This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Mode selectivity of dynamically induced conformation in many-body chain-like bead-spring model

Yoshiyuki Y. Yamaguchi [email protected] Department of Applied Mathematics and Physics, Graduate School of Informatics, Kyoto University, Kyoto 606-8501, Japan
Abstract

We consider conformation of a chain consisting of beads connected by stiff springs, where the conformation is determined by the bending angles between the consecutive two springs. A conformation is stabilized or destabilized not only by a given bending potential but also the fast spring motion, and stabilization by the spring motion depends on their excited normal modes. This stabilization mechanism has been named the dynamically induced conformation in a previous work on a three-body system. We extend analyses of the dynamically induced conformation in many-body chain-like bead-spring systems. The normal modes of the springs depend on the conformation, and the simple rule of the dynamical stabilization is that the lowest eigenfrequency mode contributes to the stabilization of the conformation. The high the eigenfrequency is, the more the destabilization emerges. We verify theoretical predictions by performing numerical simulations.

I Introduction

Conformation has a deep relationship to function as found in isomerization. Maintenance of a conformation requires stability, and the stability is usually associated with landscape of the given potential energy function. We however underline that dynamics also contributes to the stability.

A typical example of the dynamical stabilization is the Kapitza pendulum, which is the inverted pendulum under the uniform gravity stephenson-08 ; kapitza-51 . The inverted pendulum is stabilized by applying fast vertical oscillation of the pivot. This highly unintuitive stabilization is applied in a wide variety of fields due to importance of the mechanism bukov-dalessio-polkovnikov-15 ; grifoni-hanggi-96 ; wickenbrock-etal-12 ; chizhevsky-smeu-giacomelli-03 ; chizhevsky-14 ; cubero-etal-06 ; bordet-morfu-13 ; weinberg-14 ; uzuntarla-etal-15 ; buchanan-jameson-oedjoe-62 ; baird-63 ; jameson-66 ; apffel-20 ; bellman-mentsman-meerkov-86 ; shapiro-zinn-97 ; borromeo-marchesoni-07 ; richards-etal-18 .

In the Kapitza pendulum the stabilization is realized by adding fast motion externally. We stress that the fast oscillation is not necessarily external, and an internal fast oscillation can also stabilize a conformation in an autonomous system. Indeed, in a chain-like bead-spring model consisting of beads connected by stiff springs, the straight conformation is stabilized by fast spring motion, whereas the potential energy function contains only the spring energy and does not depend on the bending angles yanagita-konishi-jp .

The above dynamical stabilization in a bead-spring model is firstly observed in numerical simulations, and then theoretically analyzed in the three-body model yamaguchi-etal-22 with the aids of the multiple-scale analysis bender-orszag-99 and the averaging method krylov-bogoliubov-34 ; krylov-bogoliubov-47 ; guckenheimer-holmes-83 . A surprising result is that the stability of a conformation depends on the excited normal modes of the springs. Suppose that the system has the equal masses and the identical springs. The straight (the fully bent) conformation is stabilized (destabilized) by the in-phase mode and is destabilized (stabilized) by the antiphase mode. Here, the in-phase (the antiphase) mode is a normal mode of the springs, and is defined as the two springs expand and contract simultaneously (alternatively). The stabilization is obtained from dynamics of the springs, and the obtained conformation is called the dynamically induced conformation (DIC).

The previous analysis is performed in the three-body model, and several natural questions are induced in many-body systems: Is DIC ubiquitous? How does the stability of a conformation depend on the excited normal modes? Is there a simple rule in the dependence? The aim of this paper is to answer these questions in chain-like bead-spring models.

The present study has another importance in the context of conformational isomerization in flexible molecules. It is experimentally observed in NN-acetyl-tryptophan methyl amide that population of isomers is modified by exciting vibration in a bond, and the modified population depends on the excited bond dian-longarte-zweier-02 , whereas the Rice-Ramsperger-Kassel-Marcus theory RRKM1 ; RRKM2 ; RRKM3 states that the destination is determined statistically. This mode selectivity may have a deep connection with the mode dependence of DIC.

This paper is organized as follows. The chain-like bead-spring model is introduced in Sec. II with the equations of motion. We extract the equations of motion for the slow bending motion in Sec. III. Assuming absence of the bending potential for observing the simplest case, we theoretically exhibit the excited mode dependence of stability with concentrating on one-dimensional conformations, whose bending angles are 0 or π\pi. The theoretical predictions are examined through numerical simulations in Sec. V with applying the Lennard-Jones potential as the bending potential. Finally, Sec. VI is devoted to summary and discussions.

II Model

We consider the NN-body chain-like bead-spring model on 2\mathbb{R}^{2}. The model consists of NN beads connected by N1N-1 springs. The iith bead is characterized by the mass mi(0,)m_{i}\in(0,\infty), the position 𝒓i2\boldsymbol{r}_{i}\in\mathbb{R}^{2}, and the velocity 𝒓˙i=d𝒓i/dt2\dot{\boldsymbol{r}}_{i}={\rm d}\boldsymbol{r}_{i}/{\rm d}t\in\mathbb{R}^{2}, where tt\in\mathbb{R} is the time and 𝒓i\boldsymbol{r}_{i} are column vectors. See Fig. 1 for a schematic diagram of the system.

Refer to caption
Figure 1: The bead-spring model on 2\mathbb{R}^{2}. This diagram shows an example of N=5N=5 (five beads connected by four springs).

II.1 Lagrangian in Cartesian coordinates

The Lagrangian of the system is

L(𝒓,𝒓˙)=K(𝒓˙)V(𝒓)L(\boldsymbol{r},\dot{\boldsymbol{r}})=K(\dot{\boldsymbol{r}})-V(\boldsymbol{r}) (1)

in the Cartesian coordinates, where

𝒓=(𝒓1𝒓N),𝒓˙=(𝒓˙1𝒓˙N)2N.\boldsymbol{r}=\begin{pmatrix}\boldsymbol{r}_{1}\\ \vdots\\ \boldsymbol{r}_{N}\end{pmatrix},\quad\dot{\boldsymbol{r}}=\begin{pmatrix}\dot{\boldsymbol{r}}_{1}\\ \vdots\\ \dot{\boldsymbol{r}}_{N}\\ \end{pmatrix}\quad\in\mathbb{R}^{2N}. (2)

The kinetic energy function K(𝒓˙)K(\dot{\boldsymbol{r}}) is

K(𝒓˙)=12i=1Nmi𝒓˙i2.K(\dot{\boldsymbol{r}})=\dfrac{1}{2}\sum_{i=1}^{N}m_{i}||\dot{\boldsymbol{r}}_{i}||^{2}. (3)

The potential energy function V(𝒓)V(\boldsymbol{r}) consists of the spring potential VspringV_{\rm spring} and the bending potential UbendU_{\rm bend} as

V(𝒓)=Vspring(𝒓)+Ubend(𝒓),V(\boldsymbol{r})=V_{\rm spring}(\boldsymbol{r})+U_{\rm bend}(\boldsymbol{r}), (4)

where we assume that VspringV_{\rm spring} depends on only the lengths of the springs. When we consider a molecule, the spring potential represents stronger bonds, and the bending potential corresponds to weaker bonds. A chain-like model is expressed by the nearest neighbor interactions in the spring potential as

Vspring(𝒓)=i=1N1Vi(𝒓i+1𝒓i).V_{\rm spring}(\boldsymbol{r})=\sum_{i=1}^{N-1}V_{i}(||\boldsymbol{r}_{i+1}-\boldsymbol{r}_{i}||). (5)

II.2 Lagrangian in internal coordinates

The system described by the Lagrangian (1) has the two-dimensional translational symmetry. To reduce the two degrees of freedom, we introduce 2N22N-2 internal coordinates 𝒚2N2\boldsymbol{y}\in\mathbb{R}^{2N-2} as

𝒚=(𝒍ϕ)2N2,{𝒍=(l1,,lN1)TN1ϕ=(ϕ1,,ϕN1)TN1\boldsymbol{y}=\begin{pmatrix}\boldsymbol{l}\\ \boldsymbol{\phi}\\ \end{pmatrix}\in\mathbb{R}^{2N-2},\quad\left\{\begin{array}[]{l}\boldsymbol{l}=(l_{1},\cdots,l_{N-1})^{\rm T}\in\mathbb{R}^{N-1}\\ \boldsymbol{\phi}=(\phi_{1},\cdots,\phi_{N-1})^{\rm T}\in\mathbb{R}^{N-1}\\ \end{array}\right. (6)

where the superscript T represents transposition. The internal coordinates play an important role to extract coupling between bending motion and spring motion (see Ref. yanao-etal-07 for instance).

The vector 𝒍\boldsymbol{l} contains the lengths of the springs as

li=𝒓i+1𝒓i,(i=1,,N1).l_{i}=||\boldsymbol{r}_{i+1}-\boldsymbol{r}_{i}||,\quad(i=1,\cdots,N-1). (7)

The Euclidean norm is defined by 𝒙=x2+y2||\boldsymbol{x}||=\sqrt{x^{2}+y^{2}} for 𝒙=(x,y)2\boldsymbol{x}=(x,y)\in\mathbb{R}^{2}.

The angle ϕi(i=1,,N2)\phi_{i}~{}(i=1,\cdots,N-2) represents the bending angle between the vectors 𝒓i+2𝒓i+1\boldsymbol{r}_{i+2}-\boldsymbol{r}_{i+1} and 𝒓i+1𝒓i\boldsymbol{r}_{i+1}-\boldsymbol{r}_{i}, and satisfies

cosϕi=(𝒓i+2𝒓i+1)(𝒓i+1𝒓i)𝒓i+2𝒓i+1𝒓i+1𝒓i,(i=1,,N2)\cos\phi_{i}=\dfrac{(\boldsymbol{r}_{i+2}-\boldsymbol{r}_{i+1})\cdot(\boldsymbol{r}_{i+1}-\boldsymbol{r}_{i})}{||\boldsymbol{r}_{i+2}-\boldsymbol{r}_{i+1}||~{}||\boldsymbol{r}_{i+1}-\boldsymbol{r}_{i}||},\quad(i=1,\cdots,N-2) (8)

where 𝒙𝒚\boldsymbol{x}\cdot\boldsymbol{y} is the Euclidean inner product between 𝒙\boldsymbol{x} and 𝒚\boldsymbol{y}. The last angle variable ϕN1\phi_{N-1} associates to the total angular momentum. The system has the rotational symmetry and ϕN1\phi_{N-1} is a cyclic coordinate, but we keep it for later convenience.

Rewriting the kinetic energy in the variables 𝒚\boldsymbol{y} and 𝒚˙\dot{\boldsymbol{y}}, and assuming that the total angular momentum is zero, we have the Lagrangian

L(𝒚,𝒚˙)=12α,β=12N2Bαβ(𝒚)y˙αy˙βV(𝒚).L(\boldsymbol{y},\dot{\boldsymbol{y}})=\dfrac{1}{2}\sum_{\alpha,\beta=1}^{2N-2}B^{\alpha\beta}(\boldsymbol{y})\dot{y}_{\alpha}\dot{y}_{\beta}-V(\boldsymbol{y}). (9)

We used the same symbol V(𝒚)V(\boldsymbol{y}) for the potential energy function to simplify the notation. The function Bαβ(𝒚)B^{\alpha\beta}(\boldsymbol{y}) is the (α,β)(\alpha,\beta) element of the matrix 𝐁(𝒚)Mat(2N2)\boldsymbol{\rm B}(\boldsymbol{y})\in{\rm Mat}(2N-2). Here Mat(n){\rm Mat}(n) represents the set of real square matrices of size nn. The explicit form of 𝐁(𝒚)\boldsymbol{\rm B}(\boldsymbol{y}) is given in Appendix A. Note that an alphabetic index runs from 11 to N1N-1, and a Greek index runs from 11 to 2N22N-2.

II.3 Euler-Lagrange equations

From now on, we adopt the Einstein notation for the sum: We take the sum over an index if it appears twice in a term. The Euler-Lagrange equation for the variable yα(α=1,,2N2)y_{\alpha}~{}(\alpha=1,\cdots,2N-2) is expressed as

Bαβ(𝒚)y¨β+Dαβγ(𝒚)y˙βy˙γ+Vyα(𝒚)=0B^{\alpha\beta}(\boldsymbol{y})\ddot{y}_{\beta}+D_{\alpha}^{\beta\gamma}(\boldsymbol{y})\dot{y}_{\beta}\dot{y}_{\gamma}+\dfrac{\partial V}{\partial y_{\alpha}}(\boldsymbol{y})=0 (10)

where

Dαβγ(𝒚)=Bαβyγ(𝒚)12Bβγyα(𝒚).D_{\alpha}^{\beta\gamma}(\boldsymbol{y})=\dfrac{\partial B^{\alpha\beta}}{\partial y_{\gamma}}(\boldsymbol{y})-\dfrac{1}{2}\dfrac{\partial B^{\beta\gamma}}{\partial y_{\alpha}}(\boldsymbol{y}). (11)

III Theory

This section provides a general theory to derive the equations of motion which describe the slow bending motion. We introduce two assumptions in Sec. III.1 with a dimensionless small parameter ϵ(0<ϵ1)\epsilon~{}(0<\epsilon\ll 1). The small parameter induces expansions of the variables 𝒚\boldsymbol{y} and tt, where the expansion of tt follows the multiple-scale analysis. The spring and bending potentials are also expanded in Sec. III.2. The slow bending motion is captured in O(ϵ2)O(\epsilon^{2}) of the expanded Euler-Lagrange equations as shown in Sec. III.3. To eliminate the fast timescale included in O(ϵ2)O(\epsilon^{2}), we perform the averaging over the fast timescale in Sec. III.4. The averaged equations are not closed due to autonomy of the system, and we will overcome this difficulty by introducing a hypothesis in Sec. III.5 and the energy conservation law in Sec. III.6. Finally, we will obtain closed equations for the slow bending motion in Sec. III.7

The theory can be simplified in the chain-like model with the aid of the explicit form of the spring potential (5). However, we develop the theory as general as possible in this section.

III.1 Assumptions and expansions of variables

Let 𝒍=(l1,,,lN1,)T\boldsymbol{l}_{\ast}=(l_{1,\ast},\cdots,l_{N-1,\ast})^{\rm T} be the natural length vector of the springs, namely 𝒍\boldsymbol{l}_{\ast} solves

Vspring𝒍(𝒍)=𝟎.\dfrac{\partial V_{\rm spring}}{\partial\boldsymbol{l}}(\boldsymbol{l}_{\ast})=\boldsymbol{0}. (12)

We introduce the two assumptions with a small parameter ϵ\epsilon:

  • (A1)

    The amplitudes of the springs are sufficiently small comparing with the natural lengths. The ratio is of O(ϵ)O(\epsilon).

  • (A2)

    Large bending motion is sufficiently slow than the spring motion. The ratio of the two timescales is of O(ϵ)O(\epsilon).

We express these assumptions by the expansions of t,𝒍t,\boldsymbol{l}, and ϕ\boldsymbol{\phi} as

ddt=t0+ϵt1\dfrac{{\rm d}}{{\rm d}t}=\dfrac{\partial}{\partial t_{0}}+\epsilon\dfrac{\partial}{\partial t_{1}} (13)

and

𝒍(t0,t1)=𝒍+ϵ𝒍(1)(t0,t1),ϕ(t0,t1)=ϕ(0)(t1)+ϵϕ(1)(t0,t1),\begin{split}&\boldsymbol{l}(t_{0},t_{1})=\boldsymbol{l}_{\ast}+\epsilon\boldsymbol{l}^{(1)}(t_{0},t_{1}),\\ &\boldsymbol{\phi}(t_{0},t_{1})=\boldsymbol{\phi}^{(0)}(t_{1})+\epsilon\boldsymbol{\phi}^{(1)}(t_{0},t_{1}),\\ \end{split} (14)

which is summarized as

𝒚(t0,t1)=𝒚(0)(t1)+ϵ𝒚(1)(t0,t1).\boldsymbol{y}(t_{0},t_{1})=\boldsymbol{y}^{(0)}(t_{1})+\epsilon\boldsymbol{y}^{(1)}(t_{0},t_{1}). (15)

The two timescales t0=tt_{0}=t and t1=ϵtt_{1}=\epsilon t correspond to the fast spring motion and the slow bending motion, respectively. The vector 𝒍=(l1,,,lN1,)T\boldsymbol{l}_{\ast}=(l_{1,\ast},\cdots,l_{N-1,\ast})^{\rm T} is constant. We are interested in the large and slow motion of the bending angles described by ϕ(0)(t1)\boldsymbol{\phi}^{(0)}(t_{1}).

Two remarks are in order. First, the leading order of the expanded equations of motion is of O(ϵ)O(\epsilon), since the leading order of the velocities and the accelerations is of O(ϵ)O(\epsilon). Second, the velocities l˙i\dot{l}_{i} and ϕ˙i\dot{\phi}_{i} are of the same order of O(ϵ)O(\epsilon), and the fastness of motion in the assumption (A2) connotes a short period. Indeed, the distance of a normal mode orbit is of O(ϵ)O(\epsilon) and the period is of O(ϵ0)O(\epsilon^{0}), while the distance of the large bending motion is of O(ϵ0)O(\epsilon^{0}) and the period is of O(ϵ1)O(\epsilon^{-1}).

