Skip to main content
Log in

Analytical aspects of the Brownian motor effect in randomly flashing ratchets

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

Abstract

The muscle contraction, operation of ATP synthase, maintaining the shape of a cell are believed to be secured by motor proteins, which can be modelled using the Brownian ratchet mechanism. We consider the randomly flashing ratchet model of a Brownian motor, where the particles can be in two states, only one of which is sensitive the applied spatially periodic potential (the mathematical setting is a pair of weakly coupled reaction-diffusion and Fokker–Planck equations). We prove that this mechanism indeed generates unidirectional transport by showing that the amount of mass in the wells of the potential decreases/increases from left to right. The direction of transport is unambiguously determined by the location of each minimum of the potential with respect to the so-called diffusive mean of its adjacent maxima. The transport can be generated not only by an asymmetric potential, but also by a symmetric potential and asymmetric transition rates, and as a consequence of the general result we derive explicit conditions when the latter happens. When the transitions are localized on narrow active sites in the protein conformation space, we find a more explicit characterization of the bulk transport direction, and infer that some common preconditions of the motor effect are redundant.

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.

Similar content being viewed by others

References

  • Ait-Haddou R, Herzog W (2003) Brownian ratchet models of molecular motors. Cell Biochem Biophys 38:191–214

    Article  Google Scholar 

  • Amengual P, Allison A, Toral R, Abbott D (2004) Discrete-time ratchets, the Fokker-Planck equation and Parrondo’s paradox. Proc R Soc Lond Ser A Math Phys Eng Sci 460:2269–2284

    Article  MATH  MathSciNet  Google Scholar 

  • Astumian RD (1997) Thermodynamics and kinetics of a Brownian motor. Science 276:917–922

    Article  Google Scholar 

  • Chipot M, Hastings S, Kinderlehrer D (2004) Transport in a molecular motor system. M2AN Math Model Numer Anal 38:1011–1034

    Google Scholar 

  • Chipot M, Kinderlehrer D, Kowalczyk M (2003) A variational principle for molecular motors. Meccanica 38:505–518

    Article  MATH  MathSciNet  Google Scholar 

  • Diestel J (1991) Uniform integrability: an introduction. School on measure theory and real analysis (Grado, 1991). Rend Istit Mat Univ Trieste 23:41–80

    MATH  MathSciNet  Google Scholar 

  • Friedman A, Hu B (2007) Uniform convergence for approximate traveling waves in linear reaction-hyperbolic systems. Indiana Univ Math J 56:2133–2158

    Article  MATH  MathSciNet  Google Scholar 

  • Harmer G, Abbott D, Taylor P (2000) The paradox of Parrondo’s games. Proc R Soc Lond Ser A Math Phys Eng Sci 456:247–259

    Article  MATH  MathSciNet  Google Scholar 

  • Hastings S, Kinderlehrer D, McLeod JB (2007/2008) Transport in multiple state systems. SIAM J Math Anal 39:1208–1230

    Google Scholar 

  • Jordan R, Kinderlehrer D, Otto F (1998) The variational formation of the Fokker-Planck equation. SIAM J Math Anal 29:1–17

    Article  MATH  MathSciNet  Google Scholar 

  • Jülicher F (1999) Force and motion generation of molecular motors: a generic description. In: Müller SC, Parisi J, Zimmermann W (eds) Transport and structure, their competitive roles in biophysics and chemistry, Lecture Notes in Physics, vol 532. Springer, New York, pp 46–74

    Google Scholar 

  • Jülicher F, Ajdari A, Prost J (1997) Modeling molecular motors. Rev Modern Phys 69:1269–1281

    Article  Google Scholar 

  • Kinderlehrer D, Kowalczyk M (2002) Diffusion-mediated transport and the flashing ratchet. Arch Rat Mech Anal 161:149–179

    Article  MATH  MathSciNet  Google Scholar 

  • Littman W (1963) Generalized subharmonic functions: Monotonic approximations and an improved maximum principle. Ann Scuola Norm Sup Pisa 17(3):207–222

    MATH  MathSciNet  Google Scholar 

  • Mennerat-Robilliard C (1999) Atomes froids dans des réseaux optiques - Quelques facettes surprenantes d’un système modèle. Thèse de Doctorat de l’Université Paris-VI, LKB/ENS - Université Paris-VI

  • Mirrahimi S, Souganidis PE (2013) A homogenization approach for the motion of motor proteins. Nonlinear Differ Equ Appl 20:129–147

    Article  MATH  MathSciNet  Google Scholar 

  • Parrondo JMR, Blanco JM, Cao FJ, Brito R (1998) Efficiency of Brownian motors. Europhys Lett 43:248–254

    Article  Google Scholar 

  • Parrondo JMR, de Cisneros BJ (2002) Energetics of Brownian motors: a review. Appl Phys A 75:179–191

    Article  Google Scholar 

  • Paxton WF, Sundararajan S, Mallouk TE, Sen A (2006) Chemical locomotion. Angew Chem Int Ed 45:5420–5429

    Article  Google Scholar 

  • Perthame B, Souganidis PE (2009a) Asymmetric potentials and motor effect: a homogenization approach. Ann Inst Henri Poincaré Anal Non Linéaire 26:2055–2071

