Localization, Escape Rate and Delocalization in Kinked Ratchet Potentials by Sang Hyun Choi Submitted to the Department of Physics in partial fulfillment of the Requirements for the Degree of BACHELOR OF SCIENCE at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY ARCHI\ES June, 2015 MASSEME T @2015 SANG HYUN CHOI All Rights Reserved INSTITUTE LG Y LO-HO S, AUG 10 2015 LIBRARIES The author hereby grants to MIT permission to reproduce and to distribute publicly paper and electronic copies of this thesis document in whole or in part Author Signature redacted Certified by Signature redacted - Department of Physics ,*ay 8, 20,15 Alfredo Alexander-Katz Thesis Supervisor, Department of Materials Science and Engineering Signature redacted /11 Jeremy England Thesis Co-Supervisor, Department of Physics Accepted by Signature redacted ___ 'Professor Nergis Mavalvala Senior Thesis Coordinator, Department of Physics I Localization, Escape Rate and Delocalization in Kinked Ratchet Potentials by Sang Hyun Choi Submitted to the Department of Physics on May 8, 2015, in partial fulfillment of the requirements for the degree of Bachelor of Science in Physics Abstract The particle localization in ratchet potentials with segments of reverse directions, or kinked ratchets, is computationally studied. Kinks localize a particle on on-off pulsating ratchet potentials, forming stable points in the effective potential. Analogous Kramers rate for transition betwen kinks is derived through simulations under different values of parameters defining the system. Adding tilting to the system with a proper choice of tilting frequency induces stochastic resonance. Delocalization of a particle is observed in the resonant activation regime. Thesis Supervisor: Alfredo Alexander-Katz Title: Associate Professor of Materials Science and Engineering Thesis Co-Supervisor: Jeremy England Title: Assistant Professor of Physics 3 4 Acknowledgments I would like to thank Professor Alfredo Alexander-Katz for giving me the opportunity to work in his group, and for his guidance and support. I appreciate all the insightful discussions we had and his consistent encouragement. I would like to thank all the members of the Alexander-Katz group for being wonderful colleagues and making my research experience fruitful. I would also like to thank Professor Jeremy England for co-supervising my thesis and Bobby Marsland from the England group for answering my questions on non-equilibrium statistical physics. Finally, I would like to thank Professor Udo Seifert and David Hartich from the Institute for Theoretical Physics at the University of Stuttgart for introducing me to the research on non-equilibrium dynamics. 5 6 Contents 1 Introduction 11 11 Study of Transport by Thermal Noise Basics of Stochastic Dynamics ..... 1.2.1 Langevin Equation....... 1.2.2 Fokker-Planck Equation . . . Overview of the Ratchet System . . 1.3.1 Pulsating Ratchet . . . . . . 1.3.2 Tilting Ratchet . . . . . . . . 1.3.3 Application of Ratchets . . . Outline of the Paper . . . . . . . . . 12 12 13 14 14 15 17 18 Method and System Description 19 3 Localization in Kinked Ratchets 3.1 Single Kink . . . . . . . . . . . . . . 3.1.1 Effective Potential . . . . . . 21 21 21 . . . . . . . . . . . . . . 22 23 24 . . 2 . 1.4 . . . . 1.3 . . 1.1 1.2 3.3 3.2.1 Effective Potential . . . . . . Periodic Kink Lattice . . . . . . . . . Double Kink Effective Rate Theory 4.1 Kramers Problem . . . . . . . . . . . . . . . . . . 4.2 Parameter Dependence of Effective Energy Barrier AG.. 4.2.1 Dependence on the Fraction of Time during which the Ratchet is Turned Off 4.2.2 Dependence on the Ratchet Height . . . . 4.2.3 Dependence on the Ratchet Asymmetry . 4.2.4 Dependence on the Distance between Kinks 4.3 Effective Diffusion Coefficient . . . . . . . . . . . 4.4 Mean First Passage Time . . . . . . . . . . . . . 25 Stochastic Resonance and Delocalization by Resonant Activation 5.1 Stochastic Resonance . . . . . . . . . . . . . . . . . . . 5.2 Stochastic Resonance in Kinked Ratchets . . . . . . . 5.3 Resonant Activation . . . . . . . . . . . . . . . . . . . 5.4 Double Kink . . . . . . . . . . . . . . . . . . . . . . . 5.5 Periodic Kink Lattice . . . . . . . . . . . . . . . . . . 35 Conclusion and Future work 43 . 29 . . . . 31 32 38 39 . 7 30 35 37 38 . 6 27 27 28 . 5 26 . . 4 3.2 Appendix A MATLAB code 47 A.1 M ain Simulation Code ................................... A.2 Ratchet Potentials ........ ..................................... 47 49 A.2.1 A.2.2 Single Kink ............................................. Double Kink ............................................. 49 50 A.2.3 Periodic Kink Lattice ...................................... 50 8 List of Figures 1.1 1.2 The potential of a pulsating on-off ratchet system. When a potential is turned off, there is no underlying potential in the system like the dot-dashed horizontal line at zero potential. In this case, a particle, drawn as a circle in the figure, undergoes free diffusion. When a potential is turned on, the particle is driven by the ratchet potential drawn in a solid line. Net drift to the right direction is resulted in this case as shown by the blue arrow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The tilting ratchet potential. (a) Before tilting or at the half period, the ratchet potential is the same as the one in Figure 1.1, resembling a sawtooth. (b) The ratchet potential with the maximum tilting in the positive direction. (c) The ratchet potential with the maximum tilting in the negative direction. Adapted from Astumian 1997 [2] 3.1 3.2 3.3 3.4 3.5 4.1 15 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 The ratchet potential with single kink. Dotted line at x = 0 denotes the location of the kink. As indicated in the arrow, a particle under this potential is driven toward the kink, being trapped there. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (a) The particle distribution under the single kink ratchet potential in linear scale. Dashed line at x = 0 indicates the location of the kink. (b) The distribution in logarithmic scale. Dashed line shows the envelope of the overall distribution. The distribution was generated with Ag = 0.2, L = 0.7, and f = 0.95. . . . . . . . . . The ratchet potential with two kinks. Dashed vertical lines indicate the location of kinks. The direction of the sawtooth changes at the kinks and the halfway between the kinks. Because of the direction of the segments of the ratchet, particles get trapped at the kinks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (a) The particle distribution in linear scale for Ag = 0.2, L = 0.7, f = 0.95 and Ax = 3.7. Dashed vertical lines at x = 3.7 denote the location of kinks, and it is where the particle has most frequently existed. (b) The distribution in logarithmic scale. Dashed lines are an envelope of the distribution. . . . . . . . . . . . . . . . . . . . (a) The particle distribution in a periodic kink lattice in linear scale for Ag = 0.2, L = 0.7, f = 0.95 and Ax = 3.7. Dashed line indicates the location of a kink. The system is in a periodic boundary condition, and one period is shown. (b) The logarithmic scale of (a). Dashed lines are the linear envelope of the distribution. (c) The distribution showing five periods. Dashed lines indicate kink locations. . . . . . 21 . 22 . 23 . 23 . 24 The effective potential of a double kink ratchet system. Due to thermal fluctuation, a particle is able to transfer from one stable point to another. Since the location of a minimum point and that of a kink coincide, the half-distance Ax between the kinks is the half-distance between the minima. AG is the effective energy barrier. . . . . 25 9 4.2 4.3 4.4 4.5 4.6 4.7 4.8 5.1 5.2 5.3 5.4 (a) The particle distribution for various fractions of time of turning off the ratchet f. Dashed lines are the logarithmic scale of the particle distribution and the solid lines are linear envelopes of the distributions. (b) The relation between AG/Ax and 1 - f. Dashed line is the linear regression of the slopes of the envelopes in (a). Error bars indicate 95% confidence bound. . . . . . . . . . . . . . . . . . . . . . . (a) The particle distribution under various ratchet heights Ag. Dashed lines show the logarithmic scale of the distribution and the solid lines show their linear envelopes. (b) The relation between 3AG/Ax and Ag. Dashed line is the linear regression of the slopes of the envelopes in (a). Error bars indicate 95% confidence bound. . . . (a) The particle distribution under various ratchet asymmetry parameters L. Dashed lines show the logarithmic scale of the distribution and the solid lines show their linear envelopes. (b) The relation between slope of the envelopes and 2L - 1. The dot-dashed line is a quadratic fitting and the dashed line is a linear fitting. Error bars indicate 95% confidence bound . . . . . . . . . . . . . . . . . . . . . . . . . (a) The particle distribution under various values of Ax. Dashed lines are the logarithmic scale of the distribution while solid lines are their linear envelopes. (b) The relation between 3AG/Ax and Ax. Error bars indicate 95% confidence bound. For comparison, the data from a single kink ratchet is also plotted as Ax = 0 case. . . Effective diffusion coefficients relative to the free diffusion coefficient for under different values of (a) f and (b) Ag. Error bars indicate one standard deviation obtained by error propagation [6]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The time series of the location of a particle for Ag = 0.2, Ax = 3.7, f = 0.95, and L = 0.7. The horizontal solid line marks the kink locations or the minimum points of Veff while the horizontal dashed lines mark the boundary of the kink region. . . (a) The distribution of time intervals for Ag = 0.2, L = 0.7, f = 0.95 and Ax = 3.7. (b) Linearly fitting MFPT on the collective variable (1 - f)Ag(2L - 1)Ax/Teff. Error bars indicate one standard deviation. . . . . . . . . . . . . . . . . . . . . . . 