Skip to main content
Log in

Assessing the Impact of Electrostatic Drag on Processive Molecular Motor Transport

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

Abstract

The bidirectional movement of intracellular cargo is usually described as a tug-of-war among opposite-directed families of molecular motors. While tug-of-war models have enjoyed some success, recent evidence suggests underlying motor interactions are more complex than previously understood. For example, these tug-of-war models fail to predict the counterintuitive phenomenon that inhibiting one family of motors can decrease the functionality of opposite-directed transport. In this paper, we use a stochastic differential equations modeling framework to explore one proposed physical mechanism, called microtubule tethering, that could play a role in this “co-dependence” among antagonistic motors. This hypothesis includes the possibility of a trade-off: weakly bound trailing molecular motors can serve as tethers for cargoes and processing motors, thereby enhancing motor–cargo run lengths along microtubules; however, this introduces a cost of processing at a lower mean velocity. By computing the small- and large-time mean-squared displacement of our theoretical model and comparing our results to experimental observations of dynein and its “helper protein” dynactin, we find some supporting evidence for microtubule tethering interactions. We extrapolate these findings to predict how dynein–dynactin might interact with the opposite-directed kinesin motors and introduce a criterion for when the trade-off is beneficial in simple systems.

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

Similar content being viewed by others

References

  • Abramowitz M, Stegun IA (1964) Handbook of mathematical functions: with formulas, graphs, and mathematical tables, vol 55. Courier Corporation, North Chelmsford

    MATH  Google Scholar 

  • Ali MY, Lu H, Bookwalter CS, Warshaw DM, Trybus KM (2008) Myosin V and kinesin act as tethers to enhance each others’ processivity. Proc Nat Acad Sci 105(12):4691–4696

    Article  Google Scholar 

  • Ayloo S, Lazarus JE, Dodda A, Tokito M, Ostap EM, Holzbaur EL (2014) Dynactin functions as both a dynamic tether and brake during dynein-driven motility. Nat Commun 5:4807

    Article  Google Scholar 

  • Berezuk MA, Schroer TA (2007) Dynactin enhances the processivity of kinesin-2. Traffic 8(2):124–129

    Article  Google Scholar 

  • Berger F, Müller M, Lipowsky R (2009) Enhancement of the processivity of kinesin-transported cargo by myosin V. Europhys Lett EPL 87(2):28,002

    Article  Google Scholar 

  • Berman A, Plemmons RJ (1994) Nonnegative matrices in the mathematical sciences. SIAM, Philadelphia

    Book  MATH  Google Scholar 

  • Bieling P, Telley IA, Piehler J, Surrey T (2008) Processive kinesins require loose mechanical coupling for efficient collective motility. EMBO Rep 9(11):1121–1127

    Article  Google Scholar 

  • Bouzat S (2016) Models for microtubule cargo transport coupling the Langevin equation to stochastic stepping motor dynamics: caring about fluctuations. Phys Rev E 93(1):012,401

    Article  MathSciNet  Google Scholar 

  • Bruno L, Salierno M, Wetzler DE, Despósito MA, Levi V (2011) Mechanical properties of organelles driven by microtubule-dependent molecular motors in living cells. PLoS ONE 6(4):e18,332

    Article  Google Scholar 

  • Burgess SA, Walker ML, Sakakibara H, Knight PJ, Oiwa K (2003) Dynein structure and power stroke. Nature 421(6924):715–718

    Article  Google Scholar 

  • Carter NJ, Cross R (2005) Mechanics of the kinesin step. Nature 435(7040):308–312

    Article  Google Scholar 

  • Chen A, McKinley SA, Wang S, Shi F, Mucha PJ, Forest MG, Lai SK (2014) Transient antibody-mucin interactions produce a dynamic molecular shield against viral invasion. Biophys J 106(9):2028–2036

    Article  Google Scholar 

  • Culver-Hanlon TL, Lex SA, Stephens AD, Quintyne NJ, King SJ (2006) A microtubule-binding domain in dynactin increases dynein processivity by skating along microtubules. Nat Cell Biol 8(3):264–270

