Astumian

Embed Size (px)

Citation preview

  • 7/30/2019 Astumian

    1/6

    Thermodynamics and Kinetics of aBrownian Motor

    R. Dean Astumian

    Nonequilibrium fluctuations, whether generated externally or by a chemical reaction far

    from equilibrium, can bias the Brownian motion of a particle in an anisotropic mediumwithout thermal gradients, a net force such as gravity, or a macroscopic electric field.

    Fluctuation-driven transport is one mechanism by which chemical energy can directlydrive the motion of particles and macromolecules and may find application in a wide

    variety of fields, including particle separation and the design of molecular motors andpumps.

    A small particle in a liquid is subject torandom collisions with solvent molecules.The resulting erratic movement, or Brown-ian motion, has been described theoretical-ly by Einstein (1) and independently by

    Langevin (2). Langevin hypothesized thatthe forces on the particle due to the solventcan be split into two components: (i) afluctuating force that changes direction andmagnitude frequently compared to any oth-er time scale of the system and averages tozero over time, and (ii) a viscous drag forcethat always slows the motions induced bythe fluctuation term. These two forces arenot independent: The amplitude of thefluctuating force is governed by the viscos-ity of the solution and by temperature, sothe fluctuation is often termed thermalnoise.

    At equilibrium, the effect of thermal

    noise is symmetric, even in an anisotropicmedium. The second law of thermodynamicsrequires this: Structural features alone, nomatter how cleverly designed, cannot biasBrownian motion (3, 4). To illustrate thispoint, Feynman discussed the possibility ofusing thermal noise in conjunction with an-isotropy to drive a motor in the context of aratchet and pawl device shrunk to micro-scopic size (4). He showed that when allcomponents of such a device are treatedconsistently, net motion is not achieved inan isothermal system, despite the anisotropyof the ratchets teeth. However, a thermal

    gradient in synergy with Brownian motioncan cause directed motion of a ratchet andcan be used to do work. As a practical mat-ter, large thermal gradients are essentiallyimpossible to maintain over small distances.Particularly in biology and chemistry, thethermal gradients necessary to drive signifi-cant motion are not realistic.

    It might seem then that, despite its per-

    vasive nature, Brownian motion cannot beused to any advantage in separating or mov-ing particles, either in natural systems (suchas biological ion pumps and biomolecularmotors) or by artificial devices. Recent work

    has focused, however, on the possibility of anenergy source other than a thermal gradientto power a microscopic motor. If energy issupplied by external fluctuations (58) or anonequilibrium chemical reaction (9, 10),Brownian motion can be biased if the medi-um is anisotropic, even in an isothermalsystem. Thus, directed motion is possiblewithout gravitational force, macroscopicelectric fields, or long-range spatial gradientsof chemicals.

    In devices based on biased Brownianmotion, net transport occurs by a combi-nation of diffusion and deterministic mo-tion induced by externally applied time-

    dependent electric fields. Because parti-

    cles of different sizes experience differentlevels of friction and Brownian motion, an

    appropriately designed external modula-

    tion can be exploited to cause particles ofslightly different sizes to move in oppositedirections (1116). One can imagine anapparatus where a mixture is fed into thesystem from the middle and purified frac-tions are continuously removed from ei-ther side. Because part of the energy re-quired for transport over energy barriers isprovided by thermal noise and because theexternal forces are exerted on a smalllength scale, such devices may be able tooperate with small external voltages. Incontrast, many conventional techniquessuch as electrophoresis, centrifugation

    and chromatography, must be turned offand on each time a new batch of particlesis added. These methods rely on motioncaused by long-range gradients, where themajor influence of thermal noise is todegrade the quality of separation by diffu-sive broadening of the bands.

    Biased Brownian Motion

    As a specific example of biased Brownianmotion (Fig. 1A), picture a charged parti-cle moving over an array of interdigitatedelectrodes with a spatial period of 10 m.When a voltage is applied, the potential

    energy of the particle is approximately ananisotropic sawtooth function (Fig. 1C)Although the electric generator is certain-

    The author is in the Departments of Surgery and of Bio-chemistry and Molecular Biology, University of Chicago,MC6035, Chicago, IL 60637, USA. E-mail: [email protected]

    ( 1)L0

    LLL

    U(x)

    x

    C

    A

    B

    -

    V

    x

    L ~ 10-8m

    + ++

    U

    HS H+ + S

    +10-5 m

    -

    Fig. 1.Two specific models by which an anisotro-

    picperiodicpotential canarise. (A) When a voltage

    difference Vis applied to interdigitated electrodes

    deposited on a glass slide with anisotropically po-

    sitioned teeth, the resulting potential is spatially

    periodic but anisotropic. With photolithography it

    is possible to achieve a spatial period L for such a

    structure as small as 105 m or even somewhat

    less. (B) A linear array of dipoles aligned head to

    tail on which a charged Brownian particle (possi-

    bly a protein) moves. The individual dipoles couldbe macromolecular monomers that aggregate to

    form an extended linear polymer. The length of

    the individual monomer is L 108 m, a reason-able value for many aggregating proteins. If the

    particle catalyzes the reaction HS 3H S,the charge on the particle, and hence its electro-

    static potential energy of interaction with the di-

    pole array U(x, t), fluctuates depending on its

    chemical state. (C) The potential U(x) for both (A) and (B) is approximately an anisotropic sawtooth

    function Usaw, drawn for simplicity as a piece-wise linear function. The amplitude is U, and the wells(potential minima) are spaced periodically at positions iL. The anisotropy is parameterized by . Theamplitude Uof the sawtooth potential can be modulated in (A) by using an external switching device tocontrol the applied voltage.