27 . 28 . 29 . 31 . 32 . 33 . 33 The schematic of the stochastic resonance. A particle is in a bistable potential connected to a heat bath. A suitable amount of noise in the system causes the particle to hop to a global minimum. Adapted from Gammaitoni 1998 [15] . . . . . The trajectory of a particle in a double kink ratchet with tilting with frequency W = rr/T. The parameters are Ag = 0.2, L = 0.7, f = 0.95, Ax = 3.7 and A 0 = 0.4Ag/L. The horizontal lines near x = 0 indicate the location of kinks. . . . . . . . Finding the tilting frequency that induces rapid transition. Simulations were performed for Ag = 0.2, L = 0.7, Ax = 3.7, f = 0.95 and Ao = 0.4 Ag/L. Error bars indicate one standard deviation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The particle trajectory in the presence of tilting. Tilting frequency W = 0.01 and tilting amplitude A 0 = 0.4 Ag/L. Other parameters are set as Ag = 0.2, L = 0.7, f 5.5 . = 0.95, and A x = 3.7. 36 37 38 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 (a)The particle distribution in a periodic kink lattice with resonant activation for Ag = 0.2, L = 0.7, f = 0.95, Ax = 3.7, w = 0.01 and A 0 = 0.4 Ag/L. It is showing one period. (b) The particle distribution in a periodic kink lattice with the same condition for five periods. (c) The particle distribution in a double kink ratchet for the same condition. Dashed lines in each figure indicate the location of kinks. Solid lines in (a) and (b) are the distributions for the same condition without tilting. . . 10 41 Chapter 1 Introduction 1.1 Study of Transport by Thermal Noise Imagine a pollen fallen on a surface of water. Even though you keep everything intact, the pollen would drift around with a random trajectory. The movement is called Brownian motion after a botanist Robert Brown who observed the phenomenon. Einstein and Langevin independently worked on the theory of Brownian motion [12] [21]. Langevin hypothesized that a particle, or the pollen in the example, is subject to two forces, namely a fluctuating force with frequent change in direction and magnitude and a viscous drag force. While the fluctuating force drives the random motion of the pollen, the drag force slows it down. The fluctuating force is often called thermal noise. As its name suggests, thermal noise originates from a heat bath in contact with the system of interest. Although the amplitude of thermal noise is in the order of kBT, which is very small, it has a non-negligible effect on a small system like a pollen suspended on water. At equilibrium, it is obvious that useful work cannot be extracted from thermal noise since the fluctuation is symmetric and thus the effect averages out to zero. However, when the system is driven far from equilibrium, one may suspect the possibility of doing useful work out of the thermal fluctuation. Indeed, this motivation led many physicists to working on the problem of biased Brownian motion, especially driven by unbiased perturbation. Such directional transport by Brownian motion turns out to be realized by breaking the spatial inversion symmetry of the system [27]. The system with broken symmetry is called a ratchet system, and is our central interest in this paper. 11 1.2 Basics of Stochastic Dynamics As mentioned in the previous section, small systems like a pollen on water are subject to thermal fluctations from a heat bath. In this regime, usual Newtonian mechanics and equilibrium statistical mechanics are not sufficient to describe the system. Therefore, stochastic dynamics involving random fluctuations is introduced. The dynamics of individual systems like a particle in Brownian motion follows a Langevin equation while the ensemble probability follows a FokkerPlanck equation [30]. Since the system studied in this paper is one-dimensional, the equations introduced are one-dimensional as well. 1.2.1 Langevin Equation If a small particle of mass m is in a fluid, a frictional force or a damping force due to the medium will act on the particle. Friction is written as f = -myv where -y is the damping constant and v is the velocity of the particle. However this expression is valid only when the velocity due to thermal fluctuation is negligible. Since the particle is assumed to be very small, the velocity due to thermal energy Vthermal = VkBT/m is no longer negligible. In other words, when the mass of the particle m is not much bigger than that of molecules of the medium, the effect of each collision between the particle and the fluid molecule becomes important. Because there are too many molecules in the fluid, it is impossible to solve all the coupled equations exactly. So the force by the collisions is treated as a stochastic term. As a result, the equation of motion for the particle contains the random force term as well as the friction term. v = -yv - V'(x)/m + Here, V(x) is the potential of the system and (t) (1.1) (t) is the random force, or the Langevin force, which is the fluctuating force divided by the mass m. Eq 1.1 is called a stochastic differential equation or the Langevin equation. It is identical to Newtonian force except that it contains a stochastic term. The Langevin force is assumed to be white noise, which has zero average and the 6 correlation [29] as < (t) >= 0, < (t)(t') >= (2-ykBT/m)6(t - t') 12 (1.2) where kB is the Boltzmann constant and T is the temperature. When the particle is very small and submerged in a liquid medium, the inertial force on the particle becomes negligible [26]. So the system becomes overdamped. For this overdamped limit, the inertial term 'b is neglected, and Eq 1.1 is reduced to -yv -V'(x) + = (1.3) (t) The Langevin equation is used to simulate the movement of the particle under the ratchet potential in this paper. 1.2.2 Fokker-Planck Equation As mentioned in the previous section, the microscopic interaction in the system is treated as a fluctuating stochastic term. The Fokker-Planck equation describes how the distribution function of fluctuating macroscopic variables evolves. One-dimensional Brownian motion is expressed as aw at =[ aD(1)(x) ax + D(2)(X) W ax2 (1.4) X where W(x, t) is the distribution function for the Brownian motion, D(2) (x) > 0 is the diffusion coefficient, and DC')(x) is the drift coefficient [29]. When Brownian particles are in an external field, the time evolution of the distribution function in position and velocity space W(x, v, t) can be expressed as aw= at -- 09x V+ a 9V v M + yW M av2I (1.5) where F(x) is the force from the potential of the system. This special form of the Fokker-Planck equation is called the Klein-Kramers or Kramers equation [29]. In the large friction limit where the Langevin equation reduces to Eq 1.3, the Fokker-Planck equation for the distribution function in position W(x, t) becomes t= at m-Y _9 - -F(x) + kBT a2 axx2 W (1.6) Eq 1.6 is called the Smoluchowski equation [29]. Since the system we are working on is assumed to be in the large friction limit, Eq 1.6 is used to the analytical expression for W(x, t) in this paper. 13 By solving for W(x, t) from the Fokker-Planck equation, one can get the mean value of macroscopic variables. The Fokker-Planck equation is applicable to not only systems near the equilibrium but also systems far from the equilibrium. Since the equation deals with fluctuations, it is mostly used for small systems. When the system becomes large enough to be able to neglect small fluctuations, the stochastic part, which is the diffusion term, becomes negligible, and the equation, as well as the system, becomes deterministic. 1.3 Overview of the Ratchet System At equilibrium, no work can be done from a noise. Feynman designed a microscopic "ratchet and pawl" device to discuss the possibility of utilizing thermal noise along with the anisotropy to drive a motor [14] and showed that the anisotropy of the ratchet's teeth alone can not generate net motion. Although a thermal gradient can make the device do net work, it is unrealistic to implement the thermal gradient enough to do a non-negligible work in the real system [2]. Theoretical work by Magnasco showed that if the particle is subject to an external force with time correlations, detailed balance is broken, which enables the particle to have a nonzero net drift [23]. This led to many theoretical works on the "ratchet potential", which is the asymmetric periodic potential that resembles saw teeth as shown in Figure 1.1. The ratchet system, or the Brownian motor, requires satisfaction of all three conditions to generate the net drift of a particle, which are presence of thermal noise, anisotropy of the medium, and time-dependent energy supplied either by external variations of the constraints on the system or by chemical reactions far from equilibrium [2]. The diffusive transport by thermal noise is called the ratchet effect [27]. The third condition for the ratchet effect, the external perturbations of the system, is achieved in various ways, but the most common way is to either pulsate or tilt the ratchet. 