    Article  Google Scholar 

  • Feng Q, Mickolajczyk KJ, Chen GY, Hancock WO (2018) Motor reattachment kinetics play a dominant role in multimotor-driven cargo transport. Biophys J 114(2):400–409

    Article  Google Scholar 

  • Gill SR, Schroer TA, Szilak I, Steuer ER, Sheetz MP, Cleveland DW (1991) Dynactin, a conserved, ubiquitously expressed component of an activator of vesicle motility mediated by cytoplasmic dynein. J Cell Biol 115(6):1639–1650

    Article  Google Scholar 

  • Gross SP, Welte MA, Block SM, Wieschaus EF (2000) Dynein-mediated cargo transport in vivo a switch controls travel distance. J Cell Biol 148(5):945–956

    Article  Google Scholar 

  • Hackney DD (1994) Evidence for alternating head catalysis by kinesin during microtubule-stimulated ATP hydrolysis. Proc Nat Acad Sci 91(15):6865–6869

    Article  Google Scholar 

  • Hancock WO (2014) Bidirectional cargo transport: moving beyond tug of war. Nat Rev Mol Cell Biol 15(9):615–628

    Article  Google Scholar 

  • Hancock WO (2018) Mechanics of bidirectional cargo transport. In: King SM (ed) Dyneins: structure, biology and disease: dynein mechanics, dysfunction, and disease, chapter 7, vol 2. Academic Press, New York, pp 152–171

    Google Scholar 

  • Hariharan V, Hancock WO (2009) Insights into the mechanical properties of the kinesin neck linker domain from sequence analysis and molecular dynamics simulations. Cell Mol Bioeng 2(2):177–189

    Article  Google Scholar 

  • Hughes J, Hancock WO, Fricks J (2012) Kinesins with extended neck linkers: a chemomechanical model for variable-length stepping. Bull Math Biol 74(5):1066–1097

    Article  MathSciNet  MATH  Google Scholar 

  • Johnson RW (1998) Handbook of fluid dynamics. CRC Press, Boca Raton

    MATH  Google Scholar 

  • Jun Y, Tripathy SK, Narayanareddy BR, Mattson-Hoss MK, Gross SP (2014) Calibration of optical tweezers for in vivo force measurements: how do different approaches compare? Biophys J 107(6):1474–1484

    Article  Google Scholar 

  • Kardon JR, Reck-Peterson SL, Vale RD (2009) Regulation of the processivity and intracellular localization of saccharomyces cerevisiae dynein by dynactin. Proc Nat Acad Sci 106(14):5669–5674

    Article  Google Scholar 

  • Kincaid MM, King SJ (2006) Motors and their tethers: the role of secondary binding sites in processive motility. Cell Cycle 5(23):2733–2737

    Article  Google Scholar 

  • King SJ, Schroer TA (2000) Dynactin increases the processivity of the cytoplasmic dynein motor. Nat Cell Biol 2(1):20

    Article  Google Scholar 

  • Klumpp S, Lipowsky R (2005) Cooperative cargo transport by several molecular motors. Proc Nat Acad Sci USA 102(48):17,284–17,289

    Article  Google Scholar 

  • Kojima H, Muto E, Higuchi H, Yanagida T (1997) Mechanics of single kinesin molecules measured by optical trapping nanometry. Biophys J 73(4):2012–2022

    Article  Google Scholar 

  • Korn CB, Klumpp S, Lipowsky R, Schwarz US (2009) Stochastic simulations of cargo transport by processive molecular motors. J Chem Phys 131(24):12B624

    Article  Google Scholar 

  • Kubo R (1966) The fluctuation–dissipation theorem. Rep Prog Phys 29(1):255

    Article  MATH  Google Scholar 

  • Kunwar A, Vershinin M, Xu J, Gross SP (2008) Stepping, strain gating, and an unexpected force-velocity curve for multiple-motor-based transport. Curr Biol 18(16):1173–1183

    Article  Google Scholar 

  • Kunwar A, Tripathy SK, Xu J, Mattson MK, Anand P, Sigua R, Vershinin M, McKenney RJ, Clare CY, Mogilner A et al (2011) Mechanical stochastic tug-of-war models cannot explain bidirectional lipid-droplet transport. Proc Nat Acad Sci 108(47):18,960–18,965

