Significance

The myoneme-based contractile apparatus, responsible for the ultrafast motions of certain protozoa, presents an excellent opportunity for engineering control over subcellular biomechanics. However, its basic dynamical properties remain poorly understood. Unlike actomyosin contraction, which is ATP-driven and has been extensively modeled, myoneme-based contraction is Ca2+-driven and lacks a comprehensive mechanochemical description that covers its manifestations in various organisms. In this study, we derive a minimal continuum model that semiquantitatively explains experimental measurements and use it to uncover distinct operating regimes of myoneme-based contraction. We demonstrate that the dynamics rely on three timescales representing chemical driving, material stiffness, and viscous drag. These findings carry broad implications for mechanochemical force generation, protozoan evolution, and synthetic biology.

Abstract

In nature, several ciliated protists possess the remarkable ability to execute ultrafast motions using protein assemblies called myonemes, which contract in response to Ca2+ ions. Existing theories, such as actomyosin contractility and macroscopic biomechanical latches, do not adequately describe these systems, necessitating development of models to understand their mechanisms. In this study, we image and quantitatively analyze the contractile kinematics observed in two ciliated protists (Vorticella sp. and Spirostomum sp.), and, based on the mechanochemistry of these organisms, we propose a minimal mathematical model that reproduces our observations as well as those published previously. Analyzing the model reveals three distinct dynamic regimes, differentiated by the rate of chemical driving and the importance of inertia. We characterize their unique scaling behaviors and kinematic signatures. Besides providing insights into Ca2+-powered myoneme contraction in protists, our work may also inform the rational design of ultrafast bioengineered systems such as active synthetic cells.
A well-known molecular mechanism for motion in biological systems is actomyosin contractility, which involves the relative sliding of actin filaments due to ATP hydrolysis by bound molecular motors (1). However, the fastest motions currently known to occur in the biological world (measured as the proportional change in length per unit time) are due not to actomyosin assemblies but to lesser-known protein assemblies called myonemes (24). We use “myonemes” in the broadest sense (5), encompassing terms used for specific organisms such as “spasmonemes” and “M-bands.” Myonemes are common in ciliated protists: For example, they are found in the wine glass-shaped genus Vorticella (6, 7), the cigar-shaped genus Spirostomum (811), the trumpet-shaped genus Stentor (12, 13), the tear drop-shaped genus Lacrymaria (14), and the chandelier-shaped genus Zoothamnium (15, 16). Similar Ca2+-responsive supramolecular assemblies are proposed to occur in a range of other single-celled organisms (1724), and possibly even plants (2527). The mechanism of ultrafast myoneme contraction is qualitatively different from actomyosin contraction in that ATP is not directly consumed during the contractile motion (6, 28). Instead, a local increase in Ca2+ concentration acts as a mechanical actuator to initiate contraction (2932). ATP is subsequently consumed to pump Ca2+ away from the myoneme fibers and reset the process (15, 33).
While several experimental studies have been conducted on myoneme-based contraction dynamics (710, 15, 3442), it is difficult to draw general conclusions without a theoretical framework to interpret the observations. Whereas actomyosin contractility has been extensively studied theoretically and computationally (4348), Ca2+-powered myoneme-based contractility is relatively unexplored from a modeling point of view. An existing computational model of Vorticella (49) is specific to the geometry of the organism and thus does not apply across species. Previous mechanochemical analyses also do not address the dynamics of contraction (6, 9). Questions that a mathematical model of myoneme contraction might address are: What synchronizes contraction across the organism’s length, what sets the rate of chemical driving, and what are the key geometric and mechanical factors that affect the contraction dynamics?
Here, we develop a minimal mathematical model of myoneme contraction and show that it applies across species (Fig. 1). In our model, an elastic medium in a dissipative environment is driven out of equilibrium by a traveling wave of shrinking rest lengths, which captures the effect of a traveling Ca2+ wave. This simple setup omits several system-specific details but can fit experimental contraction measurements for Spirostomum and Vorticella. Mathematically, the model comprises a wave equation with dissipation and a nonautonomous source term. We derive an approximate form for the nonautonomous function by analyzing an auxiliary kinetic model for a traveling Ca2+ wave. Using the model, we show that the balance between chemical driving, mechanical stiffness, and viscous drag results in three dynamic regimes. These regimes are distinguished by the relative importance of inertia and Ca2+ wave speed, and they exhibit kinematic signatures, including asymmetric contraction and different scaling behavior of the maximum contraction rate. This work provides a starting point for extracting general mechanochemical principles of myoneme contraction in biological and engineered systems.
Fig. 1.
Myoneme-based contraction in ciliated protists. (A) Images of Spirostomum undergoing contraction, with the time in milliseconds labeled on the bottom left of each image. Rapid contraction begins at T = 0 ms and is followed by a slower reelongation phase (not depicted). Note that the left side of the organism begins contracting (between 0.5 and 1.0 ms) before the right (between 1.0 and 1.5 ms). (B) Same as (A) for Vorticella. (C) Schematic of the model. The contractile body of Spirostomum and stalk of Vorticella are abstracted as a 1D elastic medium in a viscous environment, represented as a chain of springs. The local spring rest length (depicted as the number of ridges) depends on the number of bound Ca2+ ions (shown as blue spheres bound to receptors on the springs). A rightward traveling Ca2+ wave thus shrinks the spring rest lengths, exerting forces in the chain which then contracts against viscous drag. The bottom graph schematically illustrates several quantities in our mechanochemical model at a fixed time, which we elaborate on in the main text.

