Research article

A limiting model for a low Reynolds number swimmer with N passive elastic arms

  • Received: 02 March 2023 Revised: 18 April 2023 Accepted: 18 April 2023 Published: 11 May 2023
  • We consider a low Reynolds number artificial swimmer that consists of an active arm followed by N passive springs separated by spheres. This setup generalizes an approach proposed in Montino and DeSimone, Eur. Phys. J. E, vol. 38, 2015. We further study the limit as the number of springs tends to infinity and the parameters are scaled conveniently, and provide a rigorous proof of the convergence of the discrete model to the continuous one. Several numerical experiments show the performances of the displacement in terms of the frequency or the amplitude of the oscillation of the active arm.

    Citation: François Alouges, Aline Lefebvre-Lepot, Jessie Levillain. A limiting model for a low Reynolds number swimmer with N passive elastic arms[J]. Mathematics in Engineering, 2023, 5(5): 1-20. doi: 10.3934/mine.2023087

    Related Papers:

    [1] Arthur. J. Vromans, Fons van de Ven, Adrian Muntean . Homogenization of a pseudo-parabolic system via a spatial-temporal decoupling: Upscaling and corrector estimates for perforated domains. Mathematics in Engineering, 2019, 1(3): 548-582. doi: 10.3934/mine.2019.3.548
    [2] Yu Leng, Lampros Svolos, Dibyendu Adak, Ismael Boureima, Gianmarco Manzini, Hashem Mourad, Jeeyeon Plohr . A guide to the design of the virtual element methods for second- and fourth-order partial differential equations. Mathematics in Engineering, 2023, 5(6): 1-22. doi: 10.3934/mine.2023100
    [3] Lars Eric Hientzsch . On the low Mach number limit for 2D Navier–Stokes–Korteweg systems. Mathematics in Engineering, 2023, 5(2): 1-26. doi: 10.3934/mine.2023023
    [4] Diogo Gomes, Julian Gutierrez, Ricardo Ribeiro . A mean field game price model with noise. Mathematics in Engineering, 2021, 3(4): 1-14. doi: 10.3934/mine.2021028
    [5] Pablo Blanc, Fernando Charro, Juan J. Manfredi, Julio D. Rossi . Games associated with products of eigenvalues of the Hessian. Mathematics in Engineering, 2023, 5(3): 1-26. doi: 10.3934/mine.2023066
    [6] Emilio N. M. Cirillo, Giuseppe Saccomandi, Giulio Sciarra . Compact structures as true non-linear phenomena. Mathematics in Engineering, 2019, 1(3): 434-446. doi: 10.3934/mine.2019.3.434
    [7] Ludovica Cicci, Stefania Fresca, Stefano Pagani, Andrea Manzoni, Alfio Quarteroni . Projection-based reduced order models for parameterized nonlinear time-dependent problems arising in cardiac mechanics. Mathematics in Engineering, 2023, 5(2): 1-38. doi: 10.3934/mine.2023026
    [8] Marco Cirant, Kevin R. Payne . Comparison principles for viscosity solutions of elliptic branches of fully nonlinear equations independent of the gradient. Mathematics in Engineering, 2021, 3(4): 1-45. doi: 10.3934/mine.2021030
    [9] Daniele Del Santo, Francesco Fanelli, Gabriele Sbaiz, Aneta Wróblewska-Kamińska . On the influence of gravity in the dynamics of geophysical flows. Mathematics in Engineering, 2023, 5(1): 1-33. doi: 10.3934/mine.2023008
    [10] Gennaro Ciampa, Gianluca Crippa, Stefano Spirito . Propagation of logarithmic regularity and inviscid limit for the 2D Euler equations. Mathematics in Engineering, 2024, 6(4): 494-509. doi: 10.3934/mine.2024020
  • We consider a low Reynolds number artificial swimmer that consists of an active arm followed by N passive springs separated by spheres. This setup generalizes an approach proposed in Montino and DeSimone, Eur. Phys. J. E, vol. 38, 2015. We further study the limit as the number of springs tends to infinity and the parameters are scaled conveniently, and provide a rigorous proof of the convergence of the discrete model to the continuous one. Several numerical experiments show the performances of the displacement in terms of the frequency or the amplitude of the oscillation of the active arm.



    As stated by Purcell's Scallop Theorem [16], reciprocal shape changes in a swimmer never leads to a net displacement of the system in a low Reynolds number setting. Indeed, a microscopic scallop opening and closing its valve would be completely unable to swim, due to negligible inertial forces in this situation [8]. Several simple mechanisms have then been introduced (see e.g., [11]) to overcome this obstruction, most of them using two degrees of freedom in order to create closed curves with nonzero surface in the shape space of the swimmer.

    One of the simplest mechanisms introduced in the literature is probably Najafi and Golestanian's three-sphere swimmer [14], which consists in three spheres linked by two extensible arms of negligible thickness, moving along a single direction. This model is much simpler than Purcell's original three-link swimmer [16], or Purcell's rotator [10], as there is no rotational motion involved. This swimmer has two degrees of freedom, activated periodically in time with a phase lag in order to produce the loop. Both Purcell's and Najafi and Golestanian's swimmers have been extensively studied in [1,2,3,4,5,9].