    Article  Google Scholar 

  • Kurtz Thomas J (1981) Approximation of population processes. SIAM, Philadelphia

    Book  MATH  Google Scholar 

  • Kutys ML, Fricks J, Hancock WO (2010) Monte carlo analysis of neck linker extension in kinesin molecular motors. PLoS Comput Biol 6(11):e1000,980

    Article  MathSciNet  Google Scholar 

  • Mallik R, Carter BC, Lex SA, King SJ, Gross SP (2004) Cytoplasmic dynein functions as a gear in response to load. Nature 427(6975):649–652

    Article  Google Scholar 

  • McKinley SA, Athreya A, Fricks J, Kramer PR (2012) Asymptotic analysis of microtubule-based transport by multiple identical molecular motors. J Theor Biol 305:54–69

    Article  MathSciNet  Google Scholar 

  • Minoura I, Katayama E, Sekimoto K, Muto E (2010) One-dimensional brownian motion of charged nanoparticles along microtubules: a model system for weak binding interactions. Biophys J 98(8):1589–1597

    Article  Google Scholar 

  • Müller MJ, Klumpp S, Lipowsky R (2008) Tug-of-war as a cooperative mechanism for bidirectional cargo transport by molecular motors. Proc Nat Acad Sci 105(12):4609–4614

    Article  Google Scholar 

  • Müller MJ, Klumpp S, Lipowsky R (2010) Bidirectional transport by molecular motors: enhanced processivity and response to external forces. Biophys J 98(11):2610–2618

    Article  Google Scholar 

  • Muthukrishnan G, Zhang Y, Shastry S, Hancock WO (2009) The processivity of kinesin-2 motors suggests diminished front-head gating. Curr Biol 19(5):442–447

    Article  Google Scholar 

  • Okada Y, Hirokawa N (2000) Mechanism of the single-headed processivity: diffusional anchoring between the K-loop of kinesin and the C terminus of tubulin. Proc Nat Acad Sci 97(2):640–645

    Article  Google Scholar 

  • Posta F, D’Orsogna MR, Chou T (2009) Enhancement of cargo processivity by cooperating molecular motors. Phys Chem Chem Phys 11(24):4851–4860

    Article  Google Scholar 

  • Segal M, Soifer I, Petzold H, Howard J, Elbaum M, Reiner O (2012) Ndel1-derived peptides modulate bidirectional transport of injected beads in the squid giant axon. Biol Open 1:220–231

    Article  Google Scholar 

  • Shubeita GT, Tran SL, Xu J, Vershinin M, Cermelli S, Cotton SL, Welte MA, Gross SP (2008) Consequences of motor copy number on the intracellular transport of kinesin-1-driven lipid droplets. Cell 135(6):1098–1107

    Article  Google Scholar 

  • Svoboda K, Schmidt CF, Schnapp BJ, Block SM (1993) Direct observation of kinesin stepping by optical trapping interferometry. Nature 365(6448):721

    Article  Google Scholar 

  • Toba S, Watanabe TM, Yamaguchi-Okimoto L, Toyoshima YY, Higuchi H (2006) Overlapping hand-over-hand mechanism of single molecular motility of cytoplasmic dynein. Proc Nat Acad Sci 103(15):5741–5745

    Article  Google Scholar 

  • Urbakh M, Klafter J, Gourdon D, Israelachvili J (2004) The nonlinear nature of friction. Nature 430(6999):525

    Article  Google Scholar 

  • Urnavicius L, Zhang K, Diamant AG, Motz C, Schlager MA, Yu M, Patel NA, Robinson CV, Carter AP (2015) The structure of the dynactin complex and its interaction with dynein. Science 347(6229):1441–1446

    Article  Google Scholar 

  • Visscher K, Schnitzer MJ, Block SM (1999) Single kinesin molecules studied with a molecular force clamp. Nature 400(6740):184–189

    Article  Google Scholar 

  • Waterman-Storer CM, Karki S, Holzbaur E (1995) The p150glued component of the dynactin complex binds to both microtubules and the actin-related protein centractin (Arp-1). Proc Nat Acad Sci 92(5):1634–1638

    Article  Google Scholar 

  • Wiersema UF (2008) Brownian motion calculus. Wiley, New York

    MATH  Google Scholar 

  • Witten J, Ribbeck K (2017) The particle in the spiders web: transport through biological hydrogels. Nanoscale 9:8080–8095

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to thank Will Hancock, John Fricks, Pete Kramer, Veronica Ciocanel, and Erkan Tüzel for their thoughtful and very informative feedback in the development of this work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. Darby Smith.

