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

Dynamically induced conformation depending on excited normal modes of fast oscillation

Yoshiyuki Y. Yamaguchi [email protected] Department of Applied Mathematics and Physics, Graduate School of Informatics, Kyoto University, Kyoto 606-8501, Japan    Tatsuo Yanagita Department of Engineering Science, Osaka Electro-Communication University, Neyagawa 572-8530, Japan    Tetsuro Konishi General Education Division, College of Engineering, Chubu University, Kasugai 487-8501, Japan    Mikito Toda Faculty Division of Natural Sciences, Research Group of Physics, Nara Women’s University, Kita-Uoya-Nishimachi, Nara 630-8506, Japan Present position: Research fellow, Nara Women’s University Graduate School of Information Science , University of Hyogo, 7-1-28 Minatojima-minamimachi, Chuo-ku, Kobe, Hyogo 650-0047, Japan Research Institute for Electronic Science, Hokkaido University, Kita 20 Nishi 10, Kita-Ku, Sapporo 001-0020, Japan
Abstract

We present dynamical effects on conformation in a simple bead-spring model consisting of three beads connected by two stiff springs. The conformation defined by the bending angle between the two springs is determined not only by a given potential energy function depending on the bending angle, but also fast motion of the springs which constructs the effective potential. A conformation corresponding with a local minimum of the effective potential is hence called the dynamically induced conformation. We develop a theory to derive the effective potential by using multiple-scale analysis and the averaging method. A remarkable consequence is that the effective potential depends on the excited normal modes of the springs and amount of the spring energy. Efficiency of the obtained effective potential is numerically verified.

I Introduction

Conformation is deeply connected with function. A typical example is a biomolecule whose conformation is crucial for binding a ligand koshland-58 ; monod-wyman-changeux-65 ; okazaki-takada-08 ; seeliger-degroot-10 ; fuchigami-etal-11 . Morphological computation hauser-etal-11 ; special-issue-13 ; muller-hoffmann-17 is another example, which can be found for instance as walking robots collins-ruina-tedrake-wisse-05 ; hermans-schrauwen-bienstman-dambre-14 . Mechanical metamaterial mechanical-metamaterials-19 provides several examples like the Miura fold which exhibits negative Poisson’s ratio wei-etal-13 .

Realization of conformations is usually associated with the minimum of a potential energy function. In addition to the potential function, dynamics sometimes contributes to construct an effective potential. A well-known example is the Kapitza pendulum stephenson-08 ; kapitza-51 : An inverted pendulum persists against the gravity by applying a rapidly oscillating external force, since the effective potential provides a local minimum at the inverted position.

We present another dynamical effect on conformation realized in autonomous Hamiltonian systems containing fast and slow motion. Consider a bead-spring model rouse-53 consisting of three beads connected by two stiff springs. The conformation of this system can be identified with the bending angle between the two springs. If the system has a bending potential which depends only on the bending angle, one may imagine that the bending angle goes to a local minimum of the bending potential. The conformation of this system is, however, determined by the effective potential which consists of the bending potential and contribution from the fast spring motion.

More precise description of the above phenomenon is as follows. First of all, the dynamical effect is comparable with the bending potential under the condition that large bending motion is sufficiently slow than the spring motion. The bending potential dominates the effective potential when the spring energy is sufficiently small. However, the dynamical effect enlarges as the spring energy increases, and a local minimum of the effective potential does not necessarily coincide with a local minimum of the bending potential. We call a conformation corresponding to a local minimum of the effective potential a dynamically induced conformation (DIC). Further interesting fact is that the effective potential depends on the excited normal modes of the springs in addition to the spring energy. The three-body bead-spring model has the two normal modes of the springs and each mode makes a different valley.

The three-body system is quite simple, and hence it is theoretically tractable and clearly shows DIC. The aim of this paper is to present the dynamical contribution to conformation in the three-body system. It is worth noting that the bead-spring model mimics several systems: a polymer rouse-53 ; zimm-56 , a microscopic artificial swimmer gauger-stark-06 , a soft magnetic nanowire mirzae-etal-20 , and a semi-flexible macromolecule saadat-khomami-16 .

A theoretical analysis reveals that the essence of DIC is existence of multiple timescales, which is realized in the bead-spring model by stiff springs and slow bending motion. Appearance of multiple timescales is generic in nature. For instance, biomolecules have several forces of diverse strength as strong covalent bonds, intermediate hydrogen bonds, and weak van der Waals forces, and each of them has a characteristic timescale. DIC therefore enriches understanding of the origin of conformation change and its function.

A celebrated example of the dynamically constructed effective potential is found in the aforementioned Kapitza pendulum, and it has been studied in a wide range of fields 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 . The development of the Kapitza pendulum suggests that DIC will provide a large spectrum of applications. Nevertheless, we underline three crucial differences between DIC and the Kapitza pendulum. (i) The bead-spring model is autonomous and the effective potential is intrinsically determined, while one in the Kapitza pendulum can be controlled by the applied external force. (ii) The effective potential depends on the excited normal modes in the bead-spring model. (iii) The local minimum points of the effective potential may continuously move depending on the spring energy in the bead-spring model, while the local minimum created by the external vertical oscillation is fixed at the inverted position in the Kapitza pendulum.

This paper is organized as follows. The three-body bead-spring model is presented in Sec. II with the two important assumptions to have DIC. Following the assumptions, we develop in Sec. III a theory to describe slow bending motion by using a multiple-scale analysis bender-orszag-99 and the averaging method krylov-bogoliubov-34 ; krylov-bogoliubov-47 ; guckenheimer-holmes-83 . The theory provides the effective potential depending on the excited normal modes of the springs and amount of the spring energy. Examples of the effective potential are exhibited in Sec. IV so as to reveal the above dependency. Efficiency of the effective potential is examined through numerical simulations in Sec. V. The final section VI is devoted to summary and discussions.

II Model

The three-body bead-spring model is sketched in Fig. 1. We assume that the beads move on a two-dimensional plane. The mass and the position of the jjth bead are respectively denoted by mjm_{j} and 𝒓j2\boldsymbol{r}_{j}\in\mathbb{R}^{2}. The Lagrangian of the model is expressed by

L=12j=13mj𝒓˙j2V(𝒓1,𝒓2,𝒓3),L=\dfrac{1}{2}\sum_{j=1}^{3}m_{j}||\dot{\boldsymbol{r}}_{j}||^{2}-V(\boldsymbol{r}_{1},\boldsymbol{r}_{2},\boldsymbol{r}_{3}), (1)

where 𝒓˙j:=d𝒓j/dt\dot{\boldsymbol{r}}_{j}:=\mathrm{d}\boldsymbol{r}_{j}/\mathrm{d}t and ||||||\cdot|| is the Euclidean norm: 𝒓=x2+y2||\boldsymbol{r}||=\sqrt{x^{2}+y^{2}} for 𝒓=(x,y)2\boldsymbol{r}=(x,y)\in\mathbb{R}^{2}. The jjth and the (j+1)(j+1)th beads are connected by a stiff spring, and we assume that the two springs have the identical potential. Further, for simplicity, we focus on the symmetric masses: m1=m3=mm_{1}=m_{3}=m. The term VV represents the potential energy function, which will be specified later.

We assume that the system described by Eq. (1) has the two-dimensional translational symmetry and the rotational symmetry, which induce the conservation of the two-dimensional total momentum vector and of the total angular momentum, respectively. The total momentum vector can be set as the zero vector without loss of generality, while the total angular momentum is assumed to be zero. The three integrals reduces the system and the reduced Lagrangian is

L=12α,β=13Cαβ(𝒚)y˙αy˙βV(𝒚),L=\dfrac{1}{2}\sum_{\alpha,\beta=1}^{3}C^{\alpha\beta}(\boldsymbol{y})\dot{y}_{\alpha}\dot{y}_{\beta}-V(\boldsymbol{y}), (2)

where

𝒚=(y1,y2,y3)T=(l1,l2,ϕ)T\boldsymbol{y}=(y_{1},y_{2},y_{3})^{\rm T}=(l_{1},l_{2},\phi)^{\rm T} (3)

and the superscript T represents transposition. The variables l1l_{1} and l2l_{2} are the lengths of the two springs,

lj=𝒓j+1𝒓j,(j=1,2)l_{j}=||\boldsymbol{r}_{j+1}-\boldsymbol{r}_{j}||,\quad(j=1,2) (4)

and ϕ\phi is the bending angle,

cosϕ=(𝒓3𝒓2)(𝒓2𝒓1)𝒓3𝒓2𝒓2𝒓1,\cos\phi=\dfrac{(\boldsymbol{r}_{3}-\boldsymbol{r}_{2})\cdot(\boldsymbol{r}_{2}-\boldsymbol{r}_{1})}{||\boldsymbol{r}_{3}-\boldsymbol{r}_{2}||||\boldsymbol{r}_{2}-\boldsymbol{r}_{1}||}, (5)

where \cdot is the Euclidean inner product. The function Cαβ(𝒚)C^{\alpha\beta}(\boldsymbol{y}) is the (α,β)(\alpha,\beta) element of the size-33 symmetric matrix 𝐂(𝒚)\boldsymbol{\rm C}(\boldsymbol{y}), whose explicit form is given in Appendix A. The potential energy function V(𝒚)V(\boldsymbol{y}) consists of the two parts as

V(𝒚)=Vspring(l1,l2)+Vbend(ϕ).V(\boldsymbol{y})=V_{\rm spring}(l_{1},l_{2})+V_{\rm bend}(\phi). (6)

We call VspringV_{\rm spring} and VbendV_{\rm bend} the spring potential and the bending potential, respectively.

Refer to caption
Figure 1: The three-body bead-spring model. We assume the symmetric masses m1=m3=mm_{1}=m_{3}=m and the two springs having the idential potential.

We introduce the two assumptions to realize DIC in the above model. Let ϵ\epsilon be a dimensionless small parameter as |ϵ|1|\epsilon|\ll 1. The two assumptions are:

  • (A1)

    The amplitudes of the springs are sufficiently small comparing with the natural length. 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).

These two assumptions lead the effective potential for the bending angle ϕ\phi. A local minimum point of the effective potential does not necessarily coincide with a local minimum point of the bending potential Vbend(ϕ)V_{\rm bend}(\phi). That is, the bending angle oscillates around an angle at which the bending potential Vbend(ϕ)V_{\rm bend}(\phi) does not take a local minimum. We develop a theory to derive the effective potential in Sec.III.

III Theory

From now on, we use the Einstein notation for the sum: We take the sum over an index if it appears twice in a term. We derive the equation of motion for the slow bending motion, and construct the effective potential induced by the fast spring motion. A review of the Kapitza pendulum is provided in Appendix B, which might be helpful to understand the theory.

III.1 Multiscale analysis and averaging

The Euler-Lagrange equations derived from Eq. (2) are

Cαβ(𝒚)y¨β+[Cαβyγ(𝒚)12Cβγyα(𝒚)]y˙βy˙γ+Vyα(𝒚)=0,C^{\alpha\beta}(\boldsymbol{y})\ddot{y}_{\beta}+\left[\dfrac{\partial C^{\alpha\beta}}{\partial y_{\gamma}}(\boldsymbol{y})-\dfrac{1}{2}\dfrac{\partial C^{\beta\gamma}}{\partial y_{\alpha}}(\boldsymbol{y})\right]\dot{y}_{\beta}\dot{y}_{\gamma}+\dfrac{\partial V}{\partial y_{\alpha}}(\boldsymbol{y})=0, (7)

where α,β,γ{1,2,3}\alpha,\beta,\gamma\in\{1,2,3\}. These equations are the starting point of our theory.