    ARTICLE

    www.sciencemag.org SCIENCE VOL. 276 9 MAY 1997 917

  • 7/30/2019 Astumian

    2/6

    ly a macroscopic device, the electric fieldin the x direction averaged over a spatialperiod is zero no matter what the voltage,and so there is no net macroscopic force.A nonequilibrium fluctuation can be pro-duced by using a switching device thatimposes an externally defined but possiblyrandom modulation of the voltage. Recentexperiments have shown that unidirec-

    tional motion of microscopic particles canbe induced by modulating the amplitudeof such an anisotropic sawtooth potential(17, 18). Theory shows that the directionof flow is governed by a combination ofthe local spatial anisotropy of the appliedpotential, the diffusion coefficient of theparticle, and the specific details of howthe external modulation is carried out(1116). Under some conditions, particlesof slightly different sizes can move in op-posite directions, possibly providing thebasis for a continuous separation process.Several technological hurdles remain,

    however, before a practical device can beconstructed.In a second example (Fig. 1B), a

    Brownian particle moves along a lattice ofelectric dipoles arranged head to tail (7).The individual dipoles could be, for exam-ple, macromolecular monomers that ag-gregate to form an extended linear poly-mer. Imagine that the particle has an ac-tive site to catalyze a hydrolysis reactionHS 3 H S, where S denotes somesubstrate molecule. The charge on an in-

    dividual particle will then fluctuate, de-

    pending on whether H or S is bound.The potential energy of the particle alongthe axis of the dipole lattice is approxi-mately described by a sawtooth function(Fig. 1C), but the amplitude depends onthe charge and hence on the chemicalstate. When the chemical reaction HS 3H S is far from equilibrium, thefluctuation of the charge on the particledue to the reaction can cause unidirec-tional transport. This effect may be impor-tant in the chemomechanical energy con-version of biomolecular motors and

    pumps, proteins that convert energy froma chemical reaction such as adenosinetriphosphate (ATP) hydrolysis to drive

    unidirectional transport. Although it isunlikely that any biological motor oper-ates exactly as a simple Brownian ratchetusing chemical energy to bias thermalnoise, some of the basic principles of op-eration may be the same.

    Recent experiments have shown that,consistent with a ratchet mechanism (5,19, 20), external oscillating (21) or fluctu-

    ating (22) electric fields can drive transportby a molecular ion pump, the sodium-po-tassium adenosine triphosphatase. Energyfrom the field substitutes for the energynormally provided by ATP hydrolysis eventhough the average value of the field is zero.Fluctuation-driven transport provides anexplicit mechanism for coupling energyfrom a nonequilibrium chemical reaction tocause local fluctuations similar to those im-posed externally to drive unidirectionaltransport (20). Ultimately, it may be possi-ble to construct microscopic motors andpumps that use nonequilibrium chemical

    reactions as their fuel.The examples in Fig. 1 are provided tomake our discussion more concrete. Thegeneral idea of diffusion on a periodic po-tential arises in many other contexts (23),including enzyme catalysis, where an en-zyme cycles through several intermediatestates in carrying out its function (19); thesolid-state physics of transistors (24), wherean electron moves through several p-n junc-tions; models of Josephson junctions (25);and generalized models of computation(26). The details of the interactions givingrise to the potential are not as critical as themechanisms by which spatial anisotropy

    conspires with thermal noise to allow ener-

    gy from external fluctuations or a nonequi-librium chemical reaction to drive unidirec-tional flow. Anisotropic potentials such asthat shown in Fig. 1C are known as ratchetpotentials, in analogy to macroscopic ratch-ets such as a car jack that allow motion inonly one direction by a series of asymmetricgears.

    A Fluctuating Potential, orFlashing Ratchet

    To see how thermal noise combined with

    an externally modulated anisotropic poten-

    tial can lead to unidirectional transport,consider the model (7, 27) shown in Fig.2A. A particle subject to viscous dampingmoves along a track where an anisotropicpotential with periodically spaced wells atpositions iL cyclically turns, or flashes, onand off. To show that work can be doneagainst an external force, the potential pro-files are illustrated with a net force Fextacting from right to left.

    If the external force is not too large, theparticle is pinned near the bottom of one of

    0 L(1)L LL

    U

    x

    tF/

    (4Dt)

    A

    B

    C

    5 10 15 20

    0.1

    0.1

    0.2

    0.2

    0.3

    toff (s)

    U

    Uon = Usaw xFext

    Uon = xFext

    V

    Fgrav sin

    Smallparticles

    Largeparticles

    +-

    (m/s)

    Fig. 2. How a fluctuating potential causes uphill

    transport. (A) Schematic representation of a fluc-

    tuating potential energy profile. Whenthe potential

    is on, the potential energy Uon of a particle is a

    sawtooth function Usaw xFext with periodicallyspaced wells at positions iL. The anisotropy of

    Usaw results in two legs of the potential, one of

    length L on which the force is U/(L) Fext,and the other of length(1 )L on which theforceis U/[(1 )L] Fext. When the potential is off,theenergy profile is flat with force Fext everywhere.