    As an extension of this three-sphere swimmer, Montino and DeSimone then introduced a three-sphere swimmer with a passive elastic arm [13]. This adaptation has only one degree of freedom, which is the length of the non-elastic arm. Thanks to a resonant effect at natural frequency of the system (which depends on the viscosity of the fluid, the masses and the spring constant), an out-of-phase oscillation of the spring is created, which ultimately leads to a net motion of the swimmer. However, at very low or very high frequency, no net motion is possible after a stroke. Having this passive elastic arm also confines net motion to only one direction on the swimming axis, swimming direction is thus limited, and the swimmer can only move with its passive arm ahead. This was also denoted by Passov [15], when looking at Purcell's three-link swimmer with a passive elastic tail.

    In this paper, Montino and DeSimone's swimmer is extended by adding a large number N of passive elastic arms to their one-dimensional swimmer, thus turning it into an (N+2)-sphere swimmer. This simple swimmer then leads to a limit model with an elastic tail resembling a one-dimensional flagella along which compressive waves propagate.

    The paper is organized as follows. In Section 2, we describe the N-spring swimmer, and its equations of motion, before looking at the limit model, when the number of springs tends to infinity, in Section 2.3. We prove the convergence of the discrete model to the continuous one in Section 3, using the fact that it is found to be a non-conventional mass lumping discretization of the limit model. Section 4 introduces two formulas in order to compute the net displacement of both swimmers, discrete and continuous. Finally, in Section 5 we study numerically the movement and displacement of our swimmer depending on various system parameters, in order to find optimal swimming parameters to obtain the largest net displacement possible.

    The swimmer studied in this paper is an extension of the three-sphere swimmer with a single passive elastic arm [13], to a swimmer with N+2 spheres and N passive elastic arms, presented in Figure 1. The first arm of this artificial swimmer is a rod of negligible thickness, surrounded by two spheres of radius a1. This arm has a prescribed periodic movement around a length at rest L, of the form L0(t)=L(1+˜εcos(ωt)) where ˜ε[0,1) is a non-dimensional parameter. ˜ε<1 so that the active arm always has a positive length. We define ε as ε=L˜ε. The rest of the swimmer has a total length at rest Λ that does not depend on N. In order to keep a constant length and have an elastic force that does not depend on N, all the other spheres have a radius a=˜a/N, the springs have each a rest length h=Λ/Na, and an elastic constant k=˜kN, with ˜k and ˜a prescribed and independent of N.

    Figure 1.  Low Reynolds number swimmer with N elastic arms.

    If the swimmer is able to control the length of the front rod with the prescribed periodic function L0(t), the length of the N remaining springs are governed by the balance of viscous and elastic forces. At any time t, the length Lj(t) of the j-th arm, j1 is written as Lj(t)=j(t)N+h. Let us then denote by μ the fluid viscosity, fFj and fRj the hydrodynamic and elastic forces on the j-th sphere. We also call xj the coordinate of its center, so that Vj=˙xj is the velocity of the j-th sphere. The geometry of the system entails ˙Lj=Vj+2Vj+1 for all j=0,,N.

    In order to effectively swim, our N-spring swimmer undergoes periodic harmonic but non-reversible deformations, just like the original swimmers from Najafi and Golestanian [14], and Montino and DeSimone [13]. However, due to the geometry, we expect a wave to propagate along the tail. This is the behaviour of this wave that we aim at describing in the remainder of the paper.

    In a first approximation, we consider the case where the hydrodynamic force on the j-th sphere only depends on the speed of that same sphere, and neglect interactions between spheres. This leads to the following set of equations on (fluid) forces and velocities:

    {fFj=6πμaVj for j3,fFj=6πμa1Vj for j=1,2. (2.1)

    The elastic forces on each sphere can be written as:

    {fR2=k(L1h)=k1NfRj=k((Lj1h)(Lj2h)),=kj1j2N for 3jN+1fRN+2=k(LNh)=kNN. (2.2)

    At low Reynolds number, inertial forces are negligible. This, together with the fact that the artificial swimmer is self-propelled, gives:

    {fF1++fFN+2=0,fRj+fFj=0 for j3. (2.3)

    Using (2.1)–(2.3), we obtain the expression of fluid forces on each sphere with respect to the length of the adjacent arms. In particular, for the first two spheres:

    {fF1fF2=6πμa1(V2V1)=6πμa1˙L0,fF1+fF2=fR3++fRN+2=k1/N, (2.4)

    which finally leads to:

    {fF1=12(+6πμa1˙L0˜k1),fF2=12(6πμa1˙L0˜k1). (2.5)

    In order to write the equations governing the system, we use Eqs (2.1)–(2.5) to find ODEs on the elongation lj(t) of the j-th arm, for j1. We first consider the case j2. Writing ˙Lj=Vj+2Vj+1=16πμa(fRj+2fRj+1), one deduces

    ˙j=Λ2Kj12j+j+1h2,2jN, (2.6)

    where we have added a fictitious variable

    N+1=0, (2.7)

    and with K=˜k6πμ˜a.