Additional information

This work was supported by NIH Grant Number R01GM122082-01.

Appendices

Appendix

A Solution for Pinned Cargo System

The SDE used for pinned diffusion is the Ornstein–Uhlenbeck equation with nonzero mean. For notational efficiency, we write it as follows

$$\begin{aligned} \mathrm {d}X(t) = - \alpha (X(t) - \mu ) \mathrm {d}t + \sigma \mathrm {d}W(t), \end{aligned}$$
(31)

with \(X(0) = x\). Because this SDE is linear, with additive noise, we can write the solution in convolution form using Duhamel’s formula

$$\begin{aligned} \begin{aligned} X(t)&= e^{-\alpha t} x + \int _0^t e^{-\alpha (t-u)} (-\alpha \mu ) \mathrm {d}u + \int _0^t e^{-\alpha (t-u)} \sigma \mathrm {d}W(u) \\&= e^{-\alpha t} x + (1 - e^{-\alpha t}) \, \mu \, + \sigma \int _0^t e^{-\alpha (t - u)} \mathrm {d}W(u). \end{aligned} \end{aligned}$$

Allowing for the possibility that the initial condition x and the center \(\mu \) can be random variables, two quantities of interest follow immediately:

$$\begin{aligned} \mathbb {E}[X(t)]&= e^{-\alpha t} \mathbb {E}[x] + (1- e^{-\alpha t}) \, \mathbb {E}[\mu ] \end{aligned}$$
(32)
$$\begin{aligned} \mathbb {E}[X(t) \mu ]&= e^{-\alpha t} \mathbb {E}[x \mu ] + (1- e^{-\alpha t}) \, \mathbb {E}[\mu ^2]. \end{aligned}$$
(33)

To compute the MSD of X(t), we apply Itô’s formula to (31) using the function \(f(x) = x^2\) to find that

$$\begin{aligned} \mathrm {d}X^2(t) = - 2 \alpha X(t) (X(t) - \mu ) + \sigma ^2 \mathrm {d}t + 2 \sigma X(t) \mathrm {d}W(t). \end{aligned}$$

It follows that

$$\begin{aligned} \frac{\mathrm {d}}{\mathrm {d}t} \mathbb {E}\big [X^2(t)\big ] = -2 \alpha \mathbb {E}\big [X^2(t)\big ] + 2 \alpha \mathbb {E}[X(t) \mu ] + \sigma ^2, \end{aligned}$$

which, again by Duhamel’s formula has the solution

$$\begin{aligned} \begin{aligned} \mathbb {E}\big [X^2(t)\big ]&= e^{-2\alpha t} \mathbb {E}[x^2] + \int _0^t e^{-2 \alpha (t - u)} (2 \alpha \mathbb {E}[X(u) \mu ] + \sigma ^2) \mathrm {d}u \\&= e^{-2\alpha t} \mathbb {E}[x^2] + 2 \alpha \int _0^t e^{-2 \alpha (t - u)} \mathbb {E}[X(u) \mu ] \mathrm {d}u + \big (1 - e^{-2 \alpha t}\big ) \frac{\sigma ^2}{2 \alpha }. \end{aligned} \end{aligned}$$

Substituting in 32, the second term on the right-hand side evaluates to

$$\begin{aligned} 2 \alpha \int _0^t e^{-2 \alpha (t - u)} \mathbb {E}[X(u) \mu ] \mathrm {d}u = 2 e^{-\alpha t} (1 - e^{-2 \alpha t}) \mathbb {E}[x\mu ] + (1 - e^{-\alpha t})^2 \mathbb {E}[\mu ^2]. \end{aligned}$$
(34)