III.2 Expansions of the potential function

We also expand the potential functions VspringV_{\rm spring} and UbendU_{\rm bend} into power series of ϵ\epsilon. The expansions of the two functions are performed in different ways, as a result of Eq. (14).

The spring potential function VspringV_{\rm spring} is expanded around 𝒍\boldsymbol{l}_{\ast} as

Vspring(𝒍)=Vspring(𝒍)+12(𝒍𝒍)T𝐊l(𝒍𝒍)+O(ϵ3),V_{\rm spring}(\boldsymbol{l})=V_{\rm spring}(\boldsymbol{l}_{\ast})+\dfrac{1}{2}(\boldsymbol{l}-\boldsymbol{l}_{\ast})^{\rm T}\boldsymbol{\rm K}_{l}(\boldsymbol{l}-\boldsymbol{l}_{\ast})+O(\epsilon^{3}), (16)

where the (i,j)(i,j) element KlijK_{l}^{ij} of the matrix 𝐊lMat(N1)\boldsymbol{\rm K}_{l}\in{\rm Mat}(N-1) is

Klij=2Vspringlilj(𝒍),(i,j=1,,N1)K_{l}^{ij}=\dfrac{\partial^{2}V_{\rm spring}}{\partial l_{i}\partial l_{j}}(\boldsymbol{l}_{\ast}),\quad(i,j=1,\cdots,N-1) (17)

and 𝐊l\boldsymbol{\rm K}_{l} is assumed to be positive definite.

The bending potential UbendU_{\rm bend} can be also expanded around a stationary point ϕ\boldsymbol{\phi}_{\ast} if it exists, but this expansion is not useful since ϕϕ||\boldsymbol{\phi}-\boldsymbol{\phi}_{\ast}|| is not necessarily of O(ϵ)O(\epsilon). We thus expand UbendU_{\rm bend} by its amplitude as

Ubend(𝒚)=Ubend(0)(𝒚)+ϵUbend(1)(𝒚)+ϵ2Ubend(2)(𝒚)+.U_{\rm bend}(\boldsymbol{y})=U_{\rm bend}^{(0)}(\boldsymbol{y})+\epsilon U_{\rm bend}^{(1)}(\boldsymbol{y})+\epsilon^{2}U_{\rm bend}^{(2)}(\boldsymbol{y})+\cdots. (18)

It is important to note that Ubend(0)(𝒚)U_{\rm bend}^{(0)}(\boldsymbol{y}) and Ubend(1)(𝒚)U_{\rm bend}^{(1)}(\boldsymbol{y}) solely depend on 𝒍\boldsymbol{l} with the conditions

Ubend(0)𝒍(𝒍)=Ubend(1)𝒍(𝒍)=𝟎\dfrac{\partial U_{\rm bend}^{(0)}}{\partial\boldsymbol{l}}(\boldsymbol{l}_{\ast})=\dfrac{\partial U_{\rm bend}^{(1)}}{\partial\boldsymbol{l}}(\boldsymbol{l}_{\ast})=\boldsymbol{0} (19)

for satisfying the assumptions (see Appendix B). Integrating them into the spring potential Vspring(𝒍)V_{\rm spring}(\boldsymbol{l}), we may set Ubend(0)(𝒚),Ubend(1)(𝒚)0U_{\rm bend}^{(0)}(\boldsymbol{y}),U_{\rm bend}^{(1)}(\boldsymbol{y})\equiv 0. The leading order of UbendU_{\rm bend} is hence of O(ϵ2)O(\epsilon^{2}), and this ordering is consistent with weaker bonds derived from the bending potential.

III.3 Expansion of the Euler-Lagrange equations

Substituting Eqs. (13), (14), (16), and (18) into Eq. (10), we have the expanded equations of motion order by order of ϵ\epsilon. As remarked in the end of Sec. III.1, the nontrivial equations of motion start from O(ϵ)O(\epsilon).

III.3.1 O(ϵ)O(\epsilon)

The equations of motion in O(ϵ)O(\epsilon) is linear and governed by the springs as

𝒚(1)2t02=𝐗(𝒚(0))𝒚(1)\dfrac{\partial{}^{2}\boldsymbol{y}^{(1)}}{\partial t_{0}^{2}}=-\boldsymbol{\rm X}(\boldsymbol{y}^{(0)})\boldsymbol{y}^{(1)} (20)

where

𝐗(𝒚)=[𝐁(𝒚)]1𝐊Mat(2N2)\boldsymbol{\rm X}(\boldsymbol{y})=[\boldsymbol{\rm B}(\boldsymbol{y})]^{-1}\boldsymbol{\rm K}\in{\rm Mat}(2N-2) (21)

and

𝐊=(𝐊l𝐎𝐎𝐎)Mat(2N2).\boldsymbol{\rm K}=\begin{pmatrix}\boldsymbol{\rm K}_{l}&\boldsymbol{\rm O}\\ \boldsymbol{\rm O}&\boldsymbol{\rm O}\end{pmatrix}\in{\rm Mat}(2N-2). (22)

The symbol 𝐎\boldsymbol{\rm O} represents the zero matrix. The variables ϕ(0)\boldsymbol{\phi}^{(0)} to observe do not appear in O(ϵ)O(\epsilon), and we progress to the next order.

III.3.2 O(ϵ2)O(\epsilon^{2})

In O(ϵ2)O(\epsilon^{2}) the equations of motion are

Bαβ(𝒚(0))(y¨β)(2)+Dαβγ(y(0))(y˙β)(1)(y˙γ)(1)+Bαβyγ(y(0))(y¨β)(1)yγ(1)+Ubend(2)yα(𝒚(0))=0,\begin{split}&B^{\alpha\beta}(\boldsymbol{y}^{(0)})\left(\ddot{y}_{\beta}\right)^{(2)}+D_{\alpha}^{\beta\gamma}(y^{(0)})\left(\dot{y}_{\beta}\right)^{(1)}\left(\dot{y}_{\gamma}\right)^{(1)}\\ &+\dfrac{\partial B^{\alpha\beta}}{\partial y_{\gamma}}(y^{(0)})\left(\ddot{y}_{\beta}\right)^{(1)}y_{\gamma}^{(1)}+\dfrac{\partial U_{\rm bend}^{(2)}}{\partial y_{\alpha}}(\boldsymbol{y}^{(0)})=0,\end{split} (23)

where

(𝒚˙)(1)=d𝒚(0)dt1+𝒚(1)t0,(𝒚¨)(1)=𝒚(1)2t02,(𝒚¨)(2)=d𝒚(0)2dt12+22𝒚(1)t0t1.\begin{split}&\left(\dot{\boldsymbol{y}}\right)^{(1)}=\dfrac{{\rm d}\boldsymbol{y}^{(0)}}{{\rm d}t_{1}}+\dfrac{\partial\boldsymbol{y}^{(1)}}{\partial t_{0}},\qquad\left(\ddot{\boldsymbol{y}}\right)^{(1)}=\dfrac{\partial{}^{2}\boldsymbol{y}^{(1)}}{\partial t_{0}^{2}},\\ &\left(\ddot{\boldsymbol{y}}\right)^{(2)}=\dfrac{{\rm d}{}^{2}\boldsymbol{y}^{(0)}}{{\rm d}t_{1}^{2}}+2\dfrac{\partial^{2}\boldsymbol{y}^{(1)}}{\partial t_{0}\partial t_{1}}.\\ \end{split} (24)

The vector (𝒚˙)(1)(\dot{\boldsymbol{y}})^{(1)} is the first order part of 𝒚˙\dot{\boldsymbol{y}} and (𝒚˙)(1)d𝒚(1)/dt(\dot{\boldsymbol{y}})^{(1)}\neq{\rm d}\boldsymbol{y}^{(1)}/{\rm d}t. The variables ϕ(0)\boldsymbol{\phi}^{(0)} appear in (𝒚˙)(1)(\dot{\boldsymbol{y}})^{(1)} and (𝒚¨)(2)(\ddot{\boldsymbol{y}})^{(2)}.

III.4 Averaging

The equations (23) depend on the fast timescale t0t_{0} through 𝒚(1)\boldsymbol{y}^{(1)}, and we eliminate it by taking the average. The average in the timescale t0t_{0} is defined by

φ=limT1T0Tφ(t0)𝑑t0.\left\langle\varphi\right\rangle=\lim_{T\to\infty}\dfrac{1}{T}\int_{0}^{T}\varphi(t_{0})dt_{0}. (25)

The averaged equations are

Bαβ(𝒚(0))dyβ(0)2dt12+Dαβγ(𝒚(0))dyβ(0)dt1dyγ(0)dt1+Ubend(2)yα(𝒚(0))=𝒜α,\begin{split}&B^{\alpha\beta}(\boldsymbol{y}^{(0)})\dfrac{{\rm d}{}^{2}y_{\beta}^{(0)}}{{\rm d}t_{1}^{2}}+D_{\alpha}^{\beta\gamma}(\boldsymbol{y}^{(0)})\dfrac{{\rm d}y_{\beta}^{(0)}}{{\rm d}t_{1}}\dfrac{{\rm d}y_{\gamma}^{(0)}}{{\rm d}t_{1}}\\ &+\dfrac{\partial U_{\rm bend}^{(2)}}{\partial y_{\alpha}}(\boldsymbol{y}^{(0)})=\mathcal{A}_{\alpha},\end{split} (26)

where the averaged term 𝒜α\mathcal{A}_{\alpha} is

𝒜α=12Tr[𝐁yα(𝒚(0))𝐗(𝒚(0))𝒚(1)𝒚(1)T].\begin{split}\mathcal{A}_{\alpha}&=\dfrac{1}{2}{\rm Tr}\left[\dfrac{\partial\boldsymbol{\rm B}}{\partial y_{\alpha}}(\boldsymbol{y}^{(0)})\boldsymbol{\rm X}(\boldsymbol{y}^{(0)})\left\langle\boldsymbol{y}^{(1)}\boldsymbol{y}^{(1){\rm T}}\right\rangle\right].\\ \end{split} (27)

The symbol Tr{\rm Tr} represents the matrix trace. We used the relation

𝒚(1)t0(𝒚(1)t0)T=𝐗(𝒚(0))𝒚(1)𝒚(1)T\left\langle\dfrac{\partial{\boldsymbol{y}}^{(1)}}{\partial t_{0}}\left(\dfrac{\partial\boldsymbol{y}^{(1)}}{\partial t_{0}}\right)^{\rm T}\right\rangle=\boldsymbol{\rm X}(\boldsymbol{y}^{(0)})\left\langle\boldsymbol{y}^{(1)}\boldsymbol{y}^{(1){\rm T}}\right\rangle (28)

proven by performing the integration by parts.

The averaged term 𝒜α\mathcal{A}_{\alpha} depends on the solution 𝒚(1)\boldsymbol{y}^{(1)} to Eq. (20), which is obtained by diagonalizing the matrix 𝐗(𝒚(0))\boldsymbol{\rm X}(\boldsymbol{y}^{(0)}) as

𝐗(𝒚(0))𝐏(𝒚(0))=𝐏(𝒚(0))𝚲(𝒚(0)).\boldsymbol{\rm X}(\boldsymbol{y}^{(0)})\boldsymbol{\rm P}(\boldsymbol{y}^{(0)})=\boldsymbol{\rm P}(\boldsymbol{y}^{(0)})\boldsymbol{\rm\Lambda}(\boldsymbol{y}^{(0)}). (29)

𝐏Mat(2N2)\boldsymbol{\rm P}\in{\rm Mat}(2N-2) is a diagonalizing matrix, and the diagonal matrix 𝚲\boldsymbol{\rm\Lambda} contains the eigenvalues of 𝐗\boldsymbol{\rm X}:

𝚲(𝒚(0))=diag(λ1,,λN1,0,,0)Diag(2N2),\boldsymbol{\rm\Lambda}(\boldsymbol{y}^{(0)})={\rm diag}(\lambda_{1},\cdots,\lambda_{N-1},0,\cdots,0)\in{\rm Diag}(2N-2), (30)

where the symbol Diag(n){\rm Diag}(n) represents the set of real diagonal matrices. The N1N-1 zeroeigenvalues come from the fact dim(Ker𝐊)=N1\dim({\rm Ker}\boldsymbol{\rm K})=N-1 in general. We assume that the corresponding N1N-1 amplitudes of 𝒚(1)\boldsymbol{y}^{(1)} are zero. Since the average of the square of a sinusoidal function is 1/21/2, we have in general

𝒚(1)𝒚(1)T=12𝐏(𝒚(0))𝐖(t1)2𝐏(𝒚(0))T,\left\langle\boldsymbol{y}^{(1)}\boldsymbol{y}^{(1){\rm T}}\right\rangle=\dfrac{1}{2}\boldsymbol{\rm P}(\boldsymbol{y}^{(0)})\boldsymbol{\rm W}(t_{1})^{2}\boldsymbol{\rm P}(\boldsymbol{y}^{(0)})^{\rm T}, (31)

where

𝐖(t1)=diag(w1,,wN1,0,,0)Diag(2N2).\boldsymbol{\rm W}(t_{1})={\rm diag}(w_{1},\cdots,w_{N-1},0,\cdots,0)\in{\rm Diag}(2N-2). (32)

The matrix 𝐖\boldsymbol{\rm W} contains the N1N-1 nontrivial amplitudes wi(i=1,,N1)w_{i}~{}(i=1,\cdots,N-1), which evolve in the slow timescale t1t_{1} through coupling with ϕ(0)(t1)\boldsymbol{\phi}^{(0)}(t_{1}), while the initial phases of 𝒚(1)\boldsymbol{y}^{(1)} are eliminated by the average.

Summarizing, the averaged term 𝒜α\mathcal{A}_{\alpha} is modified by Eqs. (29) and (31) to

𝒜α(𝒚(0))=14Tr[𝐏T𝐁yα𝐏𝚲𝐖(t1)2],(α=1,,2N2).\begin{split}&\mathcal{A}_{\alpha}(\boldsymbol{y}^{(0)})=\dfrac{1}{4}{\rm Tr}\left[\boldsymbol{\rm P}^{\rm T}\dfrac{\partial\boldsymbol{\rm B}}{\partial y_{\alpha}}\boldsymbol{\rm P}\boldsymbol{\rm\Lambda}\boldsymbol{\rm W}(t_{1})^{2}\right],\\ &(\alpha=1,\cdots,2N-2).\end{split} (33)

We can show that

𝒜1==𝒜N1=0.\mathcal{A}_{1}=\cdots=\mathcal{A}_{N-1}=0. (34)

In other words, the averaged terms survive only in the equations for ϕ(0)\boldsymbol{\phi}^{(0)}. See Appendix C.

III.5 Hypothesis

We stress that Eq. (26) is not closed, because the averaged term (33) depends on 𝐖(t1)\boldsymbol{\rm W}(t_{1}), which nontrivially evolves in the timescale t1t_{1}. We have to eliminate the N1N-1 nontrivial amplitudes wi(t1)(i=1,,N1)w_{i}(t_{1})~{}(i=1,\cdots,N-1). To this end, we introduce the hypothesis yamaguchi-etal-22 inspired from the adiabatic invariance:

wi(t1)2=νiw(t1)2(i=1,,N1).w_{i}(t_{1})^{2}=\nu_{i}w(t_{1})^{2}\quad(i=1,\cdots,N-1). (35)

This hypothesis supposes that the normal mode energy ratios are constant of time, and reduces the number of unknown variables from N1N-1 to one, although the averaged term 𝒜α\mathcal{A}_{\alpha} depends on the constants νi(i=1,,N1)\nu_{i}~{}(i=1,\cdots,N-1). We arrange the constants in the matrix

𝐍=diag(ν1,,νN1,0,,0)Diag(2N2),\boldsymbol{\rm N}={\rm diag}(\nu_{1},\cdots,\nu_{N-1},0,\cdots,0)\in{\rm Diag}(2N-2), (36)

and the amplitude matrix 𝐖\boldsymbol{\rm W} is simplified to

𝐖(t1)2=w(t1)2𝐍.\boldsymbol{\rm W}(t_{1})^{2}=w(t_{1})^{2}\boldsymbol{\rm N}. (37)

The averaged term 𝒜α\mathcal{A}_{\alpha} is then modified to

𝒜α(𝒚(0))=w(t1)24Tr[𝐏T𝐁yα𝐏𝚲𝐍].\mathcal{A}_{\alpha}(\boldsymbol{y}^{(0)})=\dfrac{w(t_{1})^{2}}{4}{\rm Tr}\left[\boldsymbol{\rm P}^{\rm T}\dfrac{\partial\boldsymbol{\rm B}}{\partial y_{\alpha}}\boldsymbol{\rm P}\boldsymbol{\rm\Lambda}\boldsymbol{\rm N}\right]. (38)

We remark that the numbering of the normal modes is important in the application of the hypothesis, since the eigenvalues of 𝐗(𝒚(0))\boldsymbol{\rm X}(\boldsymbol{y}^{(0)}) depend on the bending angles ϕ(0)\boldsymbol{\phi}^{(0)}. In N=3N=3, the two springs expand and contract simultaneously (alternatively) in the in-phase mode (the antiphase mode), which has the energy ratio νin\nu_{\rm in} (νanti\nu_{\rm anti}). Adopting ν1=νin\nu_{1}=\nu_{\rm in} and ν2=νanti\nu_{2}=\nu_{\rm anti}, the hypothesis (35) is approximately verified for any value of ϕ(0)\boldsymbol{\phi}^{(0)} yamaguchi-etal-22 .