The assumption (A2) induces the two timescales of t0=tt_{0}=t and t1=ϵtt_{1}=\epsilon t. The fast timescale t0t_{0} describes the fast spring motion, and the slow timescale t1t_{1} corresponds to the slow bending motion. The two timescales transform the time derivative into

ddt=t0+ϵt1.\dfrac{{\rm d}}{{\rm d}t}=\dfrac{\partial}{\partial t_{0}}+\epsilon\dfrac{\partial}{\partial t_{1}}. (8)

From the assumptions (A1) and (A2) the variables ljl_{j} and ϕ\phi are expanded as

{lj(t0,t1)=lj(0)+ϵlj(1)(t0,t1),lj(0)=l(j=1,2)ϕ(t0,t1)=ϕ(0)(t1)+ϵϕ(1)(t0,t1),\left\{\begin{split}&l_{j}(t_{0},t_{1})=l_{j}^{(0)}+\epsilon l_{j}^{(1)}(t_{0},t_{1}),\quad l_{j}^{(0)}=l_{\ast}\quad(j=1,2)\\ &\phi(t_{0},t_{1})=\phi^{(0)}(t_{1})+\epsilon\phi^{(1)}(t_{0},t_{1}),\end{split}\right. (9)

where ll_{\ast} is the natural length of the two springs. As we will observe later, the fast motion of ϕ(1)(t0,t1)\phi^{(1)}(t_{0},t_{1}) is induced by the fast motion of the springs and is of the same order O(ϵ)O(\epsilon) as the spring amplitudes. We are interested in ϕ(0)(t1)\phi^{(0)}(t_{1}), which represents large and slow bending motion. We denote the above expansions for simplicity 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}). (10)

We further expand the potential energy function VV. The spring potential VspringV_{\rm spring} is assumed to be expanded into the Taylor series around the natural length as

Vspring(l1,l2)=k2j=12(ljl)2+O(|ljl|3).V_{\rm spring}(l_{1},l_{2})=\dfrac{k}{2}\sum_{j=1}^{2}\left(l_{j}-l_{\ast}\right)^{2}+O(|l_{j}-l_{\ast}|^{3}). (11)

That is, the two springs have the same spring constant

k=Vspring2lj2(l,l)(j=1,2).k=\dfrac{\partial{}^{2}V_{\rm spring}}{\partial l_{j}^{2}}(l_{\ast},l_{\ast})\quad(j=1,2). (12)

The bending potential is assumed to be expanded into the series of ϵ\epsilon as

Vbend(ϕ)=Vbend(0)(ϕ)+ϵVbend(1)(ϕ)+ϵ2Vbend(2)(ϕ)+.V_{\rm bend}(\phi)=V_{\rm bend}^{(0)}(\phi)+\epsilon V_{\rm bend}^{(1)}(\phi)+\epsilon^{2}V_{\rm bend}^{(2)}(\phi)+\cdots. (13)

The two assumptions (A1) and (A2) induce

Vbend(0)(ϕ),Vbend(1)(ϕ)0V_{\rm bend}^{(0)}(\phi),~{}V_{\rm bend}^{(1)}(\phi)\equiv 0 (14)

as shown in Appendix C, and hence the leading term of VbendV_{\rm bend} is of O(ϵ2)O(\epsilon^{2}). This ording results from the assumption (A2): The force from the bending potential VbendV_{\rm bend} should be weaker than that of the spring potential VspringV_{\rm spring}.

We construct the equations of motion order by order, by substituting Eqs. (8), (10), (11), and (13) into Eq. (7). In O(ϵ0)O(\epsilon^{0}), we have no terms, because yβ(0)/t0=0\partial y_{\beta}^{(0)}/\partial t_{0}=0, y˙β,y¨β=O(ϵ)\dot{y}_{\beta},\ddot{y}_{\beta}=O(\epsilon), and V/yα=O(ϵ)\partial V/\partial y_{\alpha}=O(\epsilon).

In O(ϵ)O(\epsilon) we have

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)}. (15)

The size-33 matrix 𝐗\boldsymbol{\rm X} is defined by

𝐗(𝒚)=[𝐂(𝒚)]1𝐊,\boldsymbol{\rm X}(\boldsymbol{y})=[\boldsymbol{\rm C}(\boldsymbol{y})]^{-1}\boldsymbol{\rm K}, (16)

where

𝐊=(k000k0000).\boldsymbol{\rm K}=\begin{pmatrix}k&0&0\\ 0&k&0\\ 0&0&0\\ \end{pmatrix}. (17)

See Appendix D for the explicit form of 𝐗(𝒚(0))\boldsymbol{\rm X}(\boldsymbol{y}^{(0)}). The third column vector of 𝐗\boldsymbol{\rm X} is the zero vector, and the right-hand side of Eq. (15) has no contribution from the third element of 𝒚(1)\boldsymbol{y}^{(1)}, namely ϕ(1)\phi^{(1)}. The fast motion of ϕ(1)\phi^{(1)} is hence induced by l1(1)l_{1}^{(1)} and l2(1)l_{2}^{(1)}, as mentioned after Eq. (9).

The slow motion of ϕ(0)(t1)\phi^{(0)}(t_{1}) is described in O(ϵ2)O(\epsilon^{2}), and the equation of motion for ϕ(0)(t1)\phi^{(0)}(t_{1}) is

C3β(𝒚(0))(y¨β)(2)+[C3βyγ(𝒚(0))12Cβγϕ(𝒚(0))](y˙β)(1)(y˙γ)(1)+C3βyγ(𝒚(0))(y¨β)(1)yγ(1)+dVbend(2)dϕ(ϕ(0))=0.\begin{split}&C^{3\beta}(\boldsymbol{y}^{(0)})(\ddot{y}_{\beta})^{(2)}+\left[\dfrac{\partial C^{3\beta}}{\partial y_{\gamma}}(\boldsymbol{y}^{(0)})-\dfrac{1}{2}\dfrac{\partial C^{\beta\gamma}}{\partial\phi}(\boldsymbol{y}^{(0)})\right](\dot{y}_{\beta})^{(1)}(\dot{y}_{\gamma})^{(1)}\\ &+\dfrac{\partial C^{3\beta}}{\partial y_{\gamma}}(\boldsymbol{y}^{(0)})(\ddot{y}_{\beta})^{(1)}y_{\gamma}^{(1)}+\dfrac{{\rm d}V_{\rm bend}^{(2)}}{{\rm d}\phi}(\phi^{(0)})=0.\end{split} (18)

Here (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 explicit forms are

(𝒚˙)(1)=d𝒚(0)dt1+𝒚(1)t0,(𝒚¨)(1)=𝒚(1)2t02,(𝒚¨)(2)=d𝒚(0)2dt12+22𝒚(1)t0t1.\begin{split}&(\dot{\boldsymbol{y}})^{(1)}=\dfrac{{\rm d}\boldsymbol{y}^{(0)}}{{\rm d}t_{1}}+\dfrac{\partial\boldsymbol{y}^{(1)}}{\partial t_{0}},\\ &(\ddot{\boldsymbol{y}})^{(1)}=\dfrac{\partial{}^{2}\boldsymbol{y}^{(1)}}{\partial t_{0}^{2}},\\ &(\ddot{\boldsymbol{y}})^{(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} (19)

Equation (18) contains the fast oscillation of 𝒚(1)\boldsymbol{y}^{(1)}, and we eliminate it by taking the average over the fast timescale t0t_{0}. The average is defined by

φ(t1)=limT1T0Tφ(t0,t1)𝑑t0\left\langle\varphi\right\rangle(t_{1})=\lim_{T\to\infty}\dfrac{1}{T}\int_{0}^{T}\varphi(t_{0},t_{1})dt_{0} (20)

for an arbitrary function φ(t0,t1)\varphi(t_{0},t_{1}). After taking the average and recalling 𝒚(0)=(l,l,ϕ(0))\boldsymbol{y}^{(0)}=(l_{\ast},l_{\ast},\phi^{(0)}), Eq. (18) is simplified to

C33(𝒚(0))dϕ(0)2dt12+12C33ϕ(𝒚(0))(dϕ(0)dt1)2=dVbend(2)dϕ(ϕ(0))+𝒜.\begin{split}&C^{33}(\boldsymbol{y}^{(0)})\dfrac{{\rm d}{}^{2}\phi^{(0)}}{{\rm d}t_{1}^{2}}+\dfrac{1}{2}\dfrac{\partial C^{33}}{\partial\phi}(\boldsymbol{y}^{(0)})\left(\dfrac{{\rm d}\phi^{(0)}}{{\rm d}t_{1}}\right)^{2}\\ &=-\dfrac{{\rm d}V_{\rm bend}^{(2)}}{{\rm d}\phi}(\phi^{(0)})+\mathcal{A}.\end{split} (21)

The right-hand side represents the force, and the averaged term

𝒜=12Tr[𝐂ϕ(𝒚(0))𝒚(1)t0(𝒚(1)t0)T]\mathcal{A}=\dfrac{1}{2}{\rm Tr}\left[\dfrac{\partial\boldsymbol{\rm C}}{\partial\phi}(\boldsymbol{y}^{(0)})\left\langle\dfrac{\partial\boldsymbol{y}^{(1)}}{\partial t_{0}}\left(\dfrac{\partial\boldsymbol{y}^{(1)}}{\partial t_{0}}\right)^{\rm T}\right\rangle\right] (22)

represents the effective force yielded by the fast spring motion. Here Tr represents the matrix trace. The right-hand side of Eq. (22) depends on 𝒚(0)\boldsymbol{y}^{(0)} and 𝒚(1)\boldsymbol{y}^{(1)}. The 𝒚(0)\boldsymbol{y}^{(0)} dependence can be regarded as the ϕ(0)\phi^{(0)} dependence, since 𝒚(0)=(l,l,ϕ(0))\boldsymbol{y}^{(0)}=(l_{\ast},l_{\ast},\phi^{(0)}) and ll_{\ast} is constant. 𝒚(1)\boldsymbol{y}^{(1)} depends on t0t_{0} and t1t_{1}, and the t0t_{0} dependence is averaged out. We have to eliminate the t1t_{1} dependence to obtain the effective potential as a function of ϕ(0)\phi^{(0)}.

III.2 Explicit form of the averaged term

We compute the explicit form of the averaged term 𝒜\mathcal{A} by performing the diagonalization of Eq. (15), and observe the t1t_{1} dependence, which has to be eliminated. Let 𝐏\boldsymbol{\rm P} diagonalize 𝐗\boldsymbol{\rm X} 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)}). (23)

The diagonal matrix 𝚲(𝒚(0))\boldsymbol{\rm\Lambda}(\boldsymbol{y}^{(0)}) consists of the eigenvalues of 𝐗(𝒚(0))\boldsymbol{\rm X}(\boldsymbol{y}^{(0)}) and is denoted by

𝚲(𝒚(0))=(λ1(𝒚(0))000λ2(𝒚(0))0000),\boldsymbol{\rm\Lambda}(\boldsymbol{y}^{(0)})=\begin{pmatrix}\lambda_{1}(\boldsymbol{y}^{(0)})&0&0\\ 0&\lambda_{2}(\boldsymbol{y}^{(0)})&0\\ 0&0&0\\ \end{pmatrix}, (24)

where

λ1(𝒚(0))=k(M2M1cosϕ(0))M22M12,λ2(𝒚(0))=k(M2+M1cosϕ(0))M22M12,\begin{split}&\lambda_{1}(\boldsymbol{y}^{(0)})=\dfrac{k(M_{2}-M_{1}\cos\phi^{(0)})}{M_{2}^{2}-M_{1}^{2}},\\ &\lambda_{2}(\boldsymbol{y}^{(0)})=\dfrac{k(M_{2}+M_{1}\cos\phi^{(0)})}{M_{2}^{2}-M_{1}^{2}},\\ \end{split} (25)

and

M2=m(m+m2)2m+m2,M1=m22m+m2.M_{2}=\dfrac{m(m+m_{2})}{2m+m_{2}},\quad M_{1}=\dfrac{m^{2}}{2m+m_{2}}. (26)

A diagonalizing matrix is

𝐏(𝒚(0))=(𝒑in,𝒑anti,𝒑ϕ)=(1/21/201/21/20v(𝒚(0))01)\boldsymbol{\rm P}(\boldsymbol{y}^{(0)})=\begin{pmatrix}\boldsymbol{p}_{\rm in},&\boldsymbol{p}_{\rm anti},&\boldsymbol{p}_{\phi}\\ \end{pmatrix}=\begin{pmatrix}1/\sqrt{2}&1/\sqrt{2}&0\\ 1/\sqrt{2}&-1/\sqrt{2}&0\\ v(\boldsymbol{y}^{(0)})&0&1\\ \end{pmatrix} (27)

with

v(𝒚(0))=2lM1sinϕ(0)M2M1cosϕ(0).v(\boldsymbol{y}^{(0)})=\dfrac{\sqrt{2}}{l_{\ast}}\dfrac{M_{1}\sin\phi^{(0)}}{M_{2}-M_{1}\cos\phi^{(0)}}. (28)

The three column vectors 𝒑in,𝒑anti\boldsymbol{p}_{\rm in},\boldsymbol{p}_{\rm anti}, and 𝒑ϕ\boldsymbol{p}_{\phi} are eigenvectors of 𝐗(𝒚(0))\boldsymbol{\rm X}(\boldsymbol{y}^{(0)}), and we call the three modes as the in-phase mode, the anti-phase mode, and the zero-eigenvalue mode, respectively.