1.3.1 Pulsating Ratchet In a pulsating ratchet scheme, a set of potentials, usually two, appears alternately with certain transition rates. The alternating potentials can be oscillating between V(x) - AV and V(x) + AV or more drastically between 0 and V(x). The latter is called the on-off ratchet because the potential is "turned on" and "turned off'. The on-off ratchet is how the biased Brownian motion is induced 14 in our study. y(t) ,-( (0))[ 1+f (t) I +((t) = (1.7) Eq 1.7 shows the overdamped system with an on-off ratchet [27]. Here, V(x) is the ratchet potential and f(t) is a function of time which takes the value of either 1 or -1. f(t) = 1 describes the case when the potential is "turned on" while f(t) = -1 is the case when the potential is "turned off". In a more general case, f(t) is assumed to be an unbiased time periodic function or a stochastic stationary process [27]. 0.5an -Turned ----- Turned off 0.40.3 L 0.2 '* 0 0.1g -0 .1 -3 -3 -2 1 0 1 -1 1 1 2 3 Position Figure 1.1: The potential of a pulsating on-off ratchet system. When a potential is turned off, there is no underlying potential in the system like the dot-dashed horizontal line at zero potential. In this case, a particle, drawn as a circle in the figure, undergoes free diffusion. When a potential is turned on, the particle is driven by the ratchet potential drawn in a solid line. Net drift to the right direction is resulted in this case as shown by the blue arrow. Figure 1.1 shows the ratchet effect in the on-off ratchet. When the ratchet is turned on, a particle feels force due to the potential and moves towards a location of lower energy, tending to stay at the local minimum of the ratchet. When the potential is turned off, the particle starts to diffuse freely. As shown in Figure 1.1, the asymmetry of the sawtooth (L > 1 - L) makes the particle more likely to end up moving to the minimum on its right side than to the one on its left side when the potential is turned on again. Therefore, after a long run, the net movement of the particle is in the right direction, which is precisely the ratchet effect. 1.3.2 Tilting Ratchet In a tilting ratchet scheme, the overall slope of the ratchet is changing as if somebody is tilting the x-axis of the ratchet shown in Figure 1.2 (a). 15 (a) U = Usa saw r~p~r\ (b) (C) U= Usaw + XlFmaX I U sawXImaxI t U = Usa - x 1Fmax (t Figure 1.2: The tilting ratchet potential. (a) Before tilting or at the half period, the ratchet potential is the same as the one in Figure 1.1, resembling a sawtooth. (b) The ratchet potential with the maximum tilting in the positive direction. (c) The ratchet potential with the maximum tilting in the negative direction. Adapted from Astumian 1997 [2] In Figure 1.2, Fmax is a tilting force. When Eo/L < IFmax| < Eo/(1 - L), the potential energy V(x), or U in the figure, decreases monotonically in the case depicted in Figure 1.2 (c) because the tilting aids the drift due to the anisotropy. When the ratchet enters the regime of Figure 1.2 (b), a particle is driven to flow in the "reverse" direction of the ratchet. However, because of the constraint in the magnitude of the force, there still remain minima to trap the particle. Therefore, tilting induces net flow of particles even without thermal noise [2]. In the presence of noise, net drift of a particle occurs even for IFmaxl < Eo/L [23]. In fact, increase in thermal noise increases the flow [2]. ys(t) -V(x(t)) + y(t) + (t) (1.8) Eq 1.8 shows the overdamped system with a tilting ratchet [27]. The function y(t) is the tilting force. In general, the equation for the potential is V(x, t) = Vo(x) +xy(t). y(t) is either an unbiased periodic function, in which case it is called a "rocking ratchet", or an unbiased stationary random process, in which case it is called a "fluctuating force ratchet" [27]. In this paper, we focus on the rocking ratchet. 16 1.3.3 Application of Ratchets A ratchet potential has applications in various areas of science. Here its two most common applications are introduced. First one is a molecular motor in biology. Eukaryotic cells, which are cells with nucleus, have cytoskeleton which is a filamental structure organizing and interconnecting other organelles in a cell [27]. It is a path for intracellular transport of nutrients, wastes, proteins and others. One major type of this filament is made of proteins called microtubuli consisting of tubulin. Motor proteins such as kinesin move along this path and do the transporting job. Kinesin is processive, which means that it remains bound to the substrate for multiple rounds of activity [8], and travels only in one direction due to the asymmetric protein structure of the microtubule [27] and kinesin motor domain [18]. Surrounding environment serves as a heat bath, causing energy dissipation as well as providing reactants and taking products from chemical reactions. Kinesin experiences friction not only with the solvent but also itself and microtubule, and keeps undergoing chemical reactions with ATP to generate energy. Therefore, damping and fluctuation depend on its position, making the resulting effective potential "rough" as well as the system is coupled to non-equilibrium enzymatic chemical reactions. Working out the model with more biochemical and mathematical rigor shows that the molecular motor system of kinesin is modeled with a pulsating ratchet scheme [27] [3]. Second application is its use in modeling quantum devices. An one-dimensional rocking ratchet is realized in an asymmetric SQUID (Superconducting Quantum Interference Device) threaded by a magnetic flux [32] and an one-dimensional array of Josephson-junctions with alternating critical currents [13]. The asymmetric SQUID consists of a ring with two Josephson junctions in series in one arm and one junction in the other arm. When the ring is threaded by a magnetic flux that is not integer multiple of half the flux quanta, the total phase across the ring experiences an effective potential of a ratchet. Then when an external AC current is given to the device, only one sign of the phase velocity is favored [32], which shows the characteristics of a rocking ratchet. In a parallel array of Josephson-junctions with alternating critical currents, a quasi-particle called a fluxon moves along the pinning potential. Appropriately tuning the asymmetry of the array gives a sawtooth-like potential, and indeed fluxons show directional movement [13]. Ratchet potentials are used in numerous other systems including transport of DNA [4], transport 17 of bacteria [19] and quantum ratchets [28], not discussed in above examples. Such versatility makes the study on ratchet potentials both interesting and important. 1.4 Outline of the Paper Previous works on biased Brownian motion have been focused on the directional transport of particles [23] [25] [11] [9]. In this paper, localization of particles by biased Brownian motion is studied, which is obtained by inserting "kinks" in the ratchet. However, a particle is still able to overcome the localization and escapes the trap. Delocalization is obtained by tilting the system. I will ultimately induce stochastic resonance and resonant activation, showing the tunability of localization in a biased Brownian motion. In Chapter 2, the simulation method used in the study is reviewed. Chapter 3 discusses the localization of a particle in various kinked ratchets. Chapter 4 focuses on the escape of the particle from kinks by studying the transition rate. In Chapter 5, stochastic resonance is realized by using the escape rate obtained in Chapter 4, and delocalization by resonant activation is discussed. Finally, Chapter 6 concludes our discussion and suggests future works. 