However, the global numbering is not trivial in general, because two eigenvalues of 𝐗(𝒚(0))\boldsymbol{\rm X}(\boldsymbol{y}^{(0)}) may cross by varying ϕ(0)\boldsymbol{\phi}^{(0)} as shown in Fig. 2. Nevertheless, the hypothesis is approximately valid when the bending motion is not large, since the system is close to the integrable system, Eq. (20), and the mode numbers can be identified in a local region of ϕ(0)\boldsymbol{\phi}^{(0)}. Consequently, the hypothesis is useful to study stationarity and stability of a conformation. From now on, we locally number the modes in the ascending order of the eigenvalues (see Fig. 2) unless there otherwise stated.

Refer to caption
Figure 2: The two eigenvalues of 𝐗(𝒚(0))\boldsymbol{\rm X}(\boldsymbol{y}^{(0)}), which depend on ϕ1(0)\phi_{1}^{(0)} for N=3N=3. The in-phase mode (purple solid line) and the antiphase mode (green dashed line). m1=m2=m3=m=1m_{1}=m_{2}=m_{3}=m=1. 𝐊l=k𝐄\boldsymbol{\rm K}_{l}=k\boldsymbol{\rm E} with k=10k=10, where 𝐄\boldsymbol{\rm E} is the unit matrix. The eigenvalues are (k/m)(2cosϕ1(0))(k/m)(2\mp\cos\phi_{1}^{(0)}). The modes are locally numbered in the ascending order of the eigenvalues. The conformation symbol 𝒄\boldsymbol{c} is defined in Sec. IV.1: 𝒄=(1)\boldsymbol{c}=(1) represents the straight conformation and 𝒄=(1)\boldsymbol{c}=(-1) the fully bent conformation.

III.6 Energy conservation

The last unknown variable w(t1)w(t_{1}) is eliminated by the energy conservation. Expanding energy EE as E=ϵ2E(2)+E=\epsilon^{2}E^{(2)}+\cdots, the leading term is written as

E(2)=12Tr[𝐁(𝒚(0))(𝒚˙)(1)(𝒚˙)(1)T]+Ubend(2)(𝒚(0))+12Tr[𝐊𝒚(1)𝒚(1)T].\begin{split}E^{(2)}&=\dfrac{1}{2}{\rm Tr}\left[\boldsymbol{\rm B}(\boldsymbol{y}^{(0)})\left(\dot{\boldsymbol{y}}\right)^{(1)}\left(\dot{\boldsymbol{y}}\right)^{(1){\rm T}}\right]+U_{\rm bend}^{(2)}(\boldsymbol{y}^{(0)})\\ &+\dfrac{1}{2}{\rm Tr}\left[\boldsymbol{\rm K}\boldsymbol{y}^{(1)}\boldsymbol{y}^{(1){\rm T}}\right].\end{split} (39)

Taking the average, we have

E(2)=12Tr[𝐁(𝒚(0))d𝒚(0)dt1(d𝒚(0)dt1)T]+Ubend(2)(𝒚(0))+12Tr[𝐏T𝐊𝐏𝐖2]\begin{split}\left\langle E^{(2)}\right\rangle&=\dfrac{1}{2}{\rm Tr}\left[\boldsymbol{\rm B}(\boldsymbol{y}^{(0)})\dfrac{{\rm d}\boldsymbol{y}^{(0)}}{{\rm d}t_{1}}\left(\dfrac{{\rm d}\boldsymbol{y}^{(0)}}{{\rm d}t_{1}}\right)^{\rm T}~{}\right]+U_{\rm bend}^{(2)}(\boldsymbol{y}^{(0)})\\ &+\dfrac{1}{2}{\rm Tr}\left[\boldsymbol{\rm P}^{\rm T}\boldsymbol{\rm K}\boldsymbol{\rm P}\boldsymbol{\rm W}^{2}\right]\\ \end{split} (40)

by the relations (28) and (31). In the right-hand side of Eq. (40), the first line represents the bending energy, and the second line the spring energy.

The hypothesis (36) gives the equality

w(t1)2Tr[𝐏T𝐊𝐏𝐍]=2[E(2)Ubend(2)(𝒚(0))]Tr[𝐁(𝒚(0))d𝒚(0)dt1(d𝒚(0)dt1)T],\begin{split}w(t_{1})^{2}{\rm Tr}\left[\boldsymbol{\rm P}^{\rm T}\boldsymbol{\rm K}\boldsymbol{\rm P}\boldsymbol{\rm N}\right]&=2\left[E^{(2)}-U_{\rm bend}^{(2)}(\boldsymbol{y}^{(0)})\right]\\ &-{\rm Tr}\left[\boldsymbol{\rm B}(\boldsymbol{y}^{(0)})\dfrac{{\rm d}\boldsymbol{y}^{(0)}}{{\rm d}t_{1}}\left(\dfrac{{\rm d}\boldsymbol{y}^{(0)}}{{\rm d}t_{1}}\right)^{\rm T}~{}\right],\end{split} (41)

where we denoted E(2)\left\langle E^{(2)}\right\rangle by E(2)E^{(2)} for simplicity. Substituting Eq. (41) into Eq. (38), we have the averaged term 𝒜α\mathcal{A}_{\alpha} as

𝒜α(𝒚(0))={E(2)Ubend(2)(𝒚(0))214Tr[𝐁(𝒚(0))d𝒚(0)dt1(d𝒚(0)dt1)T]}𝒮α(𝒚(0))\begin{split}\mathcal{A}_{\alpha}(\boldsymbol{y}^{(0)})&=\left\{\dfrac{E^{(2)}-U_{\rm bend}^{(2)}(\boldsymbol{y}^{(0)})}{2}\right.\\ &\left.-\dfrac{1}{4}{\rm Tr}\left[\boldsymbol{\rm B}(\boldsymbol{y}^{(0)})\dfrac{{\rm d}\boldsymbol{y}^{(0)}}{{\rm d}t_{1}}\left(\dfrac{{\rm d}\boldsymbol{y}^{(0)}}{{\rm d}t_{1}}\right)^{\rm T}~{}\right]\right\}\mathcal{S}_{\alpha}(\boldsymbol{y}^{(0)})\end{split} (42)

where 𝒮α(α=1,,2N2)\mathcal{S}_{\alpha}~{}(\alpha=1,\cdots,2N-2) are

𝒮α(𝒚)=Tr[𝐏(𝒚)T𝐁yα(𝒚)𝐏(𝒚)𝚲(𝒚)𝐍]Tr[𝐏(𝒚)T𝐊𝐏(𝒚)𝐍].\mathcal{S}_{\alpha}(\boldsymbol{y})=\dfrac{{\rm Tr}\left[\boldsymbol{\rm P}(\boldsymbol{y})^{\rm T}\dfrac{\partial\boldsymbol{\rm B}}{\partial y_{\alpha}}(\boldsymbol{y})\boldsymbol{\rm P}(\boldsymbol{y})\boldsymbol{\rm\Lambda}(\boldsymbol{y})\boldsymbol{\rm N}\right]}{{\rm Tr}\left[\boldsymbol{\rm P}(\boldsymbol{y})^{\rm T}\boldsymbol{\rm K}\boldsymbol{\rm P}(\boldsymbol{y})\boldsymbol{\rm N}\right]}. (43)

We can show that

𝒮1==𝒮N1=0.\mathcal{S}_{1}=\cdots=\mathcal{S}_{N-1}=0. (44)

See Appendix C. We remark that in the inside traces of 𝒮α\mathcal{S}_{\alpha} the size of the matrices can be reduced from 2N22N-2 to N1N-1 as derived in Appendix C.

We have eliminated the nontrivial unknown variables wi(t1)(i=1,,N1)w_{i}(t_{1})~{}(i=1,\cdots,N-1) from the averaged term (42), which depends on only the variables ϕ(0)(t1)\boldsymbol{\phi}^{(0)}(t_{1}) and the constants 𝒍\boldsymbol{l}_{\ast}, 𝐍\boldsymbol{\rm N}, and E(2)E^{(2)}. Therefore, the equations of motion (26) for ϕ(0)\boldsymbol{\phi}^{(0)} is now closed, and dynamics depends on the excited normal modes 𝐍\boldsymbol{\rm N} and energy E(2)E^{(2)}.

III.7 Final result

Substituting Eq. (42) into Eq. (26), we have

Bαβ(𝒚(0))dyβ(0)2dt12+[Dαβγ(𝒚(0))+14Bβγ(𝒚(0))𝒮α(𝒚(0))]dyβ(0)dt1dyγ(0)dt1+Ubend(2)yα(𝒚(0))E(2)Ubend(2)(𝒚(0))2𝒮α(𝒚(0))=0\begin{split}&B^{\alpha\beta}(\boldsymbol{y}^{(0)})\dfrac{{\rm d}{}^{2}y_{\beta}^{(0)}}{{\rm d}t_{1}^{2}}+\left[D_{\alpha}^{\beta\gamma}(\boldsymbol{y}^{(0)})+\dfrac{1}{4}B^{\beta\gamma}(\boldsymbol{y}^{(0)})\mathcal{S}_{\alpha}(\boldsymbol{y}^{(0)})\right]\dfrac{{\rm d}y_{\beta}^{(0)}}{{\rm d}t_{1}}\dfrac{{\rm d}y_{\gamma}^{(0)}}{{\rm d}t_{1}}+\dfrac{\partial U_{\rm bend}^{(2)}}{\partial y_{\alpha}}(\boldsymbol{y}^{(0)})-\dfrac{E^{(2)}-U_{\rm bend}^{(2)}(\boldsymbol{y}^{(0)})}{2}\mathcal{S}_{\alpha}(\boldsymbol{y}^{(0)})=0\end{split} (45)

for α=1,,2N2\alpha=1,\cdots,2N-2. The final result for the bending variables ϕ(0)(t1)\boldsymbol{\phi}^{(0)}(t_{1}) is

dϕi(0)2dt12+Fijn(𝒚(0))dϕj(0)dt1dϕn(0)dt1+Gi(𝒚(0))=0\dfrac{{\rm d}{}^{2}\phi_{i}^{(0)}}{{\rm d}t_{1}^{2}}+F_{i}^{jn}(\boldsymbol{y}^{(0)})\dfrac{{\rm d}\phi_{j}^{(0)}}{{\rm d}t_{1}}\dfrac{{\rm d}\phi_{n}^{(0)}}{{\rm d}t_{1}}+G_{i}(\boldsymbol{y}^{(0)})=0 (46)

for i=1,,N1i=1,\cdots,N-1. The functions FijnF_{i}^{jn} and GiG_{i} are

Fijn(𝒚)=[Bϕϕ1]is(Bϕϕsjϕn12Bϕϕjnϕs+14𝒯sBϕϕjn)F_{i}^{jn}(\boldsymbol{y})=\left[B_{\phi\phi}^{-1}\right]^{is}\left(\dfrac{\partial B_{\phi\phi}^{sj}}{\partial\phi_{n}}-\dfrac{1}{2}\dfrac{\partial B_{\phi\phi}^{jn}}{\partial\phi_{s}}+\dfrac{1}{4}\mathcal{T}_{s}B_{\phi\phi}^{jn}\right) (47)

and

Gi(𝒚)=[Bϕϕ1]is(Ubend(2)ϕsE(2)Ubend(2)2𝒯s).G_{i}(\boldsymbol{y})=\left[B_{\phi\phi}^{-1}\right]^{is}\left(\dfrac{\partial U_{\rm bend}^{(2)}}{\partial\phi_{s}}-\dfrac{E^{(2)}-U_{\rm bend}^{(2)}}{2}\mathcal{T}_{s}\right). (48)

Here, we decomposed the matrix 𝐁Mat(2N2)\boldsymbol{\rm B}\in{\rm Mat}(2N-2) into four square submatrices of the size N1N-1 as

𝐁(𝒚)=(𝐁ll(ϕ)𝐁lϕ(𝒚)𝐁ϕl(𝒚)𝐁ϕϕ(𝒚)).\boldsymbol{\rm B}(\boldsymbol{y})=\begin{pmatrix}\boldsymbol{\rm B}_{ll}(\boldsymbol{\phi})&\boldsymbol{\rm B}_{l\phi}(\boldsymbol{y})\\ \boldsymbol{\rm B}_{\phi l}(\boldsymbol{y})&\boldsymbol{\rm B}_{\phi\phi}(\boldsymbol{y})\\ \end{pmatrix}. (49)

See Eq. (86) for explicit forms of the submatrices. The functions

𝒯i(𝒚)=𝒮i+N1(𝒚)=Tr[𝐏T𝐁ϕi𝐏𝚲𝐍]Tr[𝐏T𝐊𝐏𝐍](i=1,,N1)\begin{split}&\mathcal{T}_{i}(\boldsymbol{y})=\mathcal{S}_{i+N-1}(\boldsymbol{y})=\dfrac{{\rm Tr}\left[\boldsymbol{\rm P}^{\rm T}\dfrac{\partial\boldsymbol{\rm B}}{\partial\phi_{i}}\boldsymbol{\rm P}\boldsymbol{\rm\Lambda}\boldsymbol{\rm N}\right]}{{\rm Tr}\left[\boldsymbol{\rm P}^{\rm T}\boldsymbol{\rm K}\boldsymbol{\rm P}\boldsymbol{\rm N}\right]}\\ &(i=1,\cdots,N-1)\end{split} (50)

are introduced to renumber 𝒮α\mathcal{S}_{\alpha} for avoiding zerocomponents shown in Eq. (44). It is worth noting that the spring potential VspringV_{\rm spring} is included in the final equations of motion (46) up to the second order, namely only through the matrix 𝐊l\boldsymbol{\rm K}_{l}.

IV Dynamically induced stability

The equations (46) are rewritten as

dϕi(0)dt1=vi,dvidt1=Fijn(𝒚(0))vjvnGi(𝒚(0)),\dfrac{{\rm d}\phi_{i}^{(0)}}{{\rm d}t_{1}}=v_{i},\quad\dfrac{{\rm d}v_{i}}{{\rm d}t_{1}}=-F_{i}^{jn}(\boldsymbol{y}^{(0)})v_{j}v_{n}-G_{i}(\boldsymbol{y}^{(0)}), (51)

which describe dynamics on the 2N22N-2 dimensional space of

𝚽=(ϕ(0)𝒗)2N2,\boldsymbol{\Phi}=\begin{pmatrix}\boldsymbol{\phi}^{(0)}\\ \boldsymbol{v}\end{pmatrix}\in\mathbb{R}^{2N-2}, (52)

where 𝒗=(v1,,vN1)T\boldsymbol{v}=(v_{1},\cdots,v_{N-1})^{\rm T}. It is clear that

𝚽=(ϕ(0)𝒗):stationary{𝑮(𝒚(0))=𝟎,𝒗=𝟎.\boldsymbol{\Phi}_{\ast}=\begin{pmatrix}\boldsymbol{\phi}^{(0)}_{\ast}\\ \boldsymbol{v}_{\ast}\end{pmatrix}:\text{stationary}\quad\Longleftrightarrow\quad\left\{\begin{array}[]{l}\boldsymbol{G}(\boldsymbol{y}^{(0)}_{\ast})=\boldsymbol{0},\\ \boldsymbol{v}_{\ast}=\boldsymbol{0}.\end{array}\right. (53)

Stability of a stationary point 𝚽\boldsymbol{\Phi}_{\ast} is hence determined by the Jacobian of the vector field 𝑮(𝒚(0))\boldsymbol{G}(\boldsymbol{y}^{(0)}_{\ast}), which depends on the averaged terms 𝒯i\mathcal{T}_{i} and the bending potential Ubend(2)U_{\rm bend}^{(2)}.

In this section we concentrate on

Ubend(2)(𝒚)0U_{\rm bend}^{(2)}(\boldsymbol{y})\equiv 0 (54)

to clarify the dynamically induced stability by the averaged terms 𝒯i\mathcal{T}_{i}. According to Eq. (48), E(2)E^{(2)} is an overall factor of the function GiG_{i} when the bending potential is absent, and stability does not depend on E(2)E^{(2)}. We come back to the chain-like system

𝐊l=diag(k1,,kN1)Diag(N1).\boldsymbol{\rm K}_{l}={\rm diag}(k_{1},\cdots,k_{N-1})\in{\rm Diag}(N-1). (55)

Further, we restrict ourselves to the uniform setting

mi=m,kj=k,lj,=l(1iN;1jN1)m_{i}=m,~{}k_{j}=k,~{}l_{j,\ast}=l_{\ast}\quad(1\leq i\leq N;1\leq j\leq N-1) (56)

and to the one-dimensional conformations introduced in Sec. IV.1.

IV.1 One-dimensional conformations

We introduce the one-dimensional conformations whose set is denoted by

𝒞1={(ϕ1,,ϕN2)|ϕi{0,π}(i=1,,N2)}.\mathcal{C}^{1}=\left\{(\phi_{1},\cdots,\phi_{N-2})~{}|~{}\phi_{i}\in\{0,\pi\}~{}(i=1,\cdots,N-2)\right\}. (57)

A conformation in 𝒞1\mathcal{C}^{1} is stationary as proven in Appendix D. Appearance of the bending potential forbids ϕi=π\phi_{i}=\pi in general to avoid collision between beads, but we accept ϕi=π\phi_{i}=\pi in this section to discuss the simplest case. Later we will perform numerical simulations under appearance of a bending potential which forbids ϕi=π\phi_{i}=\pi.