To solve Eq. (15), we perform the change of variables as

𝒚(1)=𝐏(𝒚(0))𝜼,\boldsymbol{y}^{(1)}=\boldsymbol{\rm P}(\boldsymbol{y}^{(0)})\boldsymbol{\eta}, (29)

and 𝜼\boldsymbol{\eta} solves the diagonalized equations

𝜼2t02=𝚲(𝒚(0))𝜼.\dfrac{\partial{}^{2}\boldsymbol{\eta}}{\partial t_{0}^{2}}=-\boldsymbol{\rm\Lambda}(\boldsymbol{y}^{(0)})\boldsymbol{\eta}. (30)

Denoting the amplitudes of the in-phase and the anti-phase modes by w1(t1)w_{1}(t_{1}) and w2(t1)w_{2}(t_{1}) respectively, which evolve in the slow timescale t1t_{1} through the coupling with ϕ(0)(t1)\phi^{(0)}(t_{1}), and setting the amplitude of the zero-eigenvalue mode as zero, we introduce the diagonal matrix

𝐖(t1)=(w1(t1)000w2(t1)0000).\boldsymbol{\rm W}(t_{1})=\begin{pmatrix}w_{1}(t_{1})&0&0\\ 0&w_{2}(t_{1})&0\\ 0&0&0\\ \end{pmatrix}. (31)

Putting all together and remembering that the average of the square of a sinusoidal function is 1/21/2, we have

𝒜(ϕ(0),w1,w2)=14Tr[𝐂ϕ(𝒚(0))𝐏𝚲𝐖2𝐏T]=k4[M1sinϕ(0)M2M1cosϕ(0)w12M1sinϕ(0)M2+M1cosϕ(0)w22].\begin{split}&\mathcal{A}(\phi^{(0)},w_{1},w_{2})=\dfrac{1}{4}{\rm Tr}\left[\dfrac{\partial\boldsymbol{\rm C}}{\partial\phi}(\boldsymbol{y}^{(0)})\boldsymbol{\rm P}\boldsymbol{\rm\Lambda}\boldsymbol{\rm W}^{2}\boldsymbol{\rm P}^{\rm T}\right]\\ &=-\dfrac{k}{4}\left[\dfrac{M_{1}\sin\phi^{(0)}}{M_{2}-M_{1}\cos\phi^{(0)}}w_{1}^{2}-\dfrac{M_{1}\sin\phi^{(0)}}{M_{2}+M_{1}\cos\phi^{(0)}}w_{2}^{2}\right].\end{split} (32)

The averaged term 𝒜\mathcal{A} contains the two evolving amplitudes w1(t1)w_{1}(t_{1}) and w2(t1)w_{2}(t_{1}). The untrivial evolution of the amplitudes differs from the Kapitza pendulum, which also contains the amplitude of the external oscillation but it is explicitly given. We have to eliminate the two unknown amplitudes from the averaged term 𝒜\mathcal{A} to obtain a closed equation for ϕ(0)(t1)\phi^{(0)}(t_{1}).

III.3 Hypothesis and energy conservation

The strategy to eliminate the two unknown amplitudes w1(t1)w_{1}(t_{1}) and w2(t1)w_{2}(t_{1}) is as follows. First, we introduce a hypothesis, which is inspired from the adiabatic invariant. The hypothesis reduces the number of unknown variables from two to one. Second, we eliminate the remaining unknown variable by using the energy conservation law.

The first step is the introduction of the hypothesis expressed by

(H)w1(t1)2=ν1w(t1)2,w2(t1)2=ν2w(t1)2,({\it H})\quad w_{1}(t_{1})^{2}=\nu_{1}w(t_{1})^{2},\quad w_{2}(t_{1})^{2}=\nu_{2}w(t_{1})^{2}, (33)

where ν1\nu_{1} and ν2\nu_{2} are constants satisfying

ν1+ν2=1.\nu_{1}+\nu_{2}=1. (34)

A physical interpretation of the hypothesis (H) is that the slow bending motion exchanges energy with the fast spring motion in proportion to its normal mode energy. Validity of the hypothesis (H) is examined in Appendix E. We note that the hypothesis should be valid if the modification of ϕ(0)\phi^{(0)} is sufficiently small. The unique unknown variable is now w(t1)w(t_{1}), while the constants ν1\nu_{1} and ν2\nu_{2} have been included in the equations of motion.

The second step is the energy conservation. The leading order of the total energy is of O(ϵ2)O(\epsilon^{2}) and we expand it as E=ϵ2E(2)+O(ϵ3)E=\epsilon^{2}E^{(2)}+O(\epsilon^{3}). The leading term is

E(2)=12Tr[𝐂(𝒚(0))(𝒚˙)(1)(𝒚˙)(1)T]+12Tr[𝐊𝒚(1)𝒚(1)T]+Vbend(2)(ϕ(0)).\begin{split}E^{(2)}&=\dfrac{1}{2}{\rm Tr}\left[\boldsymbol{\rm C}(\boldsymbol{y}^{(0)})(\dot{\boldsymbol{y}})^{(1)}(\dot{\boldsymbol{y}})^{(1){\rm T}}\right]\\ &+\dfrac{1}{2}{\rm Tr}\left[\boldsymbol{\rm K}\boldsymbol{y}^{(1)}\boldsymbol{y}^{(1){\rm T}}\right]+V_{\rm bend}^{(2)}(\phi^{(0)}).\end{split} (35)

Taking the average over t0t_{0}, we have

E(2)=12Tr[𝐂(𝒚(0))d𝒚(0)dt1(d𝒚(0)dt1)T]+k2Tr𝐖2+Vbend(2)(ϕ(0)).\begin{split}\left\langle E^{(2)}\right\rangle&=\dfrac{1}{2}{\rm Tr}\left[\boldsymbol{\rm C}(\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]\\ &+\dfrac{k}{2}{\rm Tr}\boldsymbol{\rm W}^{2}+V_{\rm bend}^{(2)}(\phi^{(0)}).\end{split} (36)

Substituting Eq. (33), the unique unknown variable ww is obtained as

kw(t1)2=2[E(2)Vbend(2)(ϕ(0))]C33(𝒚(0))(dϕ(0)dt1)2,kw(t_{1})^{2}=2\left[E^{(2)}-V_{\rm bend}^{(2)}(\phi^{(0)})\right]-C^{33}(\boldsymbol{y}^{(0)})\left(\dfrac{{\rm d}\phi^{(0)}}{{\rm d}t_{1}}\right)^{2}, (37)

where we denoted E(2)\left\langle E^{(2)}\right\rangle by E(2)E^{(2)} for simplicity. Finally, we eliminate the unknown amplitudes from the averaged term 𝒜\mathcal{A} represented in Eq. (32) by substituting Eqs. (33) and (37):

𝒜(ϕ(0))=[E(2)Vbend(2)(ϕ(0))214C33(𝒚(0))(dϕ(0)dt1)2]Tν,\begin{split}&\mathcal{A}(\phi^{(0)})=\left[\dfrac{E^{(2)}-V_{\rm bend}^{(2)}(\phi^{(0)})}{2}-\dfrac{1}{4}C^{33}(\boldsymbol{y}^{(0)})\left(\dfrac{{\rm d}\phi^{(0)}}{{\rm d}t_{1}}\right)^{2}\right]T_{\nu},\\ \end{split} (38)

where

Tν=[M1sinϕ(0)M2M1cosϕ(0)ν1M1sinϕ(0)M2+M1cosϕ(0)ν2].\begin{split}T_{\nu}&=-\left[\dfrac{M_{1}\sin\phi^{(0)}}{M_{2}-M_{1}\cos\phi^{(0)}}\nu_{1}-\dfrac{M_{1}\sin\phi^{(0)}}{M_{2}+M_{1}\cos\phi^{(0)}}\nu_{2}\right].\end{split} (39)

We underline that the averaged term 𝒜\mathcal{A} depends on the constants ν1,ν2\nu_{1},\nu_{2} and E(2)E^{(2)}.

III.4 Final result

Substituting Eq. (38) into Eq. (21), we have

dϕ(0)2dt12+Fν(ϕ(0))(dϕ(0)dt1)2+Gν(ϕ(0))=0\begin{split}&\dfrac{{\rm d}{}^{2}\phi^{(0)}}{{\rm d}t_{1}^{2}}+F_{\nu}(\phi^{(0)})\left(\dfrac{{\rm d}\phi^{(0)}}{{\rm d}t_{1}}\right)^{2}+G_{\nu}(\phi^{(0)})=0\\ \end{split} (40)

where

Fν(ϕ(0))=12C33(𝒚(0))C33ϕ(𝒚(0))+14Tν,F_{\nu}(\phi^{(0)})=\dfrac{1}{2C^{33}(\boldsymbol{y}^{(0)})}\dfrac{\partial C^{33}}{\partial\phi}(\boldsymbol{y}^{(0)})+\dfrac{1}{4}T_{\nu}, (41)
Gν(ϕ(0))=1C33(𝒚(0))[dVbend(2)dϕ(ϕ(0))E(2)Vbend(2)(ϕ(0))2Tν],G_{\nu}(\phi^{(0)})=\dfrac{1}{C^{33}(\boldsymbol{y}^{(0)})}\left[\dfrac{{\rm d}V_{\rm bend}^{(2)}}{{\rm d}\phi}(\phi^{(0)})-\dfrac{E^{(2)}-V_{\rm bend}^{(2)}(\phi^{(0)})}{2}T_{\nu}\right], (42)

and

C33(ϕ(0))=l22(M2M1cosϕ(0)).C^{33}(\phi^{(0)})=\dfrac{l_{\ast}^{2}}{2}(M_{2}-M_{1}\cos\phi^{(0)}). (43)

Equation (40) is the closed equation for the slow bending motion. It is reproduced as the Euler-Lagrange equation of the effective Lagrangian

Leff(ϕ(0),dϕ(0)dt1)=12Meff(ϕ(0))(dϕ(0)dt1)2Veff(ϕ(0)).L_{{\rm eff}}\left(\phi^{(0)},\dfrac{{\rm d}\phi^{(0)}}{{\rm d}t_{1}}\right)=\dfrac{1}{2}M_{{\rm eff}}(\phi^{(0)})\left(\dfrac{{\rm d}\phi^{(0)}}{{\rm d}t_{1}}\right)^{2}-V_{{\rm eff}}(\phi^{(0)}). (44)

Here, the effective (dimensionless) mass Meff(ϕ(0))M_{{\rm eff}}(\phi^{(0)}) is

Meff(ϕ(0))=exp[20ϕ(0)Fν(z)𝑑z]=(M2M1cosϕ(0)M2M1)1ν1/2(M2+M1M2+M1cosϕ(0))ν2/2,\begin{split}&M_{{\rm eff}}(\phi^{(0)})=\exp\left[2\int_{0}^{\phi^{(0)}}F_{\nu}(z)dz\right]\\ &=\left(\dfrac{M_{2}-M_{1}\cos\phi^{(0)}}{M_{2}-M_{1}}\right)^{1-\nu_{1}/2}\left(\dfrac{M_{2}+M_{1}}{M_{2}+M_{1}\cos\phi^{(0)}}\right)^{\nu_{2}/2},\end{split} (45)

and the effective potential Veff(ϕ(0))V_{{\rm eff}}(\phi^{(0)}) is

Veff(ϕ(0))=0ϕ(0)Meff(z)Gν(z)𝑑z.V_{{\rm eff}}(\phi^{(0)})=\int_{0}^{\phi^{(0)}}M_{{\rm eff}}(z)G_{\nu}(z)dz. (46)

Note that the physical dimension of VeffV_{{\rm eff}} differs from Vbend(2)V_{\rm bend}^{(2)} due to the factor 1/C331/C^{33}.

The effective potential VeffV_{{\rm eff}} of Eq. (46) is the main product of the theory. A remarkable observation is that VeffV_{{\rm eff}} depends on energy E(2)E^{(2)} and the normal mode energy distribution (ν1,ν2)(\nu_{1},\nu_{2}) through MeffM_{{\rm eff}} [Eq. (45)] and GνG_{\nu} [Eq. (42)]. Examples of the effective potential are exhibited in Sec. IV.

IV Effective potential

We exhibit examples of the effective potential with varying the parameters E(2)E^{(2)} and ν1\nu_{1} (remember ν2=1ν1\nu_{2}=1-\nu_{1}). The equal mass condition m2=mm_{2}=m is assumed unless there is a notice. Note that the in-phase (anti-phase) mode is the mode-11 (mode-22) as defined in Sec. III.2.

IV.1 Examples without bending potential

First of all, we observe the effective potential VeffV_{{\rm eff}} without bending potential, Vbend0V_{\rm bend}\equiv 0, to observe the simplest case. We exhibit effective potentials for some values of ν1\nu_{1} (ν2=1ν1\nu_{2}=1-\nu_{1}) in Fig. 2. The dynamical contribution is completely opposite between the in-phase mode and the anti-phase mode. The in-phase mode makes a valley at ϕ=0\phi=0, while the anti-phase mode makes a valley at ϕ=π\phi=\pi. A precise analysis reveals that there are the two local minima at ϕ=0\phi=0 and π\pi in the interval of ν1(1/4,3/4)\nu_{1}\in(1/4,3/4). The coexistence interval is generalized to

ν1(M2M12M2,M2+M12M2)\nu_{1}\in\left(\dfrac{M_{2}-M_{1}}{2M_{2}},\dfrac{M_{2}+M_{1}}{2M_{2}}\right) (47)

for any value of m2m_{2}. See Appendix F for details.

Refer to caption
Figure 2: The effective potential Veff(ϕ)V_{{\rm eff}}(\phi). Vbent0V_{\rm bent}\equiv 0. The numbers in the panels represent the values of ν1\nu_{1}, while ν2=1ν1\nu_{2}=1-\nu_{1}. The total energy E(2)E^{(2)} is an overall factor of VeffV_{{\rm eff}} and is set as E(2)=1E^{(2)}=1. Graphs are shifted in the vertical direction by a graphical reason.