    Article  MATH  MathSciNet  Google Scholar 

  • Perthame B, Souganidis PE (2009b) Asymmetric potentials and motor effect: a large deviation approach. Arch Rat Mech Anal 193:153–169

    Article  MATH  MathSciNet  Google Scholar 

  • Perthame B, Souganidis PE (2011) A homogenization approach to flashing ratchets. Nonlinear Differ Equ Appl 18:45–58

    Article  MATH  MathSciNet  Google Scholar 

  • Peskin CS, Ermentrout B, Oster G et al (1994) The correlation ratchet: a novel mechanism for generating directed motion by ATP hydrolysis. In: Mow VC (ed) Cell mechanics and cellular engineering. Springer, New York, pp 479–489

    Chapter  Google Scholar 

  • Prost J, Chauwin JF, Peliti L, Ajdari A (1994) Asymmetric pumping of particles. Phys Rev Lett 72:2652–2655

    Article  Google Scholar 

  • Reimann P, Hänggi P (2002) Introduction to the physics of Brownian motors. Appl Phys A 75:169–178

    Article  Google Scholar 

  • Spirin AS (2009) The ribosome as a conveying thermal ratchet machine. J Biol Chem 284(32):21103–21119

    Article  Google Scholar 

  • Vorotnikov D (2011) The flashing ratchet and unidirectional transport of matter. Discrete Contin Dyn Syst Ser B 16:963–971

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

The research was partially supported by CMUC and FCT (Portugal), through European program COMPETE/FEDER under the project PEst-C/MAT/UI0324/2011. The author thanks David Kinderlehrer who introduced him to ratchets, and Philippe Verkerk for a discussion on the topic. He also thanks the unknown referees for detailed comments which helped to significantly improve this article

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dmitry Vorotnikov.

Appendices

Appendix A: Proof of Propositions 2.1 and 3.1

Let \(G(x,y)\) be Green’s function of the Sturm–Liouville operator

$$\begin{aligned} \mathcal L = - \frac{d^2}{dx^2}+\phi \end{aligned}$$

on \((A,B)\) with homogeneous Neumann boundary condition. Then \(U(x)=G(x,s)\) is a solution to (9) if an only if \(G(A,s)=G(B,s)\).

By the distributional maximum principle (Littman 1963, Theorem B), \(G(x,y)>0\). Observe that

$$\begin{aligned} G_x(x,A)=\int \limits _{B}^x\phi (z) G(A,z) \, dz,\end{aligned}$$
(37)
$$\begin{aligned} G_x(x,B)=\int \limits _{A}^x\phi (z) G(B,z) \, dz. \end{aligned}$$
(38)

Hence, the function \(G(x,A)\) is decreasing in \(x\), and \(G(x,B)\) is increasing. Thus, the function

$$\begin{aligned} g(x)=G(A,x)-G(B,x) \end{aligned}$$

is also (strictly) decreasing. At the ends of the segment, we have \(g(A)=G(A,A)-G(B,A)> 0\) and \(g(B)=G(A,B)-G(B,B) < 0\). Since \(g\) is a continuous function, there is unique \(s\in (A,B)\) such that \(G(A,s)=G(B,s)\). The first proposition is proven.

Now, let

$$\begin{aligned} g_1(x)=G\left( \frac{A+B}{2},x\right) -G\left( \frac{A+B}{2},A+B-x\right) . \end{aligned}$$

We claim that

$$\begin{aligned} g_1(A)>0. \end{aligned}$$
(39)

Since \(G\) is Green’s function,

$$\begin{aligned} (g_1)_x(A)=0, \end{aligned}$$
(40)

and

$$\begin{aligned} (g_1)_{xx}(x)&= \phi (x)G\left( \frac{A+B}{2},x\right) -\phi (A+B-x)G\left( \frac{A+B}{2},A+B-x\right) \nonumber \\&= \phi (x) g_1(x)+[\phi (x)-\tilde{\phi }(x)]G\left( \frac{A+B}{2},A+B-x\right) \nonumber \\&\le \phi (x) g_1(x), \ A\le x< \frac{A+B}{2}, \end{aligned}$$
(41)

and the inequality is strict at least at one point. In particular, \(g_1\) cannot be identically zero.

Assume that \(g_1(A)\le 0\). By the maximum principle, \(g_1\) cannot have non-positive minima within \((A,\frac{A+B}{2})\). But

$$\begin{aligned} g_1\left( \frac{A+B}{2}\right) =0, \end{aligned}$$
(42)

so \(A\) must be a minimum point. Let \(g_2(x)=g_1(x)-g_1(A)\). Then \(g_2\) is non-negative, and

$$\begin{aligned} (g_2)_{xx}(x) \le \phi (x) g_2(x), \ A\le x< \frac{A+B}{2}. \end{aligned}$$
(43)

Thus,

$$\begin{aligned} (g_2)_{x}(x) \le \int \limits _A^x \phi (t) g_2(t)\,dt, \ A\le x\le \frac{A+B}{2}. \end{aligned}$$
(44)

Since \(g_2(A)=0\), by the mean value theorem,

$$\begin{aligned} g_2(x)&= g_2(x)-g_2(A)= (x-A)(g_2)_x(c) \nonumber \\&\le (x-A)\int \limits _A^c \phi (t) g_2(t)\,dt \le \frac{1}{2} (B-A)\int \limits _A^x \phi (t) g_2(t)\,dt \end{aligned}$$
(45)

for some \(c, A<c<x\le \frac{A+B}{2}\). The Gronwall lemma implies \(g_2\equiv 0\), so, by (42), \(g_1\equiv 0\), and we get a contradiction.

