max{0.5,1β}\rho(\ell)=\max\left\{0.5,1-\beta\right\}) of Equation (14) for ε=0\varepsilon=0. The reason we choose ρ()=max{0.5,1β}\rho(\ell)=\max\left\{0.5,1-\beta\right\} instead of ρβ()=1β\rho_{\beta}(\ell)=1-\beta is because there is no solution continuous at x=x=\ell with ρ()<0.5\rho(\ell)<0.5 as visible in Fig. 4 (curve γ\gamma is unstable in region {0ρ<0.5}\left\{0\leq\rho<0.5\right\}).

Introduce also the corresponding fluxes:

Jα(x)=v0ρα(x)(1ρα(x)) and Jβ(x)=v0ρβ(x)(1ρβ(x)).J_{\alpha}(x)=v_{0}\rho_{\alpha}(x)(1-\rho_{\alpha}(x))\text{ and }J_{\beta}(x)=v_{0}\rho_{\beta}(x)(1-\rho_{\beta}(x)).

From the definition of function gg it follows that Jα(x)J_{\alpha}(x) and Jβ(x)J_{\beta}(x) are both monotonic functions, and function Jβ(x)J_{\beta}(x) is defined for all 0x<0\leq x<\ell. Moreover, Jβ(x)J_{\beta}(x) can be extended onto (,](-\infty,\ell] and

limxJβ(x)=Jeq, where Jeq:=v0ΩAΩD(ΩA+ΩD)2.\lim\limits_{x\to-\infty}J_{\beta}(x)=J_{\text{eq}},\text{ where }J_{\text{eq}}:=v_{0}\dfrac{\Omega_{A}\Omega_{D}}{(\Omega_{A}+\Omega_{D})^{2}}.

Consider case α0.5\alpha\geq 0.5. From Fig. 4, it follows that a trajectory emanating for initial point (α,J)(\alpha,J) for any 0<J<v0/40<J<v_{0}/4 immediately reaches γr\gamma_{r} and stays on γrΓ\gamma_{r}\cup\Gamma for 0<x0<x\leq\ell. Thus, at x=0x=0 trajectory {(ρ0(x),J0(x)):0x}\left\{(\rho_{0}(x),J_{0}(x)):0\leq x\leq\ell\right\}, describing the outer solution, jumps from (α,J0(0))(\alpha,J_{0}(0)) at t=0t=0 to γr\gamma_{r}:

ρ0(x)={α,x=0,ρβ(x),0<x.\rho_{0}(x)=\left\{\begin{array}[]{ll}\alpha,&x=0,\\ \rho_{\beta}(x),&0<x\leq\ell.\end{array}\right. (46)

In the case where α<0.5\alpha<0.5, denote by 0xJ0\leq x_{J}\leq\ell location at which fluxes Jα(x)J_{\alpha}(x) and Jβ(x)J_{\beta}(x) intersect, that is,

Jα(xJ)=Jβ(xJ).J_{\alpha}(x_{J})=J_{\beta}(x_{J}). (47)

Equality (47) implies that either ρα(xJ)=1ρβ(xJ)\rho_{\alpha}(x_{J})=1-\rho_{\beta}(x_{J}) or ρα(xJ)=ρβ(xJ)\rho_{\alpha}(x_{J})=\rho_{\beta}(x_{J}). If ρα(xJ)=ρβ(xJ)\rho_{\alpha}(x_{J})=\rho_{\beta}(x_{J}), then since ρα\rho_{\alpha} and ρβ\rho_{\beta} are solutions of the same first order ordinary differential equation, these two functions coincide ρα(x)ρβ(x)\rho_{\alpha}(x)\equiv\rho_{\beta}(x).