18 Chapter 2 Method and System Description The system of interest in this paper is composed of a Brownian particle under a given potential V(x, t). For simplicity, the system is studied only for one-dimensional case. The stochastic process of a Brownian particle is denoted as its spatial location, or its trajectory, x(t). Assuming the friction is large, as in a liquid medium, we use the overdamped Langevin equation shown in Eq 1.3 as the equation of motion for the particle. The study was done computationally through simulations in MATLAB. The random force (t) is implemented with MATLAB function randn, which generates a random number from the normal distribution. The spatial potential of the system V(x) is defined to be a ratchet with various conditions specified by parameters. The parameters used in the simulation are the height of the ratchet (Ag), the half-distance between kinks (Ax), the fraction of time the ratchet is turned off (f) and the asymmetry parameter of the ratchet (L). pictorially. Figure 1.1 and Figure 3.3 shows each parameter Here, Ax and L are dimensionless parameters since they actually mean the ratio of half-distance between kinks to the ratchet tooth length and the ratio of longer segment of the ratchet tooth to the whole ratchet tooth length respectively. However, the ratchet tooth length is set to be 1 as shown in Figure 1.1, so AZx and L are equivalently used as the half-distance and the length of the longer segment as in the figure. In the first part of this study, we are interested in a pulsating on-off ratchet. During simulations, a particle moves according to Eq 1.3 while the ratchet potential V(x, t) is turned on and off periodically. The particle undergoes free diffusion while V(x, t) is turned off and is drived while V(x, t) is turned on. 19 In the second part of the study, in addition to turning on and off the ratchet, the same system is being tilted periodically. The tilting of the ratchet is implemented by addition of a sine function in the force of the system. The frequency of tilting that induces stochastic resonance and resonant activation is explored. During both parts, the location of the particle x(t) is recorded after each time step. Each simulation is performed with 3 x 108 time steps. Constants such as -y, kB and m are set as 1. The MATLAB code for the simulations is attached in the Appendix. 20 Chapter 3 Localization in Kinked Ratchets Single Kink 3.1 As discussed in the Introduction, the ratchet potential induces an unidirectional motion of a particle. If the "grain" of the ratchet is reversed in the middle of the track, the directional flows converge at the point of reversal as shown in Figure 3.1. This point is called a "kink" throughout the paper. The presence of a kink significantly perturbs the directional motion of a particle [17]. Due to the collision of flow, the probability of a particle to exist at the kink becomes much higher than other locations on the track. Thus a particle gets trapped or localized at the kink. 0.4 - 0.3 10.2~.0.1 0L -4 -3 -2 -1 0 Position 12 3 4 Figure 3.1: The ratchet potential with single kink. Dotted line at x = 0 denotes the location of the kink. As indicated in the arrow, a particle under this potential is driven toward the kink, being trapped there. 3.1.1 Effective Potential Figure 3.2 (a) shows the histogram of the trajectory of a particle under the single kink ratchet potential. The higher the bin is, the more frequently the particle has existed at that location. 21 Therefore, this histogram is equivalent to the probability distribution of the particle. Each peak in the distribution indicates a local minimum in the ratchet track. As expected, the highest peak is at the kink, which confirms our assumption that the particle would be localized. (a) (b) 7 x106 16 14 6 61- 212 -14 305- 10- E 22 6- 1 0 -20 -A41 -10 0 Partide Position 10 20 -20 -10 0 Partide Position 10 20 Figure 3.2: (a) The particle distribution under the single kink ratchet potential in linear scale. Dashed line at x = 0 indicates the location of the kink. (b) The distribution in logarithmic scale. Dashed line shows the envelope of the overall distribution. The distribution was generated with Ag = 0.2, L = 0.7, and f = 0.95. The effective potential of the system affects the overall shape, or the envelope, of the distribution. Taking the logarithmic scale of the distribution shows linear envelopes as in Figure 3.2 (b). Solving the Smoluchowski equation (Eq 1.6) for stationary state [29] gives the expression of probability distribution W t (x) oc exp[-V(x)/kBT] (3.1) where V(x) is the underlying potential of the system. The stationary probability distribution shown in Figure 3.2 (b) is Nexp[-alxI, for some positive constant a and N. From Eq 3.1, the effective potential of a single kink ratchet system is thus deduced as Veff(x) = ajxI, having a V-shape. 3.2 Double Kink Now, as shown in Figure 3.3, the ratchet potential has two kinks. Since there are two kinks, the distance between them becomes an additional parameter of the system. In this paper, it is parametrized by the half-distance between the kinks, which is denoted as Ax. As in the single kink case, the reversal of the ratchet direction causes collision of particle flow and thus localization of the particle at the kinks. 22 0.4W 0.3- 0 L 0- -0.1 -6 4 6 4 2 0 Position -2 Figure 3.3: The ratchet potential with two kinks. Dashed vertical lines indicate the location of kinks. The direction of the sawtooth changes at the kinks and the halfway between the kinks. Because of the direction of the segments of the ratchet, particles get trapped at the kinks. 3.2.1 Effective Potential Due to the presence of two localizing points, the overall particle distribution has two peaks, each of which is located at the kinks as shown in the histogram of the particle trajectory in Figure 3.4 (a). As in the case of single kink ratchets, the effective potential has its minima at kinks. It is analogous to a two-well or bistable potential. (b) (a) 5x 106 _14- 4- 12 c3 C E E 1 0 -30 --- 16 86- -20 10 0 -10 Particle Position 20 4 -30 30 -20 10 0 -10 Particle Position 20 30 Figure 3.4: (a) The particle distribution in linear scale for Ag = 0.2, L = 0.7, f = 0.95 and Ax = 3.7. Dashed vertical lines at x = t3.7 denote the location of kinks, and it is where the particle has most frequently existed. (b) The distribution in logarithmic scale. Dashed lines are an envelope of the distribution. Figure 3.4 (b) shows that the envelope of the distribution is exponential. It makes sense because a double kink ratchet is superposition of two single kink ratchets. By Eq 3.1, each of the peaks is described by Wt oc exp[-adx/kBT]. Placing the resulting V-shaped potentials next to each other, we get a W-shaped effective potential of a double kink ratchet system. 23 3.3 Periodic Kink Lattice As a double kink ratchet system has two localizing points, a N-kink ratchet has N localizing points. It means that we can construct a lattice made of kinks in a ratchet. (a) (b) 2 "106 15 1.5 14- ( a- 0 13 E 12 z E0.5z 00 2 0 11 4 6 10 8 Partice Position 2 4 6 8 Particle Position 3 -106 (C) 0 E 0 10 20 Particle Position 30 40 Figure 3.5: (a) The particle distribution in a periodic kink lattice in linear scale for Ag = 0.2, L = 0.7, f = 0.95 and Ax = 3.7. Dashed line indicates the location of a kink. The system is in a periodic boundary condition, and one period is shown. (b) The logarithmic scale of (a). Dashed lines are the linear envelope of the distribution. (c) The distribution showing five periods. Dashed lines indicate kink locations. Figure 3.5 shows the particle distribution in a periodic kink lattice, in which kinks are placed periodically. The period of kink placement is 2Ax, and the system is constructed with a periodic boundary condition. Figure 3.5 (a) shows the distribution for one period. As in the case of a single kink ratchet and a double kink ratchet, a particle is localized at kinks, forming a highest peak in the distribution. Figure 3.5 (b) shows that the envelope of the distribution is exponential as in the previous cases. So the effective potential of the system becomes a symmetric sawtooth-shaped potential with its minima located at kinks. Figure 3.5 (c) shows five periods, which is essentially alignment of 5 copies of (a). A particle tends to be localized at kinks, forming high peaks of the distribution there. The difference among peak heights will disappear with more statistics. 24 Chapter 4 Effective Rate Theory Our discussion on the double kink ratchet concludes that the effective potential is a W-shaped bistable potential shown in Figure 4.1. As time evolves, a particle is not only confined to one minimum point of Veff (x) but also transfers to another minimum point. Indeed, such transition is obvious in Figure 3.4 where the probability of a particle to exist in either kink is almost the same. AX 00 AG Position Figure 4.1: The effective potential of a double kink ratchet system. Due to thermal fluctuation, a particle is able to transfer from one stable point to another. Since the location of a minimum point and that of a kink coincide, the half-distance Ax between the kinks is the half-distance between the minima. AG is the effective energy barrier. Tuning the parameters of a ratchet changes the effective energy barrier AG. An energy barrier is related to the transition rate of a particle to move from one potential well to another by Kramers relation discussed soon. Therefore, probing AG gives us information about the transition rate in the kinked ratchet system. 