    The curve below the two potential profiles shows aGaussian probability function resulting from turn-

    ing the potential off at t 0, with the particlestarting at x 0. The result is resolved into twocomponents: the downhill drift and the diffusive

    spreading of the probability distribution. For inter-

    mediate times, it is more likely for a particle to be

    between L and (1 )L, where it would betrapped in the well at L if the potential were turned

    back on, than between ( 1)L and ( 2)L,where it would be trapped in the well at L if thepotential were turned back on. Thus, turning the

    potential on and off cyclically can cause motion to

    the right despite the net force to the left. (B) Aver-

    age velocity of a spherical particle with radiusrp 2 m (solid line) and rp 5 m (dashed line)

    induced by cyclically turning an anisotropic saw-tooth potential ( 0.1) onand off versusthe timespent in the off state toff (with ton toff). Theperiod was L 10 m with an external force dueto gravity Fext Vp(dp dw)gsin(); Vp is the vol-ume of theparticle, dp 1.05 g/cm

    3 is thedensity

    of the particle, dw is the density of water, g 9.8 m/s2 is the acceleration due to gravity, and 45.

    The coefficient of viscous drag was 6r 4 108 Ns/m, where 1 centipoise is the viscosityof water. The Einstein relation yields D kBT/ 1 10

    13 m2/s for the diffusion coefficient. The

    amplitude Uwas taken to be greater than 2 1016 J (U/L 20 pN) so that the particle has time toreach the bottom of the well while the potential is on. Such a large U is necessary for the approxima-tions inherent in this simple picture given to be fulfilled. An exact calculation using the diffusion equation

    results in qualitatively the same behavior for smaller values of U. (C) Schematic illustration of anapparatus for separating particles with the use of biased Brownian motion combined with gravity.

    SCIENCE VOL. 276 9 MAY 1997 www.sciencemag.org918

  • 7/30/2019 Astumian

    3/6

    the wells (for example, at x 0) when thepotential is on. Without noise, the particlewould move to the left with a velocityFext/ when the potential is off, where isthe coefficient of viscous drag. Thermalnoise changes the situation dramatically.Because of Brownian motion, a randomwalk is superimposed on the deterministicdrift, and the position of the particle can be

    determined only in terms of a probabilitydistribution. While the potential is off, theprobability distribution drifts downhill witha velocity Fext/ and spreads out like aGaussian function. After a time toff, thedistribution (Fig. 2A) is (28)

    P 0x;toff

    exp x toffFext /

    2

    4Dtoff

    4Dtoff(1)

    where D is the diffusion coefficient. Whenthe potential is turned on again, the particleis trapped in the well at iL if it is between

    L(i 1 ) and L(i ). The probabilityof finding the particle in each well, PiL

    , iscalculated by integrating the probabilitydensity (Eq. 1) between these limits. Be-cause of the anisotropy of the potential, aparticle starting at iL is more likely to betrapped in the well at (i 1)L than in thewell at (i 1)L for small enough Fext andintermediate values of toff. The averagenumber of steps R in a cycle of turning thepotential on for a time ton long enough thatthe particle reaches the bottom of a well(ton L

    2/U), off for a time toff, and thenback on again, is R

    i

    iPiL

    , and theaverage velocity is v RL/(toff ton) (Fig.

    2B). Motion to the right occurs despitethe fact that the macroscopic force pushesthe particle to the left. For large U/L, theexternal force necessary to cause the flowto be identically zero, Fstop (the stoppingforce), can be evaluated exactly (21) to beFstop (L/toff)( 1/2). For toff (L/F)( 1/2), only the terms due tointegration of Eq. 1 with i 1 contrib-ute to the velocity. In this regime, veloc-ities scale with D/L and times with L2/D.

    The mechanism illustrated in Fig. 2works not despite diffusion and thermalnoise but because of diffusion. When the

    potential is turned off, the particle diffusessymmetrically. Then, when the anisotro-pic potential is turned on, it is more likelyfor the particle to feel a force to the rightthan to the left. Thus, net motion to theright occurs, even in the presence of asmall homogeneous force to the left. Theenergy to drive the transport, however,does not come from thermal noise but isprovided when the potential is turned on.This energy can be calculated by integrat-ing over space the product of Uon(x) andthe probability density P(0x; toff). Be-

    cause the potential remains on for longenough to allow the particle to reach thebottom of a well, no energy is returnedwhen the potential is turned off. The out-put energy per cycle is the number of stepsR multiplied by the energy gain per stepFextL. The difference in these energies isdissipated as heat resulting from viscousdrag of the particle with the solvent. For

    the simple model shown in Fig. 2, thethermodynamic efficiency is less than 5%(29). Other features of biased Brownianmotion make it more promising for theconstruction of microscopic motors. Witha spatial period of 10 m, velocities ofaround 0.3 m/s can be achieved for a2-m particle with toff 2 s in aqueoussolution (Fig. 2B, solid line). This motorcan do work against a force as large as 0.05pN, easily overcoming gravity, which for a2-m particle with a density of 1.05 g/cm3

    in water is 0.004 pN. Larger particles,however, cannot overcome gravity and

    drift to the left when toff is large enough(Fig. 2B, dashed line).The velocity depends nonmonotoni-

    cally on the frequency of modulation (Fig.2B). This can be understood in terms ofthe competing requirements that a parti-cle be able to diffuse at least the shortdistance L but not the longer distance(1 )L while the potential is off. Thenonmonotonic behavior allows for thepossibility of tuning a device to a frequen-cy characteristic to a particular particle.This tuning may be useful in applicationsfor separating particles on the basis of size,but the bandwidth of the response is rela-