We show now that either

there exists at most one xJ1x_{J}\leq 1 or ρα(x)ρβ(x)\rho_{\alpha}(x)\equiv\rho_{\beta}(x). (48)

Indeed, since α<0.5\alpha<0.5, trajectory (ρα(x),Jα(x))(\rho_{\alpha}(x),J_{\alpha}(x)) evolves on γl\gamma_{l} for all 0x0\leq x\leq\ell where solution ρα(x)\rho_{\alpha}(x) exists, and Jα(x)J_{\alpha}(x) monotonically increases. Trajectory (ρβ(x),Jβ(x))(\rho_{\beta}(x),J_{\beta}(x)) evolves also for all 0x0\leq x\leq\ell within either γr,+\gamma_{r,+} or γr,\gamma_{r,-}. If (ρβ(x),Jβ(x))(\rho_{\beta}(x),J_{\beta}(x)) evolves within γr,\gamma_{r,-}, then Jβ(x)J_{\beta}(x) is monotonically decreasing in xx whereas Jα(x)J_{\alpha}(x) is monotonically increasing xx, and thus equation Jα(x)=Jβ(x)J_{\alpha}(x)=J_{\beta}(x) can have at most one root in this case. If (ρβ(x),Jβ(x))(\rho_{\beta}(x),J_{\beta}(x)) evolves within γr,+\gamma_{r,+}, then both Jα(x)J_{\alpha}(x) and Jβ(x)J_{\beta}(x) increase with xx. Assume that there are at least two distinct numbers xJ(1)x_{J}^{(1)}, xJ(2)x_{J}^{(2)} such that xJ(1)<xJ(2)x_{J}^{(1)}<x_{J}^{(2)} and Jα(xJ(i))=Jβ(xJ(i))J_{\alpha}(x_{J}^{(i)})=J_{\beta}(x_{J}^{(i)}), i=1,2i=1,2. Assume also that xJ(1)x_{J}^{(1)} and xJ(2)x_{J}^{(2)} are neighbor roots of equation Jα(x)=Jβ(x)J_{\alpha}(x)=J_{\beta}(x), i.e., for all x(xJ(1),xJ(2))x\in(x_{J}^{(1)},x_{J}^{(2)}) we have Jα(x)Jβ(x)J_{\alpha}(x)\neq J_{\beta}(x). Then due to

xJ=ΩA(ΩA+ΩD)g, where J(x)=v0g(x)(1g(x))\partial_{x}J=\Omega_{A}-(\Omega_{A}+\Omega_{D})g,\text{ where }J(x)=v_{0}g(x)(1-g(x))

and ρα(xJ(i))<0.5\rho_{\alpha}(x_{J}^{(i)})<0.5 ρα(xJ(i))>0.5\rho_{\alpha}(x_{J}^{(i)})>0.5, i=1,2i=1,2, we have that xJα(xJ(i))>xJβ(xJ(i))\partial_{x}J_{\alpha}(x_{J}^{(i)})>\partial_{x}J_{\beta}(x_{J}^{(i)}), i=1,2i=1,2. Noting that a smooth function can’t have the same sign of its derivative at two successive roots we arrive to contradiction. Therefore, such xJx_{J} is at most one and (48) is shown.

If Jα(x)Jβ(x)J_{\alpha}(x)\neq J_{\beta}(x) for all 0x10\leq x\leq 1, then define xJx_{J} as follows:

xJ={0,Jβ(x)<Jα(x) for all 0<x<,1,Jα(x)<Jβ(x) for all 0<x<.x_{J}=\left\{\begin{array}[]{ll}0,&J_{\beta}(x)<J_{\alpha}(x)\text{ for all }0<x<\ell,\\ 1,&J_{\alpha}(x)<J_{\beta}(x)\text{ for all }0<x<\ell.\end{array}\right.

We note that point x=xJx=x_{J} is where the outer solution jumps from ρα(x)\rho_{\alpha}(x) to ρβ(x)\rho_{\beta}(x), thus

ρ0(x)={ρα(x),0x<xJ,ρβ(x),xJ<x<.\rho_{0}(x)=\left\{\begin{array}[]{ll}\rho_{\alpha}(x),&0\leq x<x_{J},\\ \rho_{\beta}(x),&x_{J}<x<\ell.\end{array}\right. (49)

and ρ0()=1β\rho_{0}(\ell)=1-\beta.