25 4.1 Kramers Problem When a particle is on a potential that has maxima and minima, it tends to stay in one of the minima because they are energetically more favorable. However, random fluctuation in the energy can induce the particle to jump over the energy barrier and move to another minimum. Transition rate is most commonly studied in chemical reactions, for example, from molecule A to molecule B, including biochemical reactions. Non-equilibrium dynamics is concerned about the transition rate of a Brownian particle to overcome a potential barrier. This problem is called the Kramers problem [33]. One way to approach this problem is to use the mean first-passage time (MFPT), which is defined as the average time elapsed until a stochastic process starting at some point xo leaves a prescribed domain of the state space for the first time [16]. Solving the Fokker-Planck equation gives the MFPT r as T(X) = k+] b eu(IkBT dz e-Utz)IkBT (4.1) where a is a reflecting point and b is an absorbing point in the potential U(x) [16]. Kramers problem is solved using Eq 4.1. The absorbing barrier b is Xmax where U(xmax) is the maximum of the potential. The reflecting barrier a is set to be -oo. kBT is assumed to be small, which implies a high barrier limit. In this condition, the integral over z is dominated by the potential near the minimum, whose expression is obtained by quadratic expansion as U(z) = Umin + 1w 2 .(z - xmin) 2 . Likewise, the integral over y is dominated by the potential near the maximum with quadratic expansion as U(y) Umax - 2 2 mnax(Y - Xmax) [33]. Calculating Eq 4.1 with these approximations gives the expression for T. Then the transition rate or the Kramers escape rate k is just a reciprocal of k = r- = WminWmax 7rm-Y I T, which is [20] Umax - Umin oc exp[-AU/kBT] kBT I (4.2) It is notable that k from Eq 4.2 reproduces the Arrhenius equation, k oc exp[-AE/kBT] [1] where AE is the activation energy, or the energy barrier, which is Umax - Umin in our case. 26 Parameter Dependence of Effective Energy Barrier AG 4.2 By Eq 3.1, the probability distribution is expected to be exp[-Veff(x)/kBT. From Figure 3.4, it is also exp[-3ajxj] where 3 is 1/kBT. a is the slope of the linear segment of Veff(x), so a = AG/Ax. In other words, Veff(x) = (AG/Ax)x. Using this relation, the dependence of AG on various parameters of the system, which are f, Ag, L, and Ax, is investigated by measuring the slope of Veff(x). 4.2.1 Dependence on the Fraction of Time during which the Ratchet is Turned Off When the ratchet is turned off, a particle undergoes free diffusion, which is much faster than the movement of the particle interfered by a potential. Therefore, the longer the potential is turned off (bigger the farther the particle reaches. A broader distribution of a particle is interpreted f), f, as lower AG. To verify the relation between AG and simulations are performed for 0.900, 0.925, 0.950 and 0.975. Other parameters are set as Ag (a) 18 (b) f=0.875 X f =0.90 f = 0.925 * f = 0.95 0 f = 0.975 16- 0.2, L f 0.7 and Ax = 0.875, 3.7. -- 2.5 -+ 2- 01.5 0 62 12- _ _ _ _ _ _ _ _ _ --0.5- ___fi -20 -0.5- 100- -15 -10 0.02 -5 0.04 0.06 0.08 0.1 0.12 0.14 1-f Particle Position Figure 4.2: (a) The particle distribution for various fractions of time of turning off the ratchet f. Dashed lines are the logarithmic scale of the particle distribution and the solid lines are linear the envelopes of the distributions. (b) The relation between #AG/Ax and 1 - f. Dashed line is bound. confidence 95% linear regression of the slopes of the envelopes in (a). Error bars indicate Figure 4.2 (a) is superposition of the particle distributions shown in Figure 3.4 (b) under various values of f. As expected, a higher value of f yields a broader distribution or weaker localization, 27 implying lowering of the effective potential barrier. Figure 4.2 (b) shows the linear relation between AG and of f (1 - f. In the plot, the variable is (1 - f) instead of f. Reminding ourselves that the function we are looking for is going to be an exponent of the probability distribution convinces us that f) is a better variable since f = 1, which means the potential is turned off all the time, gives us uniform distribution without biasedness. In other words, conclude that AG oc (1 - f = 1 should mean AG = 0. So we f). However, extrapolating the linear fitting in Figure 4.2 (b) gives AG = 0 before f =1. implies different physics at f ~ 1, which is a regime close to free diffusion. For now, we take the asymptotic linear relation between AG and (1 - 4.2.2 This f). Dependence on the Ratchet Height It is uncertain whether AG would depend on Ag because a particle travels mostly during the time the potential is turned off. To see the possible relation, simulations are performed for Ag 0.15, 0.2, 0.25, 0.3, 0.35, and 0.4. Other parameters are set as f = 0.95, L = 0.7 and Ax = 3.7. 18 16 14 0 X EO = E = EO = -+ *EO = 0 EO = EO = (b) 2 0.15 02 025 0.3 0.35 0,4 1.8 -- 1.6 - (a) 1.4S12 1.2 CL 10 E o. 0.8 -ii 0.6 6 0.4 4 2 0.2 -14 -12 -10 -8 -6 Particle Position -4 0 0.1 -2 0.2 0.3 0.4 0.5 Ag Figure 4.3: (a) The particle distribution under various ratchet heights Ag. Dashed lines show the logarithmic scale of the distribution and the solid lines show their linear envelopes. (b) The relation between 3AG/Ax and Ag. Dashed line is the linear regression of the slopes of the envelopes in (a). Error bars indicate 95% confidence bound. Figure 4.3 shows that a bigger value of Ag gives a steeper slope of the envelope. This implies a bigger value of AG and thus stronger localization. As shown in Figure 4.3 (b), AG depends linearly 28 on Ag. However, as in the case for f dependence, extrapolation of the linear fitting gives AG = 0 before Ag falls to zero, suggesting the possible entrance to the new regime. For a regime where a ratchet effect is still robust, it is valid to say AG oc Ag. 4.2.3 Dependence on the Ratchet Asymmetry One of the conditions necessary for the ratchet effect is asymmetry in a ratchet. As shown in Figure 1.1, if L = 1 - L, the probability for a particle to land on a trap on either side becomes identical, and the biasedness in its motion disappears. One can speculate that the more asymmetric the ratchet is, the more robust the directional motion is. More robust ratchet effect results in stronger localization. So this implies the possible correlation between AG and L. The simulations are performed for L = (a) 18 0.6, 0.65, 0.7, 0.75, and 0.8. Other parameters are set as f 0.95, Ag = 3.7. 0 (b) L =0.60 1-X L = 0.65 + L0.9 * L = 0.75 14- = 1 09 - 0.2, and Ax = 0.8- 0 L = 0.80 0.7 12 CL 10 V/ %/ 0.6 e 6 - 0.5 n ;0.3 S0.4 4 -20 0.2 21 -15 -10 Partide Position 0.11 0.1 -5 0.2 0.3 0.4 0.6 0.5 0.7 2L-1 Figure 4.4: (a) The particle distribution under various ratchet asymmetry parameters L. Dashed lines show the logarithmic scale of the distribution and the solid lines show their linear envelopes. (b) The relation between slope of the envelopes and 2L - 1. The dot-dashed line is a quadratic fitting and the dashed line is a linear fitting. Error bars indicate 95% confidence bound. Figure 4.4 shows that bigger values of L gives a steeper slope of the envelopes. This implies that the more asymmetric the ratchet is, the stronger the localization is. The variable on the x-axis is set to be 2L - 1 rather than L so that the biasedness in the motion vanishes at L = 0.5, which is the symmetric case, by the same argument why the proper variable was 1 case for the time fraction dependence of AG. 29 f instead of f in the Figure 4.4 (b) suggests either fAG oc (2L - 1)2 or /AG oc (2L - 1). The data points themselves seem that quadratic relation is better. However, collecting all the parameters together in the later section suggests the other option. Therefore, the data points for L close to 0.5, which gives free diffusion, are thought to be in the regime between free diffusion and a ratchet effect suggested for other parameters. The entrance to the regime happens earlier than other parameters, given that asymmetry is a necessary condition for a ratchet effect. There are two earlier studies supporting the linear relation in our regime of study. First, for a steep ratchet, the force to stop the movement of a particle in the reverse direction to the ratchet effect is F8 tp = (Qy/ft)/(L - 1/2) [2] [22] where t is the duration time. Hence the net force causing the directional motion is F8 tp oc (2L - 1). F8 tp comes from the effective potential of the system by Ftop = -xVeff(x). So Veff(x) oc (2L - 1) but then Veff(x) oc AG and thus AG oc (2L - 1). Second, for a two-state pulsating ratchet with low switching rate, velocity of a particle v = (L - (1 - L))/2Tir 2 [9] where -r and T2 are duration times for each potential of the two states. In an overdamped condition like our system, force F oc v, which means that F 0c (2L - 1). The same argument as before concludes that AG cc (2L - 1). 