Hence,

$$\begin{aligned} g\left( \frac{A+B}{2}\right) =g_1(A)> 0. \end{aligned}$$

We have observed above that \(g(B)<0\), so there is \(s\in (\frac{A+B}{2},B)\) such that \(g(s)=0\), and this number \(s\) is the \(\phi \)-diffusive mean.

Appendix B: Proof of Proposition 3.2

Without loss of generality, \([A,B]=[0,1]\). Denoting the solutions of (9) corresponding to \(\phi _n\) and \(s_n\) by \(U_n\), we infer that

$$\begin{aligned} \int \limits _0^1 \phi _n(x) U^2_n(x) +(U_n)_x^2\, dx = U_n(s_n). \end{aligned}$$
(46)

The function \(U_n\) is positive and thus convex on \((0,s_n)\) and \((s_n,1)\). Hence, due to the boundary conditions, the derivative \((U_n)_x(x)\) is positive/negative when \(x< s_n\) / \(x >s_n\), and tends toward its supremum/infimum as \(x\) approaches \(s_n\) from the left/right, resp. Moreover, the maximum of \(U_n\) is achieved at \(s_n\), and the minima are reached at \(0\) and \(1\).

It is easy to see that

$$\begin{aligned} \min \limits _{x\in [0,1]} U_n^2(x)\le \frac{2}{K_*} \int \limits _0^1 \phi _n(x) U^2_n(x)\,dx \end{aligned}$$
(47)

for large \(n\), and

$$\begin{aligned} \max \limits _{x\in [0,1]} U_n^2(x)\le 2\left[ \min \limits _{x\in [0,1]} U_n^2(x)+\int \limits _0^1 (U_n)_x^2\, dx\right] . \end{aligned}$$
(48)

Hence, (46) implies

$$\begin{aligned} \Vert U_n^2\Vert _{C[0,1]} \le C \Vert U_n\Vert _{C[0,1]}, \end{aligned}$$
(49)

whence

$$\begin{aligned} \Vert U_n\Vert _{C[0,1]} \le C. \end{aligned}$$
(50)

Without loss of generality, there exists a limit \(s^*\in [0,1]\) of the sequence \(\{s_n\}\). Assume that \(s^*\ne s_*\), or, more particularly, \(s_*>s^*\) (the opposite case may be examined in a similar way). This yields that the value of the integral \(\int _0^{s_n} \phi _n(x) \) tends to zero as \(n\rightarrow + \infty \). But integration of (9) implies that the left derivative \((U_n)^\prime _- (s_n)\) is equal to \(\int _0^{s_n} \phi _n(x) U_n(x)\,dx\). Due to (50), this integral goes to zero as \(n\rightarrow + \infty \). Using the information on the behaviour of \(U_n\) summarized after equality (46), we conclude that \(\Vert (U_n)_x\Vert _{C[0,s_n]}\rightarrow 0\). Due to (50), without loss of generality there exists a constant \(U_*\) such that

$$\begin{aligned} \max \limits _{x\in [0,1]} U_n(x) \rightarrow U_*. \end{aligned}$$
(51)

But

$$\begin{aligned} \max \limits _{x\in [0,1]} U_n(x)- \min \limits _{x\in [0,1]} U_n(x)&= U_n(s_n)-U_n(0)\nonumber \\&\le s_n\Vert (U_n)_x\Vert _{C[0,s_n]}\rightarrow 0. \end{aligned}$$
(52)

Hence,

$$\begin{aligned} U_n(x) \rightarrow U_* \end{aligned}$$
(53)

uniformly on \([0,1]\).

Let us now test (9) with a smooth function \(h\) such that \(h(s_*)=0, h(s^*)=1, h_x(0)=0, h_x(1)=0\), obtaining

$$\begin{aligned} \int \limits _0^1 \phi _n(x) U_n(x) h(x)-U_n(x) h_{xx}(x)\, dx = h(s_n). \end{aligned}$$
(54)

Passing to the limit, we find that the left-hand side goes to zero and the right-hand side goes to one, arriving at a contradiction.

Appendix C: Proof of Theorem 4.1

Consider the set of functions

$$\begin{aligned} B&= \left\{ y(x)\in L_1(0,1)\left| \, \int \limits _{0}^1 \nu (x) y(x) \, dx=1, \ \widehat{(\nu y)}_k \ge M, \right. \right. \\&\quad \left. y(x)\ge y(x+1/k)+M \gamma ,\quad {\mathrm{for \, a.a.}} 0\le x\le 1-1/k.\right\} \end{aligned}$$

Inverse induction shows that for any \(i=1,\dots , k\) and \(y\in B\) one has

$$\begin{aligned} \widehat{(\nu y)}_i \ge M. \end{aligned}$$
(55)

Let us define a mapping \(\mathcal A \) on \(B\). For each \(y\in B\), we let \(\mathcal A (y)=Y\), where \(Y\) is the solution of the problem

$$\begin{aligned} \left\{ \begin{array}{l@{\quad }l} \nu Y - \varsigma Y_{xx}= \sum \nolimits _{i=1}^k \widehat{(\nu y)}_i \delta _{a_i}, &{} x\in (0,1), \\ Y_{x}(0)=Y_{x}(1)=0. \end{array}\right. \end{aligned}$$
(56)