Formulas (46), (49), and (15) complete the proof of Theorem 2.1.

\square

Appendix B Examples of solutions given by (20)

To illustrate the result of Theorem 2.1 we continue with the following examples. We take v0=1v_{0}=1, =1\ell=1, ΩA=0.8\Omega_{A}=0.8 and ΩD=0.2\Omega_{D}=0.2, and we vary the boundary rates α\alpha and β\beta. The outer solution for each example, as both a trajectory in (ρ,J)(\rho,J) plane and the plot of ρ0(x)\rho_{0}(x), is depicted in Fig. 5.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: Left: The thick line represents the trajectories from Examples 1-4; it starts at ρ=α\rho=\alpha and ends at ρ=1β\rho=1-\beta, the black circle at (0.8,0.16) represents the stationary solution. Right: The thick line represents the outer solution ρ0(x)\rho_{0}(x) for Examples 1-4. In Examples 2 and 3, branches g(x;0,α)g(x;0,\alpha) and g(x;1,max{0.5,1β})g(x;1,\max\{0.5,1-\beta\}) extend slightly beyond the intervals where they are a part of the outer solution ρ0(x)\rho_{0}(x) (thin curves).

Example 1. α=0.4\alpha=0.4 and β=0.39\beta=0.39.

ρ0(x)={0.4,x=0g(x;1,0.61),0<x1.\rho_{0}(x)=\left\{\begin{array}[]{ll}0.4,&x=0\\ g(x;1,0.61),&0<x\leq 1.\end{array}\right.

Example 2. α=0.1\alpha=0.1 and β=0.4\beta=0.4.

ρ0(x)={g(x;0,0.1),0xxJ,xJ0.133g(x;1,0.6),xJ<x1.\rho_{0}(x)=\left\{\begin{array}[]{ll}g(x;0,0.1),&0\leq x\leq x_{J},\,x_{J}\approx 0.133\\ g(x;1,0.6),&x_{J}<x\leq 1.\end{array}\right.

Example 3. α=0.1\alpha=0.1 and β=0.85\beta=0.85.

ρ0(x)={g(x;0,0.1),0xxJ,xJ0.135,g(x;1,1/2),xJ<x<1,0.15,x=1.\rho_{0}(x)=\left\{\begin{array}[]{ll}g(x;0,0.1),&0\leq x\leq x_{J},\,x_{J}\approx 0.135,\\ g(x;1,1/2),&x_{J}<x<1,\\ 0.15,&x=1.\end{array}\right.

Example 4. α=0.9\alpha=0.9 and β=0.8\beta=0.8.

ρ0(x)={0.9,x=0,g(x;1,1/2),0<x<1,0.2,x=1.\rho_{0}(x)=\left\{\begin{array}[]{ll}0.9,&x=0,\\ g(x;1,1/2),&0<x<1,\\ 0.2,&x=1.\end{array}\right.

The case xJ>1x_{J}>1 corresponds to the case of fast motor proteins or, more precisely, unidirectional motion dominates attachment/detachment, and thus resulting density is low in MT, ρ0(x)<0.5\rho_{0}(x)<0.5 for x(0,1)x\in(0,1). Consider the following example:

Example 5. α=0.05\alpha=0.05, β=0.85\beta=0.85, ΩA=0.16\Omega_{A}=0.16 and ΩD=0.04\Omega_{D}=0.04.

ρ0(x)={g(x,0,α),0x<1,1β,x=1.\rho_{0}(x)=\left\{\begin{array}[]{ll}g(x,0,\alpha),&0\leq x<1,\\ 1-\beta,&x=1.\end{array}\right.

The solution is depicted in Fig. 6.

Refer to caption
Refer to caption
Figure 6: The thick line represents the trajectories from Example 5; it starts at ρ=α\rho=\alpha and ends at ρ=1β\rho=1-\beta, the black circle at (0.8,0.16) represents the stationary solution. Right: The thick line represents the outer solution ρ0(x)\rho_{0}(x) for Example 5.