1. Model Derivation

Myoneme contraction is driven by an increase in cytosolic Ca2+, which occurs as a traveling wave that traverses the organism’s length in several milliseconds (30, 4952). The mechanism by which Ca2+ physically induces contraction is believed to be a screened entropic spring effect: Ca2+ binds to the negatively charged myoneme fibers and screens their electrostatic self-repulsion, which allows for the entropic contribution to the free energy to induce collapse of the polymer (53). This system is sometimes considered a polyelectrolyte gel, in which thermodynamic balances of entropic and osmotic forces determine the gel’s equilibrium (15, 54, 55). The precise contributions to the myoneme free energy as a function of length are not needed to formulate a minimal mechanical model, and we make the phenomenological assumption that the system obeys linear elasticity (56). The key difference from the usual elastic theory is that the rest lengths of the elastic elements are dynamic quantities that depend on the local concentration of bound Ca2+ (Fig. 1C).
To formulate the model mathematically, we start from a discrete chain of harmonic springs in one dimension with externally imposed rest lengths, and we define Xi(T) as the position of the ith mass at time T. The spring stiffness parameter k, mass m, drag γ, and initial rest length δ, are the same for all springs, and the instantaneous rest length Gi+12(T) of the spring connecting masses i and i + 1 is a nonautonomous function which we specify later. The force on mass i is
mT2Xi=&kXi+1XiGi+12kXiXi1Gi12γTXi.
[1]
Rather than a discrete description in terms of the index i, we seek a continuum description of the dynamics in terms of a material coordinate S ∈ [0, L], where L is the uncontracted length of the system. By Taylor expanding the functions X and gG/δ to second order in δ and then letting δ → 0, we find the following partial differential equation (PDE) for the position X(S, T):
T2XαS2XSg+μTX=0.
[2]
Here, α=kδ2/m=EMA/m˜ is the product of Young’s modulus EM and the effective spring cross-sectional area A divided by the linear mass density m~=m/δ [where we used k = EMA/δ (56)]. The parameter α can be interpreted as the square of the speed of sound in the spring chain (57). μ=γ/δm~ is the ratio of the drag constant per unit length to the linear mass density, and g(S, T) is a nonautonomous function for the active stretch, defined as the ratio of the current spring rest length G(S, T) to the initial rest length δ. The quantities g, α, and μ remain finite as δ → 0 because Gδ, kδ−1, mδ and γδ, while EM, A, and m~ are independent of δ. Eq. 2 represents the force balance of X(S, T), where the first term is the acceleration, the second term (proportional to α) captures the force due to the difference between the local stretch and the active stretch g, and the third term captures the drag force. See SI Appendix, Supplementary Methods, section A for a detailed derivation of the model.
We find the boundary conditions by retaking the continuum limit at the terminal masses in the discrete representation Eq. 1, noting that these masses are connected to only a single spring. This yields the stress-free condition
SX(S,T)=g(S,T)atS=0,L.
[3]
In the case of Vorticella, which has one end (say S = L) attached to a fixed substrate, we use the boundary condition X(L, T)=L at that end. We further take the initial conditions
X(S,0)=SandTX(S,0)=0.
[4]
We model the active stretch g(S, T) as a linear function of the local concentration of Ca2+ bound to the myoneme fibers; see 17. Specifying the active stretch function g(S, T) requires describing the Ca2+ dynamics, for which we adapt a tractable reaction–diffusion model of a traveling Ca2+ wave (58), as described in Materials and Methods, Ca2+ dynamics and SI Appendix, Supplementary Methods, section B. By analyzing this model, we find that a good approximation to the active stretch function g(S, T)=g*(S − VT; W) is a traveling wave moving at speed V and having a sigmoidal profile g* with width W Eq. 22; Fig. 1C). This function decreases from 1 to gmin ∈ [0, 1] as the wave passes through the organism. An additional parameter O sets the time offset such that the center of the wave arrives at S = 0 at time O. As described in Results, section B, we can approximately express V and W in terms of the kinetic parameters of the Ca2+ model (Eqs. 5 and 6).
Our model neglects contributions from transverse dimensions and twisting effects, which may alter the one-dimensional dynamics through elastic coupling in the real system (59, 60). As explained in SI Appendix, Supplementary Methods, section C, we assume that this coupling is negligible as a first approximation. In that section, we also discuss the explicit assumptions needed to disregard the poroelastic contribution arising from fluid motion around the organelles inside the contracting organism. We demonstrate that under standard approximations made when treating cytoplasm poroelasticity (61), poroelasticity in our model contributes an additive term to the Stokes drag coefficient (SI Appendix, Supplementary Methods, Eq. 50). Moreover, we examine a model extension in which motion encounters resistance from auxiliary elastic elements with unchanged rest lengths during Ca2+ binding. This representation approximates, for instance, the microtubule sheath under the cortex of Spirostomum (11, 36) or the elastic stalk supporting the thin myoneme fiber in Vorticella (49). We show that a linear change of variables effectively maps the model with auxiliary elastic elements to the model presented here.