    To determine the equation for the first elastic arm, we use the fact that ˙L1=V3V2=16πμafF3+16πμa1fF2 to obtain, using Eqs (2.2) and (2.5):

    h˙1=Λ2K21hΛK˜a2a11Λ2˙L0. (2.8)

    We can easily verify that the ODE Problem (2.6), (2.7), (2.8) is well-posed using Cauchy-Lipschitz theorem, and provides a unique solution (j(t))1jN+1 for any initial configuration.

    Seeking for periodic (complex) solutions to Eq (2.6) leads to

    j(t)=(αdγj1++βdγj1)eiωt, (2.9)

    where αd,βdC and

    γ±=i/(KωN2)+2±Δ2 (2.10)

    and Δ=1K2ωN4+4iKωN2, where Kω=Kω=˜k6πμ˜aω is an adimensional number. Notice that |γ+|>1 while |γ|<1. The constants αd and βd may be determined through the boundary conditions. Namely assuming, from the linearity of the problem, 1=bdeiωt, with bdC and recalling lN+1=0 enables us to write

    {1(t)=bdeiωt=eiωt(αd+βd),N+1(t)=eiωt(αdγN++βdγN)=0, (2.11)

    to finally obtain

    αd=γNbd(γN+γN),βd=γN+bd(γN+γN). (2.12)

    Then, we use (2.8) to determine bd:

    bd=εi/2i/N+NKω(1zd)+Kω˜a2a1, (2.13)

    where zd=γN+γγNγ+γN+γN.

    As we increase the number of springs in our swimmer, a limit model arises, with an elastic-like tail, as shown in Figure 2. This elastic tail compresses and dilates itself in the same way that the springs do, following the active arm, in order to create a global displacement of our swimmer.

    Figure 2.  Continuous model of the low-Reynolds-number elastic swimmer. Color variations in the tail indicate compression and expansion of the swimmer.

    Equations (2.6) and (2.8) can be viewed as a finite element discretization of a PDE, which describes the continuous version of our swimmer. Limit expressions for this PDE model are formally derived throughout this section while the convergence of the N-spring model to the continuous model will be proven in Section 3.

    First, as h0 (N), j12j+j+1h2 formally converges to a second order derivative. More precisely, we introduce a new space variable yj=(j1)h for 1jN+1. The points yj are equally spaced and thus different from the previous xj. Since y1=0, the y variable can be seen as a local space coordinate attached to the second sphere, and we assume (yj)=j for a smooth enough function . Passing to the formal limit in (2.6) leads to a heat equation:

    t(y,t)=KΛ2yy(y,t),(y,t)[0,Λ]×R+. (2.14)

    Concerning the boundary conditions, we first notice that N+1=0 leads to (Λ,t)=0 for all t>0. As h0, the Eq (2.8) on 1 formally becomes a Fourier-type boundary condition:

    Λ2Ky(0,t)ΛK˜a2a1(0,t)=Λ2˙L0(t),t>0.

    Therefore, we finally obtain the following continuous problem:

    Find C2([0,Λ]×R+) such that (y,t)(0,Λ)×R+,

    {t(y,t)Λ2Kyy(y,t)=0,Λ2Ky(0,t)ΛK˜a2a1(0,t)=Λ2˙L0(t),(Λ,t)=0. (2.15)

    Equation (2.15) belongs to the class of problem for which the classical theory of parabolic equations applies. Namely, calling

    V={uH1((0,Λ))|u(Λ)=0},

    which is a Hilbert space with the scalar product (u,v)V=Λ0yuyvdy, the variational formulation reads:

    Let T>0, find (y,t)Lt(0,T;L2y((0,Λ)))L2t(0,T;V) such that for all t(0,T) and for all vV

    ddtΛ0vdy+Λ2KΛ0yyvdy              +ΛK˜a2a1(0,t)v(0)=Λ2˙L0(t)v(0) (2.16)

    with (y,0)=0(y)L2((0,Λ)) a given initial data.

    Defining the bilinear form κ in V×V as:

    κ:(u,v)Λ2KΛ0yu(y)yv(y)dy+ΛK˜a2a1u(0)v(0), (2.17)

    which is symmetric and coercive on V, well-posedness of the Problem (2.16) follows from standard results on parabolic equations (see e.g., [17]). Moreover, it is well known that the solution (,t) is of class C([0,Λ]) for any time t>0.

    Let us now solve the System (2.15) using the following ansatz (y,t)=_(y)eiωt. From (2.14) we deduce the following equation:

    i_=Λ2Kωyy_. (2.18)

    The characteristic polynomial associated to (2.18) has two roots, r:=1+iΛ2Kω and r, which leads to the following solutions:

    _(y)=αery+βery, (2.19)

    with α,βC.