To put it differently,

$$\begin{aligned} Y=\sum \limits _{i=1}^k \widehat{(\nu y)}_i q_{i}. \end{aligned}$$
(57)

Then, the set \(B\) is invariant for the map \(\mathcal A \). In fact, let \(y\in B\). Then (56) implies

$$\begin{aligned} \int \limits _{0}^1 \nu (x) Y(x) \, dx= \sum \limits _{i=1}^k \widehat{(\nu y)}_i = \int \limits _{0}^1 \nu (x) y(x) \, dx=1. \end{aligned}$$

Further,

$$\begin{aligned} \widehat{(\nu Y)}_k=\sum \limits _{i=1}^k \widehat{(\nu y)}_i \widehat{(\nu q_i)}_k\ge M\sum \limits _{i=1}^k \widehat{(\nu y)}_i =M. \end{aligned}$$

Finally, fix \(x_*\in [0,1-1/k]\). Then there is a number \(n\) such that \(x_*\in [x_n,x_{n+1})\). Set

$$\begin{aligned} N_*=\left\{ \begin{array}{ll} \widehat{(\nu y)}_n, &{} x_*\le a_n,\\ \widehat{(\nu y)}_{n+1}, &{} x_* > a_n. \end{array}\right. \end{aligned}$$
(58)

We claim that

$$\begin{aligned} (\widehat{(\nu y)}_i - N_*)[q_i(x_*)-q_i(x_*+1/k)]\ge 0,\ i=1,\dots ,k. \end{aligned}$$
(59)

Indeed, integration of (22) gives

$$\begin{aligned}&\varsigma {q_i}_{x}(x)=\int \limits _{0}^x\nu (z) q_i(z) \, dz, \ x< a_i, \end{aligned}$$
(60)
$$\begin{aligned}&\varsigma {q_i}_{x}(x)=\int \limits _{1}^x\nu (z) q_i(z) \, dz, \ x> a_i. \end{aligned}$$
(61)

Thus, the function \(q_i\) is increasing on the segment \([0,a_i]\) and decreasing on \([a_i,1]\). Assume first \(x_*\le a_n\). Then, if \(i<n\), we have \(q_i(x_*)>q_i(x_*+1/k)\). Since \(y\in B\), we also have \(\widehat{(\nu y)}_i >\widehat{(\nu y)}_n= N_*\), and (59) holds true. If \(i>n\), we have \(q_i(x_*)<q_i(x_*+1/k)\) and \(\widehat{(\nu y)}_i < N_*\), and (59) again holds. If \(i=n\), (59) is trivial. Now, let \(x_*> a_n\). In this case, if \(i<n+1\), \(q_i(x_*)>q_i(x_*+1/k)\) and \(\widehat{(\nu y)}_i >\widehat{(\nu y)}_{n+1}= N_*\); if \(i>n+1, q_i(x_*)<q_i(x_*+1/k)\) and \(\widehat{(\nu y)}_i < N_*\); and if \(i=n+1\), (59) is again trivial.

Formulas (57), (59), (26) and (55) yield

$$\begin{aligned} Y(x_*)- Y(x_*+1/k)&= \sum \limits _{i=1}^k \widehat{(\nu y)}_i [q_i(x_*)-q_i(x_*+1/k)]\\&\ge \sum \limits _{i=1}^k N_* [q_i(x_*)-q_i(x_*+1/k)]\\&= N_*[q(x_*)-q(x_*+1/k)]\ge M\gamma , \end{aligned}$$

so the invariance of \(B\) is confirmed.

Observe that \(\mathcal A \) is a compact linear operator in \(L_1(0,1)\). Indeed, let \(\mathbb B \) be the unit ball of the space \(L_1(0,1)\). Due to (57), its image \(\mathcal A (\mathbb B )\) is a bounded subset of the linear span of \(\{q_1,\dots ,q_k \}\), thus being a relatively compact subset of a finite-dimensional subspace of \(L_1(0,1)\).

Let us show that (25) may have at most one solution, so \(\mathcal A \) can have at most one fixed point in \(B\). If not, let \({\tilde{Q}}\) be the difference of two distinct solutions. Then

$$\begin{aligned} \int \limits _{0}^1 \nu (z) \tilde{Q}(z) \, dz=0. \end{aligned}$$
(62)

Moreover,

$$\begin{aligned} {\tilde{Q}}=\sum \limits _{i=1}^k \widehat{(\nu {\tilde{Q}})}_i q_{i}, \end{aligned}$$
(63)

whence

$$\begin{aligned} \widehat{(\nu {\tilde{Q}})}_j=\sum \limits _{i=1}^k \widehat{(\nu {\tilde{Q}})}_i \widehat{(\nu q_i)}_j,\ j=1,\dots ,k. \end{aligned}$$
(64)

From (22) we deduce

$$\begin{aligned} \int \limits _{0}^1 \nu (z) q_i(z) \, dz=1. \end{aligned}$$
(65)