IV.2 Examples with a bending potential

Next, we introduce an example of the bending potential as

Vbend(2)(ϕ)=cos2ϕ+1.V_{\rm bend}^{(2)}(\phi)=\cos 2\phi+1. (48)

This potential has the two minima at ϕ=±π/2\phi=\pm\pi/2. We set the equal mass condition, m2=mm_{2}=m. Since VeffV_{{\rm eff}} depends on the normal mode energy distribution (ν1,ν2)(\nu_{1},\nu_{2}) and the total energy E(2)E^{(2)}, we show graphs of the effective potential for (ν1,ν2)=(1,0)(\nu_{1},\nu_{2})=(1,0) (in-phase), (1/2,1/2)(1/2,1/2) (mixed), and (0,1)(0,1) (anti-phase) with varying the value of E(2)E^{(2)} in Fig. 3. The effective potential VeffV_{{\rm eff}} is similar to the bending potential Vbend(2)V_{\rm bend}^{(2)} when the total energy E(2)E^{(2)} is small. As the total energy increases, the local minimum points move from ϕ=±π/2\phi=\pm\pi/2 towards ϕ=0\phi=0 and/or ϕ=π\phi=\pi. The local minimum points of VeffV_{{\rm eff}} are the dynamically induced conformations (DICs).

Refer to caption
Figure 3: The bending potential Vbend(2)(ϕ)=2cosϕ+1V_{\rm bend}^{(2)}(\phi)=2\cos\phi+1 (a) and the effective potential Veff(ϕ)V_{{\rm eff}}(\phi) [(b)-(d)]. (b) (ν1,ν2)=(1,0)(\nu_{1},\nu_{2})=(1,0) (the in-phase mode). (c) (ν1,ν2)=(0.5,0.5)(\nu_{1},\nu_{2})=(0.5,0.5) (a mixed mode). (d) (ν1,ν2)=(0,1)(\nu_{1},\nu_{2})=(0,1) (the anti-phase mode). The numbers in the panels (b)-(d) represent values of E(2)E^{(2)}. Graphs are shifted in the vertical direction by a graphical reason. The black vertical straight lines mark the minimum points of Vbend(2)(ϕ)V_{\rm bend}^{(2)}(\phi).

The effective potential is determined at each point on the (E(2),ν1)(E^{(2)},\nu_{1}) plane, and yields the set of the local minimum points. We categorize the local minimum points into the three classes: ϕ=0,π\phi=0,\pi, and ϕ(0,π)\phi_{\sharp}(\neq 0,\pi). The three classes induce the seven types of sets as arranged in Table 1. By using the seven types, the (E(2),ν1)(E^{(2)},\nu_{1}) plane is divided into regions each of which is assigned by a type of the set as reported in Fig. 4. We stress that the seven types are realized by changing the total energy E(2)E^{(2)} and the mode energy distribution ν1\nu_{1}.

Table 1: The seven types of local minimum point sets of the effective potential VeffV_{{\rm eff}}. The symbol ϕ\phi_{\sharp} represents a conformation which is neither ϕ=0\phi=0 nor ϕ=π\phi=\pi. For each type the conformation ϕ\phi with the symbol M (O) is a local minimum point (not a local minimum point).
Conformation ϕ\phi I II0 IIπ III0 IIIπ IV V
ϕ(0,π)\phi_{\sharp}(\neq 0,\pi) M O O M M O M
0 O M O M O M M
π\pi O O M O M M M
Refer to caption
Figure 4: The phase diagram on the (E(2),ν1)(E^{(2)},\nu_{1}) plane with the equal mass condition m2=mm_{2}=m. m=1m=1. ν2=1ν1\nu_{2}=1-\nu_{1}. The effective potential takes a local minimum at ϕ=0\phi=0 over the red dotted line, and at ϕ=π\phi=\pi under the blue dashed line, where the two lines are obtained from Eqs. (107) and (108) respectively. ϕ=ϕ(0,π)\phi=\phi_{\sharp}~{}(\neq 0,\pi) is a local minimum point between the two solid black lines. See Table 1 for the types from I to V.

Finally, we present a phase diagram by varying the center mass m2m_{2} with fixing E(2)=4E^{(2)}=4 in Fig. 5. We give two remarks. First, the seven types are also realized by changing the center mass m2m_{2}. The regions assigned by the types III0, IIIπ, IV, and V are enhanced comparing with Fig. 4. This fact suggests that the mass distribution is useful to control the conformation. Second, the dynamical contribution dominates the effective potential when m2m_{2} is small. This domination can be explained from Eq. (39), which is a part of the averaged term 𝒜\mathcal{A}, and Eq. (26), which is the definitions of M2M_{2} and M1M_{1}. We have M2M1M_{2}\to M_{1} as m20m_{2}\to 0 from Eq. (26). Thus, the denominators of the function TνT_{\nu} can be close to 0 near ϕ(0)=0,π\phi^{(0)}=0,\pi, and hence contribution from the averaged term becomes large.

Refer to caption
Figure 5: The phase diagram on the plane (m2/m,ν1)(m_{2}/m,\nu_{1}) with E(2)=4E^{(2)}=4. The meanings of the lines are the same as Fig. 4. See Table 1 for the types from I to V.

V Numerical tests

We verify efficiency of the effective potential through numerical simulations of the model.

V.1 Setting

Numerical simulations are performed by using the fourth order symplectic integrator yoshida-90 for the Hamiltonian

H=12j=13𝒑j2mj+V(𝒓1,𝒓2,𝒓3),𝒑j=mj𝒓˙j,H=\dfrac{1}{2}\sum_{j=1}^{3}\dfrac{||\boldsymbol{p}_{j}||^{2}}{m_{j}}+V(\boldsymbol{r}_{1},\boldsymbol{r}_{2},\boldsymbol{r}_{3}),\quad\boldsymbol{p}_{j}=m_{j}\dot{\boldsymbol{r}}_{j}, (49)

which is the Legendre transform of Eq. (1). The time step is set as Δt=103\Delta t=10^{-3}. The relative energy error is suppressed in the reported simulations as |(EnumE0)/E0|<1010|(E_{\rm num}-E_{0})/E_{0}|<10^{-10}, where E0E_{0} and EnumE_{\rm num} are respectively the initial energy and the numerically obtained energy.

The two springs are assumed to be linear and VspringV_{\rm spring} is

Vspring(l1,l2)=k2[(l1l)2+(l2l)2],V_{\rm spring}(l_{1},l_{2})=\dfrac{k}{2}\left[(l_{1}-l_{\ast})^{2}+(l_{2}-l_{\ast})^{2}\right], (50)

because the theory includes only the linear part of the springs. The bending potential is Vbend=ϵ2Vbend(2)V_{\rm bend}=\epsilon^{2}V_{\rm bend}^{(2)}, and we use Eq. (48) as Vbend(2)V_{\rm bend}^{(2)}. The small parameter ϵ\epsilon is fixed as ϵ=0.1\epsilon=0.1. The masses are equal and m1=m2=m3=m=1m_{1}=m_{2}=m_{3}=m=1. The spring constant is k=10k=10, and the natural length is l=1l_{\ast}=1.

V.2 Initial condition

We set the initial condition as follows. All the beads have zero initial velocities in the xx- and yy-directions. The beads are once placed at the natural lengths of the springs with the bending angle ϕ\phi_{\ast}. Then, under the hypothesis (H), we give small displacements of l1,l2l_{1},l_{2}, and ϕ\phi so as to excite the normal modes of the springs for a given pair of (ν1,ν2)(\nu_{1},\nu_{2}). The initial condition, denoted by the subscript 0, is summarized as

{(l1,0l2,0ϕ0)=(llϕ)+ϵw[ν1𝒑in(𝒚(0))+ν2𝒑anti],(l˙1,0l˙2,0ϕ˙0)=(000).\left\{\begin{split}&\begin{pmatrix}l_{1,0}\\ l_{2,0}\\ \phi_{0}\end{pmatrix}=\begin{pmatrix}l_{\ast}\\ l_{\ast}\\ \phi_{\ast}\end{pmatrix}+\epsilon w\left[\sqrt{\nu_{1}}\boldsymbol{p}_{\rm in}(\boldsymbol{y}^{(0)})+\sqrt{\nu_{2}}\boldsymbol{p}_{\rm anti}\right],\\ &\begin{pmatrix}\dot{l}_{1,0}\\ \dot{l}_{2,0}\\ \dot{\phi}_{0}\\ \end{pmatrix}=\begin{pmatrix}0\\ 0\\ 0\\ \end{pmatrix}.\end{split}\right. (51)

See Eq. (27) for the definitions of 𝒑in\boldsymbol{p}_{\rm in} and 𝒑anti\boldsymbol{p}_{\rm anti}. We note that the amplitude ww is of O(ϵ0)O(\epsilon^{0}). In the Hamiltonian system of Eq. (49), a corresponding initial condition is

{𝒓1,0=(l1,0cos(ϕ0/2)l1,0sin(ϕ0/2)),𝒓3,0=(l2,0cos(ϕ0/2)l2,0sin(ϕ0/2))𝒓2,0=𝒑1,0=𝒑2,0=𝒑3,0=𝟎,\left\{\begin{split}&\boldsymbol{r}_{1,0}=\begin{pmatrix}-l_{1,0}\cos(\phi_{0}/2)\\ l_{1,0}\sin(\phi_{0}/2)\end{pmatrix},~{}\boldsymbol{r}_{3,0}=\begin{pmatrix}l_{2,0}\cos(\phi_{0}/2)\\ l_{2,0}\sin(\phi_{0}/2)\end{pmatrix}\\ &\boldsymbol{r}_{2,0}=\boldsymbol{p}_{1,0}=\boldsymbol{p}_{2,0}=\boldsymbol{p}_{3,0}=\boldsymbol{0},\end{split}\right. (52)

where 𝟎\boldsymbol{0} is the two-dimensional zero vector. The above initial condition gives the second order total energy as

E(2)=k2w2+Vbend(2)(ϕ0).E^{(2)}=\dfrac{k}{2}w^{2}+V_{\rm bend}^{(2)}(\phi_{0}). (53)

In the next section we set ϕ=π/2\phi_{\ast}=\pi/2, which is a local minimum point of the bending potential VbendV_{\rm bend}. The initial condition Eq. (51) hence has two free parameters of the amplitude ww, which is equivalent with E(2)E^{(2)} through Eq. (53), and the normal mode energy ratio ν1\nu_{1} (remember ν2=1ν1\nu_{2}=1-\nu_{1}).

V.3 Efficiency of the effective potential

We concentrate on the in-phase mode, (ν1,ν2)=(1,0)(\nu_{1},\nu_{2})=(1,0). Temporal evolution of ϕ(t)\phi(t) is exhibited in Figs. 6(d), (e), and (f) for three values of ww, corresponding to three values of E(2)E^{(2)} [see Eq. (53)]. Small and fast oscillation of ϕ(t)\phi(t) comes from ϕ(1)\phi^{(1)} and is induced from lj(1)l_{j}^{(1)}, which are governed by Eq. (15). When the amplitude ww is small, the bending angle ϕ\phi almost stays around the initial value ϕ0\phi_{0} [Fig. 6(d)], as it is predicted from the bending potential VbendV_{\rm bend}. However, the amplitude of oscillation of ϕ\phi becomes large as ww gets large, and the center of oscillation approaches to the zero [Figs. 6(e) and (f)]. We estimate the center of oscillation by the time average

ϕave=1T0Tϕ(t)𝑑t,T=1000.\phi_{\rm ave}=\dfrac{1}{T}\int_{0}^{T}\phi(t)dt,\quad T=1000. (54)

The estimated center is plotted as a function of E(2)E^{(2)} in Fig. 6(g) with the minimum ϕmin\phi_{\rm min} and the maximum ϕmax\phi_{\rm max} of ϕ(t)\phi(t) in t[0,1000]t\in[0,1000].

Two remarks are in order. First, the minimum and the maximum of ϕ\phi are well predicted by VeffV_{{\rm eff}}. There is a gap between ϕave\phi_{\rm ave} and the bottom of VeffV_{{\rm eff}} in the energy interval approximated by E(2)[6.8,10]E^{(2)}\in[6.8,10]. The gap is not a counterevidence but a supporting evidence of the theory. This gap comes from the inequality Veff(0)<Veff(ϕ0)V_{{\rm eff}}(0)<V_{{\rm eff}}(\phi_{0}), which implies that the bending angle climbs over the saddle point at ϕ=0\phi=0. This passing is confirmed by the jump of the theoretical minimum value of ϕ\phi. Second, the center of oscillation is continuously modified as the spring energy increases. The conformation is determined not by a local minimum of the bending potential VbendV_{\rm bend} but by a local minimum of the effective potential VeffV_{{\rm eff}} derived from dynamics. We conclude that the effective potential successfully predicts the slow bending motion.

We provide movies in Supplemental Material SM to show dynamics of the system. See Appendix G for explanation on the movies.

Refer to caption
Figure 6: Graphs of the effective potential [(a)-(c)] and temporal evolution of ϕ(t)\phi(t) with the reference conformation ϕ=π/2\phi_{\ast}=\pi/2 [(d)-(f)]. (ν,1,ν2)=(1,0)(\nu,_{1},\nu_{2})=(1,0). The amplitude ww is (a,d) w=0.5w=0.5, (b,e) w=1w=1, and (c,f) w=1.5w=1.5. The vertical black straight solid line represents the minimum point of VeffV_{{\rm eff}}. The two vertical gray straight dashed lines represent the two values of ϕ\phi satisfying Veff(ϕ)=Veff(ϕ0)V_{{\rm eff}}(\phi)=V_{{\rm eff}}(\phi_{0}) and belonging to the same valley. (g) Energy dependence of the time average ϕave\phi_{\rm ave} (red circles) with the minimum ϕmin\phi_{\rm min} (blue triangles) and the maximum ϕmax\phi_{\rm max} (purple inverse triangles) of ϕ\phi in the time interval t[0,1000]t\in[0,1000]. The black solid and the gray dashed thick lines corresponds respectively to the vertical black and gray straight lines presented in the panels (d), (e), and (f). The three black vertical solid lines indicate the values of energy which correspond to the panels (d), (e), and (f) from left to right.

VI Summary and discussions

We have investigated the three-body bead-spring model and demonstrated the dynamically induced conformation (DIC). The fast motion of the springs induces the effective potential for the slow bending motion, and the conformation is governed by the bottoms of the effective potential instead of the bending potential. One crucial remark is that the effective potential depends on the excited normal modes and its energy: The effective potential tends to have the minimum at ϕ=0\phi=0 (ϕ=π\phi=\pi), by exciting the in-phase (anti-phase) mode. Moreover, a mixed mode makes the two local minima at ϕ=0\phi=0 and π\pi.

We have developed a theory to derive the effective potential based on the multiple-scale analysis and the averaging method. The main idea of the theory is to introduce a hypothesis inspired from the adiabatic invariant. The hypothesis with the energy conservation law eliminate the unknown variables being unavoidable in autonomous systems, and the elimination introduces the mode dependence and the total energy dependence into the effective potential. A theory for a generic system can be found in Refs. yamaguchi-etal-21 ; yamaguchi-22 .

Efficiency of the theoretically obtained effective potential is successfully examined through numerical simulations. The bending angle oscillates in general, and the center of oscillation is not a bottom of the bending potential, but a bottom of the effective potential. An extreme example is that a local maximum point of the bending potential becomes a local minimum point of the effective potential. The amplitude of the bending angle oscillation is also predicted by the effective potential.

The studied model is quite simple as we neglected the excluded volume effect, for instance. The potential of the excluded volume effect, and any other potentials between the two beads of the ends, can be treated in the same way as the bending potential discussed in this article (see Appendix H). Therefore, the phenomenon of DIC is universal as long as the two assumptions (A1) and (A2) hold.

We give three discussions on DIC revealed in this article: universality, application to control, and the beat effect. First, DIC must be universal since the essential mechanism to have DIC is existence of multiple timescales. Indeed, numerical simulations show that NN-body bead-spring systems exhibit DIC yanagita-konishi-jp . Details will be reported elsewhere. Second, it is interesting to use DIC for controlling the conformation of proteins by changing energy. Control of a robot is also an interesting subject by changing the center mass m2m_{2} as shown in Fig. 5. Finally, we have neglected the beat effect between the eigenfrequencies of the fast springs, namely λ1=λ2\lambda_{1}=\lambda_{2} at ϕ=±π/2\phi=\pm\pi/2. The beat effect may trigger the Arnold diffusion arnold-64 since the full dynamics has the three degrees of freedom, (l1,l2,ϕ)(l_{1},l_{2},\phi) (see, for example, Refs. manikandan-keshavamurthy-14 ; firmbach-etal-18 for the recent progress on systems of three degrees of freedom). It will be interesting to observe evolution of the system in a very long time beyond the slow timescale t1=ϵtt_{1}=\epsilon t.

Acknowledgements.
Y.Y.Y. acknowledges the support of JSPS KAKENHI Grant Numbers 16K05472 and 21K03402. T.Y. acknowledges the support of JSPS KAKENHI Grant Numbers 18K03471 and 21K03411. T.K. is supported by Chubu University Grant (A). M.T. is supported by the Research Program of ”Dynamic Alliance for Open Innovation Bridging Human, Environment and Materials” in ”Network Joint Research Center for Materials and Devices”, and a Grant-in-Aid for Scientific Research (C) ( No. 22654047, 25610105, and 19K03653 ) from JSPS. The authors express their thanks to the anonymous referees who suggested to input the bending potential.

Appendix A Lagrangians of the three-body bead-spring model

The system of Eq. (1) has the two-dimensional translational symmetry and the rotational symmetry. We reduce Eq. (1) and derive Eq. (2) by introducing the internal coordinates. For the reduction we perform three changes of variables.

The first change of variables introduces the vectors along the springs, denoted by 𝒒1\boldsymbol{q}_{1} and 𝒒2\boldsymbol{q}_{2}, with the center-of-mass 𝒒G\boldsymbol{q}_{\rm G}. This change of variables is expressed as

(𝒒1𝒒2𝒒G)=(110011m/Mm2/Mm/M)(𝒓1𝒓2𝒓3)\begin{pmatrix}\boldsymbol{q}_{1}\\ \boldsymbol{q}_{2}\\ \boldsymbol{q}_{\rm G}\end{pmatrix}=\begin{pmatrix}-1&1&0\\ 0&-1&1\\ m/M&m_{2}/M&m/M\\ \end{pmatrix}\begin{pmatrix}\boldsymbol{r}_{1}\\ \boldsymbol{r}_{2}\\ \boldsymbol{r}_{3}\end{pmatrix} (55)

with the total mass

M=2m+m2.M=2m+m_{2}. (56)

Since each element of 𝒒G\boldsymbol{q}_{\rm G} is a cyclic coordinate by an assumption and 𝒒˙G\dot{\boldsymbol{q}}_{\rm G} is conserved, we set 𝒒˙G𝟎\dot{\boldsymbol{q}}_{\rm G}\equiv\boldsymbol{0} without loss of generality. This setting reduces 𝒒G\boldsymbol{q}_{\rm G} and 𝒒˙G\dot{\boldsymbol{q}}_{\rm G} from the Lagrangian, which is written as

L=12i,j=12Aij𝒒˙i𝒒˙jV(𝒒1,𝒒2),L=\dfrac{1}{2}\sum_{i,j=1}^{2}A^{ij}\dot{\boldsymbol{q}}_{i}\cdot\dot{\boldsymbol{q}}_{j}-V(\boldsymbol{q}_{1},\boldsymbol{q}_{2}), (57)

where we assumed that the potential energy function VV depends on only 𝒒1\boldsymbol{q}_{1} and 𝒒2\boldsymbol{q}_{2}. AijA^{ij} is the (i,j)(i,j) element of the size-22 matrix 𝐀\boldsymbol{\rm A}, which is defined by

𝐀=(M2M1M1M2),M2=m(m+m2)M,M1=m2M.\boldsymbol{\rm A}=\begin{pmatrix}M_{2}&M_{1}\\ M_{1}&M_{2}\\ \end{pmatrix},\quad M_{2}=\dfrac{m(m+m_{2})}{M},\quad M_{1}=\dfrac{m^{2}}{M}. (58)

The second change of variables introduces the polar coordinates (lj,θj)(l_{j},\theta_{j}), where ljl_{j} is the length of 𝒒j\boldsymbol{q}_{j} and θj\theta_{j} is the angle of 𝒒j\boldsymbol{q}_{j} measured from a fixed direction on 2\mathbb{R}^{2}. The vectors 𝒒j\boldsymbol{q}_{j} and 𝒒˙j\dot{\boldsymbol{q}}_{j} are then written as

𝒒j=lj𝒆rj,𝒒˙j=l˙j𝒆rj+ljθ˙j𝒆θj\boldsymbol{q}_{j}=l_{j}\boldsymbol{e}_{rj},\quad\dot{\boldsymbol{q}}_{j}=\dot{l}_{j}\boldsymbol{e}_{rj}+l_{j}\dot{\theta}_{j}\boldsymbol{e}_{\theta j} (59)

where 𝒆rj\boldsymbol{e}_{rj} is the unit vector to the radial direction of 𝒒j\boldsymbol{q}_{j}, and 𝒆θj\boldsymbol{e}_{\theta j} is the unit vector to the angle direction.

As the third change of variables, we define

ϕ=θ2θ1,ψ=θ2+θ1,\phi=\theta_{2}-\theta_{1},\quad\psi=\theta_{2}+\theta_{1}, (60)

where ϕ\phi represents the bending angle (see Fig. 1). The variables l1,l2,ϕ,l_{1},l_{2},\phi, and ψ\psi describes the Lagrangian

L=12α,β=14Bαβ(𝒛)z˙αz˙βV(l1,l2,ϕ)L=\dfrac{1}{2}\sum_{\alpha,\beta=1}^{4}B^{\alpha\beta}(\boldsymbol{z})\dot{z}_{\alpha}\dot{z}_{\beta}-V(l_{1},l_{2},\phi) (61)

where 𝒛=(z1,z2,z3,z4)=(l1,l2,ϕ,ψ)\boldsymbol{z}=(z_{1},z_{2},z_{3},z_{4})=(l_{1},l_{2},\phi,\psi) and VV does not depend on ψ\psi by the assumption of rotational symmetry. The four-dimensional vector is represented by 𝒛4\boldsymbol{z}\in\mathbb{R}^{4} to distinguish from the three-dimensional vector 𝒚3\boldsymbol{y}\in\mathbb{R}^{3} used in Eq. (2). BαβB^{\alpha\beta} is the (α,β)(\alpha,\beta) element of the size-44 matrix 𝐁\boldsymbol{\rm B}. The matrix 𝐁\boldsymbol{\rm B} is symmetric, and we show only the upper triangle elements. The diagonal elements are

{B11=B22=M2,B33=14M2(l12+l22)12M1l1l2cosϕ,B44=14M2(l12+l22)+12M1l1l2cosϕ,\left\{\begin{split}&B^{11}=B^{22}=M_{2},\\ &B^{33}=\dfrac{1}{4}M_{2}(l_{1}^{2}+l_{2}^{2})-\dfrac{1}{2}M_{1}l_{1}l_{2}\cos\phi,\\ &B^{44}=\dfrac{1}{4}M_{2}(l_{1}^{2}+l_{2}^{2})+\dfrac{1}{2}M_{1}l_{1}l_{2}\cos\phi,\\ \end{split}\right. (62)

and the off-diagonal elements are

{B12=M1cosϕ,B13=B14=12M1l2sinϕ,B23=B24=12M1l1sinϕ,B34=14M2(l12l22).\left\{\begin{split}&B^{12}=M_{1}\cos\phi,\\ &B^{13}=B^{14}=-\dfrac{1}{2}M_{1}l_{2}\sin\phi,\\ &B^{23}=-B^{24}=-\dfrac{1}{2}M_{1}l_{1}\sin\phi,\\ &B^{34}=-\dfrac{1}{4}M_{2}(l_{1}^{2}-l_{2}^{2}).\\ \end{split}\right. (63)

The Lagrangian of Eq. (61) does not depend on ψ\psi and ψ\psi is a cyclic coordinate. The conjugate momentum pψp_{\psi}, which corresponds to the total angular momentum, is defined by

pψ=Lz˙4=α=14B4αz˙αp_{\psi}=\dfrac{\partial L}{\partial\dot{z}_{4}}=\sum_{\alpha=1}^{4}B^{4\alpha}\dot{z}_{\alpha} (64)

and is conserved. Eliminating z˙4(=ψ˙)\dot{z}_{4}(=\dot{\psi}) from the kinetic energy, we obtain the Lagrangian

L=12α,β=13Cαβ(𝒚)y˙αy˙β+[pψ(𝒚,𝒚˙)]22B44(𝒚)V(𝒚),L=\dfrac{1}{2}\sum_{\alpha,\beta=1}^{3}C^{\alpha\beta}(\boldsymbol{y})\dot{y}_{\alpha}\dot{y}_{\beta}+\dfrac{[p_{\psi}(\boldsymbol{y},\dot{\boldsymbol{y}})]^{2}}{2B^{44}(\boldsymbol{y})}-V(\boldsymbol{y}), (65)

where we identified 𝐁(𝒛)\boldsymbol{\rm B}(\boldsymbol{z}) and 𝐁(𝒚)\boldsymbol{\rm B}(\boldsymbol{y}) since 𝐁\boldsymbol{\rm B} does not depend on ψ\psi. Assuming pψ=0p_{\psi}=0, we obtain Eq. (2) because contribution from pψp_{\psi} to the Euler-Lagrange equation vanishes. The (α,β)(\alpha,\beta) element of the size-33 symmetric matrix 𝐂\boldsymbol{\rm C} is defined by

Cαβ(𝒚)=Bαβ(𝒚)1B44(𝒚)B4α(𝒚)B4β(𝒚).C^{\alpha\beta}(\boldsymbol{y})=B^{\alpha\beta}(\boldsymbol{y})-\dfrac{1}{B^{44}(\boldsymbol{y})}B^{4\alpha}(\boldsymbol{y})B^{4\beta}(\boldsymbol{y}). (66)

The diagonal elements are

{C11(𝒙)=M2M12l22sin2ϕM2(l12+l22)+2M1l1l2cosϕ,C22(𝒙)=M2M12l12sin2ϕM2(l12+l22)+2M1l1l2cosϕ,C33(𝒙)=14M2(l12+l22)12M1l1l2cosϕM22(l12l22)24M2(l12+l22)+8M1l1l2cosϕ,\left\{\begin{split}C^{11}(\boldsymbol{x})&=M_{2}-\dfrac{M_{1}^{2}l_{2}^{2}\sin^{2}\phi}{M_{2}(l_{1}^{2}+l_{2}^{2})+2M_{1}l_{1}l_{2}\cos\phi},\\ C^{22}(\boldsymbol{x})&=M_{2}-\dfrac{M_{1}^{2}l_{1}^{2}\sin^{2}\phi}{M_{2}(l_{1}^{2}+l_{2}^{2})+2M_{1}l_{1}l_{2}\cos\phi},\\ C^{33}(\boldsymbol{x})&=\dfrac{1}{4}M_{2}(l_{1}^{2}+l_{2}^{2})-\dfrac{1}{2}M_{1}l_{1}l_{2}\cos\phi\\ &-\dfrac{M_{2}^{2}(l_{1}^{2}-l_{2}^{2})^{2}}{4M_{2}(l_{1}^{2}+l_{2}^{2})+8M_{1}l_{1}l_{2}\cos\phi},\end{split}\right. (67)

and the off-diagonal elements are

{C12(𝒙)=M1cosϕ+M12l1l2sin2ϕM2(l12+l22)+2M1l1l2cosϕ,C13(𝒙)=12M1l2sinϕ12M1M2(l12l22)l2sinϕM2(l12+l22)+2M1l1l2cosϕ,C23(𝒙)=12M1l1sinϕ+12M1M2(l12l22)l1sinϕM2(l12+l22)+2M1l1l2cosϕ.\left\{\begin{split}&C^{12}(\boldsymbol{x})=M_{1}\cos\phi+\dfrac{M_{1}^{2}l_{1}l_{2}\sin^{2}\phi}{M_{2}(l_{1}^{2}+l_{2}^{2})+2M_{1}l_{1}l_{2}\cos\phi},\\ &C^{13}(\boldsymbol{x})=-\dfrac{1}{2}M_{1}l_{2}\sin\phi-\dfrac{\frac{1}{2}M_{1}M_{2}(l_{1}^{2}-l_{2}^{2})l_{2}\sin\phi}{M_{2}(l_{1}^{2}+l_{2}^{2})+2M_{1}l_{1}l_{2}\cos\phi},\\ &C^{23}(\boldsymbol{x})=-\dfrac{1}{2}M_{1}l_{1}\sin\phi+\dfrac{\frac{1}{2}M_{1}M_{2}(l_{1}^{2}-l_{2}^{2})l_{1}\sin\phi}{M_{2}(l_{1}^{2}+l_{2}^{2})+2M_{1}l_{1}l_{2}\cos\phi}.\\ \end{split}\right. (68)

Appendix B Kapitza pendulum

We review an analysis of the Kapitza pendulum. This review is adjusted to our theory for easily capturing a road map of long computations.

We consider a pendulum on the xyxy plane where the yy axis points to the upward direction of the gravity gg. The pendulum has the length ll and a point mass mm at the tip. The angle ϕ\phi is taken from the downward direction of the yy axis to the anti-clockwise direction. An external force oscillates the pivot of the pendulum along the yy axis with the amplitude aa and the frequency ω\omega. The position (x,y)(x,y) of the point mass is then written as

x=lsinϕ,y=lcosϕacos(ωt+δ),x=l\sin\phi,\quad y=-l\cos\phi-a\cos(\omega t+\delta), (69)

where δ\delta is the initial phase of the pivot. Constructing the Lagrangian, we have the Euler-Lagrange equation for ϕ\phi as

dϕ2dt¯2=[(ω0ω)2+alcos(t¯+δ)]sinϕ,\dfrac{{\rm d}{}^{2}\phi}{{\rm d}\bar{t}^{2}}=-\left[\left(\dfrac{\omega_{0}}{\omega}\right)^{2}+\dfrac{a}{l}\cos(\bar{t}+\delta)\right]\sin\phi, (70)

where ω0=g/l\omega_{0}=\sqrt{g/l} and t¯=ωt\bar{t}=\omega t. If no external oscillation is applied to the pivot, namely a=0a=0, the unique stable stationary point is clearly ϕ=0\phi=0.

We assume that (i) the amplitude aa of the oscillating pivot is much smaller than the pendulum length ll and is of O(ϵ)O(\epsilon), (ii) the frequency ω0\omega_{0} is much smaller than ω\omega and is of O(ϵ)O(\epsilon), where ϵ\epsilon is a dimensionless small parameter. These assumptions imply

al=ϵα,ω0ω=ϵβ,|ϵ|1,\dfrac{a}{l}=\epsilon\alpha,\quad\dfrac{\omega_{0}}{\omega}=\epsilon\beta,\quad|\epsilon|\ll 1, (71)

where α\alpha and β\beta are of O(ϵ0)O(\epsilon^{0}).

We introduce two timescales of

t0=t¯,t1=ϵt¯,t_{0}=\bar{t},\quad t_{1}=\epsilon\bar{t}, (72)

which induce

ddt¯=t0+ϵt1.\dfrac{{\rm d}}{{\rm d}\bar{t}}=\dfrac{\partial}{\partial t_{0}}+\epsilon\dfrac{\partial}{\partial t_{1}}. (73)

The angle ϕ\phi is also expanded as

ϕ(t)=ϕ(0)(t1)+ϵϕ(1)(t0,t1).\phi(t)=\phi^{(0)}(t_{1})+\epsilon\phi^{(1)}(t_{0},t_{1}). (74)

Substituting the above expansions into the Euler-Lagrange equation, Eq. (70), we have the expanded equation

ϵ2ϕ(0)2t12+ϵϕ(1)2t02+2ϵ22ϕ(1)t0t1+ϵ3ϕ(1)2t12=[ϵ2β2+ϵαcos(t0+δ)]sin(ϕ(0)+ϵϕ(1)).\begin{split}&\epsilon^{2}\dfrac{\partial{}^{2}\phi^{(0)}}{\partial t_{1}^{2}}+\epsilon\dfrac{\partial{}^{2}\phi^{(1)}}{\partial t_{0}^{2}}+2\epsilon^{2}\dfrac{\partial^{2}\phi^{(1)}}{\partial t_{0}\partial t_{1}}+\epsilon^{3}\dfrac{\partial{}^{2}\phi^{(1)}}{\partial t_{1}^{2}}\\ &=-\left[\epsilon^{2}\beta^{2}+\epsilon\alpha\cos(t_{0}+\delta)\right]\sin(\phi^{(0)}+\epsilon\phi^{(1)}).\end{split} (75)

The equation to O(ϵ)O(\epsilon) is

ϕ(1)2t02=αcos(t0+δ)sinϕ(0)(t1).\dfrac{\partial{}^{2}\phi^{(1)}}{\partial t_{0}^{2}}=-\alpha\cos(t_{0}+\delta)\sin\phi^{(0)}(t_{1}). (76)

Solving the above equation with avoiding secular terms, we have

ϕ(1)(t0,t1)=αcos(t0+δ)sinϕ(0)(t1).\phi^{(1)}(t_{0},t_{1})=\alpha\cos(t_{0}+\delta)\sin\phi^{(0)}(t_{1}). (77)

The equation to O(ϵ2)O(\epsilon^{2}) is

ϕ(0)2t12+22ϕ(1)t0t1=[β2sinϕ(0)+αϕ(1)cos(t0+δ)cosϕ(0)].\dfrac{\partial{}^{2}\phi^{(0)}}{\partial t_{1}^{2}}+2\dfrac{\partial^{2}\phi^{(1)}}{\partial t_{0}\partial t_{1}}=-\left[\beta^{2}\sin\phi^{(0)}+\alpha\phi^{(1)}\cos(t_{0}+\delta)\cos\phi^{(0)}\right]. (78)

Substituting the O(ϵ)O(\epsilon) solution Eq. (77) and averaging over the fast timescale t0t_{0}, we have

ϕ(0)2t12=(β2sinϕ(0)+α24sin2ϕ(0)).\dfrac{\partial{}^{2}\phi^{(0)}}{\partial t_{1}^{2}}=-\left(\beta^{2}\sin\phi^{(0)}+\dfrac{\alpha^{2}}{4}\sin 2\phi^{(0)}\right). (79)

The effective potential Veff(ϕ(0))V_{\rm eff}(\phi^{(0)}) satisfying

ϕ(0)2t12=dVeffdϕ(0)(ϕ(0))\dfrac{\partial{}^{2}\phi^{(0)}}{\partial t_{1}^{2}}=-\dfrac{{\rm d}V_{\rm eff}}{{\rm d}\phi^{(0)}}(\phi^{(0)}) (80)

is then obtained as

Veff(ϕ(0))=(β2cosϕ(0)+α28cos2ϕ(0)).V_{\rm eff}(\phi^{(0)})=-\left(\beta^{2}\cos\phi^{(0)}+\dfrac{\alpha^{2}}{8}\cos 2\phi^{(0)}\right). (81)

This effective potential has a local minimum at ϕ(0)=π\phi^{(0)}=\pi for α2>2β2\alpha^{2}>2\beta^{2} in addition to ϕ(0)=0\phi^{(0)}=0. The inverted pendulum (i.e. ϕ=π\phi=\pi) is therefore stabilized by sufficiently fast oscillation (i.e. small β\beta) of the pivot irrespective of the initial phase δ\delta.

Appendix C Order of the bending potential

We show Vbend(0)(ϕ)0V_{\rm bend}^{(0)}(\phi)\equiv 0 and Vbend(1)(ϕ)0V_{\rm bend}^{(1)}(\phi)\equiv 0 under the assumptions (A1) and (A2).

C.1 O(ϵ0)O(\epsilon^{0})

There is no time derivative term in O(ϵ0)O(\epsilon^{0}), and we have

Vbend(0)ϕ(ϕ(0))=0.\dfrac{\partial V_{\rm bend}^{(0)}}{\partial\phi}(\phi^{(0)})=0. (82)

The zeroth order term Vbend(0)V_{\rm bend}^{(0)} is hence constant, and we can set Vbend(0)(ϕ)0V_{\rm bend}^{(0)}(\phi)\equiv 0 without loss of generality. We note that the identical zero is induced because ϕ(0)\phi^{(0)} in Eq. (82) is a variable. The spring potential VspringV_{\rm spring} is not identically zero in general because it is required to hold in O(ϵ0)O(\epsilon^{0})

Vspringlj(l,l)=0(j=1,2)\dfrac{\partial V_{\rm spring}}{\partial l_{j}}(l_{\ast},l_{\ast})=0\quad(j=1,2) (83)

at the point (l1,l2)=(l,l)(l_{1},l_{2})=(l_{\ast},l_{\ast}).

C.2 O(ϵ)O(\epsilon)

The terms of O(ϵ)O(\epsilon) constructs

𝐂(𝒚(0))2t02(l1(1)l2(1)ϕ(1))=(kl1(1)kl2(1)(Vbend/ϕ)(1)),\boldsymbol{\rm C}(\boldsymbol{y}^{(0)})\dfrac{\partial{}^{2}}{\partial t_{0}^{2}}\begin{pmatrix}l_{1}^{(1)}\\ l_{2}^{(1)}\\ \phi^{(1)}\\ \end{pmatrix}=\begin{pmatrix}-kl_{1}^{(1)}\\ -kl_{2}^{(1)}\\ -\left(\partial V_{\rm bend}/\partial\phi\right)^{(1)}\\ \end{pmatrix}, (84)

where

(Vbendϕ)(1)=Vbend(0)2ϕ2(ϕ(0))ϕ(1)+Vbend(1)ϕ(ϕ(0)).\left(\dfrac{\partial V_{\rm bend}}{\partial\phi}\right)^{(1)}=\dfrac{\partial{}^{2}V_{\rm bend}^{(0)}}{\partial\phi^{2}}(\phi^{(0)})\phi^{(1)}+\dfrac{\partial V_{\rm bend}^{(1)}}{\partial\phi}(\phi^{(0)}). (85)

The first term of the right-hand side in Eq. (85) is zero since Vbend(0)0V_{\rm bend}^{(0)}\equiv 0, and there is no restoring force for the variable ϕ(1)\phi^{(1)}, while Eq. (83) does not imply the zero restoring force for the springs. The second term (Vbend(1)/ϕ)(ϕ(0))(\partial V_{\rm bend}^{(1)}/\partial\phi)(\phi^{(0)}) is constant in the timescale t0t_{0}. If the second term is not zero, ϕ(1)\phi^{(1)} has a secular term, and the secular term breaks the perturbation expansion Eq. (10), which assumes |ϕ(0)||ϵϕ(1)||\phi^{(0)}|\gg|\epsilon\phi^{(1)}|. Therefore, the second term must be zero and we can set Vbend(1)0V_{\rm bend}^{(1)}\equiv 0 without loss of generality.