    We then determine α and β using boundary conditions:

    {(α+β)˜a2a1+Λr(αβ)=iε2Kω,αerΛ+βerΛ=0,

    i.e.,

    {α=iε2Kω(˜a2a1(e2rΛ1)+Λr(e2rΛ+1)),β=e2rΛα. (2.20)

    We notice that rΛ=1+i2Kω only depends on Kω.

    We first notice that the discrete Problem (2.6) is a kind of non conventional mass-lumped version of a finite element discretization of the continuous one (2.15). In order to clarify this statement, we introduce the finite element setting. Let VhV the space of continuous, piecewise linear functions g on the one-dimensional partition Th={y1,,yN+1} of (0,Λ), and that verify the Dirichlet boundary condition g(Λ)=0. Let {Φj}j=1,N be the standard basis for Vh consisting of the hat functions defined by Φj(yk)=δj,k for 1j,kN.

    Let hVh be the continuous, piecewise linear function such that for 1jN+1, t>0, h(yj,t)=j(t). Using the basic semi-discrete Galerkine method would lead to the discretization of (3.1) in the matrix form:

    d(MhLh)dt+KhLh=˜f(t), (3.1)

    with Lh(t)=(1(t),,N(t))T. Similarly, ˜f=(Λ2˙L0,0,,0), (Mh)i,j=Λ0Φi(y)Φj(y)dy and (Kh)i,j=κ(Φi,Φj), where κ is defined in Eq (2.17).

    Computing explicitly the coefficients of the matrices Kh and Mh gives

    (Kh)ij={Λ2K/hfor |ij|=1,2Λ2K/hfor i=j2,Λ2K/h+ΛK˜a/(2a1)for i=j=1,

    and

    (Mh)ij={h/6for |ij|=1,2h/3for i=j2,h/3for i=j=1.

    The key observation is that Eqs (2.6) and (2.8) are nothing but a mass-lumped discretization of (2.15) where the mass matrix Mh has been replaced by the diagonal version

    ˜Mh=(h00h).

    Hence, h actually solves

    d(˜MhLh)dt+KhLh=˜f(t), (3.2)

    together with the initial condition

    h(0)=0,hVh. (3.3)

    The classical mass-lumped method, on the other hand, would have consisted in replacing the tridiagonal mass matrix Mh by a diagonal matrix ˉMh using an integration formula on the vertices of the partition. Namely, using the trapezoidal formula

    Λ0g(12g(y1)+Nj=2g(yj)+12g(yN+1))h=(12g(y1)+Nj=2g(yj))h,

    for a function g satisfying g(Λ)=0 leads to the mass-lumped matrix

    ˉMh=(h/20h0h) (3.4)

    which differs from ˜Mh.

    We shall then study the ODE (3.2), (3.3) using the method presented in [18] which provides us with a convergence result for the mass-lumped method with ˉMh.

    We introduce the two following inner products on Vh associated with ˉMh and ˜Mh respectively. Namely, for (uh,vh)Vh

    uh,vhh=h2uh(y1)vh(y1)+hNj=2uh(yj)vh(yj)

    and

    (uh,vh)h=hNj=1uh(yj)vh(yj).

    We also call h the norm associated to (,)h, while the L2 norm and inner products are denoted by and (,) respectively. Gerschgörin Theorem applied to Mh shows the equivalence of the norms and h on Vh uniformly in h, and, more precisely, we have the estimate, valid for all vhVh

    16(vh,vh)h(vh,vh)(vh,vh)h,

    from which we also deduce

    hvh(y1)2vh2h6vh2. (3.5)

    Finally, we introduce, for uh,vhVh, δh(uh,vh)=(uh,vh)h(uh,vh) the quadrature error.

    Lemma 3.1. Let uh,vhVh. We have, for h sufficiently small:

    |δh(uh,vh)|Chyuhyvh, (3.6)
    |δh(uh,vh)|Chyuhvh (3.7)

    for a constant C that does not depend on uh, vh or h.

    Proof. In all what follows, C denotes a constant that may vary from line to line, being always independent of h. Let uh,vhVh. We write |δh(uh,vh)||(uh,vh)huh,vhh|+|uh,vhh(uh,vh)|. Thomée [18] provides us with an estimate of the error between uh,vhh and (uh,vh), namely,

    |uh,vhh(uh,vh)|Ch2yuhyvh

    and

    |uh,vhh(uh,vh)|Chyuhvh

    for some constant C>0 that does not depend on uh, vh or h. The latter estimate is obtained by an inverse inequality.

    It remains to estimate the term ˜δh(uh,vh)=(uh,vh)huh,vhh.

    We notice that:

    |˜δh(uh,vh)|=h2|uh(y1)vh(y1)| (3.8)
    =h2|Λ0yuh(y)dy||Λ0yvh(y)dy|hΛ2yuhyvh. (3.9)

    Similarly, (3.8) together with (3.5) gives:

    |˜δh(uh,vh)|Chyuhvh. (3.10)

    This yields (3.6) and (3.7).