Therefore, the matrix \([\mathcal P _{ij}]=[\widehat{(\nu q_i)}_j ]\) is ergodic, i.e. it has positive entries, and the sum of the elements in every row is equal to one. By the Perron–Frobenius theorem, it has an eigenvector \([\xi _i]\) corresponding to the simple eigenvalue \(1\), so that \(\xi _j=\sum _{i=1}^k \xi _i \mathcal P _{ij}\), and all the components \(\xi _i\) are positive. On the other hand, by (64), \([\Xi _i]=[\widehat{(\nu \tilde{Q})}_i ] \) is another eigenvector of \([\mathcal P _{ij}]\) corresponding to the same eigenvalue. The sum of its components is zero due to (62), so it cannot be collinear with \([\xi _i]\) unless it is a zero vector. Since \(1\) is a simple eigenvalue, all \(\widehat{(\nu \tilde{Q})}_i\) are zeros, so \(\tilde{Q}\equiv 0\) by virtue of (63).

The set \(B\) is closed, convex and bounded in \(L_1(0,1)\). By Schauder’s fixed point principle, \(\mathcal A \) has a fixed point \(Q\) in \(B\), which is automatically a solution to (25). It remains to notice that \(Q\) is continuous as a linear combination of \(q_i\), so (27) holds for all \(0\le x\le 1-1/k\).

Appendix D: Proof of Lemma 4.1

Let us notice that

$$\begin{aligned} q=u+v, \end{aligned}$$
(66)

where \(u\) and \(v\) are the (unique) solutions to the following problems

$$\begin{aligned}&\left\{ \begin{array}{ll} \nu u - \varsigma {u}_{xx}= \sum \nolimits _{i=1}^k\delta _{S_i}, &{} x\in (0,1), \\ {u}_{x}(0)={u}_{x}(1)=0, \end{array}\right. \end{aligned}$$
(67)
$$\begin{aligned}&\left\{ \begin{array}{ll} \nu v - \varsigma {v}_{xx}= \sum \nolimits _{i=1}^k(\delta _{a_i}-\delta _{S_i}), &{} x\in (0,1), \\ {v}_{x}(0)={v}_{x}(1)=0, \end{array}\right. \end{aligned}$$
(68)

and

$$\begin{aligned} S_i= S+x_i,\ i=1,\dots , k. \end{aligned}$$
(69)

Since \(S\) is the \(\nu /\varsigma \)-diffusive mean of \(0\) and \(1/k\), there exists a solution \(U_1\) to the problem

$$\begin{aligned} \left\{ \begin{array}{ll} \nu U_1 - \varsigma {U_1}_{xx}= \delta _{S_1}, &{} x\in (0,1/k), \\ {U_1}_{x}(0)={U_1}_{x}(1/k)=0,\\ U_1(0)=U_1(1/k). \end{array}\right. \end{aligned}$$
(70)

The solution \(u\) to (67) can be constructed in the following way:

$$\begin{aligned} u(x)=U_1(x-x_i),\ x_i\le x\le x_i+1/k,\ i=1,\dots , k. \end{aligned}$$
(71)

Thus, \(u\) is \(1/k\)-periodic, i.e.

$$\begin{aligned} u(x)- u(x+1/k)=0, \ 0\le x\le x_k. \end{aligned}$$
(72)

Set

$$\begin{aligned} d(x)=v(x)- v(x+1/k), \ 0\le x\le x_k. \end{aligned}$$
(73)

Then it suffices to show that

$$\begin{aligned} \gamma =\min _{0\le x\le x_k} d(x)>0. \end{aligned}$$
(74)

Note that

$$\begin{aligned} \nu d - \varsigma {d}_{xx}= 0, \ x\in (0,x_k). \end{aligned}$$
(75)

By the maximum principle, if the minimum of \(d\) is non-positive, it is attained at \(0\) or \(x_k\). To ascertain that this cannot happen, we are going to prove that

$$\begin{aligned} d_x(0)<0, \ d_x(x_k)>0. \end{aligned}$$
(76)

Set

$$\begin{aligned} V(x)=\int \limits _{0}^x\nu (z) v(z) \, dz, \end{aligned}$$
(77)

and let \(\theta \) be the solution of the Cauchy problem

$$\begin{aligned} \left\{ \begin{array}{ll} \theta _x= \sum \nolimits _{i=1}^k(\delta _{a_i}-\delta _{S_i}), &{} x\in (0,1), \\ \theta (0)=0. \end{array}\right. \end{aligned}$$
(78)

Note that \(\theta \) is non-negative and \(1/k\)-periodic.

Integration of (68) gives

$$\begin{aligned} V-\varsigma v_x= \theta . \end{aligned}$$
(79)

Therefore

$$\begin{aligned} d_x=\frac{V(x)-V(x+1/k)}{\varsigma }. \end{aligned}$$
(80)

From (79) we deduce

$$\begin{aligned} V(0)=V(1)=0, \end{aligned}$$
(81)

and

$$\begin{aligned} V-\varsigma \left( \frac{V_x}{\nu }\right) _x= \theta \ge 0. \end{aligned}$$
(82)

Using the distributional maximum principle (Littman 1963, Theorem B), we conclude that

$$\begin{aligned} V(x)> 0, \ 0<x<1, \end{aligned}$$
(83)

so

$$\begin{aligned} d_x(0)=-\frac{V(1/k)}{\varsigma }<0, \ d_x(x_k)=\frac{V(1-1/k)}{\varsigma }>0. \end{aligned}$$
(84)

Appendix E: Proof of Lemma 5.1

It suffices to prove that the set

$$\begin{aligned} O=\left\{ \varphi +\kappa b\varphi _x\Big | \varphi \in C^2[0,1], \varphi _x(0)=\varphi _x(1)=0\right\} \end{aligned}$$

is dense in \(C[0,1]\).