We conclude that

$$\begin{aligned} \mathbb {E}[X^2(t)] = \mathbb {E}\left[ \left( e^{-\alpha t} x + (1 - e^{-\alpha t}) \mu \right) ^2\right] + (1 - e^{-2\alpha t}) \frac{\sigma ^2}{2 \alpha }. \end{aligned}$$
(35)

Expressing this in terms of the parameters used in the main text

$$\begin{aligned} \alpha = \frac{\kappa }{\gamma }, \, \sigma = \sqrt{\frac{2 k_B T}{\gamma }}, \, x = 0, \mu \sim N\left( 0, \frac{k_B T}{\kappa }\right) , \end{aligned}$$

we have

$$\begin{aligned} \mathbb {E}[X^2(t)] = 2 (1 - e^{-\kappa t /\gamma }) \frac{k_B T}{\kappa }. \end{aligned}$$
(36)

B Proof Omitted from the Main Text

Proof

(of Lemma 1) Since the rows of M sum to zero, \(\lambda _0 = 0\) is an eigenvalue of M with right eigenvector \(\mathbf {v}_0 = \left( 1,1,\ldots ,1\right) ^\top \). Let \(\mathbf {u}_0\) be the left eigenvector associated with 0 for M.

Since the diagonal of M is negative, there exists some positive real number r such that \(M' = M + r\mathbb {I}\) is a nonnegative matrix. Note that \(\lambda \), \(\mathbf {v}_\lambda \) are an eigenvalue-eigenvector pair for M if and only if \(\lambda +r\) and \(\mathbf {v}_{\lambda }\) are an eigenvalue-eigenvector pair for \(M'\).

Now we show that \(M'\) is irreducible. \(M'\) differs from M only on the diagonal. Therefore, the associated directed graph of \(M'\) will have the same edge values between distinct vertices as that of M. Since the associated graph for M is strongly connected, the one for \(M'\) is strongly connected as well and consequently \(M'\) is irreducible.

Since \(M'\) is a nonnegative, irreducible matrix, the Perron–Frobenius theorem holds (Berman and Plemmons 1994). Thus, there exists a simple real eigenvalue \(\lambda _p>0\) for \(M'\) such that the spectral radius of \(M'\) is \(\lambda _p\), all other eigenvalues of \(M'\) must have real part strictly less that \(\lambda _p\), and the only eigenvectors whose components are all real and positive are those eigenvectors associated with \(\lambda _p\). Note that

$$\begin{aligned} M'\mathbf {v}_0 = M\mathbf {v}_0 + r\mathbf {v}_0 = r\mathbf {v}_0. \end{aligned}$$

Ergo, since \(\mathbf {v}_0\) is a eigenvector whose components are all real and positive, we must have that \(\lambda _p =r\). Further,

$$\begin{aligned} \mathbf {u}_0 M' = \mathbf {u}_0M + \mathbf {u}_0r = \mathbf {u}_0r. \end{aligned}$$

Since \(\mathbf {u}_0\) is an eigenvector associated with \(\lambda _p = r\), its components are all real and positive.

As r is a simple eigenvalue for \(M'\) and all other eigenvalues of \(M'\) have real part strictly less than r, it now follows that 0 is a simple eigenvalue for M and all other eigenvalues of M have real part strictly less than 0.

Write the remaining eigenvalues of M as

$$\begin{aligned} 0=\lambda _0 < \left| {\lambda _1}\right| \le \left| {\lambda _2}\right| \le \cdots \le \left| {\lambda _k}\right| , \end{aligned}$$

where we assume that \(\lambda _i\) has multiplicity \(m_i\). Write M in Jordan normal form. That is write \(M = PJP^{-1}\) where

$$\begin{aligned} J = \begin{pmatrix} 0&{}\quad &{}\quad &{}\quad \\ &{}\quad J_1 &{}\quad &{}\quad \\ {} &{}\quad &{}\quad \ddots &{}\quad \\ &{}\quad &{}\quad &{}\quad J_k\end{pmatrix}, \end{aligned}$$

and \(J_i\) is an \(m_i\times m_i\) matrix of the form

$$\begin{aligned} J_i = \begin{pmatrix} \lambda _i &{}\quad 1 &{}\quad &{}\quad &{}\quad &{}\quad \\ &{}\quad \lambda _i &{}\quad 1&{}\quad &{}\quad &{}\quad \\ &{}\quad &{}\quad \ddots &{}\quad \ddots &{}\quad \\ &{}\quad &{}\quad &{}\quad \lambda _i&{}\quad 1\\ &{}\quad &{}\quad &{}\quad &{}\quad \lambda _i\end{pmatrix}. \end{aligned}$$