    Theorem 3.1. If and h are solution to (2.16) and (3.2), (3.3) respectively, and 0H2((0,Λ)), we have, for all t0,

    h(t)(t)C0,h0+Ch2(yy0+yy(t))Ch(t0yt2ds)1/2.

    Proof. Let Rh be the Ritz projector from V on Vh, associated with κ(,). Namely, for gV, Rhg is defined by

    κ(Rhg,vh)=κ(g,vh)

    for all vhVh. We write h=(hRh)+(Rh)=θh+ρ (Notice that θhVh). Standard estimations show that ρ(t) satisfies RhCh2yy. In order to estimate θh, we write, for all χhV

    (tθh,χh)h+κ(θh,χh)=(th,χh)h+κ(h,χh)(tRh,χh)hκ(Rh,χh)=(f,χh)(tRh,χh)hκ(,χh)=(t,χh)(tRh,χh)h=(tρ,χh)δh(tRh,χh). (3.11)

    Setting χh=θh, we obtain

    12ddtθh2h+κ(θh,θh)=(tρ,θh)δh(tRh,θh).

    Here, we have at once, using Cauchy-Schwarz and Poincaré inequalities:

    |(tρ,θh)|t(Rh)θhChytθhChytyθh.

    Using the first equation of Lemma 3.1, and the fact that yRhuCyu, we also obtain

    |δh(tRh,θh)|ChytRhyθhChytyθh,

    from which we deduce that

    12ddtθh2h+κ(θh,θh)Chytyθhκ(θh,θh)+Ch2yt2,

    using the coercivity of κ(,) on V. We therefore infer

    θh(t)2hθh(0)2h+Ch2t0yt2ds.

    We now recall that h and are equivalent norms on Vh, uniformly in h, and hence

    θh(t)Cθh(0)+Ch(t0yt2ds)1/2.

    Here θh(0)=0,hRh0 and

    0,hRh00,h0+0Rh00,h0+Ch2yy0,

    whence θh(t) is bounded as desired.

    Theorem 3.2. If and h are solution to (2.16) and (3.2), (3.3) respectively we have, for t0,

    y(h)(t)Ch(yy0+yy(t))+Cy(0,h0)+Ch(t0yt2ds)1/2.

    Proof. We now set χh=tθh in Eq (3.11) for θh to obtain:

    tθh2h+12ddtκ(θh,θh)=(tρ,tθh)δh(Rht,tθh).

    Here, as in the proof of Theorem 3.1,

    |(tρ,tθh)|t(Rh)tθhChyttθh.

    Further, by the second line of Lemma 3.1,

    |δh(tRh,tθh)|ChytRhtθhChyttθh.

    Using again the equivalence between the norms h and on Vh, we conclude:

    tθh2h+12ddtκ(θh,θh)Chyttθhhtθh2h+Chyt2,

    so that, after integration, and using the coercivity of κ(,) on V

    yθh(t)Cyθh(0)+Ch(t0yt2ds)1/2y(0,h0)+Chyy0+Ch(t0yt2ds)1/2.

    This, together with the standard estimate for yρ(t) completes the proof.

    We proved the convergence of the discrete N-spring swimmer to the continuous model we formally derived in the previous section. Note that we obtain only a first-order (resp. half order) convergence in L2 norm (resp. H1 norm) while the standard estimations for the mass-lumping method leads to a second-order (resp. first order) convergence. This is due to the Fourier-type boundary condition at 0 which differs from the classical Dirichlet boundary condition used in [18].

    We seek the swimmer's displacement by looking at the displacement of the first of the largest sphere, meaning we only compute V1=˙x1, and integrate over a period (0,2π/ω).

    Taking into account the hydrodynamic interactions due to the ith-sphere with i{2,,N+2} on the first sphere, we have

    V1=16πμa1fF1+14πμL0fF2+14πμN+2i=3fFiL0+L1++Li2

    Using expressions (2.2) and (2.5), we obtain

    V1=12˙L0˜a2a1K13a1˙L04L03K˜a14L0+3˜aK2Nj=1jj+1ji=0Li, (4.1)

    where we recall that, by convention, N+1=0.

    Finally, by integrating over one period, and noticing that both 2 and ˙L1/L1 have a vanishing time-average, we obtain, for any value of h=Λ/N, the displacement of the corresponding N-spring swimmer:

    Δhx1=2π/ω0[3K˜a14L0+3˜aK2Nj=1jj+1ji=0Li]dt (4.2)

    We may find an expression for the displacement of the limit model as h tends to 0, by passing to the limit in the preceding expression.

    Indeed, for h and y given, we define jh(y) the unique integer such that jh(y)hy(jh(y)+1)h. Then, defining χh the function

    χh(y,t)=1L0(t)++Ljh(y)+1(t),

    we may write

    2π/ω0N1j=0h(jh,t)h((j+1)h,t)j+1i=0Lidt=2π/ω0Λ0yh(y,t)χh(y,t)dydt,

    where h is the piecewise linear function defined in the previous section.