A conformation in 𝒞1\mathcal{C}^{1} is symbolized by a sequence of 11 and 1-1: 11 represents the straight joint (ϕ=0\phi=0), and 1-1 the fully bent joint (ϕ=π\phi=\pi). The conformation symbol is denoted by 𝒄=(c1,,cN2)\boldsymbol{c}=(c_{1},\cdots,c_{N-2}). We identify two symmetric conformations like (1,1,1)(1,-1,-1) and (1,1,1)(-1,-1,1), because each of which is mapped to the other by changing the starting end of the chain. All the possible one-dimensional conformations for N=5N=5 are illustrated in Fig. 3 with their conformation symbols.

Refer to caption
Figure 3: Illustration of the one-dimensional conformations for N=5N=5. The stabilizing mode is characterized by the two types of bonds: A red thin (a blue thick) bond represents that the spring is longer (shorter) than the natural length. Conformation symbols are (a) 𝒄=(1,1,1)\boldsymbol{c}=(1,1,1), (b) 𝒄=(1,1,1)\boldsymbol{c}=(1,1,-1), (c) 𝒄=(1,1,1)\boldsymbol{c}=(1,-1,1), (d) 𝒄=(1,1,1)\boldsymbol{c}=(1,-1,-1), (e) 𝒄=(1,1,1)\boldsymbol{c}=(-1,1,-1), and (f) 𝒄=(1,1,1)\boldsymbol{c}=(-1,-1,-1).

IV.2 Stability

Let the point 𝚽\boldsymbol{\Phi}_{\ast} be stationary. Stability of this stationary point is obtained from the eigenvalues of the Jacobian matrix for Eq. (51),

𝐉(𝚽)=(𝐎𝐄D𝑮(𝒚(0))𝐎),\boldsymbol{\rm J}(\boldsymbol{\Phi}_{\ast})=\begin{pmatrix}\boldsymbol{\rm O}&\boldsymbol{\rm E}\\ -D\boldsymbol{G}(\boldsymbol{y}_{\ast}^{(0)})&\boldsymbol{\rm O}\\ \end{pmatrix}, (58)

where

D𝑮=(G1ϕ1G1ϕN1GN1ϕ1GN1ϕN1)D\boldsymbol{G}=\begin{pmatrix}\dfrac{\partial G_{1}}{\partial\phi_{1}}&\cdots&\dfrac{\partial G_{1}}{\partial\phi_{N-1}}\\ \vdots&\ddots&\vdots\\ \dfrac{\partial G_{N-1}}{\partial\phi_{1}}&\cdots&\dfrac{\partial G_{N-1}}{\partial\phi_{N-1}}\\ \end{pmatrix} (59)

is the Jacobian matrix of the vector field 𝑮(𝒚)\boldsymbol{G}(\boldsymbol{y}). We do not take the derivatives with respect to the variables 𝒍\boldsymbol{l}, because Eq. (51) includes only the constant lengths 𝒍\boldsymbol{l}_{\ast}.

Let us assume that the matrix D𝑮(𝒚(0))D\boldsymbol{G}(\boldsymbol{y}_{\ast}^{(0)}) is diagonalizable and has the eigenvalues g1,,gN1g_{1},\cdots,g_{N-1}. One eigenvalue, denoted by gN1g_{N-1}, should be zero from the rotational symmetry, and we remove it from the stability criterion. The nontrivial eigenvalues of 𝐉(𝚽)\boldsymbol{\rm J}(\boldsymbol{\Phi}_{\ast}) are ±g1,,±gN2\pm\sqrt{-g_{1}},\cdots,\pm\sqrt{-g_{N-2}}. See Appendix E for a proof and Fig. 4 for a schematic explanation of the relation between the eigenvalues of D𝑮D\boldsymbol{G} and 𝐉\boldsymbol{\rm J}.

Refer to caption
Figure 4: Eigenvalues of D𝑮D\boldsymbol{G} and 𝐉\boldsymbol{\rm J} on the complex plane. The blue circle in the panel (a) produces the two blue circles in the panel (b), and the same for the red crosses. A conformation is stable if and only if all the eigenvalues are on the blue thick line.

An eigenvalue gig_{i} is called a stable eigenvalue if gi(0,)g_{i}\in(0,\infty), and a zeroeigenvalue if gi=0g_{i}=0, and an unstable eigenvalue if gi[0,)g_{i}\not\in[0,\infty). The conformation represented by ϕ(0)\boldsymbol{\phi}^{(0)}_{\ast} is (neutrally) stable holm-etal-85 if and only if there is no unstable eigenvalue.

IV.3 Mode selectivity of dynamical stabilization

We first excite only one normal mode of the springs: All the diagonal elements of 𝐍\boldsymbol{\rm N} are zero except for one element. Stability of the one-dimensional conformations is summarized in Table 1 with dependency on the excited normal mode, where the normal modes are numbered in the ascending order of the eigenfrequencies around the considering conformation, as mentioned after Eq. (35). Stability is symbolized by S, Z, and U, and the number after S (Z, U) represents the number of the stable (zero, unstable) eigenvalues of D𝑮D\boldsymbol{G}, whose sum is N2N-2. The symbol is omitted if the number of corresponding eigenvalues is zero. For instance, the symbol S2U1 represents that the conformation has 22 stable eigenvalues and 11 unstable eigenvalue, and the conformation is unstable.

NN Conformation Stability / Square of eigenfrequency
𝒄\boldsymbol{c} Mode-11 Mode-22 Mode-33 Mode-44 Mode-55
33 S1 / 1 U1 / 3
(1)(1) (+,+)(+,+) (+,)(+,-)
(1)(-1) (+,)(+,-) (+,+)(+,+)
44 S2 / 0.585786 Z2 / 2 U2 / 3.41421
(1,1)(1,1) (+,+,+)(+,+,+) (+,0,)(+,0,-) (+,,+)(+,-,+)
(1,1)(1,-1) (+,+,)(+,+,-) (+,0,+)(+,0,+) (+,,)(+,-,-)
(1,1)(-1,-1) (+,,+)(+,-,+) (+,0,)(+,0,-) (+,+,+)(+,+,+)
55 S3 / 0.381966 S2U1 / 1.38197 S1U2 / 2.61803 U3 / 3.61803
(1,1,1)(1,1,1) (+,+,+,+)(+,+,+,+) (+,+,,)(+,+,-,-) (+,,,+)(+,-,-,+) (+,,+,)(+,-,+,-)
(1,1,1)(1,1,-1) (+,+,+,)(+,+,+,-) (+,+,,+)(+,+,-,+) (+,,,)(+,-,-,-) (+,,+,+)(+,-,+,+)
(1,1,1)(1,-1,1) (+,+,,)(+,+,-,-) (+,+,+,+)(+,+,+,+) (+,,+,)(+,-,+,-) (+,,,+)(+,-,-,+)
(1,1,1)(1,-1,-1) (+,+,,+)(+,+,-,+) (+,+,+,)(+,+,+,-) (+,,+,+)(+,-,+,+) (+,,,)(+,-,-,-)
(1,1,1)(-1,1,-1) (+,,,+)(+,-,-,+) (+,,+,)(+,-,+,-) (+,+,+,+)(+,+,+,+) (+,+,,)(+,+,-,-)
(1,1,1)(-1,-1,-1) (+,,+,)(+,-,+,-) (+,,,+)(+,-,-,+) (+,+,,)(+,+,-,-) (+,+,+,+)(+,+,+,+)
66 S4 / 0.267949 S2Z2 / 1 Z4 / 2 Z2U2 / 3 U4 / 3.73205
(1,1,1,1)(1,1,1,1) (+,+,+,+,+)(+,+,+,+,+) (+,+,0,,)(+,+,0,-,-) (+,0,,0,+)(+,0,-,0,+) (+,,0,+,)(+,-,0,+,-) (+,,+,,+)(+,-,+,-,+)
(1,1,1,1)(1,1,1,-1) (+,+,+,+,)(+,+,+,+,-) (+,+,0,,+)(+,+,0,-,+) (+,0,,0,)(+,0,-,0,-) (+,,0,+,+)(+,-,0,+,+) (+,,+,,)(+,-,+,-,-)
(1,1,1,1)(1,1,-1,1) (+,+,+,,)(+,+,+,-,-) (+,+,0,+,+)(+,+,0,+,+) (+,0,,0,)(+,0,-,0,-) (+,,0,,+)(+,-,0,-,+) (+,,+,+,)(+,-,+,+,-)
(1,1,1,1)(1,1,-1,-1) (+,+,+,,+)(+,+,+,-,+) (+,+,0,+,)(+,+,0,+,-) (+,0,,0,+)(+,0,-,0,+) (+,,0,,)(+,-,0,-,-) (+,,+,+,+)(+,-,+,+,+)
(1,1,1,1)(1,-1,1,1) (+,+,,,)(+,+,-,-,-) (+,+,0,+,+)(+,+,0,+,+) (+,0,+,0,)(+,0,+,0,-) (+,,0,,+)(+,-,0,-,+) (+,,,+,)(+,-,-,+,-)
(1,1,1,1)(1,-1,1,-1) (+,+,,,+)(+,+,-,-,+) (+,+,0,+,)(+,+,0,+,-) (+,0,+,0,+)(+,0,+,0,+) (+,,0,,)(+,-,0,-,-) (+,,,+,+)(+,-,-,+,+)
(1,1,1,1)(1,-1,-1,1) (+,+,,+,+)(+,+,-,+,+) (+,+,0,,)(+,+,0,-,-) (+,0,+,0,+)(+,0,+,0,+) (+,,0,+,)(+,-,0,+,-) (+,,,,+)(+,-,-,-,+)
(1,1,1,1)(1,-1,-1,-1) (+,+,,+,)(+,+,-,+,-) (+,+,0,,+)(+,+,0,-,+) (+,0,+,0,)(+,0,+,0,-) (+,,0,+,+)(+,-,0,+,+) (+,,,,)(+,-,-,-,-)
(1,1,1,1)(-1,1,1,-1) (+,,,,+)(+,-,-,-,+) (+,,0,+,)(+,-,0,+,-) (+,0,+,0,+)(+,0,+,0,+) (+,+,0,,)(+,+,0,-,-) (+,+,,+,+)(+,+,-,+,+)
(1,1,1,1)(-1,1,-1,-1) (+,,,+,)(+,-,-,+,-) (+,,0,,+)(+,-,0,-,+) (+,0,+,0,)(+,0,+,0,-) (+,+,0,+,+)(+,+,0,+,+) (+,+,,,)(+,+,-,-,-)
(1,1,1,1)(-1,-1,-1,-1) (+,,+,,+)(+,-,+,-,+) (+,,0,+,)(+,-,0,+,-) (+,0,,0,+)(+,0,-,0,+) (+,+,0,,)(+,+,0,-,-) (+,+,+,+,+)(+,+,+,+,+)
Table 1: Stability of the one-dimensional conformations in the NN-body chain-like bead-spring model under the equal masses and the identical springs. The second column from the left represents the conformation symbol 𝒄\boldsymbol{c}. The mode number is defined in the ascending order of the eigenfrequencies. Stability of a conformation with a given mode is indicated by S (stable), Z(zero), and U (unstable), and the number after S (Z, U) represents the number of stable (zero, unstable) eigenvalues of D𝑮D\boldsymbol{G}. After the slash, the value of (m/k)ωj2(m/k)\omega_{j}^{2} is shown, where ωj\omega_{j} is the eigenfrequency of the mode. Sequences of +,0+,0, and - represent the eigenmode symbol 𝒔\boldsymbol{s}. See the text for the definitions of 𝒄\boldsymbol{c} and 𝒔\boldsymbol{s}.

The (N1)(N-1)-dimensional eigenvector of a normal mode is characterized by a sequence of +,0+,0, and -. The symbol +(0,)+~{}(0,-) represents that the corresponding spring is longer than (equal to, shorter than) the natural length. That is, for N=3N=3, the eigenmode (+,+)(+,+) implies that the two springs are initially longer than the natural length and the mode is the in-phase mode. Similarly, the eigenmode (+,)(+,-) represents the antiphase mode. We denote the eigenmode symbol by 𝒔=(s1,,sN1)\boldsymbol{s}=(s_{1},\cdots,s_{N-1}).

We have two observations in Table 1. First, each conformation is stabilized by the lowest eigenfrequency mode of the springs. The number of unstable directions increases as the eigenfrequency gets larger. Second, the stabilizing eigenmode is obtained by pulling the left (right) end of the chain to the left (right) as illustrated in Fig. 3.

Stability analysis can be extended to mixed modes. Analyses for N=3,4N=3,4 and 55 suggest that the dynamical stabilization is ubiquitous in a larger system having multimode excitation. Indeed, the dynamical stabilization is realized with an approximate probability of 0.80.8 up to N=5N=5, whereas higher eigenfrequency modes contribute to destabilization. See Appendix F.

V Numerical tests

We demonstrate dynamical stabilization and destabilization through numerical simulations of the system under the uniform setting (56) with

m=1,k=10,l=1.m=1,\quad k=10,\quad l_{\ast}=1. (60)

We use the Hamiltonian written in the Cartesian coordinate to use an explicit fourth-order symplectic integrator yoshida-90 with the time step Δt=103\Delta t=10^{-3}. The Hamiltonian associated with the Lagrangian (1) is

H(𝒓,𝒑)=12mi=1N𝒑i2+Vspring(𝒓)+Ubend(𝒓),H(\boldsymbol{r},\boldsymbol{p})=\dfrac{1}{2m}\sum_{i=1}^{N}||\boldsymbol{p}_{i}||^{2}+V_{\rm spring}(\boldsymbol{r})+U_{\rm bend}(\boldsymbol{r}), (61)

and the canonical equations of motion are

d𝒓idt=H𝒑i,d𝒑idt=H𝒓i,(i=1,,N).\dfrac{{\rm d}\boldsymbol{r}_{i}}{{\rm d}t}=\dfrac{\partial H}{\partial\boldsymbol{p}_{i}},\quad\dfrac{{\rm d}\boldsymbol{p}_{i}}{{\rm d}t}=-\dfrac{\partial H}{\partial\boldsymbol{r}_{i}},\quad(i=1,\cdots,N). (62)

V.1 System setting

The theory includes the spring potential VspringV_{\rm spring} up to the quadratic order, and we use the linear springs. The spring potential VspringV_{\rm spring} is defined in Eq.(5), and each spring ViV_{i} is

Vi(𝒓)=k2(𝒓i+1𝒓il)2,(i=1,,N1).V_{i}(\boldsymbol{r})=\dfrac{k}{2}(||\boldsymbol{r}_{i+1}-\boldsymbol{r}_{i}||-l_{\ast})^{2},\quad(i=1,\cdots,N-1). (63)

The bending potential UbendU_{\rm bend} is

Ubend(𝒓)=i<jULJ(𝒓j𝒓i),U_{\rm bend}(\boldsymbol{r})=\sum_{i<j}U_{\rm LJ}(||\boldsymbol{r}_{j}-\boldsymbol{r}_{i}||), (64)

and ULJU_{\rm LJ} is the Lennard-Jones potential

ULJ(r)=4ϵLJ[(σr)12(σr)6].U_{\rm LJ}(r)=4\epsilon_{\rm LJ}\left[\left(\dfrac{\sigma}{r}\right)^{12}-\left(\dfrac{\sigma}{r}\right)^{6}\right]. (65)

The parameter ϵLJ\epsilon_{\rm LJ} is of O(ϵ2)O(\epsilon^{2}), namely

ϵLJ=ϵ0ϵ2,ϵ0=O(ϵ0),\epsilon_{\rm LJ}=\epsilon_{0}\epsilon^{2},\quad\epsilon_{0}=O(\epsilon^{0}), (66)

to satisfy Ubend=O(ϵ2)U_{\rm bend}=O(\epsilon^{2}). We fix ϵLJ\epsilon_{\rm LJ} and σ\sigma as

ϵLJ=104,σ=1.\epsilon_{\rm LJ}=10^{-4},\quad\sigma=1. (67)

We may expect that the main contribution to the bending energy comes from pairs of second nearest beads, since the Lennard-Jones potential does not depend on the bending angles for a pair of nearest beads. For a second nearest pair with the bending angle ϕ\phi, the second-order Lennard-Jones potential is

ULJ(2)(ϕ)=4ϵ0{[σ22l2(1+cosϕ)]6[σ22l2(1+cosϕ)]3}.U_{\rm LJ}^{(2)}(\phi)=4\epsilon_{0}\left\{\left[\dfrac{\sigma^{2}}{2l_{\ast}^{2}(1+\cos\phi)}\right]^{6}-\left[\dfrac{\sigma^{2}}{2l_{\ast}^{2}(1+\cos\phi)}\right]^{3}\right\}. (68)

It takes the minimum value ϵ0-\epsilon_{0} at ±ϕmin\pm\phi_{\rm min}, where

ϕmin=|cos1[(σ21/3l)21]|1.94985.\phi_{\rm min}=\left|\cos^{-1}\left[\left(\dfrac{\sigma}{2^{1/3}l_{\ast}}\right)^{2}-1\right]\right|\simeq 1.94985. (69)

See Fig. 5.

Refer to caption
Figure 5: The second-order Lennard-Jones potential (68). The two arrows mark the minimum points ±ϕmin\pm\phi_{\rm min} (69). ϵ0=1,σ=1\epsilon_{0}=1,\sigma=1, and ł=1\l_{\ast}=1. The definition of the bending angle ϕ\phi is schematically represented at the top center.

It is worth noting that for N=3N=3 we can construct the effective potential yamaguchi-etal-22 , which is useful to understand stability of a conformation graphically. Examples are exhibited in Appendix G.

V.2 Initial conditions

The initial positions are given in the following three steps. First, we select a reference conformation from 𝒞1\mathcal{C}^{1} and put all the beads on the xx-axis. Second, we replace the bending angle ϕ=π\phi=\pi with ϕ=±ϕmin\phi=\pm\phi_{\rm min} to avoid collision of beads. The possible initial conformations for N=5N=5 are illustrated in Fig. 6. Third, we modify the lengths of the springs from the natural length to 𝒍=𝒍+ϵ𝒍(1)\boldsymbol{l}=\boldsymbol{l}_{\ast}+\epsilon\boldsymbol{l}^{(1)}, where 𝒍(1)\boldsymbol{l}^{(1)} is determined so as to excite normal modes in a desired manner approximately. Precise settings of the initial positions are described in Appendix H.

Refer to caption
Figure 6: Illustration of the possible initial conformations and their approximated bending potential energy UbendU_{\rm bend} for N=5N=5. Conformation symbols are (a) 𝒄=(1,1,1)\boldsymbol{c}=(1,1,1), (b) 𝒄=(1,1,1)\boldsymbol{c}=(1,1,-1), (c) 𝒄=(1,1,1)\boldsymbol{c}=(1,-1,1), (d) 𝒄=(1,1,1)\boldsymbol{c}=(1,-1,-1), (e) 𝒄=(1,1,1)\boldsymbol{c}=(-1,1,-1), and (f) 𝒄=(1,1,1)\boldsymbol{c}=(-1,-1,-1). The bending angle π\pi has been replaced with ±ϕmin\pm\phi_{\rm min} from Fig. 3.

The initial values of the momenta are set as follows. For the xx-direction, the initial values of the momenta are zero. For the yy-direction, they are randomly drawn from the uniform distribution on the interval [103,103][-10^{-3},10^{-3}] as perturbation to observe stability of the given conformation.

V.3 Dynamical stabilization and destabilization

We examine dynamical stabilization of high energy conformations. The first example is the straight conformation, which is illustrated in Fig. 6(a). Temporal evolution of ϕi(t)\phi_{i}(t) is exhibited in Fig. 7 by exciting the mode-11 (the stabilization mode) and varying the amplitude ϵ\epsilon of the mode. For ϵ=0.03\epsilon=0.03 the bending angles oscillate between the two points ±ϕmin\pm\phi_{\rm min}. However, for ϵ=0.04\epsilon=0.04 which provides larger E(2)E^{(2)}, the straight conformation is stabilized and the bending angles stay around 0. This observation is consistent with the expression of GiG_{i} (48), because larger E(2)E^{(2)} enhances contribution from the averaged term 𝒯s\mathcal{T}_{s}.

Refer to caption
Figure 7: Temporal evolution of the bending angles. N=5N=5. Conformation is 𝒄=(1,1,1)\boldsymbol{c}=(1,1,1), schematically represented at the top of each panel, and the mode-11 is excited. (a) ϵ=0.03\epsilon=0.03. E0.149ϵLJE\simeq-0.149\epsilon_{\rm LJ} (b) ϵ=0.04\epsilon=0.04. E0.070ϵLJE\simeq 0.070\epsilon_{\rm LJ}. ϵLJ=104\epsilon_{\rm LJ}=10^{-4}. ϕ1(t)\phi_{1}(t) (purple dotted line), ϕ2(t)\phi_{2}(t) (green broken line), and ϕ3(t)\phi_{3}(t) (blue solid line). Two black horizontal lines are shown at ±ϕmin\pm\phi_{\min}.

The stabilization by the mode-11 is also realized for partially straight conformations. For N=5,N=5, temporal evolution of ϕi(i=1,2,3)\phi_{i}~{}(i=1,2,3) are reported in Fig. 8 for conformations symbolized by 𝒄=(1,1,1)\boldsymbol{c}=(1,1,-1) [Fig. 6(b)] and 𝒄=(1,1,1)\boldsymbol{c}=(1,-1,-1) [Fig. 6(d)]. The bending angles stay around the initial values.

Refer to caption
Figure 8: Temporal evolution of the bending angles. N=5N=5. The mode-11 is excited. (a) Conformation 𝒄=(1,1,1)\boldsymbol{c}=(1,1,-1), ϵ=0.12\epsilon=0.12, and E5.22ϵLJE\simeq 5.22\epsilon_{\rm LJ}. (b) Conformation 𝒄=(1,1,1)\boldsymbol{c}=(1,-1,-1), ϵ=0.10\epsilon=0.10, and E2.45ϵLJE\simeq 2.45\epsilon_{\rm LJ}. ϵLJ=104\epsilon_{\rm LJ}=10^{-4}. The conformation is schematically represented at the top of each panel. ϕ1(t)\phi_{1}(t) (purple dotted line), ϕ2(t)\phi_{2}(t) (green broken line), and ϕ3(t)\phi_{3}(t) (blue solid line). Two black horizontal lines are shown at ±ϕmin\pm\phi_{\min}.

Finally, we demonstrate destabilization of the bent conformation 𝒄=(1,1,1)\boldsymbol{c}=(-1,-1,-1) [Fig. 6(f)], which is stable if dynamical instability does not kick in. The destabilization is realized by the mode-22 for instance as shown in Fig. 9(a), which is consistent with Table 1, although larger spring energy is necessary to destabilize the bent conformation. We stress that the destabilization is not induced only by largeness of energy, because the bent conformation is not destabilized by the mode-11 as shown in Fig. 9(b), while the values of energy are almost equal between the two cases.

Refer to caption
Figure 9: Excited mode dependence of the bent conformation 𝒄=(1,1,1)\boldsymbol{c}=(-1,-1,-1), schematically represented at the top of each panel. N=5N=5. ϵ=0.16\epsilon=0.16. (a) The mode-22 is excited and E9.6ϵLJE\simeq 9.6\epsilon_{\rm LJ}. (b) The mode-11 is excited and E10.2ϵLJE\simeq 10.2\epsilon_{\rm LJ}. ϵLJ=104\epsilon_{\rm LJ}=10^{-4}. ϕ1(t)\phi_{1}(t) (purple dotted line), ϕ2(t)\phi_{2}(t) (green broken line), and ϕ3(t)\phi_{3}(t) (blue solid line). Two lines are almost collapsed around ϕmin\phi_{\rm min} in the panel (b). Two black horizontal lines are shown at ±ϕmin\pm\phi_{\min}.

VI Summary

We have studied the dynamically induced conformation (DIC) in NN-body chain-like bead-spring models. We have extended a theory, which is developed for N=3N=3 in a previous work yamaguchi-etal-22 , to a general NN. The theory predicts that the dynamical stability depends on the excited normal modes of the springs and on the value of energy.

As the simplest case we have studied a system without the bending potential to clearly exhibit dynamical effects. Concentrating on the so-called one-dimensional conformations, which are stationary, We have investigated the mode dependent stability up to N=6N=6 under the condition of the equal masses and the identical springs. A simple rule of the mode dependency has been discovered: A conformation is stabilized by exciting the lowest eigenfrequency mode, and destabilization emerges as the eigenfrequency of the exited pure normal mode becomes higher.

We stress that DIC is ubiquitous. The theory is also applicable for mixed modes, and the stabilization of a conformation is realized with an approximate probability of 0.80.8 up to N=5N=5, when we randomly choose a mixed mode. The probability 0.80.8 is notable, because, among four normal modes in N=5N=5, only one mode contributes to the stabilization and the other three modes contribute to the destabilization. Moreover, the uniform setting of the equal masses and the identical springs is not essential for DIC yamaguchi-etal-22 .

The stabilization and destabilization of conformations have been demonstrated numerically in a system having the bending potential consisting of the Lennard-Jones potentials for each pair of beads. As the theory predicts, any conformation can be stabilized by exciting the lowest eigenfrequency mode which depends on the conformation, whereas a straight joint corresponds to the local maximum of a Lennard-Jones potential. Destabilization of the fully bend conformation, which corresponds to a local minimum point of the bending potential, has been also demonstrated by exciting a higher eigenfrequency mode.

We note that excitation of a normal mode is a nonequilibrium phenomenon, because the law of equipartition of energy holds among the normal modes in thermal equilibrium. Nevertheless, separation of the two timescales suggests that importance of DIC survives in a long time by the Boltzmann-Jeans conjecture boltzmann-95 ; jeans-03 ; jeans-05 ; landau-teller-36 ; benettin-galgani-giorgilli-89 ; baldin-benettin-91 . An important message of DIC is that the conformation is not determined by the bending potential only, and we have to input the dynamical (de)stabilization. This message sheds light on a new aspect of conformation and conformation change.

Acknowledgements.
The author thanks T. Yanagita, T. Konishi, and M. Toda for valuable discussions. The author acknowledges the support of JSPS KAKENHI Grant Numbers 16K05472 and 21K03402.

Appendix A Lagrangian in the internal coordinates

We rewrite the Lagrangian Eq. (1) into the internal coordinates through three changes of variables. We mainly consider modifications of the kinetic energy

K(𝒓˙)=12𝒓˙T𝐌𝒓˙,K(\dot{\boldsymbol{r}})=\dfrac{1}{2}\dot{\boldsymbol{r}}^{\rm T}\boldsymbol{\rm M}\dot{\boldsymbol{r}}, (70)

where

𝐌=diag(m1,,mN)Diag(N)\boldsymbol{\rm M}={\rm diag}(m_{1},\cdots,m_{N})\in{\rm Diag}(N) (71)

and the symbol Diag(n){\rm Diag}(n) represents the set of the real diagonal matrices of size nn.

The first change of variables is

(𝒒1𝒒N)=𝐒M(𝒓1𝒓N)\begin{pmatrix}\boldsymbol{q}_{1}\\ \vdots\\ \boldsymbol{q}_{N}\end{pmatrix}=\boldsymbol{\rm S}_{M}\begin{pmatrix}\boldsymbol{r}_{1}\\ \vdots\\ \boldsymbol{r}_{N}\end{pmatrix} (72)

where the matrix 𝐒MMat(N)\boldsymbol{\rm S}_{M}\in{\rm Mat}(N) is

𝐒M=(11000011000010000011m1/Mm2/Mm3/MmN1/MmN/M)\boldsymbol{\rm S}_{M}=\begin{pmatrix}-1&1&0&\cdots&0&0\\ 0&-1&1&\ddots&0&0\\ 0&0&-1&\ddots&0&0\\ \vdots&\ddots&\ddots&\ddots&\ddots&\vdots\\ 0&0&0&\cdots&-1&1\\ m_{1}/M&m_{2}/M&m_{3}/M&\cdots&m_{N-1}/M&m_{N}/M\\ \end{pmatrix} (73)

and

M=Tr𝐌=i=1Nmi.M={\rm Tr}\boldsymbol{\rm M}=\sum_{i=1}^{N}m_{i}. (74)

In the new variables 𝒒i\boldsymbol{q}_{i}, the kinetic energy KK is

K(𝒒˙)=12i,j=1N1Aij𝒒˙i𝒒˙j+μ2𝒒˙N2.K(\dot{\boldsymbol{q}})=\dfrac{1}{2}\sum_{i,j=1}^{N-1}A^{ij}\dot{\boldsymbol{q}}_{i}\cdot\dot{\boldsymbol{q}}_{j}+\dfrac{\mu}{2}||\dot{\boldsymbol{q}}_{N}||^{2}. (75)

The symmetric constant matrix 𝐀Mat(N1)\boldsymbol{\rm A}\in{\rm Mat}(N-1) is defined by

𝐒MT𝐌𝐒M1=(𝐀𝟎𝟎Tn2),\boldsymbol{\rm S}_{M}^{-{\rm T}}\boldsymbol{\rm M}\boldsymbol{\rm S}_{M}^{-1}=\begin{pmatrix}\boldsymbol{\rm A}&\boldsymbol{0}\\ \boldsymbol{0}^{\rm T}&n^{2}\\ \end{pmatrix}, (76)

where the superscript -T represents transposition of the inverse matrix and 𝟎\boldsymbol{0} is the zero column vector. The variable 𝒒N\boldsymbol{q}_{N} is a cyclic coordinate corresponding to the total momentum conservation due to the translational symmetry. We set the total momentum as zero, and we drop the last term of the right-hand side of Eq. (75).

The second change of variables introduce the polar coordinates. Denoting 𝒒i=(qxi,qyi)2\boldsymbol{q}_{i}=(q_{xi},q_{yi})\in\mathbb{R}^{2}, we introduce lil_{i} and θi\theta_{i} by

qxi=licosθi,qyi=lisinθi.q_{xi}=l_{i}\cos\theta_{i},\quad q_{yi}=l_{i}\sin\theta_{i}. (77)

In the vector form, the polar coordinates are expressed by

𝒒i=li𝒆li,𝒒˙i=l˙i𝒆li+liθ˙i𝒆θi,\boldsymbol{q}_{i}=l_{i}\boldsymbol{e}_{li},\quad\dot{\boldsymbol{q}}_{i}=\dot{l}_{i}\boldsymbol{e}_{li}+l_{i}\dot{\theta}_{i}\boldsymbol{e}_{\theta i}, (78)

where 𝒆li\boldsymbol{e}_{li} and 𝒆θi\boldsymbol{e}_{\theta i} are the unit vectors of the radial and the angle directions, respectively. The polar coordinates rewrite the spring potential as Vspring(𝒍)V_{\rm spring}(\boldsymbol{l}), and the kinetic energy KK as

K=12(𝒍˙T𝜽˙T)(𝐀C(𝜽)𝐀S(𝜽)𝐋(𝒍)𝐋(𝒍)𝐀S(𝜽)𝐋(𝒍)𝐀C(𝜽)𝐋(𝒍))(𝒍˙𝜽˙).K=\dfrac{1}{2}\begin{pmatrix}\dot{\boldsymbol{l}}^{\rm T}&\dot{\boldsymbol{\theta}}^{\rm T}\end{pmatrix}\begin{pmatrix}\boldsymbol{\rm A}_{C}(\boldsymbol{\theta})&\boldsymbol{\rm A}_{S}(\boldsymbol{\theta})\boldsymbol{\rm L}(\boldsymbol{l})\\ -\boldsymbol{\rm L}(\boldsymbol{l})\boldsymbol{\rm A}_{S}(\boldsymbol{\theta})&\boldsymbol{\rm L}(\boldsymbol{l})\boldsymbol{\rm A}_{C}(\boldsymbol{\theta})\boldsymbol{\rm L}(\boldsymbol{l})\\ \end{pmatrix}\begin{pmatrix}\dot{\boldsymbol{l}}\\ \dot{\boldsymbol{\theta}}\\ \end{pmatrix}. (79)

The diagonal matrix 𝐋\boldsymbol{\rm L} is defined by

𝐋(𝒍)=diag(l1,,lN1)Diag(N1).\boldsymbol{\rm L}(\boldsymbol{l})={\rm diag}(l_{1},\cdots,l_{N-1})\in{\rm Diag}(N-1). (80)

The symmetric matrix 𝐀C\boldsymbol{\rm A}_{C} and the antisymmetric matrix 𝐀S\boldsymbol{\rm A}_{S} are defined by

ACij(𝜽)=Aijcos(θiθj),ASij(𝜽)=Aijsin(θiθj).A_{C}^{ij}(\boldsymbol{\theta})=A^{ij}\cos(\theta_{i}-\theta_{j}),\quad A_{S}^{ij}(\boldsymbol{\theta})=A^{ij}\sin(\theta_{i}-\theta_{j}). (81)

The third change of variables introduces the bending angles ϕ\boldsymbol{\phi} as

(ϕ1ϕN1)=𝐒(θ1θN1),\begin{pmatrix}\phi_{1}\\ \vdots\\ \phi_{N-1}\\ \end{pmatrix}=\boldsymbol{\rm S}\begin{pmatrix}\theta_{1}\\ \vdots\\ \theta_{N-1}\\ \end{pmatrix}, (82)

where the constant matrix 𝐒\boldsymbol{\rm S} is

𝐒=(110000110000100000111N11N11N11N11N1)Mat(N1).\boldsymbol{\rm S}=\begin{pmatrix}-1&1&0&\cdots&0&0\\ 0&-1&1&\ddots&0&0\\ 0&0&-1&\ddots&0&0\\ \vdots&\ddots&\ddots&\ddots&\ddots&\vdots\\ 0&0&0&\cdots&-1&1\\ \frac{1}{N-1}&\frac{1}{N-1}&\frac{1}{N-1}&\cdots&\frac{1}{N-1}&\frac{1}{N-1}\\ \end{pmatrix}\in{\rm Mat}(N-1). (83)

The bending angles are from ϕ1\phi_{1} to ϕN2\phi_{N-2}, and we added an additional angle ϕN1\phi_{N-1} for later convenience. Performing the change of variables of Eq. (82), the kinetic energy is modified to

K(𝒚,𝒚˙)=12𝒚˙T𝐁(𝒚)𝒚˙,K(\boldsymbol{y},\dot{\boldsymbol{y}})=\dfrac{1}{2}\dot{\boldsymbol{y}}^{\rm T}\boldsymbol{\rm B}(\boldsymbol{y})\dot{\boldsymbol{y}}, (84)

where the matrix 𝐁\boldsymbol{\rm B} is

𝐁(𝒚)=(𝐁ll(ϕ)𝐁lϕ(𝒚)𝐁ϕl(𝒚)𝐁ϕϕ(𝒚))Mat(2N2)\boldsymbol{\rm B}(\boldsymbol{y})=\begin{pmatrix}\boldsymbol{\rm B}_{ll}(\boldsymbol{\phi})&\boldsymbol{\rm B}_{l\phi}(\boldsymbol{y})\\ \boldsymbol{\rm B}_{\phi l}(\boldsymbol{y})&\boldsymbol{\rm B}_{\phi\phi}(\boldsymbol{y})\\ \end{pmatrix}\in{\rm Mat}(2N-2) (85)

and the size-(N1)(N-1) submatrices are defined by

(𝐁ll𝐁lϕ𝐁ϕl𝐁ϕϕ)=(𝐀C(ϕ)𝐀S(ϕ)𝐋𝐒1𝐒T𝐋𝐀S(ϕ)𝐒T𝐋𝐀C(ϕ)𝐋𝐒1).\begin{pmatrix}\boldsymbol{\rm B}_{ll}&\boldsymbol{\rm B}_{l\phi}\\ \boldsymbol{\rm B}_{\phi l}&\boldsymbol{\rm B}_{\phi\phi}\end{pmatrix}=\begin{pmatrix}\boldsymbol{\rm A}_{C}(\boldsymbol{\phi})&\boldsymbol{\rm A}_{S}(\boldsymbol{\phi})\boldsymbol{\rm L}\boldsymbol{\rm S}^{-1}\\ -\boldsymbol{\rm S}^{-{\rm T}}\boldsymbol{\rm L}\boldsymbol{\rm A}_{S}(\boldsymbol{\phi})&\boldsymbol{\rm S}^{-{\rm T}}\boldsymbol{\rm L}\boldsymbol{\rm A}_{C}(\boldsymbol{\phi})\boldsymbol{\rm L}\boldsymbol{\rm S}^{-1}\\ \end{pmatrix}. (86)

The (i,j)(i,j) elements of the matrices 𝐀C(ϕ),𝐀S(ϕ)Mat(N1)\boldsymbol{\rm A}_{C}(\boldsymbol{\phi}),\boldsymbol{\rm A}_{S}(\boldsymbol{\phi})\in{\rm Mat}(N-1) are written respectively as

ACij(ϕ)=Aijcosϕi,j,ASij(ϕ)=Aijsinϕi,j,A_{C}^{ij}(\boldsymbol{\phi})=A^{ij}\cos\phi_{i,j},\quad A_{S}^{ij}(\boldsymbol{\phi})=A^{ij}\sin\phi_{i,j}, (87)

where

ϕi,j=θiθj={ϕi1++ϕj(i>j),0(i=j),(ϕj1++ϕi)(i<j).\phi_{i,j}=\theta_{i}-\theta_{j}=\left\{\begin{array}[]{ll}\phi_{i-1}+\cdots+\phi_{j}&(i>j),\\ 0&(i=j),\\ -(\phi_{j-1}+\cdots+\phi_{i})&(i<j).\\ \end{array}\right. (88)

The variable ϕN1\phi_{N-1} does not appear in 𝐁(𝒚)\boldsymbol{\rm B}(\boldsymbol{y}).

After the three changes of variables, we obtain the Lagrangian in the internal coordinates as Eq. (9). The variable ϕN1\phi_{N-1} is a cyclic coordinate corresponding to the rotational symmetry. We keep it for later convenience of computations.

Appendix B Ordering of the bending potential

Since 𝒚˙,𝒚¨=O(ϵ)\dot{\boldsymbol{y}},\ddot{\boldsymbol{y}}=O(\epsilon), the zeroth order equations of motion are

(Vyα(𝒚))(0)=0,(α=1,,2N2)\left(\dfrac{\partial V}{\partial y_{\alpha}}(\boldsymbol{y})\right)^{(0)}=0,\quad(\alpha=1,\cdots,2N-2) (89)

which implies

(Vyα(𝒚))(0)=Ubend(0)yα(𝒚(0))0,(α=1,,2N2).\left(\dfrac{\partial V}{\partial y_{\alpha}}(\boldsymbol{y})\right)^{(0)}=\dfrac{\partial U_{\rm bend}^{(0)}}{\partial y_{\alpha}}(\boldsymbol{y}^{(0)})\equiv 0,\quad(\alpha=1,\cdots,2N-2). (90)

Remembering 𝒚(0)=(𝒍,ϕ(0)(t1))T\boldsymbol{y}^{(0)}=(\boldsymbol{l}_{\ast},\boldsymbol{\phi}^{(0)}(t_{1}))^{\rm T}, we conclude that Ubend(0)U_{\rm bend}^{(0)} has no ϕ\boldsymbol{\phi} dependence and Ubend(0)(𝒍)U_{\rm bend}^{(0)}(\boldsymbol{l}) satisfies

Ubend(0)𝒍(𝒍)=𝟎.\dfrac{\partial U_{\rm bend}^{(0)}}{\partial\boldsymbol{l}}(\boldsymbol{l}_{\ast})=\boldsymbol{0}. (91)

We can put Ubend(0)U_{\rm bend}^{(0)} as a part of the spring potential Vspring(𝒍)V_{\rm spring}(\boldsymbol{l}) and neglect it. The included Ubend(0)U_{\rm bend}^{(0)} modifies the matrix 𝐊\boldsymbol{\rm K} in O(ϵ0)O(\epsilon^{0}).

The first order force is

(Ubendyα(𝒚))(1)=Ubend(1)yα(𝒚(0)).\left(\dfrac{\partial U_{\rm bend}}{\partial y_{\alpha}}(\boldsymbol{y})\right)^{(1)}=\dfrac{\partial U_{\rm bend}^{(1)}}{\partial y_{\alpha}}(\boldsymbol{y}^{(0)}). (92)

This force is constant in the fast timescale t0t_{0}. The first order equations of motion are

2t02(𝒍(1)ϕ(1))+(𝐁ll𝐁lϕ𝐁ϕl𝐁ϕϕ)1(𝐊l𝐎𝐎𝐎)(𝒍(1)ϕ(1))=(𝐁ll𝐁lϕ𝐁ϕl𝐁ϕϕ)1(𝒍Ubend(1)(𝒚(0))ϕUbend(1)(𝒚(0))).\begin{split}&\dfrac{\partial{}^{2}}{\partial t_{0}^{2}}\begin{pmatrix}\boldsymbol{l}^{(1)}\\ \boldsymbol{\phi}^{(1)}\end{pmatrix}+\begin{pmatrix}\boldsymbol{\rm B}_{ll}&\boldsymbol{\rm B}_{l\phi}\\ \boldsymbol{\rm B}_{\phi l}&\boldsymbol{\rm B}_{\phi\phi}\\ \end{pmatrix}^{-1}\begin{pmatrix}\boldsymbol{\rm K}_{l}&\boldsymbol{\rm O}\\ \boldsymbol{\rm O}&\boldsymbol{\rm O}\\ \end{pmatrix}\begin{pmatrix}\boldsymbol{l}^{(1)}\\ \boldsymbol{\phi}^{(1)}\end{pmatrix}\\ &=-\begin{pmatrix}\boldsymbol{\rm B}_{ll}&\boldsymbol{\rm B}_{l\phi}\\ \boldsymbol{\rm B}_{\phi l}&\boldsymbol{\rm B}_{\phi\phi}\\ \end{pmatrix}^{-1}\begin{pmatrix}\partial_{\boldsymbol{l}}U_{\rm bend}^{(1)}(\boldsymbol{y}^{(0)})\\ \partial_{\boldsymbol{\phi}}U_{\rm bend}^{(1)}(\boldsymbol{y}^{(0)})\\ \end{pmatrix}.\end{split} (93)

Focusing on the second line of the second term in the left-hand side, we find no restoring force in ϕ(1)\boldsymbol{\phi}^{(1)}. Therefore, if the gradient of Ubend(1)U_{\rm bend}^{(1)} at 𝒚(0)\boldsymbol{y}^{(0)} is not zero, secular terms are yielded and they break the perturbation expansion (14). Therefore, Ubend(1)U_{\rm bend}^{(1)} depends on 𝒍\boldsymbol{l} only, and satisfies

Ubend(1)𝒍(𝒍)=𝟎.\dfrac{\partial U_{\rm bend}^{(1)}}{\partial\boldsymbol{l}}(\boldsymbol{l}_{\ast})=\boldsymbol{0}. (94)

As discussed in O(ϵ0)O(\epsilon^{0}), Ubend(1)(𝒍)U_{\rm bend}^{(1)}(\boldsymbol{l}) is also put in the spring potential Vspring(𝒍)V_{\rm spring}(\boldsymbol{l}) and modifies 𝐊\boldsymbol{\rm K} in O(ϵ)O(\epsilon).

Consequently, the leading order term depending on ϕ\boldsymbol{\phi} is Ubend(2)(𝒚)U_{\rm bend}^{(2)}(\boldsymbol{y}). The slow motion of ϕ\boldsymbol{\phi} is hence determined by the second-order bending potential, which is the same order as the dynamical effects associated with the averaged term 𝒜α\mathcal{A}_{\alpha}.

Appendix C Simplifications

We can simplify expressions of the term 𝒜α\mathcal{A}_{\alpha} and the spring energy, which help to analyze stability of a stationary state. The idea is to decompose a size-(2N2)(2N-2) matrix into four half-size submatrices.

C.1 Decomposition of matrices

We consider the eigenvalue problem

𝐗𝐏=𝐏𝚲,\boldsymbol{\rm X}\boldsymbol{\rm P}=\boldsymbol{\rm P}\boldsymbol{\rm\Lambda}, (95)

where 𝐗=𝐁1𝐊\boldsymbol{\rm X}=\boldsymbol{\rm B}^{-1}\boldsymbol{\rm K}. The inverse matrix 𝐁1\boldsymbol{\rm B}^{-1} is obtained as

𝐁1=(𝐁~ll𝐁~lϕ𝐁~ϕl𝐁~ϕϕ)Mat(2N2),\boldsymbol{\rm B}^{-1}=\begin{pmatrix}\widetilde{\boldsymbol{\rm B}}_{ll}&\widetilde{\boldsymbol{\rm B}}_{l\phi}\\ \widetilde{\boldsymbol{\rm B}}_{\phi l}&\widetilde{\boldsymbol{\rm B}}_{\phi\phi}\\ \end{pmatrix}\in{\rm Mat}(2N-2), (96)

where

𝐁~ll=(𝐀C+𝐀S𝐀C1𝐀S)1,𝐁~lϕ=𝐀C1𝐀S𝐁~ll𝐋1𝐒T,𝐁~ϕl=𝐒𝐋1𝐀C1𝐀S𝐁~ll,𝐁~ϕϕ=𝐒𝐋1𝐁~ll𝐋1𝐒T.\begin{split}&\widetilde{\boldsymbol{\rm B}}_{ll}=\left(\boldsymbol{\rm A}_{C}+\boldsymbol{\rm A}_{S}\boldsymbol{\rm A}_{C}^{-1}\boldsymbol{\rm A}_{S}\right)^{-1},\\ &\widetilde{\boldsymbol{\rm B}}_{l\phi}=-\boldsymbol{\rm A}_{C}^{-1}\boldsymbol{\rm A}_{S}\widetilde{\boldsymbol{\rm B}}_{ll}\boldsymbol{\rm L}^{-1}\boldsymbol{\rm S}^{\rm T},\\ &\widetilde{\boldsymbol{\rm B}}_{\phi l}=\boldsymbol{\rm S}\boldsymbol{\rm L}^{-1}\boldsymbol{\rm A}_{C}^{-1}\boldsymbol{\rm A}_{S}\widetilde{\boldsymbol{\rm B}}_{ll},\\ &\widetilde{\boldsymbol{\rm B}}_{\phi\phi}=\boldsymbol{\rm S}\boldsymbol{\rm L}^{-1}\widetilde{\boldsymbol{\rm B}}_{ll}\boldsymbol{\rm L}^{-1}\boldsymbol{\rm S}^{\rm T}.\\ \end{split} (97)

See Appendix A for the definitions of the matrices 𝐋,𝐒,𝐀C\boldsymbol{\rm L},\boldsymbol{\rm S},\boldsymbol{\rm A}_{C}, and 𝐀S\boldsymbol{\rm A}_{S}. Note that 𝐁~ll𝐁ll1\widetilde{\boldsymbol{\rm B}}_{ll}\neq\boldsymbol{\rm B}_{ll}^{-1} in general and that 𝐁1\boldsymbol{\rm B}^{-1} is symmetric.

The decomposition of 𝐁1\boldsymbol{\rm B}^{-1} gives

𝐗=𝐁1𝐊=(𝐁~ll𝐊l𝐎𝐁~ϕl𝐊l𝐎).\boldsymbol{\rm X}=\boldsymbol{\rm B}^{-1}\boldsymbol{\rm K}=\begin{pmatrix}\widetilde{\boldsymbol{\rm B}}_{ll}\boldsymbol{\rm K}_{l}&\boldsymbol{\rm O}\\ \widetilde{\boldsymbol{\rm B}}_{\phi l}\boldsymbol{\rm K}_{l}&\boldsymbol{\rm O}\\ \end{pmatrix}. (98)

The diagonal matrix 𝚲\boldsymbol{\rm\Lambda} and a diagonalizing matrix 𝐏\boldsymbol{\rm P} are also decomposed as

𝚲=(𝚲l𝐎𝐎𝐎)Diag(2N2)\boldsymbol{\rm\Lambda}=\begin{pmatrix}\boldsymbol{\rm\Lambda}_{l}&\boldsymbol{\rm O}\\ \boldsymbol{\rm O}&\boldsymbol{\rm O}\\ \end{pmatrix}\in{\rm Diag}(2N-2) (99)

and

𝐏=(𝐏l𝐎𝐏ϕ𝐄)Mat(2N2),\boldsymbol{\rm P}=\begin{pmatrix}\boldsymbol{\rm P}_{l}&\boldsymbol{\rm O}\\ \boldsymbol{\rm P}_{\phi}&\boldsymbol{\rm E}\\ \end{pmatrix}\in{\rm Mat}(2N-2), (100)

where all the submatrices are of size-(N1)(N-1) and 𝐄\boldsymbol{\rm E} is the unit matrix. The submatrix 𝐏l\boldsymbol{\rm P}_{l} solves the eigenvalue problem

(𝐁~ll𝐊l)𝐏l=𝐏l𝚲l,\left(\widetilde{\boldsymbol{\rm B}}_{ll}\boldsymbol{\rm K}_{l}\right)\boldsymbol{\rm P}_{l}=\boldsymbol{\rm P}_{l}\boldsymbol{\rm\Lambda}_{l}, (101)

and the submatrix 𝐏ϕ\boldsymbol{\rm P}_{\phi} is determined from 𝐏l\boldsymbol{\rm P}_{l} as

𝐏ϕ=𝐒𝐋1𝐀C1𝐀S𝐏l.\boldsymbol{\rm P}_{\phi}=\boldsymbol{\rm S}\boldsymbol{\rm L}^{-1}\boldsymbol{\rm A}_{C}^{-1}\boldsymbol{\rm A}_{S}\boldsymbol{\rm P}_{l}. (102)

We further decompose the diagonal matrix 𝐖\boldsymbol{\rm W} as

𝐖=(𝐖l𝐎𝐎𝐎)Diag(2N2).\boldsymbol{\rm W}=\begin{pmatrix}\boldsymbol{\rm W}_{l}&\boldsymbol{\rm O}\\ \boldsymbol{\rm O}&\boldsymbol{\rm O}\\ \end{pmatrix}\in{\rm Diag}(2N-2). (103)

C.2 Simplification of the spring energy

The averaged spring energy Espring\left\langle E_{\rm spring}\right\rangle is

Espring=12Tr[𝐏T𝐊𝐏𝐖2]=12Tr[𝐏lT𝐊l𝐏l𝐖l2].\left\langle E_{\rm spring}\right\rangle=\dfrac{1}{2}{\rm Tr}\left[\boldsymbol{\rm P}^{\rm T}\boldsymbol{\rm K}\boldsymbol{\rm P}\boldsymbol{\rm W}^{2}\right]=\dfrac{1}{2}{\rm Tr}\left[\boldsymbol{\rm P}_{l}^{\rm T}\boldsymbol{\rm K}_{l}\boldsymbol{\rm P}_{l}\boldsymbol{\rm W}_{l}^{2}\right]. (104)

The matrices of the inside trace are reduced from size 2N22N-2 to size N1N-1. This expression is modified to

Espring=w(t1)22Tr[𝐏lT𝐊l𝐏l𝐍].\left\langle E_{\rm spring}\right\rangle=\dfrac{w(t_{1})^{2}}{2}{\rm Tr}\left[\boldsymbol{\rm P}_{l}^{\rm T}\boldsymbol{\rm K}_{l}\boldsymbol{\rm P}_{l}\boldsymbol{\rm N}\right]. (105)

in the use of the hypothesis (36).

C.3 Simplification of the averaged terms

Substituting the decomposition of the matrices 𝐁,𝚲,𝐏\boldsymbol{\rm B},\boldsymbol{\rm\Lambda},\boldsymbol{\rm P}, and 𝐖\boldsymbol{\rm W} into Eq. (33) and computing straightforwardly, we have

𝒜α=14Tr[𝐏lT𝐁~ll1yα𝐏l𝚲l𝐖l2],(α=1,,2N2).\mathcal{A}_{\alpha}=\dfrac{1}{4}{\rm Tr}\left[\boldsymbol{\rm P}_{l}^{\rm T}\dfrac{\partial\widetilde{\boldsymbol{\rm B}}_{ll}^{-1}}{\partial y_{\alpha}}\boldsymbol{\rm P}_{l}\boldsymbol{\rm\Lambda}_{l}\boldsymbol{\rm W}_{l}^{2}\right],\quad(\alpha=1,\cdots,2N-2). (106)

In the way we used the relation

(𝐀C𝐀C1)yα=𝐎𝐀C1𝐀Cyα𝐀C1=𝐀C1yα.\dfrac{\partial(\boldsymbol{\rm A}_{C}\boldsymbol{\rm A}_{C}^{-1})}{\partial y_{\alpha}}=\boldsymbol{\rm O}\quad\Longleftrightarrow\quad\boldsymbol{\rm A}_{C}^{-1}\dfrac{\partial\boldsymbol{\rm A}_{C}}{\partial y_{\alpha}}\boldsymbol{\rm A}_{C}^{-1}=-\dfrac{\partial\boldsymbol{\rm A}_{C}^{-1}}{\partial y_{\alpha}}. (107)

Similarly, the function 𝒮α\mathcal{S}_{\alpha} is also simplified to

𝒮α=Tr[𝐏lT𝐁~ll1yα𝐏l𝚲l𝐍l]Tr[𝐏lT𝐊l𝐏l𝐍l],(α=1,,2N2)\mathcal{S}_{\alpha}=\dfrac{{\rm Tr}\left[\boldsymbol{\rm P}_{l}^{\rm T}\dfrac{\partial\widetilde{\boldsymbol{\rm B}}_{ll}^{-1}}{\partial y_{\alpha}}\boldsymbol{\rm P}_{l}\boldsymbol{\rm\Lambda}_{l}\boldsymbol{\rm N}_{l}\right]}{{\rm Tr}\left[\boldsymbol{\rm P}_{l}^{\rm T}\boldsymbol{\rm K}_{l}\boldsymbol{\rm P}_{l}\boldsymbol{\rm N}_{l}\right]},\quad(\alpha=1,\cdots,2N-2) (108)

where

𝐍=(𝐍l𝐎𝐎𝐎),𝐍l=diag(ν1,,νN1).\boldsymbol{\rm N}=\begin{pmatrix}\boldsymbol{\rm N}_{l}&\boldsymbol{\rm O}\\ \boldsymbol{\rm O}&\boldsymbol{\rm O}\\ \end{pmatrix},\quad\boldsymbol{\rm N}_{l}={\rm diag}(\nu_{1},\cdots,\nu_{N-1}). (109)

The expressions (106) and (108) prove respectively

𝒜1==𝒜N1=0\mathcal{A}_{1}=\cdots=\mathcal{A}_{N-1}=0 (110)

and

𝒮1==𝒮N1=0,\mathcal{S}_{1}=\cdots=\mathcal{S}_{N-1}=0, (111)

since the matrix 𝐁~ll1\widetilde{\boldsymbol{\rm B}}_{ll}^{-1} does not depend on 𝒍\boldsymbol{l}.

C.4 Further simplifications

The average spring energy and the denominator of 𝒮α\mathcal{S}_{\alpha} are further simplified by choosing a special diagonalizing matrix 𝐏l\boldsymbol{\rm P}_{l}, where the matrix 𝐏l\boldsymbol{\rm P}_{l} satisfies the eigenvalue problem (101). The symmetric matrix 𝐊l\boldsymbol{\rm K}_{l} is positive definite, and hence we can define the real symmetric matrix 𝐊l\sqrt{\boldsymbol{\rm K}_{l}}. Introducing the matrix 𝐐l\boldsymbol{\rm Q}_{l} as

𝐐l=𝐊l𝐏lMat(N1),\boldsymbol{\rm Q}_{l}=\sqrt{\boldsymbol{\rm K}_{l}}\boldsymbol{\rm P}_{l}\in{\rm Mat}(N-1), (112)

the eigenvalue problem is rewritten to

(𝐊l𝐁~ll𝐊l)𝐐l=𝐐l𝚲l.\left(\sqrt{\boldsymbol{\rm K}_{l}}~{}\widetilde{\boldsymbol{\rm B}}_{ll}~{}\sqrt{\boldsymbol{\rm K}_{l}}\right)\boldsymbol{\rm Q}_{l}=\boldsymbol{\rm Q}_{l}\boldsymbol{\rm\Lambda}_{l}. (113)

The matrix 𝐊l𝐁~ll𝐊l\sqrt{\boldsymbol{\rm K}_{l}}~{}\widetilde{\boldsymbol{\rm B}}_{ll}~{}\sqrt{\boldsymbol{\rm K}_{l}} is real symmetric, and hence we can choose 𝐐lO(N1)\boldsymbol{\rm Q}_{l}\in O(N-1), where O(N1)O(N-1) is the set of the orthogonal matrices of size N1N-1. Using 𝐐lO(N1)\boldsymbol{\rm Q}_{l}\in O(N-1), we have

𝐏lT𝐊l𝐏l=𝐄.\boldsymbol{\rm P}_{l}^{\rm T}\boldsymbol{\rm K}_{l}\boldsymbol{\rm P}_{l}=\boldsymbol{\rm E}. (114)

This equality simplifies the averaged spring energy to

Espring=12Tr𝐖l2=i=1N1wi22.\left\langle E_{\rm spring}\right\rangle=\dfrac{1}{2}{\rm Tr}~{}\boldsymbol{\rm W}_{l}^{2}=\sum_{i=1}^{N-1}\dfrac{w_{i}^{2}}{2}. (115)

Energy of the mode ii is wi2/2w_{i}^{2}/2 accordingly. The denominator of 𝒮α\mathcal{S}_{\alpha} becomes

Tr[𝐏lT𝐊l𝐏l𝐍l]=Tr𝐍l.{\rm Tr}\left[\boldsymbol{\rm P}_{l}^{\rm T}\boldsymbol{\rm K}_{l}\boldsymbol{\rm P}_{l}\boldsymbol{\rm N}_{l}\right]={\rm Tr}~{}\boldsymbol{\rm N}_{l}. (116)

Moreover, we can introduce the normalization of the constants νi(i=1,,N)\nu_{i}~{}(i=1,\cdots,N) as

Tr𝐍l=i=1N1νi=1,{\rm Tr}~{}\boldsymbol{\rm N}_{l}=\sum_{i=1}^{N-1}\nu_{i}=1, (117)

since the scale of the spring energy is controled by w(t1)w(t_{1}). Thus, Espring\left\langle E_{\rm spring}\right\rangle is further simplified to

Espring=w(t1)22\left\langle E_{\rm spring}\right\rangle=\dfrac{w(t_{1})^{2}}{2} (118)

under the hypothesis (35), and the function 𝒮α\mathcal{S}_{\alpha} is to

𝒮α=Tr[𝐏lT𝐁~ll1yα𝐏l𝚲l𝐍l],\mathcal{S}_{\alpha}={\rm Tr}\left[\boldsymbol{\rm P}_{l}^{\rm T}\dfrac{\partial\widetilde{\boldsymbol{\rm B}}_{ll}^{-1}}{\partial y_{\alpha}}\boldsymbol{\rm P}_{l}\boldsymbol{\rm\Lambda}_{l}\boldsymbol{\rm N}_{l}\right], (119)

when we use 𝐐lO(N1)\boldsymbol{\rm Q}_{l}\in O(N-1) and the normalization (117).

Appendix D Stationarity and stability of one-dimensional conformations

We first note that

𝐀S(ϕ𝒞1)=𝐀Cϕi(ϕ𝒞1)=𝐎(i=1,,N1),\boldsymbol{\rm A}_{S}(\boldsymbol{\phi}\in\mathcal{C}^{1})=\dfrac{\partial\boldsymbol{\rm A}_{C}}{\partial\phi_{i}}(\boldsymbol{\phi}\in\mathcal{C}^{1})=\boldsymbol{\rm O}\quad(i=1,\cdots,N-1), (120)

because all the elements depend on sinϕi,j\sin\phi_{i,j} in 𝐀S\boldsymbol{\rm A}_{S} and 𝐀C/ϕi\partial\boldsymbol{\rm A}_{C}/\partial\phi_{i}, and sinϕi,j=0\sin\phi_{i,j}=0 for a conformation belonging to 𝒞1\mathcal{C}^{1}. This fact implies that

𝐁~ll1ϕi(ϕ𝒞1)=[𝐀Cϕi+(𝐀S𝐀C1𝐀S)ϕi]ϕ𝒞1=𝐎\dfrac{\partial\widetilde{\boldsymbol{\rm B}}_{ll}^{-1}}{\partial\phi_{i}}(\boldsymbol{\phi}\in\mathcal{C}^{1})=\left[\dfrac{\partial\boldsymbol{\rm A}_{C}}{\partial\phi_{i}}+\dfrac{\partial(\boldsymbol{\rm A}_{S}\boldsymbol{\rm A}_{C}^{-1}\boldsymbol{\rm A}_{S})}{\partial\phi_{i}}\right]_{\boldsymbol{\phi}\in\mathcal{C}^{1}}=\boldsymbol{\rm O} (121)

and

𝒯i(ϕ𝒞1)=0\mathcal{T}_{i}(\boldsymbol{\phi}\in\mathcal{C}^{1})=0 (122)

for i=1,,N1i=1,\cdots,N-1. Thus, we have 𝑮(ϕ𝒞1)=𝟎\boldsymbol{G}(\boldsymbol{\phi}\in\mathcal{C}^{1})=\boldsymbol{0} for Ubend(2)0U_{\rm bend}^{(2)}\equiv 0 from Eq. (48).

Similarly, the Jacobian matrix D𝑮D\boldsymbol{G} is simplified as

Giϕj(ϕ𝒞1)=E(2)2(Bϕϕ1)inTr[𝐏lT𝐘nj𝐏l𝚲l𝐍l]Tr[𝐏l𝐊l𝐏l𝐍l]\dfrac{\partial G_{i}}{\partial\phi_{j}}(\boldsymbol{\phi}\in\mathcal{C}^{1})=-\dfrac{E^{(2)}}{2}(B_{\phi\phi}^{-1})^{in}\dfrac{{\rm Tr}\left[\boldsymbol{\rm P}_{l}^{\rm T}\boldsymbol{\rm Y}_{nj}\boldsymbol{\rm P}_{l}\boldsymbol{\rm\Lambda}_{l}\boldsymbol{\rm N}_{l}\right]}{{\rm Tr}\left[\boldsymbol{\rm P}_{l}\boldsymbol{\rm K}_{l}\boldsymbol{\rm P}_{l}\boldsymbol{\rm N}_{l}\right]} (123)

where

𝐘ij=2𝐀Cϕiϕj+𝐀Sϕi𝐀C1𝐀Sϕj+𝐀Sϕj𝐀C1𝐀Sϕi.\boldsymbol{\rm Y}_{ij}=\dfrac{\partial^{2}\boldsymbol{\rm A}_{C}}{\partial\phi_{i}\partial\phi_{j}}+\dfrac{\partial\boldsymbol{\rm A}_{S}}{\partial\phi_{i}}\boldsymbol{\rm A}_{C}^{-1}\dfrac{\partial\boldsymbol{\rm A}_{S}}{\partial\phi_{j}}+\dfrac{\partial\boldsymbol{\rm A}_{S}}{\partial\phi_{j}}\boldsymbol{\rm A}_{C}^{-1}\dfrac{\partial\boldsymbol{\rm A}_{S}}{\partial\phi_{i}}. (124)

We remark that each of 𝐘ij(i,j=1,,N1)\boldsymbol{\rm Y}_{ij}~{}(i,j=1,\cdots,N-1) is a size-(N1)(N-1) matrix. Further, the matrix 𝐁~ll1\widetilde{\boldsymbol{\rm B}}_{ll}^{-1} is also simplified to

𝐁~ll1(ϕ𝒞1)=𝐀C(ϕ𝒞1),\widetilde{\boldsymbol{\rm B}}_{ll}^{-1}(\boldsymbol{\phi}\in\mathcal{C}^{1})=\boldsymbol{\rm A}_{C}(\boldsymbol{\phi}\in\mathcal{C}^{1}), (125)

and the matrices 𝚲l\boldsymbol{\rm\Lambda}_{l} and 𝐐l=𝐊l𝐏l\boldsymbol{\rm Q}_{l}=\sqrt{\boldsymbol{\rm K}_{l}}\boldsymbol{\rm P}_{l} are obtained from the eigenvalue problem

(𝐊𝐀C𝐊)𝐐l=𝐐l𝚲l.\left(\sqrt{\boldsymbol{\rm K}}\boldsymbol{\rm A}_{C}\sqrt{\boldsymbol{\rm K}}\right)\boldsymbol{\rm Q}_{l}=\boldsymbol{\rm Q}_{l}\boldsymbol{\rm\Lambda}_{l}. (126)

Note that we can choose 𝐐l\boldsymbol{\rm Q}_{l} from O(N1)O(N-1).

Appendix E Eigenvalues of the linearized equations

We consider the eigensystem of the matrix

𝐉=(𝐎𝐄𝐙𝐎),\boldsymbol{\rm J}=\begin{pmatrix}\boldsymbol{\rm O}&\boldsymbol{\rm E}\\ -\boldsymbol{\rm Z}&\boldsymbol{\rm O}\\ \end{pmatrix}, (127)

where all the submatrices are of size-nn. We assume that 𝐙\boldsymbol{\rm Z} is diagonalizable. Suppose that the nn eigenvalues of 𝐙\boldsymbol{\rm Z} are real, nonzero, and denoted by ziz_{i}. The associated nn eigenvectors are 𝒗i\boldsymbol{v}_{i} satisfying

𝐙𝒗i=zi𝒗i,(i=1,,n).\boldsymbol{\rm Z}\boldsymbol{v}_{i}=z_{i}\boldsymbol{v}_{i},\quad(i=1,\cdots,n). (128)

It is straightforward to show that the matrix 𝐉\boldsymbol{\rm J} has the eigenvalues ±zi\pm\sqrt{-z_{i}} and the associated eigenvectors 𝒗i±\boldsymbol{v}_{i}^{\pm} defined by

𝒗i±=(𝒗i±zi𝒗i).\boldsymbol{v}_{i}^{\pm}=\begin{pmatrix}\boldsymbol{v}_{i}\\ \pm\sqrt{-z_{i}}\boldsymbol{v}_{i}\\ \end{pmatrix}. (129)

Appendix F Dynamical stability of one-dimensional conformations by mixed modes

Refer to caption
Figure 10: Stable (purple circles) and unstable (green crosses) regions on the parameter plane (ν1,ν2)(\nu_{1},\nu_{2}) for N=4N=4. All the one-dimensional conformations share this diagram. The parameter ν3\nu_{3} is determined by ν3=1ν1ν2\nu_{3}=1-\nu_{1}-\nu_{2}. The critical point on the line ν2=0\nu_{2}=0 is ν1c0.1464466\nu_{1{\rm c}}\simeq 0.1464466.

We study stability of a one-dimensional conformation with exciting multiple modes under the condition of the equal masses and the identical springs expressed in Eq. (56). The normal mode energy ratios νi(i=1,,N1)\nu_{i}~{}(i=1,\cdots,N-1) are set as

Tr𝐍=i=1N1νi=1,0νi1,{\rm Tr}~{}\boldsymbol{\rm N}=\sum_{i=1}^{N-1}\nu_{i}=1,\qquad 0\leq\nu_{i}\leq 1, (130)

and the number of the free parameters is N2N-2. We compute NN dependence of the stable probability ps(N)p_{\rm s}(N) with which a considering one-dimensional conformation is stabilized.

A necessary and sufficient condition of the stability for N=3N=3 is

the conformation is stable1/4<ν11\text{the conformation is stable}\quad\Longleftrightarrow\quad 1/4<\nu_{1}\leq 1 (131)

for the conformations 𝒄=(1)\boldsymbol{c}=(1) (straight) and 𝒄=(1)\boldsymbol{c}=(-1) (bent). The condition implies that the probability is

ps(3)=0.75.p_{\rm s}(3)=0.75. (132)

This stable probability for the two conformations is not a contradiction, because multi-stability of the two conformations is realized in the interval

1/4<ν1<3/4.1/4<\nu_{1}<3/4. (133)

The condition of Eq. (131) is in agreement with the conclusion reported previously yamaguchi-etal-22 . The agreement suggests that stability can be obtained by the current theory, although it does not reduce the rotational symmetry while the previous theory does.