    tively broad. One can exploit this depen-

    dence, however, by imposing an externalforce so that some particles will move tothe right and others to the left. The sim-plest example (Fig. 2C) is to tilt an appa-ratus such as that shown in Fig. 1A so thatthe particles experience a net gravitation-al force Fgravsin(), where is the anglerelative to a horizontal surface. Small par-ticles will feel only a small force due togravity and a larger amount of Brownianmotion, which can be biased by the ratch-et to cause motion uphill to the right. Thelarger particles have less Brownian motion

    and feel a greater force due to gravity andso move downhill to the left. In such anapparatus, a mixture of particles of differ-ent sizes can be introduced in the middle,and purified fractions can be collectedcontinuously on either end. Because of theclose spacing of the electrodes, only asmall voltage is necessary to achieve suf-ficiently large energy barriers.

    The approach shown in Fig. 2C is notappropriate for smaller particles becausegravity is not sufficiently strong. However,other similar strategies can be adopted, in-

    cluding the use of a constant electric field asthe homogeneous force (16). It has evenbeen shown that with a more complicatedpulse protocol than a simple square wave,particles of different sizes can be induced tomove in opposite directions without anyhomogeneous long-range force at all (12,13, 15). By exploiting the different scalingfor biased Brownian motion and motion

    induced by deterministic gravitational andelectric forces, approaches for continuousseparation of particles ranging from macro-molecular (108 m) to mesoscopic (105

    m) size are possible.The basic principles of motion induced

    by a fluctuating potential have recentlybeen tested in simple model systems. Rous-selet et al. (17) applied a time-dependentratchet potential to colloidal particles usinga series of interdigitated Christmas treeelectrodes deposited on a glass slide byphotolithography. Cyclically turning thepotential on and off resulted in net flow

    Despite the complexity of dealing with asuspension of many interacting particlesthe data agreed at least semiquantitativelywith the predictions of the simple theory.A similar effect was demonstrated by Fau-cheux et al. (18), who used a single col-loidal particle in an optical trap modulat-ed to generate a sawtooth potential. Thissetup has the advantage that there is onlyone particle, thus eliminating interparticlehydrodynamic interactions.

    Modulating the potential certainly re-quires work, and there is no question ofthese devices being perpetual motion ma-chines. The surprising aspect is that f low is

    induced without a macroscopic forceallof the forces involved are local and act on alength scale of the order of a single period ofthe potentialyet the motion persists in-definitely, for many periods. In contrastconventional techniques for particle sepa-ration such as centrifugation, chromatogra-phy, or electrophoresis all rely on macro-scopic gradients and can only operate asbatch separators.

    A Fluctuating Force, orRocking Ratchet

    In the mechanism shown in Fig. 2, thespatial average of the force is independentof time, and the temporal modulation onlychanges the shape of the potential locallyAnother well-studied approach to drivingflow on a ratchet potential involves apply-ing a fluctuating net force (6, 11, 14, 23,30, 31). This approach can be visualized asrocking the sawtooth potential shown inFig. 1C back and forth between the limitsFmax (Fig. 3). IfU/[(1 )L] FmaxU/(L), the potential energy decreasesmonotonically to the left when the force is

    ARTICLE

    www.sciencemag.org SCIENCE VOL. 276 9 MAY 1997 919

  • 7/30/2019 Astumian

    4/6

    Fmax, but when the force is Fmax, thereremain minima that trap a particle as itmoves to the right in response to the ap-plied force (Fig. 3A). Thus, even withoutthermal noise, a slow oscillation of the forcebetween Fmax causes net flow to the left.Fast oscillation does not result in net flowbecause the particle does not have time tomove a period L before the force reverses

    sign. Applying a fluctuating or oscillatingforce is analogous to the standard way ofdriving an electrical rectifier or a macro-scopic mechanical ratchet such as a car jackor ratchetting screwdriver.

    The presence of thermal noise allows asubthreshold fluctuating force {Fmax U/[(1 )L]} to cause flow (6). For aslow, zero-average square wave modula-tion of the force between Fmax, the av-erage rate can be calculated using theanalytic formula given by Magnasco (6).In Fig. 3B, the average velocity is plottedas a function of Fmax where the dashed

    lines indicate the location of the thresholdforces U/[(1 )L] and U/(L). InFig. 3C, the average velocity is plotted asa function of the thermal noise strengthkBT(where kB is Boltzmanns constant andT is temperature) for a subthreshold ap-plied force (Fmax 0.4 pN). We see that,to a point, increasing the noise can actu-ally increase the flow induced by the fluc-tuating force, which suggests that in sometechnological applications, it may be use-ful to electronically add noise to the sys-tem. For forces near the optimum (Fmax

    1.5 pN in Fig. 3B), however, the velocitydecreases almost monotonically with in-creasing noise.