2. Results

A. A Minimal Continuum Model Reproduces Contraction Kinematics.

Despite the model’s simplicity, we find that it reproduces the contraction kinematics of Spirostomum and Vorticella semiquantitatively. We illustrate this by comparing the trajectories of normalized length λ(T)≡(X(L,T)−X(0,T))/L from experimental measurements to fitted model curves. Contraction events for Spirostomum and Vorticella are shown in SI Appendix, Movies S1 and S2, and the experimental methods are described in Materials and Methods, Experimental methods. The second-order nonautonomous PDE model does not have an exact closed-form solution, though we describe exact series solutions in two limiting cases in Results, section C. To fit the model, we numerically integrate the PDE with different choices of parameters and compare the results to the experimental traces. See Materials and Methods, Computational methods for details on the computational methods. In Fig. 2A, we show five contraction measurements in Spirostomum along with a model solution fitted to their average, illustrating the feasibility of our model in capturing the contraction dynamics. The corresponding contraction velocities |∂Tλ| are shown in Fig. 2B. We also fit the model with a fixed boundary condition at S = L to data for five Vorticella contraction events (Fig. 2 C and D).
Fig. 2.
Comparison of the model to experimental data. (A) Five measurements of Spirostomum contraction, collapsed to the same scale by normalizing length using the organism’s uncontracted length L and by offsetting time so that contraction begins around T = 10 ms. Thin lines represent individual contraction events, and we color each differently. The thick black curve shows the model fit to all of these trajectories at once; see SI Appendix, Supplementary Results, section B for fits to the individual curves. Vertical dashed lines indicate the times in the model solution when the center of the active stretch wave arrives at S = 0 and S = L. Time is shifted by choosing O = 10 ms in Eq. 22 to achieve numerical agreement between the boundary and initial conditions (see Material and Methods, Computational methods). (B) The corresponding rate trajectories, obtained by numerically differentiating the data in (A). Colors of individual events are the same as in (A). (C and D) The same as (A) and (B) but for Vorticella contraction. (E) Simulated length trajectories with V = 1 L/ms, α = 2 (L/ms)2, gmin = 0, O = 10 ms, and μ = 10 ms−1. The width W of the wave is varied from 0.1 L to 1.0 L in steps of 0.1 L as the colors range from blue to red. The inset shows how W depends on the binding rate K+ in the Ca2+ model. The total myoneme concentration Btot ranges from 500 to 4, 000 μM in multiples of 2 as the colors range from light green to purple. The dashed line in the inset represents the analytically predicted scaling WK+−1/2. For the remaining parameters used in the Inset, see SI Appendix, Table S4.
Movie S1.
Contraction of Spirostomum, along with output from the contour-tracking Morphometrics software used to measure the change in length over time.
Movie S2.
Contraction of Vorticella, along with output from the contour-tracking Morphometrics software used to measure the change in length over time.
In SI Appendix, Figs. S6 and S7, we present fits for each of the ten individual trajectories shown in Fig. 2 A and C, demonstrating that we can obtain excellent fits for all trajectories using fairly consistent parameters for each organism. As we explain in Results, section C, we find that these contraction events occur in a regime where inertia is unimportant, meaning that the dynamics are approximately determined only by the ratio μ/α. We fit the simplified “slow-wave” model (cf. Eq. 10) and obtain a typical set of parameters for Spirostomum: L = 0.58 mm, gmin = 0.35, μ/α = 5 ms/mm2, V = 0.8 mm/ms, and W = 0.18 mm. For Vorticella, a typical set of parameters is L = 0.27 mm, gmin = 0.32, μ/α = 1 ms/mm2, V = 0.23 mm/ms, and W = 0.08 mm. In SI Appendix, Supplementary Results, section B, we demonstrate that the fitted parameters for Vorticella are consistent in order of magnitude with physiological parameters used in a detailed computational model in ref. 49. According to a two-tailed t test, the distributions of quantities μ/α, V, and W each have means that are statistically different for the two organisms (t test P-value ≪0.05). In the following section, we investigate the kinetic factors that underlie the rate of chemical driving captured by the parameters V and W.