    Finally, the displacement Δhx1 of the N-spring swimmer during one time period can be rewritten as

    Δhx1=2π/ω0[3K˜ah(0,t)4L0(t)3˜aK2Λ0yh(y,t)χh(y,t)dy]dt.

    Now, using the fact that jh(y)hy when h0, together with the L2 and H1 convergence of h to , we obtain that, for any y and t,

    χh(y,t)=1L0(t)+(jh(y)+1)h+hΛji=0h(ih,t)h01L0(t)+y+y0(t)Λ=:χ(y,t)

    Moreover 0χh(y,t)maxt1L0(t)=1L(1˜ε), shows that χh is uniformly bounded. Therefore, using dominated convergence theorem, we deduce that χh converges to χ in L2(0,2π/ω;(0,Λ)) as h tends to 0.

    Using the convergence theorems proven in the preceding section, we may pass to the limit h0 in Δhx1, and obtain the following expression for the displacement during one period for the limit model

    Δx1=2π/ω0Λ03K˜a2y(y,t)(L0(t)+y+y0Λ)1dydt2π/ω03K˜a(0,t)4L0dt. (4.3)

    In this section, we numerically study the discrete model's convergence towards the continuous one. Then, we investigate the influence of the two parameters ω and ˜ε on the system and on its displacement, while the rest of the swimmer is determined by the values in Table 1. All simulations are achieved using Matlab. We consider here that the default length L of the active arm is small compared to the rest of the swimmer. The first sphere thus acts like the head of a sperm cell, and the active arm like a link between the head and the flagella, which gives a signal so that the rest of the system oscillates.

    Table 1.  Values of the parameters used in the numerical simulations, matching those of [13]. We have taken for μ the dynamic viscosity of water at 25C.
    ˜a 1105m
    a1 1105m
    Λ 4104m
    L 3105m
    ˜k 1108Nm1
    μ 8.9104Pas

     | Show Table
    DownLoad: CSV

    We investigate numerically the convergence estimations obtained in Section 3. We recall that the continuous solution solves the heat equation PDE with the Fourier-type boundary conditions (2.15). We consider, in this section, periodic forcing for which explicit solutions are given by (2.19) and (2.20).

    We recall that the discrete solution h is the P1 discrete function based on the (i)i solution to the N-spring ODE System (2.6), (2.7) and (2.8). This discrete system corresponds to a semi-discretization in space of the continuous model, based on a non conventional mass-lumping method. The solution (i)i of the discrete problem in the periodic setting is given in Eqs (2.9), (2.10), (2.12) and (2.13).

    The space step h (or equivalently the number of springs N) being given, the discrete error is defined as the error between h and the P1 interpolation of . We plot in Figure 3, the L2 (resp. H1) error denoted by eh,L2 (resp. eh,H1).

    Figure 3.  L2 and H1 errors between the N-spring discrete model and the continuous one as a function of the number of springs in log scale, in the (2π/ω)-periodic case, for ˜ε=0.7 and ω=1rads1.

    We observe that the L2 error converges with order one, as expected from Theorem 3.1. Concerning the H1 error, we observe a superconvergence phenomenon: as the L2 error, it converges at order 1, while Theorem 3.2 predicts a convergence at order 1/2. This can be explained by the regularity of the considered periodic solution.

    As mentioned earlier, the N-spring model turns out to be a discretization in space of the continuous Problem (2.15), based on an unconventional mass-lumping method. The convergence proof that we proposed in Section 3 is based on the results of Thomée[18]. He shows that, for a standard mass-lumping discretization, the usual order of convergence for finite elements is obtained: convergence of order 2 for the L2 error and 1 for the H1 error.

    We investigate here the influence of the space discretization, by comparing the N-spring model (3.2), solved numerically this time, to the classical mass-lumping method (3.4) and the standard Galerkin finite element method (3.1). Again we consider the periodic framework for which the exact solution is available. The time discretization of the three ODE systems is achieved using a Crank-Nicolson scheme for which the time step is chosen to be small enough so that the error due to the time discretization is negligible.

    The corresponding L2 (resp. H1) error is given on Figure 4 (resp. Figure 5). We can see that, as expected, the L2 error converges at order 1 for the N-spring model, while it converges at order 2 for both the classical mass-lumping method and the standard Galerking discretization. Again, due to the regularity of the solutions, a super-convergence phenomenon of the H1 error is observed for all three methods: as the L2 error, it converges at order 1 for the N-spring model and order 2 for the other two discretizations.

    Figure 4.  L2 error between the continuous model and our mass-lumping method, as a function of the number of springs, in log scale.
    Figure 5.  H1 error between the continuous model and our mass-lumping method, as a function of the number of springs, in log scale.

    In this section, we investigate the swimming ability of the N-spring swimmer. The stroke being periodic, we use the explicit solutions given in Section 2.2. The computations are achieved for N=2,000 springs.

    Figure 6 shows a full stroke of the swimmer, in which we notice that a wave is propagating along its tail. Remember that this wave is a contraction wave along the horizontal tail. This tail appears to be oscillating fairly efficiently for the side close to the head, while the amplitude of the contraction decays considerably on the second half of the tail.