    A More General Approach

    A more general approach to investigatingfluctuation-driven transport involves solv-ing the diffusion equation

    P x,t

    t

    x

    U

    x P x,t D

    2P x,t

    x2

    (2)

    At equilibrium, the probability density isgiven by a Boltzmann distribution P(x) exp(U/kBT). Any external temporal mod-ulation ofU, whether accomplished period-ically or randomly in time, requires energyinput and drives the system away from equi-librium, disturbing the Boltzmann distribu-tion. The time dependence ofU for a fluc-tuating potential ratchet is U(x, t) f(t)

    Usaw(x), and for a fluctuating force ratchet,is U(x, t) Usaw(x) xf(t). Equation 2 canbe solved for P(x, t) for any explicit timedependence of f(t) with appropriate (typi-cally periodic) boundary conditions, andthe result can be used to calculate the av-erage velocity. It turns out, however, that inmany cases, it is simpler to solve Eq. 2 when

    f(t) is modeled as a position-independentstochastic variable (31). This situation mustnot be confused with an equilibrium fluctu-ation. At equilibrium, the probability toundergo a transition from one state to an-

    other depends on the energy difference be-tween the states according to a Boltzmannrelation (also known as detailed balance),and the energy difference for a ratchet po-tential clearly depends on the position. Anexternal modulation, whether random ornot, does not conform to detailed balanceand cannot model equilibrium fluctuations.

    Unlike motion driven by forces such as

    gravity or macroscopic electric fields, wherethe direction of motion is determined bythe sign of the force, the direction of flowinduced by zero-average modulation of apotential or force depends on the details ofhow the modulation is carried out. Recent-ly, several schemes have been proposed thatallow for flow reversal to occur as a functionof the modulation frequency, with differentreversal frequencies for particles with differ-ent diffusion coefficients (1318).

    Chemically Driven Transport

    Consider the one-

    dimensional motion of acharged particle along a periodic lattice ofdipoles arranged head to tail (Fig. 1B), andlet the interaction between the particle andlattice be purely electrostatic (7). The po-tential energy profile then consists of aperiodic series of minima (wells) and max-ima (barriers). For simplicity, this profilecan be represented as a sawtooth function(Fig. 1C). Because of thermal noise, theparticle undergoes Brownian motion. Occa-sionally, it will have enough energy to passover one of the two barriers surrounding thewell in which it starts and move to the wellon the right or on the left. Despite the

    anisotropy, without an energy supply, thetwo probabilities are exactly equal.

    Now, imagine that this particle catalyzesa chemical reaction HS ^ H S. Be-cause the product molecules H and S arecharged, the amplitude of the interactionpotential between the particle and the di-pole lattice depends on whether H or S

    are bound to the particle, resulting in acoupling between the chemical reactionand the diffusion of the particle along thedipole lattice. Because of the charge on H

    and S , their local concentrations vary asa function of position along the axis of the

    dipole lattice (the hydrogen ion concentra-

    tion [H] is larger near the negative end ofthe dipole than near the positive end, andthe opposite is true for [S]). The radii ofthe ions H and S are very small com-pared to the radius of the particle, so theirlocal concentrations equilibrate quickly,leading to [S] [S]bulkexp[U(x)/kBT]and [H] [H]bulkexp[U(x)/kBT],where the subscript bulk denotes the con-centration far from the surface of the dipolelattice. For the specific mechanism shownin Fig. 4A, k21, k23, and k31 are first-order

    0.2

    0.02

    0.01

    0.1

    1.0 2.0 3.00

    4 8 16 20 2400

    0

    U= UsawxIFmaxI

    U= Usaw + xIFmaxI

    U= Usaw

    A

    B

    C

    Fmax(pN)

    kBT(10-21 J)

    (m/s)

    (m/s)

    Fig. 3. The basic mechanism by which a fluctuating oroscillating force cancausedirectedmotion along a ratch-

    et potential Usaw. The parameters U 4 1018 J, L

    10 m, and 0.2 were used. (A) Because of theanisotropy of the potential Usaw, starting at the bottom of

    any well, the force required for the particle to move to the

    right, U/(L) 2 pN, is greater than the force neces-sary to move to the left, U/[(1 )L] 0.5 pN. Whenan external force between these values (such as 1.5pN)is applied, thenet potential U(x) U

    sawx(1.5pN)

    decreases monotonically to the left, and a particle acted

    on by this potential drifts to the left with no impediment.

    When an external force of 1.5 pN is applied, the netpotential U(x) U

    saw x(1.5 pN) decreases on aver-

    age to the right, but there remain wells (local potential

    minima, indicated by the vertical arrows) that tend to trap

    theparticle. Thus a square wave modulation of theexter-nal force Fmax

    between 1.5 pN and 1.5 pN results innet flow to the left. (B) For a low-frequency square wave

    modulation, the average velocity can be calculated as a

    function of Fmax

    with the analytic formula given by Mag-

    nasco (6). The average velocity v depends nonmono-tonically on the amplitude F

    maxand is maximum between

    Fmax

    U/[(1 )L] 0.5 pN and Fmax

    U/(L) 2pN (indicated by the dashed lines). Because of thermal

    noise, a subthreshold force Fmax

    U/[(1 )L] caninduce some flow. (C) The dependence of the flow on the

    thermal noise amplitude kBT induced by a 0.4-pN force.

    The curve is nonmonotonic, reminiscent of a phenome-

    non known as stochastic resonance (32, 33).

    SCIENCE VOL. 276 9 MAY 1997 www.sciencemag.org920

  • 7/30/2019 Astumian

    5/6

    rate constants and do not depend on theposition along the dipole, and k12 k12[SH], k13 k13[H

    ], and k32 k32[S]

    are pseudofirst-order rate coefficients intowhich the concentrations of the reactantshave been incorporated. Because [H] and[S] depend on the position x along thedipole lattice, the rate coefficients k13 andk32 also depend on position.

    If the chemical reaction is at equilibri-

    um, a direct transition from the unchargedstate i 3 to the charged state i 2 is morelikely near the positive end of the dipole,where [S] is relatively high, than near thenegative end of the dipole, where [S] isrelatively low. Similarly, direct transitionfrom the charged state i 1 to the un-charged state i 3 is more likely near thenegative end of the dipole, where [H] isrelatively high, than near the positive endof the dipole, where [H] is relatively low.The net effect is that transitions from thecharged to the uncharged state are more

    likely near the negative end of the dipole,and transitions from the uncharged tocharged state are more likely near the pos-itive end of the dipole. In this case (and inthe absence of an external force), the prob-ability density of the particle is distributedeverywhere according to a Boltzmann dis-tribution (Fig. 4B, dashed line), and theaverage velocity of the particle is zero.

    If the chemical reaction is away fromequilibrium, this relation between the posi-tion of the particle and transition probabil-ities does not necessarily hold (Fig. 4B, solidline). In the extreme case where [H] [S] 0, direct transitions from state 1 to

    state 3 and from state 3 to state 2 cannotoccur at all. None of the rate coefficientsfor the remaining transitions depend on theposition of the particle, and so the proba-bility of a transition from the charged to theuncharged state is independent of the posi-tion along the axis of the dipole lattice, as isthe probability for a transition from theuncharged to the charged state. This situa-tion is essentially identical to that describedas a fluctuating potential ratchet (Fig. 2),and as before, directed motion can result.

    Thus, we have a rudimentary motordriven by a chemical reaction. The mech-

    anism shown in Fig. 4A is a tremendousoversimplification as a description of anyactual molecular motor. The interaction be-tween the motor and dipole lattice and theeffect of the catalyzed reaction were as-sumed to be purely electrostatic, and theresulting periodic potential was grossly sim-plified. Nevertheless, with reasonable val-ues of the rate constants and other param-eters, the motor moves pretty well (Fig.4C), with a maximal velocity of severalmicrometers per second. The stoichiometryis poor: It takes more than 3 HS molecules

    on average to cause a single step (displace-ment by a period L), and less than 1 pN ofapplied force stops the motion. What thissimplified motor lacks is what a real biomo-lecular motor can certainly provide: morecomplex interaction between the motorand track, provided by conformational flex-ibility, and better regulated coupling be-tween chemical and mechanical events,

    possibly provided by allosteric interactionbetween the motor and track.

    Perspective and Outlook

    Noise is unavoidable for any system in ther-mal contact with its surroundings. In tech-nological devices, it is typical to incorporatemechanisms for reducing noise to an abso-lute minimum. An alternative approach isemerging, however, in which attempts aremade to harness noise for useful purposes.

    One example involves signal detection bythreshold detectors (32). Typically, suchdevices have been designed so that thesystem is isolated from noise, and thethreshold is as small as possible. Thesegoals, however, are often difficult and ex-pensive to attain. It may instead be possibleto operate the system with a threshold larg-er than the signal to be measured and to use

    added noise to provide a boost that willallow the signal to be detected above thenoise. The phenomenon of noise-enhancedsignal-to-noise ratio is known as stochasticresonance and holds for a wide variety ofnonlinear systems (32, 33). Similarly, fluc-tuation-driven transport uses noise to ac-complish mass transport without macro-scopic forces or gradients. Three ingredientsare essential: (i) thermal noise to causeBrownian motion; (ii) anisotropy arisingfrom the structure of the medium in which

    L/2 L/2

    0.5

    0.6

    0.7

    L L

    G/RT1 1 2 3 4

    1

    2

    3

    0

    -

    -

    -

    -

    SH

    i= 1

    i= 2

    i= 3

    i= 1

    k13k31

    k32k23

    k21k12

    Position

    Chemicalstate

    P3(x)/Pi(x)

    A

    B

    C

    S

    H+

    H+

    SH

    (m/s)

    Fig. 4.Transport driven by a chemical reaction. (A)

    Schematic illustration of how the potential energyprofile of a Brownian particle that catalyzes a reac-

    tion SH ^ S H changes depending on itschemical state, even in a constant electric potential

    due toa dipolearrayas shown inFig. 1B. (B) Prob-

    ability of the particle being uncharged P3(x)/Pi(x)as a function of position when the chemical reac-

    tion SH^S H is at equilibrium (dashed curve)and far fromequilibrium (solid curve). For simplicity,

    we treat the case thatthe particle bears a charge of1 and that the interaction potential between theparticle and dipole track (Fig. 1B) is purely electro-

    static. Because S and H are charged, their local

    concentrations depend on the local potential ac-

    cording to [S] [S]bulkexp[U(x)/kBT] and [H]

    [H]bulkexp[U(x)/kBT]. The rate constants used

    were k12 10[SH], k21 106, k23 106, k32 10[S], k13 100[H

    ], and k31 100, all in inverse

    milliseconds, and the concentrations were [SH] 4,

    [H]bulk 2, and [S]bulk 2 for the equilibrium

    case, and [SH] 20, [H]bulk 0.05, and [S]bulk

    0.05for the far fromequilibrium case, all in millimolar.

    The amplitude of the potential U was 1020 J. (C)

    Average velocity as a function of the G of the chem-

    ical reaction using the rateconstants and field ampli-

    tude as in (B). This result can be calculated in general

    by using reaction-diffusion equations to describe the

    combined physical motion and chemical reaction of

    a catalytic Brownian particle as described elsewhere

    (9, 10). Here, k12 and k23 were taken to be large so

    that state 2 could be treated as a steady-state inter-

    mediate. Then, there are effectively only two states,

    off and on, as in Fig. 2, with the effective timeconstants off {5[S

    ]bulkexp[U(x)/kBT] 100}1

    and on {5[SH] 100[H]bulkexp[U(x)/kBT]}

    1,

    as described in (9). The value ofG RTln{[SH]/

    ([S]bulk[H]bulk)} (R is the universal gas constant)

    was varied by varying [SH] with constant [S]bulk

    [H]bulk 1. The period L was taken to be 108 m,

    and the viscosity at the interface was taken to be

    greater than in bulk water (int 100 cP), resulting in 6r 2 108 Ns/m for a particle with a

    Stokes radius of r 108 m and a diffusion coeffi-

    cient D 2 1013 m2/s. The inset shows that there is a linear relation between flow and G close to

    equilibrium. The nonmonotonic dependence on G arises because G is changed by varying the chemica

    concentrations, which also varies the time scales.

    ARTICLE

    www.sciencemag.org SCIENCE VOL. 276 9 MAY 1997 921

  • 7/30/2019 Astumian

    6/6

    the particle diffuses; and (iii) energy sup-plied either by an external variation of theconstraints on the system or by a chemicalreaction far from equilibrium.

    It is often overlooked that noise-assist-ed processes incorporate features analo-gous to chemical kinetics (34). Transi-tions along a chemical pathway are de-scribed in terms of rate constants that

    reflect the probability that thermal noisewill provide sufficient energy to surmountan energy barrier separating chemicalstates. Experiments on ion pumps haveshown that energy from externally im-posed electric oscillations (21) and fluctu-ations (22) can substitute for energy fromhydrolysis of ATP to power uphill trans-port of ions. The likely mechanism in-volves biasing thermally activated steps inthe reaction pathwaya Brownian ratch-et mechanism (5, 19). As Lauger summa-rized, Ion pumps do not function by apower-stroke mechanism; instead, pump

    operation involves transitions betweenmolecular states, each one of which is veryclose to thermal equilibrium with respectto its internal degrees of freedom, even atlarge overall driving force (35).

    It is not yet clear whether molecular mo-tors such as muscle (myosin) or kinesinmove by using an ATP-driven powerstrokea viscoelastic relaxation process inwhich the protein starts from a nonequilib-rium conformation after phosphate releaseand which does not require thermal activa-tionor whether energy from hydrolysis ofATP is used to bias thermally activatedsteps. The recent work on Brownian ratchets

    shows that a thermally activated mechanismis not inconsistent with a process in which aprotein moves a distance of 10 nm in a singlestep and generates a force of several pico-newtons. This question may be resolved soonwith the use of recently developed tech-niques for studying molecular motors at thelevel of a single molecule (36).

    Incorporation of thermal motion as anessential element in the design of micro-scopic machines is essentially the equiva-lent of adopting design principles bor-rowed from chemistry rather than fromclassical macroscopic physics. A remark-

    able feature is that the fundamentally sto-

    chastic nature of noise-assisted activated

    transitions can give rise to mechanismsthat rival purely deterministic processes inthe predictability of specific outcomes.This possibility has been discussed by Ben-net in a thermodynamic comparison ofcomputers based on Brownian versus bal-listic principles (37). Recent work showsthat coupling many particles in a modu-lated ratchet potential gives rise to ex-

    traordinarily rich behavior and thermody-

    namic efficiencies up to 50% (38).The use of noise in technological appli-

    cations is still in its infancy, and it is farfrom clear what the future holds. Neverthe-less, there is already much active research inthe area of noise-enhanced magnetic sens-ing (39) and electromagnetic communica-tion (40). The recent work on fluctuation-driven transport leads to optimism that sim-ilar principles can be used to design micro-scopic pumps and motorsmachines thathave typically relied on deterministicmechanisms involving springs, cogs, and

    leversfrom stochastic elements modeledon the principles of chemical reactions andnoise-assisted processes. Such devices wouldbe consistent with the behavior of mole-cules, including enzymes, and could pavethe way to construction of true molecularmotors and pumps.

    REFERENCES AND NOTES___________________________

    1. A. Einstein, Ann. Phys. 19, 371 (1906).2. P. Langevin, Comptes Rendues 146, 530 (1908).3. M. v. Smoluchowski, Phys. Z. 13, 1069 (1912).4. R. P. Feynman, R. B. Leighton, M. Sands, The Feyn-

    man Lectures on Physics (Addison-Wesley, Read-ing, MA, 1966), vol. 1, chap. 46.

    5. R. D. Astumian, P. B. Chock, T. Y. Tsong, Y.-D.

    Chen, H. V. Westerhoff, Proc.Natl. Acad.Sci. U.S.A.84, 434 (1987).

    6. M. Magnasco, Phys. Rev. Lett. 71, 1477 (1993).7. R. D. Astumian and M. Bier, ibid. 72, 1766 (1994).8. J. Prost, J. F. Chawin, L. Peliti, A. Ajdari, ibid., p.

    2652; C. S. Peskin, B. Ermentrout, G. Oster, in CellMechanics and Cellular Engineering, V. C. Mow et

    al., Eds. (Springer-Verlag, New York, 1994).9. R. D. Astumian and M. Bier, Biophys. J. 70, 689

    (1996).10. H.-X.Zhou andY.-D.Chen, Phys.Rev. Lett. 77, 194

    (1996).11. C. R. Doering, W. Horsthemke, J. Riordan, ibid. 72,

    2984 (1994).12. J. F. Chawin, A. Ajdari, J. Prost, Europhys. Lett. 32,

    373 (1995).13. M. Bier and R. D. Astumian, Phys. Rev. Lett. 76,

    4277 (1996).14. M. Bier, Phys. Lett. A 72, 12 (1996).

    15. A. Mielke,Ann. Phys. (Leipzig) 4, 476(1995);ibid., p.721.

    16. M. Tarlie and R. D. Astumian, in preparation.

    17. J. Rousselet, L. Salome, A. Ajdari, J. Prost, Nature

    370, 446 (1994); see also S. Leibler, ibid., p. 412.

    18. L. P. Faucheux, L. S. Bourdieu, P. D. Kaplan, A. J.

    Libchaber, Phys. Rev. Lett. 74, 1504 (1995).

    19. R. D. Astumian, P. B. Chock, T. Y. Tsong, H. V.

    Westerhoff, Phys. Rev. A 39, 6416 (1989); R. D.

    Astumian and B. Robertson, J. Chem. Phys. 91,

    4891 (1990).

    20. T. Y. Tsong and R. D. Astumian, Bioelectrochem.Bioenerg. 15, 457 (1986); A. Fulinski, Phys. Lett. A

    193, 267 (1994).21. D. S.Liu, R. D.Astumian, T.Y. Tsong,J. Biol. Chem.

    265, 7260 (1990).

    22. T. D. Xie, P. Marszalek, Y. D. Chen, T. Y. Tsong,Biophys. J. 67, 1247(1994);T. D.Xie, Y.D. Chen, P.Marszalek, T. Y. Tsong, Biophys. J., in press.

    23. H. Risken, The Fokker-Planck Equation (Springer-Verlag, Berlin, 1989), chap. 11; see also P. Hanggi,P. Talkner, M. Borkovec, Rev. Mod. Phys. 62, 251(1990).

    24. M. Buttiker and R. Landauer, in Nonlinear Phenom-ena at Phase Transitions and Instabilities, T. Riste,Ed. (Plenum, New York, 1982), p. 111.

    25. M. Buttiker, E. P. Harris, R. Landauer, Phys. Rev. B28, 1268 (1983).

    26. R. Landauer and M. Buttiker, Phys. Scr. T9, 155(1985).

    27. A. Ajdari and J. Prost, C. R. Acad. Sci. Paris 315,1635 (1992).

    28. H. C. Berg, Random Walks in Biology (PrincetonUniv. Press, Princeton, NJ, 1983).

    29. M. Bier and R. D. Astumian, Bioelectrochem. Bioen-erg. 39, 67 (1996).

    30. M. Millonas and M. M. Dyckman, Phys. Lett. A 185,65 (1994); R. Bartussek, P. Hanggi, J. G. Kissner,Europhys. Lett. 28, 459 (1994); R. Bartussek, P.Reimann, P. Hanggi, Phys. Rev. Lett. 76, 1166(1996).

    31. C. R. Doering, Nuovo Cimento D 17, 685 (1995).

    32. K. Wiesenfeld and F. Moss, Nature 373, 33 (1995).

    33. A. R. Bulsara and L. Gammaitoni, Phys. Today49,39(March 1996);R. D.Astumian, J.C. Weaver,R. K.

    Adair, Proc. Natl. Acad. Sci. 92, 3740 (1995).

    34. R. D. Astumian and B. Robertson, J. Am. Chem.Soc. 115, 11063 (1993); R. D. Astumian, J. Phys.Chem. 100, 19075 (1996).

    35. P. Lauger, Electrogenic Ion Pumps (Sinauer, Sun-

    derland, MA, 1991), p. 56.36. J. Finer, R. M. Simmons, J. A. Spudich, Nature 368,113 (1994); K. Svoboda, C. F. Schmidt, B. J.Schnapp, S. M. Block, ibid. 365, 721 (1994).

    37. C. H. Bennet, Int. J. Theor. Phys. 21, 905 (1982).

    38. I. Derenyi and T. Vicsek, Phys. Rev. Lett. 75, 374(1995); F. Julicher and J. Prost, ibid., p. 2618; I.Derenyi and A. Ajdari, Phys. Rev. E54, R5 (1996).

    39. A. Hibbs et al., J. Appl. Phys. 77, 2582 (1995); R.Rouse,S. Han, J. Luckens,Appl. Phys. Lett.66, 108(1995).

    40. V. S. Anishchenko, M. A. Safanova, L. O. Chua, Int.J. Bifurcation Chaos 4, 441 (1994).

    41. I thank B. Adair, S. Berry, C. Bagdassarian, M. Bier,I. Derenyi, G. Fried, B. Glick, D. Grier, M. Hoch-strasser, S. Kron, M. Makinen, J. Ross, T. Steck, M.

    Tarlie, and T. Tsong for useful and enjoyable discus-sions. This work was supported in part by NIH grantR29ES06620 and the Materials Research Science

    and Engineering Center (MRSEC) program of theNSF (award DMR-9400379).

    SCIENCE VOL. 276 9 MAY 1997 www.sciencemag.org922