Since zero is a simple eigenvalue, the first column of P is \(\mathbf {v}_0\) and the first row of \(P^{-1}\) is \(\mathbf {u}_0\).

The solution of Eq. 16 is given by

$$\begin{aligned} \begin{aligned} \mathbf {Y}(t)&= e^{Mt}Y_0 + \int _0^te^{M(t-s)}\,\mathrm {d}s \cdot F\\&= Pe^{J t} P^{-1}Y_0 + Pe^{J t}\int _0^te^{-J s}\,\mathrm {d}s\, P^{-1}F. \end{aligned} \end{aligned}$$
(37)

Consider \(e^{Jt}\). Write \(J = D + N\), where D is a diagonal matrix with diagonal coinciding with J. Then N is a matrix with zeros everywhere except for some 1’s on the superdiagonal. Specifically, N is the block diagonal matrix

$$\begin{aligned} N = \begin{pmatrix} 0 &{}\quad &{}\quad &{}\quad \\ {} &{}\quad N_1&{}\quad &{}\quad \\ {} &{}\quad &{}\quad \ddots &{}\quad \\ {} &{}\quad &{}\quad &{}\quad N_k\end{pmatrix}, \end{aligned}$$

where each \(N_i\) is an \(m_i\times m_i\) nilpotent matrix of the form

$$\begin{aligned} N_i = \begin{pmatrix}0&{}\quad 1 &{}\quad &{}\quad \\ {} &{}\quad 0&{}\quad 1&{}\quad \\ {} &{}\quad &{}\quad \ddots &{}\quad \ddots \\ {} &{}\quad &{}\quad &{}\quad 0\end{pmatrix}. \end{aligned}$$

Notably, N commutes with D. Hence

$$\begin{aligned} e^{Jt} = e^{(D+N)t} = e^{Dt}e^{Nt}. \end{aligned}$$

Further, \(e^{Nt}\) has the form

$$\begin{aligned} e^{Nt} = \begin{pmatrix}1 &{}\quad &{}\quad &{}\quad \\ &{}\quad E_1(t) &{}\quad &{}\quad \\ {} &{}\quad &{}\quad \ddots &{}\quad \\ {} &{}\quad &{}\quad &{}\quad E_k(t)\end{pmatrix}, \end{aligned}$$

where

$$\begin{aligned} E_i(t) = \begin{pmatrix} 1 &{}\quad t &{}\quad t^2/2&{}\quad \cdots &{}\quad \frac{t^{m_i-1}}{(m_i-1)!}\\ &{}\quad 1&{}\quad t&{}\quad \cdots &{}\quad \frac{t^{m_i-2}}{(m_i-2)!}\\ {} &{}\quad &{}\quad 1&{}\quad \cdots &{}\quad \frac{t^{m_i-3}}{(m_i-3)!}\\ {} &{}\quad &{}\quad &{}\quad \ddots &{}\quad \vdots \\ {} &{}\quad &{}\quad &{}\quad &{}\quad 1\end{pmatrix}. \end{aligned}$$

Hence,

$$\begin{aligned} e^{Jt} = \begin{pmatrix}1 &{}\quad &{}\quad &{}\quad \\ &{}\quad e^{\lambda _1 t}E_1(t)&{}\quad &{}\quad \\ {} &{}\quad &{}\quad \ddots &{}\quad \\ {} &{}\quad &{}\quad &{}\quad e^{\lambda _kt}E_k(t)\end{pmatrix}. \end{aligned}$$