4.2.4 Dependence on the Distance between Kinks When the distance between the two kinks is long, the transition between them is expected to be more difficult simply because a particle has to travel farther. Then, the effective potential barrier is larger. In order to verify the hypothesis, simulations are performed under various half-distances between kinks Ax = 1.7, 2.7, 3.7, 4.7, and 5.7. Other parameters are set as f = 0.95, Ag = 0.2 and L = 0.7. Figure 4.5 (a) shows that the envelopes of particle distributions are almost parallel to each other. The data from a single kink ratchet is included in the plots as Ax = 0 case for comparison, and its slope is parallel, too. Figure 4.5 (b) suggests independence of the slopes of the envelopes from Ax. Here again, the data points of double kink ratchets have similar values as a single kink ratchet, which confirms that the slopes are indeed independent of Ax. The slopes indicate OAG/Ax. Therefore, we conclude that AG oc Ax. 30 (a) 16 (b) 0.44 V dx = 0.0 0 dx =1.7 X 15 1 14 dx = 2.7 0.43 I +| dx = 3.7 dx = 4.7t0.2 0 dx = 5-7 04 X - 13- 0 012- 1 1 0.41 0.4- 10.39 10 0.38 9 0.37 0o.36~ 8 -15 -10 0 0 -5 Figure 4.5: (a) The particle distribution mic scale of the distribution while solid 3AG/Ax and Ax. Error bars indicate single kink ratchet is also plotted as Ax 4.3 4 2 6 AX Particle Position under various values of Ax. Dashed lines are the logarithlines are their linear envelopes. (b) The relation between 95% confidence bound. For comparison, the data from a = 0 case. Effective Diffusion Coefficient Before collecting all the parameters for AG, we should consider the effective temperature of the system because, as Eq 4.2 and our discussion on the logarithmic slope of particle distributions show, AG accompanies kBT. Diffusion coefficient D is proportional to temperature T by the Einstein relation D = kBT-y [12]. Likewise, Teff oc Deff where the subscript "eff" means "effective". In the presence of a ratchet, Deff is expected to be different from D, the free diffusion coefficient, in a sense that the potential landscape is "rough" due to the sawtooth of the ratchet compared to the complete flat one for free diffusion. In order to verify the effect of a ratchet on diffusion, Deff is measured through simulations according to the following equation [29]. D= 1 1 lim - < [x(t) 2 t-+oo t - - x(O)] 2 > Figure 4.6 shows the effective diffusion coefficient Deff under different values of Ag and (4.3) f. Def f is measured in a symmetric ratchet (L = 0.5) not to consider the drift due to the ratchet effect. Since the roughness of the potenital landscape is what affects diffusion, Deff does not depend on the asymmetry parameter L. 31 (a) (b) 14 I 0.9 - 12 0.80 0.7- 0 ) 0.8 0.6- 0 0.6 0.5- 0.40.2 0.86 0.40.88 0.9 0.92 0.94 0.96 0.1 0.98 0.0033.4 0.2 0.3 0.4 0. 0.5 fAg Figure 4.6: Effective diffusion coefficients relative to the free diffusion coefficient for under different values of (a) f and (b) Ag. Error bars indicate one standard deviation obtained by error propagation [6]. Figure 4.6 shows that Def f depends linearly on f and Ag. The dependence makes sense because the movement of a particle would be disturbed more as the potential persists longer (f decreases) and as the ratchet becomes higher (Ag increases). So the presence of a ratchet indeed decreases Deff. The value of Deff/D for f = 0.975 in Figure 4.6 (a) is larger than 1, but this data point has a large error bar, which is presumably caused by that the behavior at this point is effectly free diffusion. So Deff ID ~1 at this point. However, Deff 4 D for most of the conditions in the presence of a ratchet. In the later section, we collect all the parameters and fit the data into the Kramers reaction theory model. While the Kramers theory deals with quasi-equilibrium dynamics, our system is in non-equilibrium. Therefore, temperature is modified from the "true temperature" T in effective temperature Teff = TDeff /D. # to the It turns out to help fitting our data into the equilibrium picture of Kramers. 4.4 Mean First Passage Time Collecting all the parameter dependences gives the relation AG oc (1 - f)Ag(2L - 1)Ax. By Eq 4.2, mean first passage time (MFPT) T oc exp[AG/kBTeff]. Therefore, fitting the model for AG in directly measured MFPT is expected to confirm our model. Figure 4.7 is a trajectory of a particle under a double kink system. Distributions like Figure 3.4 (a) are histograms of trajectories like this. The trajectory shows that a particle stays at one kink- 32 I IIIII I I 15 10 0 Lit 5 LLKLi~L 7 77 0 a- -5 L~6 hkiI~LI 7~7~- V WL f-, I -10 -15 -20 0 1 2 6 5 Time 4 3 7 10 9 8 5 x10 Figure 4.7: The time series of the location of a particle for Ag = 0.2, Ax = 3.7, f = 0.95, and L = 0.7. The horizontal solid line marks the kink locations or the minimum points of Veff while the horizontal dashed lines mark the boundary of the kink region. trap for certain amount time and then transfers to the other trap. Therefore, taking the average of the duration time the particle stays at one trap gives MFPT. (a) (b) 16 70 o Parameter: Ag * Parameter: 1-f g Parameter: 2L-1 X Parameter: Ax 15 60 14 --4 50 13 '40 12 a, - 30 20 1 9C X 10 U -- 10 5 Time Interval M I - M 0 0 15 0.02 0.04 0.06 0.08 0.1 0.12 (1-f)(Ag)(2L-1)(Ax)/Teff ,104 Figure 4.8: (a) The distribution of time intervals for Ag = 0.2, L = 0.7, f = 0.95 and Ax = 3.7. (b) Linearly fitting MFPT on the collective variable (1- f)Ag(2L - 1)Ax/Teff. Error bars indicate one standard deviation. Figure 4.8 (a) is the distribution of time intervals during which a particle stays in one kink before moving to another kink. As shown with the dashed line, the histogram of the intervals is well fitted in an exponential distribution. MFPT is obtained by taking the mean of the distribution. Figure 4.8 (b) shows the relation between MFPT and all the parameters discussed before. Here the collective variable (1 - f)Ag(2L - 1)Ax is divided by the relative effective diffusion coefficients 33 Deff /D because by Eq 4.2, the exponent of r is not just AG but AG/kBTeff as discussed in the previous section. By Einstein's relation [12], Teff O( Deff. The linear fit in Figure 4.8 (b) is reasonably well fitted in the "middle" and "upper" region. The data is not as well fitted in the small (1 - f)Ag(2L - 1)Ax region, where the constraint due to ratchets is weak enough that the particle movement is close to free diffusion. From the divergence of the data points, it is suspected that this is a separate regime that requires another further study. Especially, Ax dependence in this regime needs reexamination. The error bars in Figure 4.8 (b) are obtained by the standard deviation of mean and the error propagation. The standard deviation of a mean is gmean = o/V(N) [6] where o is standard deviations of the interval distribution from Figure 4.8 (a) and N is the number of samples. Since the y-axis is the logarithm of r , the uncertainty in ln(T) is gin(,) = 01,/ [6]. Overall, logarithmic scale of MFPT has a linear relation with the collective variables. So this establishes an analogous expression for the Kramers escape rate k by taking reciprocal of MFPT T k =T k cx exp[-AG/kBTeff] = exp[-Go(1 - f)(2L - 1)AxAg/kBTeff] where Go is a constant factor for AG. 34 (4.4) Chapter 5 Stochastic Resonance and Delocalization by Resonant Activation 5.1 Stochastic Resonance So far the system is subject to periodic turning on and off of the ratchet potential. It becomes curious what happens if another periodic external force, which is tilting in this case, is applied to the system. A careful choice of the tilting frequency is expected to realize stochastic resonance. Stochastic resonance is a phenomenon that a signal is amplified by noise in the system. It was originally introduced by Benzi to study the climage change of periodic coming of ice age [5]. The climate was modeled as a two-well potential and one of the wells represented ice age. Short-term climate fluctuations are Gaussian white noise and if they are synchronized with weak purturbations the Earth is subject to, stochastic resonance occurs and a giant change in the climate comes. This study showed that the coupling between an internal stochastic mechanism and an external periodic forcing causes the large amplitude, long-term alternations in temperature and the absense of either fails to reproduce the phenomenon. In a more general setting, if a particle is under a double-well potential that is connected to a heat bath, it is constantly subject to random fluctuations. The fluctuations cause a particle in one well to jump over an energy barrier and move to the other well, and its transition rate follows the Kramers rate from Eq 4.2. Until now, the system is the same as our previous systems. 