Let \(h\in C^2[0,1]\) be an arbitrary function which is locally constant near the zeros of \(b\). These functions constitute a dense subset \(O_1\) of \(C[0,1]\). Let

$$\begin{aligned} \varphi (x)&= h(x)+ \int \limits _{x}^{a_i} \exp \left( \int \limits _x^y\frac{1}{\kappa b(t)}\,dt \right) h_y(y)\,dy,\nonumber \\&\qquad \,\,\,\,\,\,x_i< x< x_{i+1}, \ i=1,\dots , k, \nonumber \\&\qquad \,\,\,\,\,\,\varphi (x_i)=h(x_i),\ i=1,\dots , k+1. \end{aligned}$$
(85)

Clearly, \(\varphi \) is equal to a constant \(c^-_i\) (resp. \(c^+_i\)) in a left (resp. right) neighbourhood of the point \(x_i\). But

$$\begin{aligned} h=\varphi +\kappa b\varphi _x, \end{aligned}$$
(86)

so \(c^-_i=c^+_i=h(x_i)\). Thus, \(\varphi \) is \(C^2\)-smooth and \(\varphi _x(0)=\varphi _x(1)=0\). By virtue of (86), \(O_1\) is contained in \(O\).

Appendix F: Proof of Lemma 5.2

The solution \(r\) to (30) can be written explicitly:

$$\begin{aligned} r(x)&= -\frac{1}{\kappa b(x)}\int \limits _{x_i}^x \exp \left( \int \limits _s^x\frac{1}{\kappa b(t)}\,dt \right) R(s)\,ds,\quad x_i<x < a_i,\end{aligned}$$
(87)
$$\begin{aligned} r(x)&= \frac{1}{\kappa b(x)}\int \limits _{x}^{x_{i+1}} \exp \left( \int \limits _s^x\frac{1}{\kappa b(t)}\,dt \right) R(s)\,ds,\quad a_i<x < x_{i+1}. \end{aligned}$$
(88)

Note that it is unique in \(L_1(0,1)\). Indeed, if \(r_1\) is another solution, then \(\omega =r-r_1\) satisfies the conditions of Lemma 5.1.

We need to show that \(r\rightarrow \sum \limits _{i=1}^k\widehat{R}_i\delta _{a_i}\) weakly-\(*\), uniformly with respect to \(R\in \mathcal R \). It suffices to prove that, for each \(i, r\rightarrow \widehat{R}_i\delta _{a_i}\) weakly-\(*\) on the interval \((x_i,x_{i+1})\), uniformly in \(R\in \mathcal R \). We restrict ourselves to the case \(i=1\), and the others are analogous.

We calculate, integrating by parts,

$$\begin{aligned} \widehat{r}_1&= \int \limits _0^{1/k} r(x)\,dx \nonumber \\&= -\int \limits _0^{a} \frac{1}{\kappa b(x)}\int \limits _{0}^x \exp \left( \int \limits _s^x\frac{1}{\kappa b(t)}\,dt \right) R(s)\,ds\,dx \nonumber \\&+\int \limits _{a}^{1/k}\frac{1}{\kappa b(x)}\int \limits _{x}^{1/k} \exp \left( \int \limits _s^x\frac{1}{\kappa b(t)}\,dt \right) R(s)\,ds\,dx\nonumber \\&= \left[ \int \limits _{0}^x \exp \left( \int \limits _s^x\frac{1}{\kappa b(t)}\,dt \right) R(s)\,ds\right] _a^0+\int \limits _0^{a} R(x)\,dx\nonumber \\&+\left[ \int \limits _{1/k}^x \exp \left( \int \limits _s^x\frac{1}{\kappa b(t)}\,dt \right) R(s)\,ds\right] _{1/k}^a+\int \limits _a^{1/k} R(x)\,dx\nonumber \\&= \widehat{R}_1-\int \limits _{0}^{1/k} \exp \left( \int \limits _s^a\frac{1}{\kappa b(t)}\,dt \right) R(s)\,ds= \widehat{R}_1. \end{aligned}$$
(89)

Let us show that for every \(x_*\in (0,a)\)

$$\begin{aligned} \lim \limits _{\kappa \rightarrow +\infty } \int \limits _0^{x_*} r(x)\,dx= 0, \end{aligned}$$
(90)

uniformly in \(R\in \mathcal R \). Indeed, let \(s_\kappa <x_*\) be such that

$$\begin{aligned} \int \limits _{x_*}^{s_\kappa }\frac{1}{ b(t)}\,dt= \sqrt{\kappa }. \end{aligned}$$
(91)