Since \(\text {Re}\left( \lambda _i\right) < 0\) for all \(i\ne 0\), we have

$$\begin{aligned} \lim _{t\rightarrow \infty }e^{J t} = \begin{pmatrix} 1&{}\quad &{}\quad &{}\quad \\ {} &{}\quad 0&{}\quad &{}\quad \\ {} &{}\quad &{}\quad \ddots &{}\quad \\ {} &{}\quad &{}\quad &{}\quad 0\end{pmatrix}. \end{aligned}$$

It follows that

$$\begin{aligned} \lim _{t\rightarrow \infty }\frac{1}{t}Pe^{J t}P^{-1}Y_0 = \mathbf {0}. \end{aligned}$$
(38)

We now turn our attention to the integral \(\int _0^te^{J(t-s)}\,\mathrm {d}s\). Note that for any integer \(a\ge 1\),

$$\begin{aligned} \int _0^t (t-s)^ae^{\lambda _i(t-s)}\,\mathrm {d}s = \frac{\varGamma \left( a+1,-\lambda _it\right) - a(a!)}{\lambda _i^{a+1}}, \end{aligned}$$

where \(\varGamma (s,x)\) represents the incomplete gamma function. For large x, \(\varGamma (s,x) \sim x^{s-1}e^{-x}\) (Abramowitz and Stegun 1964). So, for sufficiently large t,

$$\begin{aligned} \int _0^t(t-s)^ae^{\lambda _i(t-s)}\,\mathrm {d}s \sim \frac{\left( -\lambda _it\right) ^a e^{\lambda _i t}-a(a!)}{\lambda _i^{a+1}}. \end{aligned}$$

In particular, since Re\((\lambda _i)<0\),

$$\begin{aligned} \lim _{t\rightarrow \infty }\frac{1}{t}\int _0^t(t-s)^ae^{\lambda _i(t-s)} \mathrm {d}s= 0. \end{aligned}$$
(39)

Now,

$$\begin{aligned} \int _0^te^{J(t-s)}\,\mathrm {d}s = \begin{pmatrix}t &{}\quad &{}\quad &{}\quad \\ {} &{}\quad \tilde{E}_1(t) &{}\quad &{}\quad \\ {} &{}\quad &{}\quad \ddots &{}\quad \\ {} &{}\quad &{}\quad &{}\quad \tilde{E}_k(t)\end{pmatrix}, \end{aligned}$$

where \(\tilde{E}_i(t)\) is an upper triangular matrix for which the entry on the jth row and \((j+k)\)th column (where \(k \ge 0\)) is

$$\begin{aligned} \tilde{E}_i(t) = \int _0^t (t-s)^{k} e^{\lambda _i (t-s)} \mathrm {d}s \end{aligned}$$

It follows now from Eq. 39 that

$$\begin{aligned} \lim _{t\rightarrow \infty }\frac{1}{t}\int _0^te^{J(t-s)}\,\mathrm {d}s = \begin{pmatrix} 1 &{}\quad &{}\quad &{}\quad \\ {} &{}\quad 0 &{}\quad &{}\quad \\ &{}\quad &{}\quad \ddots &{}\quad \\ &{}\quad &{}\quad &{}\quad 0 \end{pmatrix}. \end{aligned}$$
(40)

Equations 38 and 40, together with the fact that \(\mathbf {v}_0 = (1,1,\ldots ,1)^\top \) is the first column of P and \(\mathbf {u}_0\) is the first row of \(P^{-1}\), imply

$$\begin{aligned} \lim _{t\rightarrow \infty } \frac{1}{t}\left( Pe^{Jt}P^{-1}Y_0 + P\int _0^te^{J(t-s)}\,\mathrm {d}s\,P^{-1}F\right) = \mathbf {v}_0\mathbf {u}_0F. \end{aligned}$$

Hence, \(\lim _{t\rightarrow \infty }\frac{1}{t}Y_j(t) = \mathbf {u}_0 F\) for any \(j\in \{0,\ldots ,n-1\}\) as claimed. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Smith, J.D., McKinley, S.A. Assessing the Impact of Electrostatic Drag on Processive Molecular Motor Transport. Bull Math Biol 80, 2088–2123 (2018). https://doi.org/10.1007/s11538-018-0448-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-018-0448-9

Keywords

Navigation