35 Then a [N ago #10 *go Figure 5.1: The schematic of the stochastic resonance. A particle is in a bistable potential connected to a heat bath. A suitable amount of noise in the system causes the particle to hop to a global minimum. Adapted from Gammaitoni 1998 [15] weak periodic force is applied. yv = -V'(x) + Aosin(wt) + (t) (5.1) Eq 5.1 describes the overdamped system with periodic tilting where W is a tilting frequency and AO is a tilting amplitude [27] [15]. With this periodic force, the once symmetric bistable potential becomes asymmetric. As shown in Figure 5.1, the energy barrier between two wells raises and lowers periodically in an antisymmetric manner [15]. In fact, the lowering of the energy barrier due to tilting itself is too small for a particle to overcome the barrier because we are assuming the force in Eq 5.1 to be weak. However, noiseinduced transtion described by Eq 4.2 can be synchronized with this force. The synchronization happens when the time-scale matching condition for stochastic resonance [15] [24] is satisfied as T, = 2T 36 (5.2) where T is MFPT from Eq 4.2 for noise-induced transtion and T, is the period of forcing, which is obtained by T, = 2ir/w. 5.2 Stochastic Resonance in Kinked Ratchets 1400 V 1200 1000- 800 - 0 0 600- 400 - 200 0 2 4 6 Time 8 12 10 x104 Figure 5.2: The trajectory of a particle in a double kink ratchet with tilting with frequency w =r/T. The parameters are Ag = 0.2, L = 0.7, f = 0.95, Ax = 3.7 and AO = O.4Ag/L. The horizontal lines near x = 0 indicate the location of kinks. and Adding tilting to the system according to Eq 5.1 gives a combination of a pulsating ratchet observe a rocking ratchet. Especially, tilting with frequency w according to Eq 5.2 enables us to on a tilted stochastic resonance in kinked ratchet potentials. Figure 5.2 is a trajectory of a particle gigantic double kink ratchet potential with w = 7r/T. The trajectory looks like a sine function with at amplitude. The amplitude is so big that the location of kinks seems not to affect the trajectory all. In fact, the particle behaves as if a ratchet does not exist. The ignorance of the ratchet suggests that the system is undergoing stochastic resonance. The input tilting amplitude was 0.4 Ag/L = without 0.11, which is small enough that a particle would be still trapped in a ratchet minimum noise. However, coupling the tilting with thermal noise results in amplification of the output signal, which means that the system behaves as if huge tilting is applied. The effective tilting is so large force that the underlying ratchet potential becomes relatively flat and the particle does not feel the is indeed by the ratchet. Inspecting the period of the sinusoidal signal in Figure 5.2 confirms that it the external tilting that we applied. Therefore, we conclude that the coupling between thermal noise in the system and the signal, or tilting we applied, has given rise to stochastic resonance. 37 Resonant Activation 5.3 Having a periodic perturbation in the system can not only amplify the input perturbation through stochastic resonance but also increase the Kramers transition rate. Doering and Gadoua named the latter as resonant activation [10]. It is not to be confused with stochastic resonance. While stochastic resonance is concerned with the amplification of input signal by interference with the noise in the system, resonant activation is about the increase in the Kramers escape rate. In this scheme, the height of energy barrier fluctuates between E + AE and E - AE. A careful choice of the fluctuation frequency brings about "resonance" between the fluctuation and the noiseinduced transtion, leading to decrease in MFPT. The exact analytical expression for MFPT is very demanding to get, so it has been studied with approximations [10] [7] [31].The exact expression for MFPT will enable us to get the "resonance frequency" wr which induces the resonant activation with minimum MFPT. Double Kink 5.4 The effective potential for a double kink ratchet is W-shaped or bistable. Tilting the ratchet raises and lowers the effective energy barrier AG in an antisymmetric manner as in Figure 5.1, and a particle is in one of the potential wells. In the perspective of the particle, therefore, the barrier AG is fluctuating, which is the setting for resonant activation. For various values of tilting frequency w, MFPT is measured to see how the transition rate of a particle depends on it. 12000 -- 10000- - 8000 E- uL 600040002000 ____ 104 _ 103 __ __10-2 101' 100 Tilting Frequency Figure 5.3: Finding the tilting frequency that induces rapid transition. Simulations were performed for Ag = 0.2, L = 0.7, Ax = 3.7, f = 0.95 and AO = 0.4 Ag/L. Error bars indicate one standard deviation. 38 Figure 5.3 shows MFPT for various tilting frequencies w. The value of MFPT decreases and increases again as it passes a certain value of w of order 10-2. MFPT at w = 0.01 is 399.57 which is significantly smaller than MFPT for the same condition without tilting, which is r = 1.9418 x 104. Such huge decrease in MFPT signifies that the resonance frequency w, is of order 10-2. 30 -20 10 -30 0 2 4 6 lime 8 10 12 x104 Figure 5.4: The particle trajectory in the presence of tilting. Tilting frequency w = 0.01 and tilting amplitude Ao = 0.4 Ag/L. Other parameters are set as Ag = 0.2, L = 0.7, f = 0.95, and Ax = 3.7. Rapid transition between the kinks implies that delocalization of a particle in a kinked ratchet has been realized. As shown in Figure 5.4, a particle rapidly moves between the kinks. It is not confined to one trap but rather delocalized between them. Tilting the system with different w far from the resonance frequency does not delocalize the particle. Therefore, it shows the tunability of localization of a particle in kinked ratchet systems. 5.5 Periodic Kink Lattice In a periodic kink lattice, resonant activation induces delocalization of a particle not only around the region between two kinks but at everywhere. Figure 5.5 shows the comparison between the particle distribution with resonant activation for a double kink ratchet and a periodic kink lattice. Figure 5.5 (a) is the distribution for a periodic kink lattice for one period. Since there are multiple kinks in the system, the particle is completely delocalized. The probability for a particle to exist at the local minima of a ratchet is still higher than other places, which is demonstrated by the peaks at the local ratchet minima. However, the particle is delocalized in a sense that kinks are no longer the location for the highest peak in the distribution. All the minima in the 39 ratchet are equally stable points. Delocalization is more obvious when compared to the distribution without tilting which is drawn with a solid line on the same plot. It is more apparent in Figure 5.5 (b) which shows five periods of the kink lattice. Indeed, due to delocalization, the particle is effectively uniformly distributed throughout the five periods. Note that it is almost impossible to distinguish kinks from other minima on the ratchet without the dashed lines indicating them. It is contrasting with the distribution without resonant activation drawn with solid lines, which shows localization of the particle at kinks. On the other hand, Figure 5.5 (c) shows that a particle in a double kink ratchet is mostly around the kinks and the region between the kinks. Although the particle occasionally goes farther than the kink region, possibly due to its ballistic-like motion, it moves between kinks for the most of the time. So the overall shape of the distribution looks like a Gaussian distribution. It is contrasting to the distribution in the absence of tilting which is highly peaked at kinks. Comparison between Figure 5.5 (a), (b) and Figure 5.5 (c) shows the difference in the extent of delocalization. Overall, Figure 5.5 suggests the tunability of delocalization in kinked ratchet systems. 40 (a) 2 6 x 10 1 0 0 UJ Q) 1 2 4 5 6 7 8 20 25 30 35 40 10 0 10 Particle Position 20 30 40 3 C3 ~ (b) 3 x 10 I a.. t+-- 0 2 E 1 ..... Q) ..c ::I z 0 0 5 10 15 (c) 6 x 10 4 2 oL__ _.___.-.aiitm• -40 30 20 Figure 5.5: (a)The particle distribution in a periodic kink lattice with resonant activation for ~g = 0.2, L = 0.7, f = 0.95, ~x = 3.7, w = 0.01 and Ao = 0.4 ~g/ L. It is showing one period. (b) The particle distribution in a periodic kink lattice with the same condition for five periods. (c) The particle distribution in a double kink ratchet for the same condition. Dashed lines in each figure indicate the location of kinks. Solid lines are the distributions for the same condition without tilting. 41 42 Chapter 6 Conclusion and Future work In conclusion, localization, escape rate and delocalization in kinked ratchets have been discussed. When the direction of a ratchet potential reverses, the point of reversal, which is called a kink, becomes the most probable region for a particle to exist because of the ratchet effect. Therefore, localization of a partcle is achieved. All the three kinds of kinked ratchets, a single kink ratchet, a double kink ratchet and a periodic kink lattice, exhibit localization. The particle distribution shows that the effective potential at kinks is V-shaped. In the case of a double kink ratchet, the dependence of escape rate k from one kink to the other on various parameters, Ag, L, studied and the analogous Kramers rate turns out to be k oc exp[-Go(1 - f)(2L - f and Ax, is 1)AxAg/kBTeff]. Then, tilting the kinked ratchet with frequency w = 7rk gives rise to stochastic resonance. In this regime, the originally very weak tilting amplitude is magnified by coupling with the thermal noise in the system. The output "signal" is so large that the underlying ratchet is ignored. In both a double kink ratchet and a periodic kink lattice, delocalization of a particle around kinks is observed for the resonant activation regime. Because kinks are placed periodically in a kink lattice, it is more apparent that the preference on kinks disappears and a particle becomes equally likely to exist at all minima in the ratchet. There still remain unsolved questions. First of all, our model of the analogous Kramers rate does not work well for small (1 - f)(2L - 1)AxAg, which is a regime close to free diffusion. It seems that the transition between free diffusion and a robust ratchet effect does not occur smoothly with our model. Therefore, further study on this regime is needed to understand the whole system. Second, we do not have an exact analytical expression for MFPT when tilting is added. The order 43 of the resonance frequency wr was obtained by simulations, but derivation of the expression will give us the exact value of wr for the minimum MFPT. Finally, it will be worth exploring the ignorance of ratchets in the stochastic resonance regime in other conditions. Our study has demonstrated that placing kinks on a ratchet potential localizes particles in the system. However, additionally tilting the potential causes stochastic resonance, resonant activation and consequential delocalization. It suggests tunability of localization and thus more control over the system. Experimental demonstration of this work is needed but is expected to be feasible. Kinked ratchet potentials will open more possibilities on applications and studies on the systems with ratchet potentials. 