Observe that \(s_\kappa \rightarrow 0\) as \(\kappa \rightarrow +\infty \). We have

$$\begin{aligned} \int \limits _0^{x_*} r(x)\,dx&= -\int \limits _0^{x_*} \frac{1}{\kappa b(x)}\int \limits _{0}^x \exp \left( \int \limits _s^x\frac{1}{\kappa b(t)}\,dt \right) R(s)\,ds\,dx\nonumber \\&= \left[ \int \limits _{0}^x \exp \left( \int \limits _s^x\frac{1}{\kappa b(t)}\,dt \right) R(s)\,ds\right] _{x_*}^0+\int \limits _0^{x_*} R(x)\,dx\nonumber \\&= \int \limits _{0}^{x_*} \left[ 1-\exp \left( \int \limits _s^{x_*}\frac{1}{\kappa b(t)}\,dt \right) \right] R(s)\,ds\nonumber \\&\le \int \limits _{0}^{s_\kappa } \left[ 1-\exp \left( \int \limits _s^{x_*}\frac{1}{\kappa b(t)}\,dt \right) \right] R(s)\,ds \nonumber \\&+ \int \limits _{s_\kappa }^{x_*} \left[ 1-\exp \left( \int \limits _{s_\kappa }^{x_*}\frac{1}{\kappa b(t)}\,dt \right) \right] R(s)\,ds\nonumber \\&\le \int \limits _{0}^{s_\kappa } R(s)\,ds + [1-\exp (-\kappa ^{-1/2})]\int \limits _{s_\kappa }^{x_*}R(s)\,ds\rightarrow 0 \end{aligned}$$
(92)

as \(\kappa \rightarrow +\infty \).

Due to (90), for any \(f_0\in C[0,1/k]\),

$$\begin{aligned} \lim \limits _{\kappa \rightarrow +\infty } \int \limits _0^{x_*} r(x)f_0(x)\,dx= 0, \end{aligned}$$
(93)

uniformly in \(R\in \mathcal R \).

Similarly, for all \(x^*\in (a,1/k)\) and \(f_0\in C[0,1/k]\),

$$\begin{aligned} \lim \limits _{\kappa \rightarrow +\infty } \int \limits _{x^*}^{1/k} r(x)f_0(x)v= 0, \end{aligned}$$
(94)

uniformly in \(R\in \mathcal R \).

Fix \(\varepsilon >0\) and \(f\in C[0,1/k]\). Let \(x_*\) and \(x^*\) be so close to \(a\) that \(|f(x)-f(a)|\le \varepsilon /2\) provided \(x_*\le x \le x^*\). Then

$$\begin{aligned} \left| \,\,\int \limits _{x_*}^{x^*} r(x)[f(x)-f(a)]\,dx\right| \le \varepsilon /2. \end{aligned}$$
(95)

Due to (93) and (94) with \(f_0=f-f(a)\),

$$\begin{aligned} \left| \int \limits _{0}^{1/k} r(x)[f(x)-f(a)]\,dx-\int \limits _{x_*}^{x^*} r(x)[f(x)-f(a)]\,dx\right| \le \varepsilon /2 \end{aligned}$$
(96)

for sufficiently large \(\kappa \). Thus,

$$\begin{aligned} \langle r- \widehat{R}_1\delta _{a_1},f\rangle =\langle r- \widehat{r}_1\delta _{a},f\rangle =\int \limits _{0}^{1/k} r(x)[f(x)-f(a)]\,dx\le \varepsilon . \end{aligned}$$
(97)

Appendix G: Proof of Lemma 5.3

Let \(r_{\sigma ,\kappa }\) be the solution of the system

$$\begin{aligned} \left\{ \begin{array}{ll} r_{\sigma ,\kappa } -\kappa (br_{\sigma ,\kappa })_x= R_{\sigma ,\kappa }, &{} x\in (0,1)\\ r_{\sigma ,\kappa }\ge 0. \end{array}\right. \end{aligned}$$
(98)

Then, by Lemma 5.2,

$$\begin{aligned} \lim \limits _{\kappa \rightarrow +\infty } d(r_{\sigma ,\kappa }, \sum \limits _{i=1}^k\widehat{(R_{\sigma ,\kappa })}_i\delta _{a_i}) =0, \end{aligned}$$
(99)

uniformly in \(\sigma \). Thus, it suffices to prove that for every \(\kappa \) there is \(\epsilon _\kappa >0\) such that

$$\begin{aligned} \lim \limits _{\kappa \rightarrow +\infty ,\ \sigma \le \epsilon _\kappa } d(w_{\sigma ,\kappa }, r_{\sigma ,\kappa }) =0. \end{aligned}$$

This would follow from the claim that for every \(\kappa \) there is \(\epsilon _\kappa >0\) so that for \(\sigma \le \epsilon _\kappa \) we have \(d(w_{\sigma ,\kappa }, r_{\sigma ,\kappa }) < 1/\kappa \). If it is not true, then for some \(\kappa \) there exists a sequence \(\sigma _n\rightarrow 0\) such that

$$\begin{aligned} d(w_{\sigma _n,\kappa }, r_{\sigma _n,\kappa }) \ge 1/\kappa . \end{aligned}$$