    Figure 6.  Movement of the whole 2,000-springs swimmer during a full stroke, at different time stamps T, for ω=1rads1 and ˜ε=0.7.

    The movement shown corresponds to the stretch of j, and not to the actual deformation which would be j/N, for all 1jN. We thus remark that this deformation is relatively small compared to the size of the artificial swimmer, which matches the approximation of small deformations that we made in the first place.

    In this section, we study the influence of the parameters ˜ε and Kω on the swimmer's displacement (4.2), in order to maximize its absolute value.

    In Figure 7, we plot the displacement of the swimmer as a function of time, for different values of ˜ε. The displacement is computed through numerical integration of Eq (4.2). The graph shows that the swimmer globally swims backwards, and we recognize the back and forth motion which is characteristic of low Reynolds number artificial swimmers. A larger amplitude ˜ε of the forcing leads to a larger displacement and we observe (see Figure 8), that Δx1 is proportional to ˜ε2, which is what is expected in theory (similar behaviors are observed, e.g., in [6,7,10] and explained as the surface of loops in the space of shapes [1]).

    Figure 7.  Displacement of the 2,000-spring swimmer against time t, for different values of ˜ε.
    Figure 8.  Displacement of the 2,000-spring swimmer depending on ε for an arm oscillating frequency ω=1rad/s compared to y=˜ε2, in log-scale.

    As we want to maximize Δx1 while having ˜ε<1, we choose a fixed value ˜ε=0.7 which, although arbitrary, allows for an easier comparison to Montino and DeSimone's results [13], as they made a similar parameter choice.

    Figure 9 shows Δx1 depending on Kω, for different values of ˜ε. At any fixed Kω, we observe once again that larger ˜ε leads to larger Δx1. We first observe that, if Kω, the net displacement of the swimmer vanishes. According to the expression of Kω, this is the case for example when ω0: the oscillation disappears, immobilizing the artificial swimmer. This can also happen when ˜k: the springs become so rigid that the tail of the swimmer can no longer deform. In that case, the swimmer has only one degree of freedom left to deform and faces Purcell's Scallop theorem's obstruction. Similarly, letting Kω0 immobilizes the swimmer. An optimal value Koptω for the non-dimensional parameter is reached between these two limiting cases, in order to maximize the displacement on one time period. According to the figure, Koptω0.3765. A complete mathematical expression of Koptω does not seem available, due to the largely nonlinear nature of the problem contrarily to the final expression obtained in [13]. A pair of optimal values for ω and k to obtain this Koptω are ω=1rads1 and ˜k6.207108Nm1. Moreover, the expression of Kω guarantees that ω must vary proportionally to ˜k for the pair (˜k,ω) to remain at the optimum. Indeed, the softer the spring, the slower the first arm needs to oscillate in order to generate a large movement.

    Figure 9.  Displacement of the 2,000-spring swimmer depending on Kω, for different values of ˜ε.

    Looking at the other parameters separately, we can also clearly see from Eq (4.2), that the displacement depends linearly on ˜a, which is predictable. However, this parameter has a direct consequence on the size of the artificial swimmer and must stay in a reasonable range (in our case no more than 1e5m) so that the swimmer stays at microscopic scale.

    Finally, we notice that the value of Λ and the ratio a1/˜a has little to no influence on the previous analysis. We therefore keep for those parameters values that seem coherent with the scale we are working at, and that match with numerical experiments provided in [13].

    We analyzed the dynamics of two low-Reynolds-number swimmers. The first one, which is an extension of [13], is made of N passive springs, and the second one is the corresponding limit model with an elastic tail. Both are activated by an active arm that elongates and retracts periodically with amplitude ε and angular frequency ω.

    Noting that the N-spring swimmer is a non-conventional mass lumping discretization of the limit model, we proved its convergence, when N tends to infinity, to the continuous model, by extending the results of Thomée [18] to the case of a Fourier-type boundary condition. For both swimmers, a phase difference between the oscillations of the active arm and the tail is created by the interaction between elastic and hydrodynamic forces. Then, both models undergo non-reciprocal shape changes and thus circumvent Scallop Theorem's obstruction [16]. Numerical simulations indeed show a wave propagating along the swimmers' tails. Similarly to what was shown in [13], our models are able to swim but there is no control over the swimming direction.

    Then, we focused on computing the net displacement over a time period of the swimmer in both cases, in view of its optimization. We obtain explicit formulae for this displacement as a function of the local elongation during the stroke. We numerically recover the classical back and forth swimming and the second-order scaling of the displacement as a function of the maximum elongation of the forcing active arm. Moreover, we highlight a dimensionless parameter Kω, driving the movement of the swimmer when its geometry (Λ, a, a1) is given. Some optimal values for this parameter can be estimated by numerical experiments.

    Lastly, we noticed that, similarly to the behavior of Machin's swimming rod [12], the deformations of both our swimmers is rapidly attenuating along their passive parts, which suggests that some form of activation is needed in order to mimic the type of behavior observed in nature.