44 Bibliography [1] S. Arrhenius. Uber die Reaktionsgeschwindigkeit bei der Inversion von Rohrzucker durch Sduren. Z. Phys. Chem., 4:226-248, 1889. [2] R. Astumian. Thermodynamics and kinetics of a brownian motor. Science, 276:917-922, 1997. [3] R. Astumian and I. Der6nyi. A chemically reversible brownian motor: Application to kinesin and ncd. Biophys. J., 77:993-1002, 1999. [4] J. Bader et al. DNA transport by a micromachined brownian ratchet device. Proc. Natl. Acad. Sci. U.S.A., 96:13165-13169, 1999. [5] R. Benzi, G. Parisi, A. Sutera, and A. Vulpiani. Stochastic resonance in climatic changes. Tellus, 34:10-16, 1982. [6] P. Bevington and D. Robinson. Data Reduction and Error Analysis for the Physical Sciences. McGraw-Hill, 2003. [7] M. Bier and R. Astumian. Matching a diffusive and a kinetic approach for escape over a fluctuating barrier. Phys. Rev. Lett., 71:1649-1652, 1993. [8] S. Block. Leading the procession: New insights into kinesin motors. J. Cell. Biol., 140:1281- 1284, 1998. [9] J. Chauwin, A. Ajdari, and J. Prost. Current reversal in asymmetric pumping. Europhys. Lett., 32:373-378, 1995. [10] C. Doering and J. Gadoua. Resonant activation over a fluctuating barrier. Phys. Rev. Lett., 69:2318-2321, 1992. [11] C. Doering, W. Horsthemke, and J. Riordan. Nonequilibrium fluctuation-induced transport. Phys. Rev. Lett., 72:2984-2987, 1994. [12] A. Einstein. Uber die von der molekularkinetischen Theorie der Wdrme geforderte Bewegung von in ruhenden Fliissigkeiten suspendierten Teilchen. Ann. Phys., 17:549-560, 1905. [13] F. Falo, P. Martinez, J. Mazo, and S. Cilla. Ratchet potential for fluxons in Josephson-junction arrays. Europhys. Lett., 45:700-706, 1999. [14] R. Feynman, R. Leighton, and M. Sands. The Feynman Lectures on Physics. Addison-Wesley, 1966. [15] L. Gammaitoni, P. Hdnggi, P. Jung, and F. Marchesoni. Phys., 70:223-287, 1998. 45 Stochastic resonance. Rev. Mod. [16] P. Hinggi and P. Talkner. Reaction-rate theory: fifty years after Kramers. Rev. Mod. Phys., 62:251-341, 1990. [17] T. Harms and R. Lipowsky. Driven ratchets with disordered tracks. Phys. Rev. Lett., 79:2895- 2898, 1997. [18] U. Henningsen and M. Schliwa. Reversal in the direction of movement of a molecular motor. Nature, 389:93-96, 1997. [19] S. Hulme, W. DiLuzio, S. Shevkoplyas, L. Turner, M. Mayer, H. Berg, and G. Whitesides. Using ratchets and sorters to fractionate motile cells of Escherichia coli by length. Lab Chip, 8:1888-1895, 2008. [20] H. Kramers. Brownian motion in a field of force and the diffusion model of chemical reactions. Physica, 7:284-304, 1940. [21] P. Langevin. Sur la thdorie du mouvement brownien. Comptes Rendues, 146:530-533, 1908. [22] D. Liu, D. Astumian, and T. Tsong. Activation of Na+ and K+ pumping modes of (Na,K)ATPase by an oscillating electric field. J. Biol. Chem, 265:7260-7267, 1990. [23] M. Magnasco. Forced thermal ratchets. Phys. Rev. Lett., 71:1477-1481, 1993. [24] B. McNamara and K. Wiesenfeld. Theory of stochastic resonance. Phys. Rev. A, 39:4854-4869, 1989. [25] J. Prost, J. Chauwin, L. Peliti, and A. Ajdari. Asymmetric pumping of particles. Phys. Rev. Lett., 72:2652-2655, 1994. [26] E. Purcell. Life at low reynolds number, 1976. [27] P. Reimann. Brownian motors: noisy transport far from equilibrium. Phys. Rep., 361:57-265, 2002. [28] P. Reimann, M. Grifoni, and P. Hsinggi. Quantum ratchets. Phys. Rev. Lett., 79:10-13, 1997. [29] H. Risken. The Fokker-Planck Equation. Springer, 1996. [30] U. Seifert. Stochastic thermodynamics, fluctuation theorems and molecular machines. Rep. Prog. Phys., 75, 2012. [31] C. Van den Broeck. Simple stochastic model for resonant activation. Phys. Rev. E, 47:4579- 4580, 1993. [32] I. Zapata, R. Bartussek, F. Sols, and P. Hdnggi. Voltage rectification by a SQUID ratchet. Phys. Rev. Lett., 77:2292-2295, 1996. [33] R. Zwanzig. Nonequilibrium Statistical Mechanics. Oxford Universiy Press, 2001. 46 Appendix A MATLAB code A.1 Main Simulation Code % Ratchet potential random walk %% Parameters and Variables t = 0.0; dt = 0.01; tmax = 1000000; E0 = 0.2; % barrier height of the ratchet LO = 0.7; % adjust slope deltax = 3 + LO; % distance from the origin to the "minimum" x = 0.0; % position of the particle v = 0.0; % velocity of the particle i = 0; timeseries = zeros (1,tmax/dt); timestep = zeros (1,tmax/dt); passage = [J; turnOff = 0.95; % ratio of the time the potential is turned off turningPeriod = 500; where = 0; AO = 0.4*EO/LO; w = 0.01; %w = pi/(1.9285*10^4); % tilting freq. %tau = 1.9285*10^4 for EO = 0.2, LO = 0.7, turnOff = 0.95, deltax = 3.7; %% Simulation while t < tmax i = i + 1; if mod((100*t),turningPeriod) < turnOff*turningPeriod % potential turn off. Free diffusion. Flip according to the flip % general case + randn; % dv = -v % % assume gamma = 1, 2*ganuma*kB*T = 1 for v =v + dv*dt; 47 simplicity rate % overdamped case without tilting v = randn; % overdamped case with tilting v = randn + AO*sin(w*t); x = x + v*dt; timeseries(i) = x; . else % potential turn on. Force present % general case dv = -v + ratchet-force(x,EO,LO, deltax) + randn; % v =v + dv*dt; % % overdamped case without tilting % v = ratchet-force (x,EO,LO, deltax) randn + % overdamped case with tilting v = ratchet-force (x,EO,LO,deltax) x = x + v*dt; timeseries(i) = x; + randn; + A*sin(w*t ); end t = t + dt; timestep(i) =t if x <= (deltax+LO) && x > (deltax-LO) && where -= 1 where = 1; passage = cat (2, passage , timestep ( i )); elseif x <= (-deltax+LO) && x > (-deltax-LO) && where ~= -1 where = -1; passage = cat (2, passage , timestep ( i )); end end %% Time Series Plot timeseries-runltrunc = timeseries (1:5:end); timestep-runl-trunc = timestep (1:5:end); figure () hold on plot (timestep-runltrunc , timeseries-runl-trunc , 'k') hlinel = refline (0, deltax); set(hlinel, 'Color', 'r','LineWidth',2) hLinell = refline(0, deltax-LO); set (hlinell , 'Color', 'b' , 'LineWidth' ,2) hline12 = refline(0, deltax+LO); set (hline12 , 'Color', 'b' , 'LineWidth' ,2) hline2 = refline (0, -deltax); 48 set (hline2 , 'Color ' , 'r ' , 'LineWidth' ,2) hline2l = refline (0, -deltax+LO); set (hline2l , 'Color ' , 'b' , 'LineWidth' ,2) hline22 = refline(0, -deltax-LO); set (hline22 , 'Color' , 'b' , 'LineWidth' ,2) xlabel ('Time') ylabel ('Particle Position ') hold off %% Particle [ counts , Distribution centers] , figure () hold on bar(centers , counts, 'w') 'r--', plot ([ deltax , deltax ] , ylim plot ([-deltax,-deltax] , ylim , 'r--' xlabel ('Particle Position ') ylabel('Number of Particles hold off % % % % % % 200); = hist (timeseries 'LineWidth', 2) , 'LineWidth',2) figure () semilogy (centers ,counts , 'LineWidth' ,2) hold on xlabel ('Position of Particles ') ylabel ('Number of Particles ') hold off %% Mean First Passage Time (Tau) t-interval = zeros(1,length(passage)-1); for j = 1:length(passage)-1 t-interval(j) = passage(j+1)-passage(j); end [tcounts, tcenters] ,30); = hist(t-interval % histogram of time intervals %MFPT is mean( t-interval). % uncertainty obtained by std (tinterval)/ figure () hold on bar(tcenters , tcounts , 'b') xlabel ('Time Interval ') ylabel ('Frequency') hold off 49 sqrt (length ( t-interval)) A.2 Ratchet Potentials A.2.1 Single Kink %% Force from one-kink ratchet function f = one-kink-ratchet-force(x,EO,LO) if abs(x)-floor(abs(x)) < LO if x > 0.0 f = -EO/LO; else f = EO/LO; end else if x > 0.0 f = EO/(1.0-LO); else f = -EO/(1.0-LO); end end end A.2.2 Double Kink %% Force from the Ratchet Potential function f = ratchet -force (x,EO, LO, deltax) if x>0.0 && x <= deltax if x-floor(x) < LO f = EO/LO; else f = -EO/(1.0-LO); end elseif x <= -deltax if (abs(x)-deltax)-floor (abs(x)-deltax) < LO f = EO/LO; else f = -EO/(1.0-LO); end elseif x<=0.O && x>-deltax if abs(x) - floor(abs(x)) < LO; f = -EO/LO; else f = EO/(1.0-LO); end else if (x-deltax)-floor (x-deltax) < LO f = -EO/LO; else f = EO/(1.0-LO); end 50 end end A.2.3 Periodic Kink Lattice %% Force of periodic kink lattice function f = kink-latticeforce(x,EO,LO,deltax) x = mod(x,2*deltax); if x < deltax if x-floor(x) < LO f = EO/LO; else f = -EO/(1-LO); end else if (x-deltax)-floor (x-deltax) < LO f = -EO/LO; else f = EO/(1-LO); end end 51