For N=4N=4, we performed numerical computations of stability at the lattice points (n1/100,n2/100)(n1,n2=0,,100)(n_{1}/100,n_{2}/100)~{}(n_{1},n_{2}=0,\cdots,100) on the parameter plane (ν1,ν2)(\nu_{1},\nu_{2}), where ν3\nu_{3} is determined from Eq. (130). The stable and the unstable regions are reported in Fig. 10, and we have the straight line boundary. The critical value ν1c\nu_{1{\rm c}} on the line ν2=0\nu_{2}=0 is in the interval [0.1464466,0.1464467][0.1464466,0.1464467], and the stable probability is

ps(4)0.8535.p_{\rm s}(4)\simeq 0.8535. (134)

The stability check on the lattice points is also performed on the parameter space (ν1,ν2,ν3)(\nu_{1},\nu_{2},\nu_{3}) for N=5N=5. Among all the researched points of 176851176851, the six conformations are stable at 141019141019 points. Thus, the stable probability is

ps(5)0.7974.p_{\rm s}(5)\simeq 0.7974. (135)

The probabilities ps(3),ps(4)p_{\rm s}(3),p_{\rm s}(4), and ps(5)p_{\rm s}(5) suggest that dynamical stability is important even if the system size is large and multiple modes are excited.

Appendix G Effective potential for N=3N=3

For N=3N=3, the number of the bending angles is two, and the second bending angle ϕ2\phi_{2} is not essential, since it is a cyclic coordinate associating with the total angular momentum. Reducing ϕ2\phi_{2}, we can construct the effective potential, which describes the slow bending motion of ϕ1\phi_{1}. For simplicity, we denote ϕ1\phi_{1} by ϕ\phi. The construction is performed in three steps.

First, instead of (ν1,ν2)(\nu_{1},\nu_{2}) defined in the ascending order of the eigenfrequencies (see the main text), we use (νin,νanti)(\nu_{\rm in},\nu_{\rm anti}), where νin\nu_{\rm in} (νanti\nu_{\rm anti}) represents the in-phase (antiphase) mode energy ratio. This change of ν\nu helps to construct the global effective potential.

Second, we consider the bending potential consisting of the interaction between the first and the third beads only. We set Ubend(2)=ULJ(2)U_{\rm bend}^{(2)}=U_{\rm LJ}^{(2)}, and the second-order Lennard-Jones potential ULJ(2)U_{\rm LJ}^{(2)} is given in Eq. (68). We set ϵ0=1,σ=1\epsilon_{0}=1,\sigma=1, and l=1l_{\ast}=1. The graph of ULJ(2)U_{\rm LJ}^{(2)} is reported in Figs. (5) and 11(a).

Third, we construct the effective potential following the previous result yamaguchi-etal-22 . The bending motion is described by the effective Lagrangian

Leff(ϕ,dϕdt1)=12Meff(ϕ)(dϕdt1)2Ueff(ϕ).L_{\rm eff}\left(\phi,\dfrac{{\rm d}\phi}{{\rm d}t_{1}}\right)=\dfrac{1}{2}M_{\rm eff}(\phi)\left(\dfrac{{\rm d}\phi}{{\rm d}t_{1}}\right)^{2}-U_{\rm eff}(\phi). (136)

The effective mass MeffM_{\rm eff} is

Meff(ϕ)=exp[20ϕF(ϕ)𝑑ϕ]M_{\rm eff}(\phi)=\exp\left[2\int_{0}^{\phi}F(\phi^{\prime})d\phi^{\prime}\right] (137)

and the effective potential UeffU_{\rm eff} is

Ueff(ϕ)=0ϕMeff(ϕ)G(ϕ)𝑑ϕ.U_{\rm eff}(\phi)=\int_{0}^{\phi}M_{\rm eff}(\phi^{\prime})G(\phi^{\prime})d\phi^{\prime}. (138)

The functions FF and GG are defined by

F(ϕ)=12C(ϕ)Cϕ(ϕ)+14𝒯(ϕ),F(\phi)=\dfrac{1}{2C(\phi)}\dfrac{\partial C}{\partial\phi}(\phi)+\dfrac{1}{4}\mathcal{T}(\phi), (139)

and

G(ϕ)=1C[dUbend(2)dϕ(ϕ)E(2)Ubend(2)(ϕ)2𝒯(ϕ)],G(\phi)=\dfrac{1}{C}\left[\dfrac{{\rm d}U_{\rm bend}^{(2)}}{{\rm d}\phi}(\phi)-\dfrac{E^{(2)}-U_{\rm bend}^{(2)}(\phi)}{2}\mathcal{T}(\phi)\right], (140)

where

C(ϕ)=ml26(2cosϕ),C(\phi)=\dfrac{ml_{\ast}^{2}}{6}(2-\cos\phi), (141)

and

𝒯(ϕ)=(νin2cosϕ+νanti2+cosϕ)sinϕ.\mathcal{T}(\phi)=\left(-\dfrac{\nu_{\rm in}}{2-\cos\phi}+\dfrac{\nu_{\rm anti}}{2+\cos\phi}\right)\sin\phi. (142)

We can see that the effective potential depends on the bending potential Ubend(2)U_{\rm bend}^{(2)}, the mode energy distribution (νin,νanti)(\nu_{\rm in},\nu_{\rm anti}), and energy E(2)E^{(2)}. Examples of the effective potential are exhibited in Fig. 11 for three pairs of (νin,νanti)(\nu_{\rm in},\nu_{\rm anti}) with varying E(2)E^{(2)}. Excitation of the in-phase mode stabilizes the straight conformation (ϕ=0\phi=0), and the antiphase mode enhances the stability around the minimum points ±ϕmin\pm\phi_{\rm min} of the Lennard-Jones potential as E(2)E^{(2)} increases.

Refer to caption
Figure 11: (a) The Lennard-Jones potential ULJ(2)U_{\rm LJ}^{(2)} as a function of the bending angle ϕ\phi. (b) The effective potential Ueff(ϕ)U_{\rm eff}(\phi) with (νin,νanti)=(1,0)(\nu_{\rm in},\nu_{\rm anti})=(1,0). (c) The effective potential Ueff(ϕ)U_{\rm eff}(\phi) with (νin,νanti)=(1/2,1/2)(\nu_{\rm in},\nu_{\rm anti})=(1/2,1/2). (d) The effective potential Ueff(ϕ)U_{\rm eff}(\phi) with (νin,νanti)=(0,1)(\nu_{\rm in},\nu_{\rm anti})=(0,1). The numbers in the panels (b)-(d) represent the values of E(2)E^{(2)}. N=3N=3, ϵ0=1\epsilon_{0}=1, σ=1\sigma=1, and l=1l_{\ast}=1. Graphs are vertically shifted for a graphical reason. The two vertical lines in each panel mark the minimum points ±ϕmin\pm\phi_{\rm min} of the Lennard-Jones potential.

Appendix H Initial positions of the beads

We start from a one-dimensional conformation symbolized by 𝒄=(c1,,cN2)\boldsymbol{c}=(c_{1},\cdots,c_{N-2}). An initial position with the bending potential UbendU_{\rm bend} is prepared by replacing the bending angle ϕ=π\phi=\pi with ϕ=±ϕmin\phi=\pm\phi_{\rm min}, where ϕmin\phi_{\rm min} is a bottom position of the Lennard-Jones potential [see Eq. (69) and Fig. 5], and by approximately exciting normal modes of the one-dimensional conformation 𝒄\boldsymbol{c}.

To describe the initial positions of the beads, we introduce the direction vectors 𝒅i(i=1,,N1)\boldsymbol{d}_{i}~{}(i=1,\cdots,N-1), where

𝒅1=(1,0)T.\boldsymbol{d}_{1}=(1,0)^{\rm T}. (143)

We define 𝒅i\boldsymbol{d}_{i} as

𝒅i=R(ρi)𝒅i1,(i=2,,N1)\boldsymbol{d}_{i}=R(\rho_{i})\boldsymbol{d}_{i-1},\quad(i=2,\cdots,N-1) (144)

where R(ρ)R(\rho) is the rotation matrix of the angle ρ\rho,

R(ρ)=(cosρsinρsinρcosρ).R(\rho)=\begin{pmatrix}\cos\rho&-\sin\rho\\ \sin\rho&\cos\rho\\ \end{pmatrix}. (145)

The angles ρi\rho_{i} are determined as

ρi={0(ci=1),ϕminj=1icj(ci=1),\rho_{i}=\left\{\begin{array}[]{ll}0&(c_{i}=1),\\ -\phi_{\rm min}\prod_{j=1}^{i}c_{j}&(c_{i}=-1),\\ \end{array}\right. (146)

which replace the bending angle ϕ=π\phi=\pi with ϕ=±ϕmin\phi=\pm\phi_{\rm min}. Let us denote the initial length of the iith spring by li,0l_{i,0}, which is determined later. The initial position of the iith bead 𝒓i,0\boldsymbol{r}_{i,0} is

𝒓1,0=𝟎,𝒓2,0=l1,0𝒅1,\boldsymbol{r}_{1,0}=\boldsymbol{0},\quad\boldsymbol{r}_{2,0}=l_{1,0}\boldsymbol{d}_{1}, (147)

and

𝒓i,0=𝒓i1,0+li1,0𝒅i1,(i=3,,N).\boldsymbol{r}_{i,0}=\boldsymbol{r}_{i-1,0}+l_{i-1,0}\boldsymbol{d}_{i-1},\quad(i=3,\cdots,N). (148)

The lengths of the springs are decided to excite normal modes in a desired manner approximately. As discussed in Appendix C.4, the matrix 𝐏l\boldsymbol{\rm P}_{l} can be constructed as 𝐏l=(𝐊l)1𝐐l\boldsymbol{\rm P}_{l}=(\sqrt{\boldsymbol{\rm K}_{l}})^{-1}\boldsymbol{\rm Q}_{l}, where 𝐐lO(N1)\boldsymbol{\rm Q}_{l}\in O(N-1) solves the eigenvalue problem (113). For simplicity, we use 𝐁~ll\widetilde{\boldsymbol{\rm B}}_{ll} defined for the one-dimensional conformation 𝒄\boldsymbol{c}. Since the jjth column vector of 𝐏l\boldsymbol{\rm P}_{l} represents the jjth normal mode of the springs of the one-dimensional conformation 𝒄\boldsymbol{c}, we create the vector

𝒑^=𝐏l(ν1νN1).\hat{\boldsymbol{p}}=\boldsymbol{\rm P}_{l}\begin{pmatrix}\sqrt{\nu_{1}}\\ \vdots\\ \sqrt{\nu_{N-1}}\end{pmatrix}. (149)

The lengths of the springs are determined as

𝒍=𝒍+ϵ𝒑^.\boldsymbol{l}=\boldsymbol{l}_{\ast}+\epsilon\hat{\boldsymbol{p}}. (150)

References

  • (1) A. Stephenson, XX. On induced stability, The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 15, 233 (1908).
  • (2) P. L. Kapitza, Dynamic stability of a pendulum when its point of suspension vibrates, Soviet Phys. JETP 21, 588 (1951); Collected papers of P. L. Kapitza, Vol.2, pp.714–737 (1965).
  • (3) M. Bukov, L. D’Alessio, and A. Polkovnikov, Universal high-frequency behavior of periodically driven systems: from dynamical stabilization to Floquet engineering, Advances in Physics 64, 139 (2015).
  • (4) M. Grifoni and P. Hänggi, Coherent and incoherent quantum stochastic resonance, Phys. Rev. Lett. 76, 1611 (1996).
  • (5) A. Wickenbrock, P. C. Holz, N. A. Abdul Wahab, P. Phoonthong, D. Cubero, and F. Renzoni, Vibrational mechanics in an optical lattice: Controlling transport via potential renormalization, Phys. Rev. Lett. 108, 020603 (2012).
  • (6) V. N. Chizhevsky, E. Smeu, and G. Giacomelli, Experimental evidence of “vibrational resonance” in an optical system, Phys. Rev. Lett. 91, 220602 (2003).
  • (7) V. N. Chizhevsky, Experimental evidence of vibrational resonance in a multistable system, Phys. Rev. E 89, 062914 (2014).
  • (8) D. Cubero, J. P. Baltanas, and J. Casado-Pascual, High-frequency effects in the Fitzhugh-Nagumo neuron model, Phys. Rev. E 73, 061102 (2006).
  • (9) M. Bordet and S. Morfu, Experimental and numerical study of noise effects in a FitzHugh–Nagumo system driven by a biharmonic signal, Chaos, Solitons & Fractals 54, 82 (2013).
  • (10) S. H. Weinberg, High frequency stimulation of cardiac myocytes: A theoretical and computational study, Chaos 24, 043104 (2014).
  • (11) M. Uzuntarla, E. Yilmaz, A. Wagemakers, and M. Ozer, Vibrational resonance in a heterogeneous scale free network of neurons, Commun. Nonlinear Sci. Numer. Simulat. 22, 367 (2015).
  • (12) R. H. Buchanan, G. Jameson, and D. Oedjoe, Cyclic migration of bubbles in vertically vibrating liquid columns, Ind. Eng. Chem. Fund. 1, 82 (1962).
  • (13) M. H. I. Baird, Resonant bubbles in a vertically vibrating liquid column, Can. J. Chem. Eng. 41, 52 (1963).
  • (14) G. J. Jameson, The motion of a bubble in a vertically oscillating viscous liquid, Chem. Eng. Sci. 21, 35 (1966).
  • (15) B. Apffel, F. Novkoski, A. Eddi, and E. Fort, Floating under a levitating liquid, Nature 585, 48 (2020).
  • (16) R. E. Bellman, J. Bentsman, and S. M. Meerkov, Vibrational control of nonlinear systems, IEEE Trans. Automat. Contr. 31, 710 (1986).
  • (17) B. Shapiro and B. T. Zinn, High-frequency nonlinear vibrational control, IEEE Trans. Automat. Contr. 42, 83 (1997).
  • (18) M. Borromeo and F. Marchesoni, Artificial sieves for quasimassless particles, Phys. Rev. Lett. 99, 150605 (2007).
  • (19) C. J. Richards, T. J. Smart, P. H. Jones, and D. Cubero, A microscopic Kapitza pendulum, Scientific Reports 8, 13107 (2018).
  • (20) T. Yanagita and T. Konishi, Numerical analysis of new oscillatory mode of bead-spring model, Journal of JSCE A2 75, I_125 (2019) (in Japanese).
  • (21) Y. Y. Yamaguchi, T. Yanagita, T. Konishi, and M. Toda, Dynamically induced conformation depending on excited normal modes of fast oscillation, Phys. Rev. E 105, 064201 (2022).
  • (22) C. M. Bender and S. A. Orszag, Advanced Mathematical Methods for Scientists and Engineers I: Asymptotic Methods and Perturbation Theory (Springer, 1999).
  • (23) N. M. Krylov and N. N. Bogoliubov, New Methods of Nonlinear Mechanics in their Application to the Investigation of the Operation of Electronic Generators. I (United Scientific and Technical Press, Moscow, 1934).
  • (24) N. M. Krylov and N. N. Bogoliubov, Introduction to Nonlinear Mechanics (Princeton University Press, Princeton, 1947).
  • (25) J. Guckenheimer and P. Holmes, Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vectgor Fields (Springer-Verlag, New York, 1983).
  • (26) B. C. Dian, A. Longarte, T. S. Zwier, Conformational dynamics in a dipeptide after single-mode vibrational excitation, Nature 296, 2369 (2002).
  • (27) O. K. Rice and H. C. Ramsperger, Theories of unimolecular gas reactions at low pressures, J. Am. Chem. Soc. 49, 1617 (1927).
  • (28) L. S. Kassel, Studies in homogeneous gas reactions I, J. Phys. Chem. 32 225 (1928).
  • (29) R. A. Marcus, Unimolecular dissociations and free radical recombination reactions, J. Chem. Phys. 20, 359 (1952).
  • (30) T. Yanao, W. S. Koon, J. E. Marsden, and I. G. Kevrekidis, Gyration-radius dynamics in structural transitions of atomic clusters, J. Chem. Phys. 126, 124102 (2007).
  • (31) H. Yoshida, Construction of higher order symlectic integrators, Phys. Lett. A 190, 262 (1990).
  • (32) D. D. Holm, J. E. Marsden, T. Ratiu, and A. Weinstein, Nonlinear stability of fluid and plasma equilibria, Phys. Rep. 123, 1 (1985).
  • (33) L. Boltzmann, On certain questions of the theory of gases, Nature 51, 413 (1895).
  • (34) J. H. Jeans, On the vibrations set up in molecules by collisions, Philos. Mag. 6, 279 (1903).
  • (35) J. H. Jeans, XI. On the partition of energy between matter and aether, Philos. Mag. 10, 91 (1905).
  • (36) L. Landau and E. Teller, On the theory of sound dispersion, Physik. Z. Sowjetaunion 10, 34 (1936), in Collected Papers of L. D. Landau, edited by D. Ter Haar (Pergamon Press, Oxford, 1965), pp. 147–153.
  • (37) G. Benettin, L. Galgani, and A. Giorgilli, Realization of holonomic constraints and freezing of high-frequency degrees of freedom in the light of classical perturbation-theory. Part II, Commun. Math. Phys. 121, 557 (1989).
  • (38) O. Baldan and G. Benettin, Classical “freezing” of fast rotations. A numerical test of the Boltzmann-Jeans conjecture, J. Stat. Phys. 62, 201 (1991).