Since \(w_{\sigma _n,\kappa }\) and \(r_{\sigma _n,\kappa }\) are solutions of the problems (32) and (98), we have

$$\begin{aligned} -&\sigma _n \langle w_{\sigma _n,\kappa },\varphi _{xx}/\eta \rangle +\langle w_{\sigma _n,\kappa },\varphi +\kappa b\varphi _x\rangle = \langle R_{\sigma _n,\kappa },\varphi \rangle ,\end{aligned}$$
(100)
$$\begin{aligned}&\langle r_{\sigma _n,\kappa },\varphi +\kappa b\varphi _x\rangle = \langle R_{\sigma _n,\kappa },\varphi \rangle , \end{aligned}$$
(101)

for any \(\varphi \in C^2[0,1], \varphi _x(0)=\varphi _x(1)=0\). Since the sequences \(w_{\sigma _n,\kappa }\) and \(r_{\sigma _n,\kappa }\) lie in the space of probability measures, which is weakly-* compact, without loss of generality there exist their weak-* limits \(w_\kappa \) and \(r_\kappa \). Clearly,

$$\begin{aligned} d(w_{\kappa }, r_{\kappa }) \ge 1/\kappa . \end{aligned}$$
(102)

On the other hand, taking the difference of (100) and (101), and passing to the limit, we find \(\langle w_{\kappa }-r_{\kappa },\varphi +\kappa b\varphi _x\rangle =0,\) so \(w_{\kappa }=r_{\kappa }\) by Lemma 5.1, and we arrive at a contradiction.

Appendix H: Proof of Lemma 5.4

Multiplying the second equation in (16) by \(P\) and integrating, we find

$$\begin{aligned} \int \limits _0^1\nu (x) P^2(x) -\varsigma P_{xx}(x)P(x)\,dx=\int \limits _0^1\eta (x) p(x)P(x)\,dx, \end{aligned}$$
(103)

whence

$$\begin{aligned} \inf \limits _{0\le x\le 1}\nu (x)\int \limits _0^1 P^2(x)\,dx +\varsigma \int \limits _0^1 P_{x}^2(x)\,dx\le \sup \limits _{0\le x\le 1} P(x). \end{aligned}$$
(104)

Hence,

$$\begin{aligned} \Vert P\Vert ^2_{W_2^1(0,1)} \le C\Vert P\Vert _{C[0,1]}\le C\Vert P\Vert _{W_2^1(0,1)}\le C. \end{aligned}$$
(105)

The pair \((w_{\sigma ,\kappa }, R_{\sigma ,\kappa })=(\eta p,\nu P)\) satisfies (32). Due to (105), the set \(\mathcal R =\{\nu P\,|\, \sigma >0,\kappa >0\}\) is uniformly bounded and thus uniformly integrable. By Lemma 5.3, for every \(\kappa \) there exists \(\epsilon _\kappa >0\) such that

$$\begin{aligned} \lim \limits _{\kappa \rightarrow +\infty ,\ \sigma \le \epsilon _\kappa } d\left( \eta p, \sum \limits _{i=1}^k\widehat{(\nu P)}_i\delta _{a_i}\right) =0. \end{aligned}$$
(106)

Assume that (34) is not true, i.e. there exist \(\delta >0\) and sequences \(\kappa _n\rightarrow \infty \) and \(\sigma _n\le \epsilon _{\kappa _n}\) such that for the corresponding solutions \((p_n, P_n)=(p_{\sigma _n,\kappa _n},P_{\sigma _n,\kappa _n})\) to (16) we have \(\Vert P_{n}-Q\Vert _{C[0,1]}>\delta \). Since the embedding \(W_2^1(0,1)\subset C[0,1]\) is compact, without loss of generality we may assume that \(P_n\) converges to some limit \(P_0\) in \(C[0,1]\). Obviously,

$$\begin{aligned} \Vert P_{0}-Q\Vert _{C[0,1]}\ge \delta . \end{aligned}$$
(107)

Passing to the limit in the second, forth and the last equations in (16)—the combination of the first two is understood in the weak sense (8)—and remembering (106), we find

$$\begin{aligned} \nu P_0 - \varsigma (P_0)_{xx}&= \sum \limits _{i=1}^k \widehat{(\nu P_0)}_i \delta _{a_i},\\ (P_0)_x(0)=(P_0)_x(1)&= 0,\\ \int \limits _0^1 \nu (x)P_0(x)\, dx&= 1. \end{aligned}$$

By Theorem 4.1, \(P_0\) coincides with \(Q\), which contradicts (107).

From (106) we deduce

$$\begin{aligned} \lim \limits _{\kappa \rightarrow +\infty ,\ \sigma \le \epsilon _\kappa } d\left( p, \sum \limits _{i=1}^k\frac{\widehat{(\nu P)}_i\delta _{a_i}}{\eta (a_i)} \right) =0. \end{aligned}$$
(108)

Due to (34) and \(1/k\)-periodicity of \(\eta \), (108) implies that

$$\begin{aligned} p \rightarrow \sum \limits _{i=1}^k\frac{\widehat{(\nu Q)}_i\delta _{a_i}}{\eta (a)}. \end{aligned}$$
(109)

weakly-* as \(\kappa \rightarrow +\infty ,\ \sigma \le \epsilon _\kappa \). Taking test functions which are equal to \(1\) in one of the wells and are zero at the minima of the potential located outside of that well, we derive (35) from (109).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Vorotnikov, D. Analytical aspects of the Brownian motor effect in randomly flashing ratchets. J. Math. Biol. 68, 1677–1705 (2014). https://doi.org/10.1007/s00285-013-0684-4

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-013-0684-4

Keywords

Mathematics Subject Classification (1991)

Navigation