Appendix D Matrices in the equations of O(ϵ)O(\epsilon)

We give the explicit forms of the matrix 𝐗(𝒚(0))\boldsymbol{\rm X}(\boldsymbol{y}^{(0)}) appearing in Eq. (15). The inverse matrix of 𝐂\boldsymbol{\rm C} at 𝒚=𝒚(0)\boldsymbol{y}=\boldsymbol{y}^{(0)} is

[𝐂(𝒚(0))]1=1M22M12×(M2M1cosϕ(0)1lM1sinϕ(0)M1cosϕ(0)M21lM1sinϕ(0)1lM1sinϕ(0)1lM1sinϕ(0)2l2(M2+M1cosϕ(0))).\begin{split}&[\boldsymbol{\rm C}(\boldsymbol{y}^{(0)})]^{-1}=\dfrac{1}{M_{2}^{2}-M_{1}^{2}}\\ &\times\begin{pmatrix}M_{2}&-M_{1}\cos\phi^{(0)}&\frac{1}{l_{\ast}}M_{1}\sin\phi^{(0)}\\ -M_{1}\cos\phi^{(0)}&M_{2}&\frac{1}{l_{\ast}}M_{1}\sin\phi^{(0)}\\ \frac{1}{l_{\ast}}M_{1}\sin\phi^{(0)}&\frac{1}{l_{\ast}}M_{1}\sin\phi^{(0)}&\frac{2}{l_{\ast}^{2}}(M_{2}+M_{1}\cos\phi^{(0)})\\ \end{pmatrix}.\end{split} (86)

The matrix 𝐗(𝒚(0))=[𝐂(𝒚(0))]1𝐊\boldsymbol{\rm X}(\boldsymbol{y}^{(0)})=[\boldsymbol{\rm C}(\boldsymbol{y}^{(0)})]^{-1}\boldsymbol{\rm K} is hence

𝐗(𝒚(0))=kM22M12(M2M1cosϕ(0)0M1cosϕ(0)M201lM1sinϕ(0)1lM1sinϕ(0)0).\boldsymbol{\rm X}(\boldsymbol{y}^{(0)})=\dfrac{k}{M_{2}^{2}-M_{1}^{2}}\begin{pmatrix}M_{2}&-M_{1}\cos\phi^{(0)}&0\\ -M_{1}\cos\phi^{(0)}&M_{2}&0\\ \frac{1}{l_{\ast}}M_{1}\sin\phi^{(0)}&\frac{1}{l_{\ast}}M_{1}\sin\phi^{(0)}&0\\ \end{pmatrix}. (87)

Appendix E Validity of the hypothesis

Under the equal mass condition m2=mm_{2}=m, we examine validity of the hypothesis (H) expressed in Eq. (33). We introduce the approximations of

lj(1)t0dljdt,ϕ(0)ϕ,\dfrac{\partial l_{j}^{(1)}}{\partial t_{0}}\to\dfrac{{\rm d}l_{j}}{{\rm d}t},\quad\phi^{(0)}\to\phi, (88)

and the amplitudes of normal modes are expressed as

{w12=12[(l1+l22l)2+1λ1(l˙1+l˙2)2],w22=12[(l1l2)2+1λ2(l˙1l˙2)2],\left\{\begin{split}w_{1}^{2}&=\dfrac{1}{2}\left[\left(l_{1}+l_{2}-2l_{\ast}\right)^{2}+\dfrac{1}{\lambda_{1}}\left(\dot{l}_{1}+\dot{l}_{2}\right)^{2}\right],\\ w_{2}^{2}&=\dfrac{1}{2}\left[\left(l_{1}-l_{2}\right)^{2}+\dfrac{1}{\lambda_{2}}\left(\dot{l}_{1}-\dot{l}_{2}\right)^{2}\right],\\ \end{split}\right. (89)

where the eigenvalues λ1\lambda_{1} and λ2\lambda_{2} are defined in Eq. (25). We compute the normal mode energy ratio defined by

R=E1E1+E2,Ej=k2wj2(j=1,2).R=\dfrac{E_{1}}{E_{1}+E_{2}},\quad E_{j}=\dfrac{k}{2}w_{j}^{2}~{}(j=1,2). (90)

The hypothesis is valid if RR is constant in time.

We use the initial condition of Eq. (51) with ϕ=π/2\phi_{\ast}=\pi/2, and the amplitude of the normal modes is w=1.5w=1.5. Temporal evolution of RR is exhibited in Fig. 7 for ν1=1,0.75,0.5,0.25\nu_{1}=1,~{}0.75,~{}0.5,~{}0.25, and 0 with ν2=1ν1\nu_{2}=1-\nu_{1}. The hypothesis (H) is valid around ν1=1\nu_{1}=1 and 0 in particular.

Refer to caption
Figure 7: Temporal evolution of the normal mode energy ratio RR. ν1=1.0\nu_{1}=1.0 (red), 0.750.75 (orange), 0.50.5 (green), 0.250.25 (blue), and 0 (magenta) from top to bottom. The amplitude of the normal modes is w=1.5w=1.5.

Appendix F Analysis of the effective potential VeffV_{{\rm eff}}

Let us study the critical points of the effective potential VeffV_{{\rm eff}}. A critical point is defined as the point at which Veff=0V_{{\rm eff}}^{\prime}=0. The derivative of VeffV_{{\rm eff}} is

Veff(ϕ)=Meff(ϕ)Gν(ϕ).V_{{\rm eff}}^{\prime}(\phi)=M_{{\rm eff}}(\phi)G_{\nu}(\phi). (91)

Thus, we have

ϕ is a critical pointGν(ϕ)=0\phi\text{ is a critical point}\quad\Longleftrightarrow\quad G_{\nu}(\phi)=0 (92)

since the effective mass MeffM_{{\rm eff}} is always positive as found in Eq. (45). The effective potential at a critical point takes a local minimum or a local maximum depending on the sign of the second derivative. At a critical point, we have Gν=0G_{\nu}=0 and the second derivative of VeffV_{{\rm eff}} is

ϕ is a critical pointVeff′′(ϕ)=Meff(ϕ)Gν(ϕ).\phi\text{ is a critical point}\quad\Longrightarrow\quad V_{{\rm eff}}^{\prime\prime}(\phi)=M_{{\rm eff}}(\phi)G_{\nu}^{\prime}(\phi). (93)

Again from Meff>0M_{{\rm eff}}>0, the sign of Veff′′V_{{\rm eff}}^{\prime\prime} is determined by the sign of GνG_{\nu}^{\prime}. Keeping in mind the above discussions, we study the critical points of the effective potential for absence and appearance of the bending potential.

F.1 Absence of the bending potential

The function GνG_{\nu} is proportional to TνT_{\nu} for Vbend0V_{\rm bend}\equiv 0, and the sinusoidal function in TνT_{\nu} gives the two critical points of ϕ=0\phi=0 and π\pi. In addition, there are the other two possible critical points ϕ=±ϕ\phi=\pm\phi_{\sharp} which solve the equation

ν1M2M1cosϕν2M2+M1cosϕ=0\dfrac{\nu_{1}}{M_{2}-M_{1}\cos\phi}-\dfrac{\nu_{2}}{M_{2}+M_{1}\cos\phi}=0 (94)

with ν2=1ν1\nu_{2}=1-\nu_{1} and exist in the interval

M2M12M2<ν1<M2+M12M2.\dfrac{M_{2}-M_{1}}{2M_{2}}<\nu_{1}<\dfrac{M_{2}+M_{1}}{2M_{2}}. (95)

The derivatives of GνG_{\nu} at ϕ=0\phi=0 and π\pi are respectively

Gν(0)=E(2)l2M1M2M1(2ν11)M2+M1M22M12G_{\nu}^{\prime}(0)=\dfrac{E^{(2)}}{l_{\ast}^{2}}\dfrac{M_{1}}{M_{2}-M_{1}}\dfrac{(2\nu_{1}-1)M_{2}+M_{1}}{M_{2}^{2}-M_{1}^{2}} (96)

and

Gν(π)=E(2)l2M1M2+M1(12ν1)M2+M1M22M12,G_{\nu}^{\prime}(\pi)=\dfrac{E^{(2)}}{l_{\ast}^{2}}\dfrac{M_{1}}{M_{2}+M_{1}}\dfrac{(1-2\nu_{1})M_{2}+M_{1}}{M_{2}^{2}-M_{1}^{2}}, (97)

where we used the relation ν2=1ν1\nu_{2}=1-\nu_{1}. The point ϕ=0\phi=0 is hence a local minimum point (Veff′′(0)>0V_{{\rm eff}}^{\prime\prime}(0)>0) if and only if

ν1>M2M12M2,\nu_{1}>\dfrac{M_{2}-M_{1}}{2M_{2}}, (98)

and the point ϕ=π\phi=\pi is a local minimum point (Veff′′(π)>0V_{{\rm eff}}^{\prime\prime}(\pi)>0) if and only if

ν1<M2+M12M2.\nu_{1}<\dfrac{M_{2}+M_{1}}{2M_{2}}. (99)

The two points ϕ=0\phi=0 and ϕ=π\phi=\pi are the local minimum points in the interval of Eq. (95). The periodicity of the effective potential Veff(ϕ)V_{{\rm eff}}(\phi) requires the same numbers of local minimum points (Veff′′>0V_{{\rm eff}}^{\prime\prime}>0) and local maximum points (Veff′′<0V_{{\rm eff}}^{\prime\prime}<0), and hence the critical points ϕ=±ϕ\phi=\pm\phi_{\sharp} are the local maximum points.

F.2 Appearance of the bending potential

We write the function Gν(ϕ)G_{\nu}(\phi) as

Gν(ϕ)=sinϕl2(M2M1cosϕ)gν1(ϕ)G_{\nu}(\phi)=\dfrac{\sin\phi}{l_{\ast}^{2}(M_{2}-M_{1}\cos\phi)}g_{\nu_{1}}(\phi) (100)

where

gν1(ϕ)=8cosϕ+(E(2)cos2ϕ1)×[ν1M2M1cosϕν2M2+M1cosϕ].\begin{split}g_{\nu_{1}}(\phi)=&-8\cos\phi+(E^{(2)}-\cos 2\phi-1)\\ &\times\left[\dfrac{\nu_{1}}{M_{2}-M_{1}\cos\phi}-\dfrac{\nu_{2}}{M_{2}+M_{1}\cos\phi}\right].\end{split} (101)

The effective potential has the critical points at ϕ=0,π\phi=0,\pi, and ϕ\phi_{\sharp} satisfying g(ϕ)=0g(\phi_{\sharp})=0. We separately discuss the second derivative

Veff′′(ϕ)=gν1(ϕ)ϕ[sinϕl2(M2M1cosϕ)]+sinϕl2(M2M1cosϕ)gν1(ϕ)\begin{split}V_{{\rm eff}}^{\prime\prime}(\phi)&=g_{\nu_{1}}(\phi)\dfrac{\partial}{\partial\phi}\left[\dfrac{\sin\phi}{l_{\ast}^{2}(M_{2}-M_{1}\cos\phi)}\right]\\ &+\dfrac{\sin\phi}{l_{\ast}^{2}(M_{2}-M_{1}\cos\phi)}g_{\nu_{1}}^{\prime}(\phi)\end{split} (102)

at a critical point.

F.2.1 The critical points ϕ=0\phi=0 and π\pi

The second term of the right-hand side of Eq. (102) is zero, and

ϕ[sinϕl2(M2M1cosϕ)]=cosϕl2(M2M1cosϕ).\dfrac{\partial}{\partial\phi}\left[\dfrac{\sin\phi}{l_{\ast}^{2}(M_{2}-M_{1}\cos\phi)}\right]=\dfrac{\cos\phi}{l_{\ast}^{2}(M_{2}-M_{1}\cos\phi)}. (103)

This factor is positive (negative) at ϕ=0\phi=0 (π\pi), and the sign of Veff′′V_{{\rm eff}}^{\prime\prime} is determined by gν1(ϕ)g_{\nu_{1}}(\phi). We have

gν1(0)=8+(E(2)2)M1M2M22M12(2ν11+M1M2)g_{\nu_{1}}(0)=-8+(E^{(2)}-2)\dfrac{M_{1}M_{2}}{M_{2}^{2}-M_{1}^{2}}\left(2\nu_{1}-1+\dfrac{M_{1}}{M_{2}}\right) (104)

and

gν1(π)=8+(E(2)2)M1M2M22M12(2ν11M1M2).g_{\nu_{1}}(\pi)=8+(E^{(2)}-2)\dfrac{M_{1}M_{2}}{M_{2}^{2}-M_{1}^{2}}\left(2\nu_{1}-1-\dfrac{M_{1}}{M_{2}}\right). (105)

We separately discuss for 0E(2)20\leq E^{(2)}\leq 2 and E(2)>2E^{(2)}>2, where the boundary E(2)=2E^{(2)}=2 comes from the bending potential at ϕ=0\phi=0 and π\pi: Vbend(2)(0)=Vbend(2)(ϕ)=2V_{\rm bend}^{(2)}(0)=V_{\rm bend}^{(2)}(\phi)=2.

If 0E(2)20\leq E^{(2)}\leq 2, the maximum value of gν1(0)g_{\nu_{1}}(0) is realized at ν1=0\nu_{1}=0, which gives

g0(0)=8+(2E(2))11/(M2/M1)M2/M11/(M2/M1)<7g_{0}(0)=-8+(2-E^{(2)})\dfrac{1-1/(M_{2}/M_{1})}{M_{2}/M_{1}-1/(M_{2}/M_{1})}<-7 (106)

for 0E(2)20\leq E^{(2)}\leq 2. Note M2/M1>1M_{2}/M_{1}>1. The point ϕ=0\phi=0 is hence a local maximum point for 0E(2)20\leq E^{(2)}\leq 2. A similar discussion states that the point ϕ=π\phi=\pi is a local maximum point for 0E(2)20\leq E^{(2)}\leq 2.