    The authors declare no conflict of interest.



    [1] F. Alouges, A. DeSimone, A. Lefebvre, Optimal strokes for low Reynolds number swimmers: an example, J. Nonlinear Sci., 18 (2008), 277–302. https://doi.org/10.1007/s00332-007-9013-7 doi: 10.1007/s00332-007-9013-7
    [2] F. Alouges, A. DeSimone, A. Lefebvre, Optimal strokes for axisymmetric microswimmers, Eur. Phys. J. E, 28 (2009), 279–284. https://doi.org/10.1140/epje/i2008-10406-4 doi: 10.1140/epje/i2008-10406-4
    [3] F. Alouges, A. DeSimone, L. Heltai, Numerical strategies for stroke optimization of axisymmetric microswimmers, Math. Mod. Meth. Appl. Sci., 21 (2011), 361–387. https://doi.org/10.1142/S0218202511005088 doi: 10.1142/S0218202511005088
    [4] F. Alouges, A. DeSimone, L. Heltai, A. Lefebvre-Lepot, B. Merlet, Optimally swimming stokesian robots, Discrete Cont. Dyn. Syst. B, 18 (2013), 1189–1215. https://doi.org/10.3934/dcdsb.2013.18.1189 doi: 10.3934/dcdsb.2013.18.1189
    [5] F. Alouges, A. DeSimone, L. Giraldi, Y. Or, O. Wiezel, Energy-optimal strokes for multi-link microswimmers: Purcell's loops and Taylor's waves reconciled, New J. Phys., 21 (2019), 043050. https://doi.org/10.1088/1367-2630/ab1142 doi: 10.1088/1367-2630/ab1142
    [6] J. E. Avron, O. Gat, O. Kenneth, Optimal swimming at low Reynolds numbers, Phys. Rev. Lett., 93 (2004), 186001. https://doi.org/10.1103/PhysRevLett.93.186001 doi: 10.1103/PhysRevLett.93.186001
    [7] J. E. Avron, O. Kenneth, D. H. Oaknin, Pushmepullyou: an efficient micro-swimmer, New J. Phys., 7 (2005), 234. https://doi.org/10.1088/1367-2630/7/1/234 doi: 10.1088/1367-2630/7/1/234
    [8] S. Childress, Mechanics of swimming and flying, Cambridge University Press, 1981. https://doi.org/10.1017/CBO9780511569593
    [9] A. DeSimone, L. Heltai, F. Alouges, A. Lefebvre-Lepot, Computing optimal strokes for low Reynolds number swimmers, In: Natural locomotion in fluids and on surfaces, New York, NY: Springer, 2012,177–184. https://doi.org/10.1007/978-1-4614-3997-4_13
    [10] R. Dreyfus, J. Baudry, H. Stone, Purcell's "rotator": mechanical rotation at low Reynolds number, Eur. Phys. J. B, 47 (2005), 161–164. https://doi.org/10.1140/epjb/e2005-00302-5 doi: 10.1140/epjb/e2005-00302-5
    [11] E. Lauga, T. R. Powers, The hydrodynamics of swimming microorganisms, Rep. Prog. Phys., 72 (2009), 096601. https://doi.org/10.1088/0034-4885/72/9/096601 doi: 10.1088/0034-4885/72/9/096601
    [12] K. E. Machin, Wave propagation along flagella, J. Exp. Biol., 35 (1958), 796–806. https://doi.org/10.1242/jeb.35.4.796 doi: 10.1242/jeb.35.4.796
    [13] A. Montino, A. DeSimone, Three-sphere low-Reynolds-number swimmer with a passive elastic arm, Eur. Phys. J. E, 38 (2015), 42. https://doi.org/10.1140/epje/i2015-15042-3 doi: 10.1140/epje/i2015-15042-3
    [14] A. Najafi, R. Golestanian, Simple swimmer at low Reynolds number: three linked spheres, Phys. Rev. E, 69 (2004), 062901. https://doi.org/10.1103/PhysRevE.69.062901 doi: 10.1103/PhysRevE.69.062901
    [15] E. Passov, Y. Or, Dynamics of Purcell's three-link microswimmer with a passive elastic tail, Eur. Phys. J. E, 35 (2012), 78. https://doi.org/10.1140/epje/i2012-12078-9 doi: 10.1140/epje/i2012-12078-9
    [16] E. Purcell, Life at low Reynolds number, Amer. J. Phys., 45 (1977), 3–11. https://doi.org/10.1119/1.10903 doi: 10.1119/1.10903
    [17] P. Raviart, J.-M. Thomas, Introduction à l'analyse numérique des équations aux dérivées partielles, Masson, 1988.
    [18] V. Thomée, Galerkin finite element methods for parabolic problems, Berlin, Heidelberg: Springer, 2006. https://doi.org/10.1007/3-540-33122-0
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1651) PDF downloads(99) Cited by(0)

Figures and Tables

Figures(9)  /  Tables(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog