Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

Embed Size (px)

Citation preview

  • 8/3/2019 Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

    1/10

    preprint

    Targeted Activation in Deterministic and Stochastic Systems

    Bryan Eisenhower and Igor Mezic

    Department of Mechanical Engineering,

    University of California Santa Barbara

    (Dated: March 6, 2009)

    Abstract

    Metastable escape and transitions are ubiquitous in many physical systems and are becoming

    a concern in engineering design as these designs become more and more inspired by dynamics of

    biological, molecular, or other natural systems (eg. swarms of vehicles, coupled building energetics,

    nanoengineering, etc). We study an example of a chain of coupled bistable oscillators which has

    two global conformation states. Specifically we investigate re-conformation between these two

    states and how specialized or targeted disturbance effects the transition process. We derive a

    multi-phase averaged approximation to these dynamics which illustrates the influence of actions in

    modal coordinates on the coarse behavior of this process. An activation condition that characterizes

    temporal rate in relation to initial energy is then derived. This result emphasizes spatial sensitivity

    and the impact of targeted activation on reconformation. We also find an unexpected inverse energy

    cascade through these actions, funneling energy to one coarse coordinate leading to the global

    transition. The prediction tools are derived for deterministic dynamics while we also present

    analogous behavior in the stochastic setting and an ultimate divergence of Kramers activation

    behavior under targeted activation conditions.

    PACS numbers: Valid PACS appear here

    1

  • 8/3/2019 Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

    2/10

    A chain of strongly coupled oscillators, each of which oscillates in a bistable potential, is a

    coarse representation of many physical systems from macromolecules to Josephson junctions

    which have multiple conformations. Switching between these conformations can occur by

    coordinated motion of the entire chain over a global energetic barrier [1]. Characterizing this

    global barrier with respect to local barrier heights is essential if this structure is to be used

    for engineering design. We present such results for a deterministic harmonically coupled

    oscillator chain with a strong backbone.

    The class of nonlinear systems of coupled oscillators that we study are close to a coupled

    chain of linear harmonic oscillators. Such near-integrable systems have been studied in [2]

    where equipartition of energy and dynamic properties related to integrable instability theory

    of partial differential equations were investigated [3].

    The Hamiltonian for an N oscillator system which has a strong linear backbone and local

    nonlinearity takes the form

    H =Ni=1

    1

    2q2i +

    1

    2B(qi1 qi)2 +U(qi) (1)

    where qi R, B is the linear coupling strength andUis a local bistable nonlinear potential.We have studied many such potentials, including a Morse potential acting on pendula (see [1],

    [4], [5]) which is affiliated with conformation change in molecular systems but its exponential

    form makes analytical progress challenging. In this study we investigate a 2-4 potential

    U(qi) =

    1

    2q2i

    1

    4q4i

    (2)

    where is a constant that separates the two nonzero equilibria (qeq = {0,}) to control thecurvature of the saddle to better capture dynamics of proteins, see [6]. This potential has

    the same behavior as the Morse potential while being easier to analyze. The backbone has

    periodic boundary (qN+1 = q1), and is much stronger than the local nonlinearity (B ).

    When this is the case the transition process is collective and coordinated as shown in Figure1 (and as described in [1]).

    In systems where the strength of the coupling is of the same order as the strength of

    the local nonlinearity, neighboring oscillators do not restrain each other from excursions

    across local barriers. These excursions or breathers initiate the transition process of the

    entire chain. In [7] it was shown that the transition process occurs when combination of

    small breathers merge into larger breathers with enough strength to pull the entire oscillator

    2

  • 8/3/2019 Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

    3/10

    5 0 50

    20

    40

    60

    80

    100T = 0

    OscillatorNumber

    2 0 20

    20

    40

    60

    80

    100T = 72.5

    2 0 20

    20

    40

    60

    80

    100T = 109

    2 0 20

    20

    40

    60

    80

    100T = 145.5

    OscillatorNumber

    Position2 0 20

    20

    40

    60

    80

    100T = 182

    Position2 0 20

    20

    40

    60

    80

    100T = 254.5

    Position

    FIG. 1: Snapshots of the evolution of 100 oscillators from one equilibrium to the other in the

    potential (2) ( = 1.0e 3, = 1, B = 1). As an initial condition, two oscillators are perturbedin the potential and because the coupling is much stronger than the nonlinearity the transition is

    collective, closely following the mean (dashed line).

    array over the barrier. In [8] the same authors developed an analytical calculation for the

    transition process considering a single oscillator diffusing through a separatrix layer and

    subsequently pulling the other oscillators along. This theory does not work in our setting

    since the coupling strength is much greater than the local nonlinearity and excursions of a

    few oscillators and formation of breathers is discouraged. We find that it is not excursions

    of individual or groups of oscillators that drives the transition process but rather dynamics

    of the mean position of all oscillators. In fact, we show that in the time-averaged sense,

    the dynamics of the mean position and velocity are sufficient to predict properties of the

    activation process.

    Since the empirical evidence of Arrhenius, noise assisted activation has been studied to a

    great extent and has led to many predictive tools [9]. In this setting, external thermal energy

    lifts the oscillators over the barrier. On the other hand, limited tools exist for the micro-canonical example where energy is conserved and escape is only realized by redistribution of

    initial energy. Such behavior is remarkable as energy is appropriately distributed throughout

    the system leading to the activation process without any external influence from a thermal

    bath. We show that such behavior is caused by an interaction between incommensurate

    linear modes due to the local nonlinearities.

    3

  • 8/3/2019 Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

    4/10

    We perform two canonical coordinate transformations to identify the coarse behavior and

    energy transfer dynamics of the transition process. Because of the translational symmetry

    on the backbone the fundamental normal basis is Fourier modes. We perform the canonical

    transformation into these coordinates using

    qi =

    2

    N

    N21

    k=1

    [q0

    2+ cos (2ik/N)qk (3)

    +cos(k)

    2qN

    2

    + sin (2ik/N)qN2+k]

    where qk are modal amplitudes and k is the wavenumber (time differentiation of (3)

    gives transform rules for velocities qi to modal velocities qk) . The frequencies k =

    2cos(2k/N) 2 are irrationally related and range from 0.0 to 2.0 with linear spacing

    at low frequencies and decreasing spacing as they approach 2.0. The zeroth mode q0 is theaveraged coordinate (scaled by

    N) and the coordinates q

    1...N21

    are Fourier modes.

    Projecting the nonzero modal dynamics onto canonical action-angle variables takes the

    system to a series of nearly integrable oscillators which interact with one highly nonlinear

    mode (we rename q0 0 and q0 J0 for notational convenience)0 = J0

    J0 = g0( J, , ),

    k = k + fk( J, , )

    Jk = gk( J, , )(4)

    k = 1, 2, . . . (N 1)where k are angle variables and Jk are the conjugate actions. To obtain a coarse repre-

    sentation we average over the angles of M nonzero modes (M N as the analysis can beperformed on a reduced order model since the transformations are canonical). The averaged

    Hamiltonian becomes

    H( J) = 1(2)M

    2

    0

    2

    0

    H( J, )d1 . . . d M (5)

    For the specific 2-4 potential the resulting averaged ODE derived from Hamiltons equationsis

    0 = J0 (6)

    J0 =

    3

    N

    Mk=1

    Jkk

    0

    3

    0

    N

    Jk = 0, k = 1, 2, . . . (M 1)

    4

  • 8/3/2019 Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

    5/10

    In an averaged sense, action in the higher modes are stationary in time and feed into

    the zeroth mode dynamics. In this sense, (6) represents a parametric 2-4 potential for the

    planar variables (0,J0) where higher order actions (Jk, k = 1, 2, . . . , M ) alter the existence

    and location of the equilibria branches.

    The rate of activation is dependent on energy in both the microcanonical and thermally

    activated settings. In the deterministic setting, there exist a dependency between the ac-

    tivation rate and the amount of initial energy (and in particular, the spatial character of

    the initial energy), and this dependency can be captured quantitatively using the averaged

    model (6). In the averaged sense, the zeroth mode has the typical figure eight phase space of

    the 2-4 potential with equilibria at 0eq =

    N in modal coordinates. With an initial con-

    dition with the zeroth mode at this location, and no energy in higher modes, no oscillations

    will occur. As soon as energy from a higher mode is introduced the equilibrium of the zeroth

    mode begins to shift, and since the oscillators are no longer on the original equilibrium they

    begin to oscillate (in a state of libration). In fact, the equilibria shift toward the origin

    bringing the separatrix closer to the initial location of the zeroth mode. With sufficient

    energy in any one of these higher order modes, the separatrix of the zeroth mode will be

    driven closer to the origin and eventually the original location of the oscillators will breach

    it. When this occurs the zeroth mode will be forced into the rotation state and transition

    to the second equilibrium will occur. This concept is illustrated in Figure 2.

    4

    2

    0

    2

    4

    0.4

    0.2

    0

    0.2

    0.40

    2

    4

    6

    8

    10

    12

    0

    J0

    J5

    FIG. 2: Impact of energy in higher order modes (eg J1) on the zeroth mode illustrating how the

    original equilibria are moved outside of the libration state to the rotation state.

    With this in mind, the minimum activation energy becomes the amount of energy needed

    5

  • 8/3/2019 Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

    6/10

    from the higher mode(s) to pull the separatrix over the original equilibrium location 0. The

    activation condition is

    U(0 = 0.0, J) =U(0 = 0eq , J) (7)which defines the minimum activation energy for transition. Performing the necessary cal-

    culations, the resulting activation condition for the 2-4 potential becomes

    Mi=1

    Jii

    =N

    6(8)

    which describes an affine hyperplane with a dimension equal to the number of nonzero

    actions. It is evident in Figure 2 that with further increase of action in the higher modes,

    the concept of having two wells disappears, and the zeroth mode will wildly circle both

    original equilibria in a state of rotation (with dissipation the masses will settle in one of the

    wells). Indeed a pitchfork bifurcation exists for these high energy orbits at

    Mi=1

    Jii

    = N3

    which is twice the minimum activation energy.

    To calculate the full activation time vs. activation energy profile we note that the when

    the oscillation of the zeroth mode is in its rotation state (which is needed for activation), the

    activation time is the time in which the trajectory crosses the origin (ie 0 = 0.0). In fact,

    since the phase space for the rotation state is symmetrical this becomes 1/4 of the period of

    a rotation state. We can analytically derive the rate of activation by calculating the period

    of rotation oscillations of the zeroth mode in (6) as it is parametrically altered by higher

    wavelength actions. To calculate the period of rotation we generalize the Hamiltonian of the

    averaged dynamics as

    H = J20

    2

    a202

    +b40

    4

    (9)

    a =

    3Ji

    iN

    , b =

    N

    This system admits an analytical solution that is different inside and outside of the figureeight separatrix, each of which uses Jacobi elliptic functions. On the outside of the separatrix

    we calculate the period as

    Tout =4K()

    out(10)

    where =0(0)

    b

    2(a + b0(0)2)

    , out =

    a

    22 1

    6

  • 8/3/2019 Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

    7/10

    where is the elliptic modulus and K(k) is the complete elliptic integral of the first kind.

    The variable 0(0) is the amplitude of the initial condition of the zeroth mode.

    For comparison, numerical data was generated using N = 100 oscillators, B = 1, =

    0.001, and = 1. The dynamics were integrated in time using a 4th order symplectic method

    [10] using a fixed stepsize of 0.005. The analytical expression captures numerical data well

    when the activation rates exceed the modal frequencies. The discrepancy in lowest mode at

    high activation rates is because the modal frequency (of which we average over) approaches

    the activation rate itself.

    0 100 200 300 400 500

    0.01

    0.02

    0.03

    0.04

    0.05

    0.06

    0.07

    U(0) / U

    Activation

    Rate[1/s]

    Mode 1Mode 2

    Mode 3Mode 4

    Mode 5Mode 6

    Mode 7Mode 8

    Mode 9Mode 10

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.70

    1

    2

    3

    4

    5

    6

    7

    Mode Frequency

    MinimumA

    ctivationEnergy

    Data

    Analytical

    FIG. 3: Activation energy vs rate (top) and minimum activation energy (bottom). Analytical -

    solid, numerical simulation - discrete points. U = N4

    is the sum of the barrier heights and U(0)

    is the initial energy.

    The activation behavior discussed above is clearly impacted by selective modal energies.

    The same dynamics were numerically analyzed in a stochastic setting using the Langevin

    framework to determine if the phenomenon persists. In Figure 4 we illustrate the stochastic

    rates using the same parameters as before with a Langevin damping value of 0 .05. In this case

    we impose an initial energy at a constant value of 5U in different spatial patterns including

    a randomly selected series of positions. The data is simulated for different temperatures

    (Boltzmanns constant k = 1.0) and compared to Kramers estimate for low damping (see

    7

  • 8/3/2019 Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

    8/10

    equation 4.49 of [9]). We find that at higher temperatures, Kramers theory predicts the

    behavior well, while at lower temperatures inertial effects dominate and the numerical data

    diverges from Kramers estimates. In addition, we find that this divergence is related to the

    spatial character of the perturbation.

    1 2 3 4 5 6 7 8

    103

    102

    U/ (kT)

    ActivationRate[1/s]

    0th Mode

    1st Mode

    2nd Mode

    3rd Mode

    Random

    Kramers

    FIG. 4: Stochastic simulation with targeted perturbation in the first 4 Fourier modes compared to

    a random initial condition. These rates are compared to Kramers estimate.

    Above we have been able to predict a static relationship between initial energy and

    activation time. However, the transfer of energy from modes which contain initial energy

    (higher modes) to the zeroth mode is a dynamical process. The equipartition of energythroughout modes in high dimensional systems is expected due to their ergodic nature but

    is not always the case. As shown initially by Fermi, Pasta, and Ulam (FPU) [11] this

    balancing of energy may not occur for even relatively simple dynamics. In fact, Ford later

    illustrated that this lack of energy balance is due to a lack of commensurate temporal

    frequencies in the set of spatial modes [12]. As we have mentioned, the frequencies of the

    linear normal modes in our model are incommensurate. Without any nonlinear perturbation

    to the backbone, we expect there to be no energy transfer at all. However as we have shown,

    there is no doubt that energy transfer occurs because activation can be observed by global

    motions of the zeroth mode and this can be realized by imposing initial energy on modes of

    high wavelength. In fact, the energy does not lead to equipartition immediately, it follows

    a unidirectional route to the lowest mode.

    By calculating the Jacobian of Hamiltons equations of (4) we obtain a matrix that reveals

    the change in action variables as influenced by other actions and angles. It turns out that

    8

  • 8/3/2019 Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

    9/10

    the change in action is mostly dependent on other actions (as opposed to angles) and so

    we present this submatrix in Figure 5. We see in this figure that the change in action in

    FIG. 5: Absolute value of the time averaged Jacobian of action variables illustrating the funneling

    of energy towards the lowest mode.

    any particular mode is dominated by influences of modes of equal or higher wavenumber

    only. This natural funneling of energy is remarkable and strengthens the robustness of the

    transition process. In fact, the funneling concept is associated with the horizontal-verticalgraph theoretic structure of the dynamics which has been shown to be a way to characterize

    the robustness of a dynamical system (see reference [13]).

    We have found that even though this high dimensional system uses a complex nonlinear

    process to achieve global transition between different states, the qualitative nature is easily

    understood as a funneling of energy from higher modes to the lowest mode whos phase

    space is manipulated by these higher modes. In fact the dynamics in the averaged sense

    are enough to predict the activation process. These results were compared in the stochastic

    setting and it is shown that the effect of targeted perturbation is an accelerant with respect

    to Kramers theory and the spatial character of the perturbation effects the divergence from

    Kramers estimate.

    This work was supported in part by DARPA DSO under AFOSR contract FA9550-07-C-

    9

  • 8/3/2019 Bryan Eisenhower and Igor Mezic- Targeted Activation in Deterministic and Stochastic Systems

    10/10

    0024 and AFOSR Grant FA9550-06-1-0088

    [1] I. Mezic, PNAS 103 (2006).

    [2] M. G. Forest, C. G. Goedde, and A. Sinha, Physical Review Letters pp. 27225 (1992).

    [3] N. Ercolani, M. G. Forest, and D. W. McLaughlin, Physica D Nonlinear Phenomena pp.

    34984 (1990).

    [4] B. Eisenhower and I. Mezic, Proc. of the 46th IEEE CDC (2007).

    [5] P. DuToit, I. Mezic, and J. Marsden, Physica D (in print) (2009).

    [6] S. Lee and W. Sung, Physical Review E 6302 (2001), ISSN 1063-651X.

    [7] S. Fugmann, D. Hennig, L. Schimansky-Geier, and P. Hanggi, Physical Review E 77 (2008),

    ISSN 1539-3755.

    [8] D. Hennig, S. Fugmann, L. Schimansky-Geier, and P. Hanggi, Acta Physica Polonica B 39

    (2008).

    [9] P. Hanggi, P. Talkner, and M. Borkovec, Reviews of Modern Physics 62, 251 (1990).

    [10] H. Yoshida, Physics Letters A 150 (1990).

    [11] E. Fermi, J. Pasta, and S. Ulam, Los Alamos report LA-1940 (1955).

    [12] J. Ford, Journal of Mathematical Physics 2 (1961).

    [13] I. Mezic, 43rd IEEE Conference on Decision and Control (2004).

    10