For E(2)>2E^{(2)}>2, the effective potential takes a local minimum at ϕ=0\phi=0 if

ν1>12(11M2/M1)+4E(2)2(M2M11M2/M1),\nu_{1}>\dfrac{1}{2}\left(1-\dfrac{1}{M_{2}/M_{1}}\right)+\dfrac{4}{E^{(2)}-2}\left(\dfrac{M_{2}}{M_{1}}-\dfrac{1}{M_{2}/M_{1}}\right), (107)

and takes a local minimum at ϕ=π\phi=\pi if

ν1<12(1+1M2/M1)4E(2)2(M2M11M2/M1).\nu_{1}<\dfrac{1}{2}\left(1+\dfrac{1}{M_{2}/M_{1}}\right)-\dfrac{4}{E^{(2)}-2}\left(\dfrac{M_{2}}{M_{1}}-\dfrac{1}{M_{2}/M_{1}}\right). (108)

By changing the inequalities into the equality, Eq. (107) gives the red dotted-line and Eq. (108) gives the blue dashed-line in Figs. 4 and 5. We remark that Eq. (108) is equivalent with

ν2>12(11M2/M1)+4E(2)2(M2M11M2/M1),\nu_{2}>\dfrac{1}{2}\left(1-\dfrac{1}{M_{2}/M_{1}}\right)+\dfrac{4}{E^{(2)}-2}\left(\dfrac{M_{2}}{M_{1}}-\dfrac{1}{M_{2}/M_{1}}\right), (109)

whose right-hand side is identical with that of Eq. (107).

F.2.2 The critical point ϕ=ϕ\phi=\phi_{\sharp}

We have g(ϕ)=0g(\phi_{\sharp})=0, and

Veff′′(ϕ)=sinϕl2(M2M1cosϕ)g(ϕ).V_{{\rm eff}}^{\prime\prime}(\phi_{\sharp})=\dfrac{\sin\phi_{\sharp}}{l_{\ast}^{2}(M_{2}-M_{1}\cos\phi_{\sharp})}g^{\prime}(\phi_{\sharp}). (110)

The effective potential takes a local minimum (maximum) if Veff′′(ϕ)>0V_{{\rm eff}}^{\prime\prime}(\phi_{\sharp})>0 (Veff′′(ϕ)<0V_{{\rm eff}}^{\prime\prime}(\phi_{\sharp})<0) at the critical point ϕ\phi_{\sharp}.

Appendix G Explanation on movies

We provide movies for the initial conformations shown in Fig. 8 in Supplemental Material SM . The bending potential Vbend(2)(ϕ)V_{\rm bend}^{(2)}(\phi) of Eq. (48) is in use. The excited spring mode is (ν1,ν2)=(1,0)(\nu_{1},\nu_{2})=(1,0), and the amplitude ww is w=0.5w=0.5 for Figs. 8 (a) and (d), w=1w=1 for (b) and (e), and w=1.5w=1.5 for (c) and (f). The other initial conditions are described in Sec. V.2, and the system parameter values (mi,km_{i},k, and ll_{\ast}) are given in Sec. V.1. Dynamics of the system corresponding to the panels from Figs. 8(a) to (f) is demonstrated in from MovieA to MovieF, respectively. We stress that temporal evolution is well understood by the effective potential VeffV_{{\rm eff}}, while the bending potential [see Fig. 3(a)] does not explain it.

Refer to caption
Figure 8: Initial conformations (red points) and effective potentials (black lines). The reference conformations are ϕ=π/2\phi_{\ast}=\pi/2 in the panels (a), (b), and (c), and ϕ=0.01\phi_{\ast}=0.01 in (d), (e), and (f), where the initial conformation ϕ0\phi_{0} is close to ϕ\phi_{\ast}. See Figs. 6(d), (e), and (f) for temporal evolution of ϕ(t)\phi(t) corresponding to the panels (a), (b), and (c), respectively.

Appendix H From a general potential to the bending potential

The bending potential Vbend(ϕ)V_{\rm bend}(\phi) represents the interaction between the two beads of the ends, because the interction between an end and the center beads results in the spring potential. We show that the bending potential energy function of ϕ\phi is derived from any potential VGV_{\rm G} which is a function of the distance r=𝒓3𝒓1r=||\boldsymbol{r}_{3}-\boldsymbol{r}_{1}||, although it is assumed to be a function of only ϕ\phi in the main text. Here rr is represented by using l1,l2l_{1},l_{2}, and ϕ\phi as

r=(𝒓3𝒓2)+(𝒓2𝒓1)=l12+l22+2l1l2cosϕ.r=||(\boldsymbol{r}_{3}-\boldsymbol{r}_{2})+(\boldsymbol{r}_{2}-\boldsymbol{r}_{1})||=\sqrt{l_{1}^{2}+l_{2}^{2}+2l_{1}l_{2}\cos\phi}. (111)

Substituting Eq. (9) into Eq. (111), we have r=r(0)+O(ϵ)r=r^{(0)}+O(\epsilon) and

r(0)=l2(1+cosϕ(0)).r^{(0)}=l_{\ast}\sqrt{2(1+\cos\phi^{(0)})}. (112)

As shown in Appendix C, the potential VGV_{\rm G} is of O(ϵ2)O(\epsilon^{2}) under the assumptions (A1) and (A2), and we expand it as

VG(r)=ϵ2VG(2)(r)+O(ϵ3)=ϵ2VG(2)(r(0))+O(ϵ3).V_{\rm G}(r)=\epsilon^{2}V_{\rm G}^{(2)}(r)+O(\epsilon^{3})=\epsilon^{2}V_{\rm G}^{(2)}(r^{(0)})+O(\epsilon^{3}). (113)

Therefore, the second order bending potential Vbend(2)V_{\rm bend}^{(2)} is derived as a function of only ϕ(0)\phi^{(0)} as

Vbend(2)(ϕ(0))=VG(2)(l2(1+cosϕ(0))).V_{\rm bend}^{(2)}(\phi^{(0)})=V_{\rm G}^{(2)}\left(l_{\ast}\sqrt{2(1+\cos\phi^{(0)})}\right). (114)

The effective potential VeffV_{{\rm eff}} is obtained from Eq. (46) by substituting the above bending potential Vbend(2)V_{\rm bend}^{(2)} into Eq. (42).

References

  • (1) D. Koshland, Application of a theory of enzyme specificity to protein synthesis, Proc. Natl. Acad. Sci. USA 44, 98 (1958).
  • (2) J. Monod, J. Wyman, and J. P. Changeux, On the nature of allosteric transitions: A plausible model, J. Mol. Biol. 12, 88 (1965).
  • (3) K. Okazaki and S. Takada, Dynamic energy landscape view of coupled binding and protein conformational change: Induced-fit versus population-shift mechanisms, PNAS 105, 11182 (2008).
  • (4) D. Seeliger and B. L. de Groot, Conformational transitions upon ligand binding: Holo structure prediction from apo conformations, PLoS Comput. Biol. 6, e1000634 (2010).
  • (5) S. Fuchigami, H. Fujisaki, Y. Matsunaga, and A. Kidera, Protein functional motions: Basic concepts and computational methodologies, in Advancing theory for kinetics and dynamics of complex, many-dimensional systems: Clusters and Proteins: Advances in Chemical Physics, Volume 145 (Wiley, New Jersey, 2011).
  • (6) H. Hauser, A. J. Ijspeert, R. M. Füchslin, R. Pfeifer, and W. Maass, Towards a theoretical foundation for morphological computation with compliant bodies, Biol. Cybern. 105, 355 (2011).
  • (7) Special issue on Morphological Computation, Artificial Life 19(1) (2013).
  • (8) V. C. Müller and M. Hoffmann, What Is Morphological Computation? On How the Body Contributes to Cognition and Control, Artificial Life 23, 1 (2017).
  • (9) S. Collins, A. Ruina, R. Tedrake, and M. Wisse, Efficient bipedal robots based on passive-dynamic walkers, Science 307, 1082 (2005).
  • (10) M. Hermans, B. Schrauwen, P. Bienstman, and J. Dambre, Automated design of complex dynamic systems, PLoS ONE 9, e86696 (2014). doi:10.1371/journal.pone.0086696
  • (11) J. U. Surjadi, L. Gao, H. Du, X. Li, X. Xiong, N. X. Fang, and Y. Lu, Mechanical metamaterials and their engineering applications, Adv. Eng. Mater. 21, 1800864 (2019).
  • (12) Z. Y. Wei, Z. V. Guo, L. Dudte, H. Y. Liang, and L. Mahadevan, Geometric mechanics of periodic pleated origami, Phys. Rev. Lett. 110, 215501 (2013).
  • (13) A. Stephenson, XX. On induced stability, The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 15, 233 (1908).
  • (14) 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).
  • (15) P. E. Rouse, A theory of the linear viscoelastic properties of dilute solutions of coiling polymers, J. Chem. Phys. 21, 1272 (1953).
  • (16) B. H. Zimm, Dynamics of polymer molecules in dilute solution: Viscoelasticity, flow birefringence and dielectric loss, J. Chem. Phys. 24, 269 (1956).
  • (17) E. Gauger and H. Stark, Numerical study of a microscopic artificial swimmer, Phys. Rev. E 74, 021907 (2006).
  • (18) Y. Mirzae, B. Y. Rubinstein, K. I. Morozov, and A. M. Leshansky, Modeling propulsion of soft magnetic nanowires, Front. Robot. AI 7, 595777 (2020).
  • (19) A. Saadat and B. Khomami, A new bead-spring model for simulation of semi-flexible macromolecules, J. Chem. Phys. 145, 024902 (2016).
  • (20) 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).
  • (21) M. Grifoni and P. Hänggi, Coherent and incoherent quantum stochastic resonance, Phys. Rev. Lett. 76, 1611 (1996).
  • (22) 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).
  • (23) V. N. Chizhevsky, E. Smeu, and G. Giacomelli, Experimental evidence of “vibrational resonance” in an optical system, Phys. Rev. Lett. 91, 220602 (2003).
  • (24) V. N. Chizhevsky, Experimental evidence of vibrational resonance in a multistable system, Phys. Rev. E 89, 062914 (2014).
  • (25) D. Cubero, J. P. Baltanas, and J. Casado-Pascual, High-frequency effects in the Fitzhugh-Nagumo neuron model, Phys. Rev. E 73, 061102 (2006).
  • (26) 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).
  • (27) S. H. Weinberg, High frequency stimulation of cardiac myocytes: A theoretical and computational study, Chaos 24, 043104 (2014).
  • (28) 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).
  • (29) R. H. Buchanan, G. Jameson, and D. Oedjoe, Cyclic migration of bubbles in vertically vibrating liquid columns, Ind. Eng. Chem. Fund. 1, 82 (1962).
  • (30) M. H. I. Baird, Resonant bubbles in a vertically vibrating liquid column, Can. J. Chem. Eng. 41, 52 (1963).
  • (31) G. J. Jameson, The motion of a bubble in a vertically oscillating viscous liquid, Chem. Eng. Sci. 21, 35 (1966).
  • (32) B. Apffel, F. Novkoski, A. Eddi, and E. Fort, Floating under a levitating liquid, Nature 585, 48 (2020).
  • (33) R. E. Bellman, J. Bentsman, and S. M. Meerkov, Vibrational control of nonlinear systems, IEEE Trans. Automat. Contr. 31, 710 (1986).
  • (34) B. Shapiro and B. T. Zinn, High-frequency nonlinear vibrational control, IEEE Trans. Automat. Contr. 42, 83 (1997).
  • (35) M. Borromeo and F. Marchesoni, Artificial sieves for quasimassless particles, Phys. Rev. Lett. 99, 150605 (2007).
  • (36) C. J. Richards, T. J. Smart, P. H. Jones, and D. Cubero, A microscopic Kapitza pendulum, Scientific Reports 8, 13107 (2018).
  • (37) C. M. Bender and S. A. Orszag, Advanced Mathematical Methods for Scientists and Engineers I: Asymptotic Methods and Perturbation Theory (Springer, 1999).
  • (38) 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).
  • (39) N. M. Krylov and N. N. Bogoliubov, Introduction to Nonlinear Mechanics (Princeton University Press, Princeton, 1947).
  • (40) J. Guckenheimer and P. Holmes, Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vectgor Fields (Springer-Verlag, New York, 1983).
  • (41) 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).
  • (42) H. Yoshida, Construction of higher order symlectic integrators, Phys. Lett. A 190, 262 (1990).
  • (43) See Supplemental Material at [URL] for movies.
  • (44) Y. Y. Yamaguchi, T. Yanagita, T. Konishi, and M. Toda, Dynamically induced conformation depending on excited normal modes of fast oscillation, arXiv:2111.10025v1.
  • (45) Y. Y. Yamaguchi, in preparation.
  • (46) V. I. Arnold, Instability of dynamical systems with several degrees of freedom, Sov. Math.-Dokl. 5, 58 (1964).
  • (47) P. Manikandan and S. Keshavamurthy, Dynamical traps lead to the slowing down of intramolecular vibrational energy flow, PNAS 111, 14354 (2014).
  • (48) M. Firmbach, S. Lange, R. Ketzmerick, and A. Bäcker, Three-dimensional billiards: Visualization of regular structures and trapping of chaotic trajectories, Phys. Rev. E 98, 022214 (2018).