B. Ca2+ Dynamics Control Contraction.

The speed V and, less intuitively, the width W of the active stretch wave g(S − VT; W) determine how rapidly the length changes over time. As W decreases, the active stretch wave approaches a traveling step function in which the springs ahead of the wave have their original rest lengths, and those behind have rest lengths which are reduced by a factor gmin. As W increases, the distribution of changing rest lengths widens, and the driving and relaxation become more spatially and temporally homogeneous. We illustrate this in Fig. 2E, where we show the length trajectory for several values of W.
We are interested in understanding which features of the Ca2+ dynamics control the quantities V and W, since these parameters control the rate at which the mechanical system is chemically driven out of equilibrium. To do this, we analyzed a kinetic model (58) for a traveling Ca2+ wave powered by calcium-induced calcium release (62, 63) from an out-of-equilibrium internal store [e.g., the endoplasmic reticulum (30, 50)] into the cytosol. It has been previously shown that in this model, the wave speed V approximately follows (58)
VJDCC0,
[5]
where J represents the Ca2+ influx of open channels, DC is the diffusion constant of Ca2+, and C0 is the concentration threshold of the Ca2+ channels (see Materials and Methods, Ca2+ dynamics and SI Appendix, Supplementary Methods, section B for more details). We incorporated the binding of cytosolic Ca2+ to myoneme fibers into the model and derived a relationship for the width W of the resulting active stretch wave:
WDCK+C0,
[6]
where K+ is the rate at which Ca2+ binds to myonemes. We provide numerical confirmation of this relationship in SI Appendix, Supplementary Methods, section B and the Inset of Fig. 2E. A seemingly restrictive assumption made in the derivation of Eq. 6 is that the total concentration of myoneme binding sites Btot is much smaller than the typical cytosolic Ca2+ concentration. However, we find numerically that the predicted scaling holds even as Btot increases to relatively large values.
This analysis reveals that features in the dynamics of contraction, such as the time taken for the length to change, can inform us about the magnitudes of the underlying biophysical parameters like the binding rate K+. Assuming DC remains the same in both organisms [though the presence of internal buffers could affect this (58)], the estimated values of V and W for Vorticella and Spirostomum suggest, through Eqs. 5 and 6, that these organisms possess different threshold Ca2+ concentrations C0 (30, 64, 65) and different release rates J or binding constants K+. This minimal model serves as a first step in connecting contraction dynamics to the organisms’ biochemistry, but more in vivo experiments are necessary to test these predictions.

C. Dimensional Analysis Reveals the Minimal Set of Model Parameters.

The model considered so far (Eq. 2 together with 22) is written in dimensional quantities and depends on the parameters α and μ, which are related to the stiffness of the material and viscosity of the environment, and V, W, and gmin, which are the speed, width, and minimum value of the active stretch wave. However, these parameters are not all independent. We now reduce the number of parameters by one through nondimensionalization. We use lowercase variables throughout to denote dimensionless quantities.
There are two length scales, W and L, and we pick the uncontracted length L as the reference length and define s = S/L, x = X/L, and w = W/L. Next, three timescales can be defined:
τw=LV,τm=Lα,andτd=1μ.
[7]
Here, τw is the wave timescale, which represents the time taken for the center of the traveling wave to traverse the length L; τm is the material timescale, which represents the time taken for a sound wave to traverse the length of the spring chain; τd is the drag timescale, characterizing the exponential deceleration of motion due to viscous damping. Unless otherwise specified, we nondimensionalize time using t = T/τw. In these units, the wave speed is equal to 1, and Eq. 2 is expressed as the following PDE for x(s, t):
t2xηw,m2s2xsf(st)+ηw,dtx=0,
[8]
where f*(s)≡g*(Ls) is the active stretch wave profile, rewritten to take a dimensionless argument s. The size of τw relative to the two remaining timescales defines the nondimensional parameters
ηw,mτwτm=αVandηw,dτwτd=μLV.
[9]
The parameter ηw, m describes the ratio of the myonemal speed of sound (which depends on the stiffness and density) to the wave traversal speed, and it is independent of length. The parameter ηw, d describes the ratio of the drag timescale (depending on viscosity and density) to the wave traversal speed and is proportional to the length.
We can further simplify these nondimensional dynamics by taking the “slow wave” limit ηw, m, ηw, d ≫ 1, in which (8) reduces to
txζs2xsf(st)=0,
[10]
where the only remaining parameter is
ζτdτwτm2=αVμL.
[11]
Eq. 10 can be viewed as an overdamped limit of the model, in which inertia has dropped out of the dynamics because only a first derivative with respect to time remains, and in which mass is no longer relevant because ζ is independent of m~.
Eq. 10 is similar to the diffusion (or heat) equation with a source term. One difference is the boundary conditions: typically, the heat equation has fixed Dirichlet or Neumann boundary conditions, yet x obeys a boundary condition that depends in time on f*(s − t) (cf. Eq. 3). However, we can define
h(s,t)=sx(s,t)f(st),
[12]
and express Eq. 10 as
thζs2h=tf(st).
[13]
The boundary conditions are now the fixed Dirichlet conditions h(0, t)=h(1, t)=0, and thus Eq. 13 is exactly the diffusion equation with a source term −∂tf, for which exact series solutions can be found. The mixed boundary conditions of Vorticella also result in mixed boundary conditions for the diffusion equation.
We also consider a “quench” limit ηw, d, ηw, m ≪ 1, in which the traveling wave moves at infinite speed, such that the rest lengths of all the springs are set to their new values before the organism can react. This implies that f(s) is independent of time. Because the timescale τw is zero if the wave is infinitely fast, we instead nondimensionalize time using τm in the quench limit. The dynamics consist of only the relaxation of the system to its equilibrium
xeq(s)=0sf(s)ds.
[14]
The deviation from equilibrium, y(s, t)≡x(s, t)−xeq(s), obeys the dynamics
t2y(s,t)s2y(s,t)+ηm,dty(s,t)=0
[15]
with the initial condition y(s, 0)) ≡ x(s, 0)−xeq(s), ∂ty(s, 0)=0, and stress-free boundary conditions ∂sy(0, t)=∂sy(1, t)=0. We note that the one remaining parameter
ηm,dτmτd=Lμα
[16]
is independent of wave speed, which is infinite. The quench limit has an exact Fourier series solution, described in SI Appendix, Supplementary Methods, section D.
In Fig. 3 A and B we compare the solutions of the full nondimensional model (Eq. 8), the slow wave model (Eq. 10), and the quench model (Eq. 15) for two sets of the nondimensional parameters ηw, d and ηw, m. In these plots, we fix w = 0.1 and gmin = 0 for simplicity. For ηw, d = ηw, m = 1 (Fig. 3A), the models all behave quite differently because neither limiting case is reached. Decaying oscillations about l = 0 are observed for the full and quench model, while the slow wave model does not oscillate because it lacks inertia. The oscillations for the quench and full model have the same frequency (which is proportional to ηm, d, reflecting the resonant frequency of the spring chain) yet different waveforms and amplitudes. For ηw, d = 20, ηw, m = 2 (Fig. 3B), the full model and slow wave model are in close agreement.
Fig. 3.
The nondimensional model and its limiting versions capture key kinematic properties of contraction. (A) Nondimensional length λ(t) as a function of time for the three versions of the nondimensional model. Parameter values are ηw, d = ηw, m = 1, w = 0.1, gmin = 0. The vertical lines at t = 2 and 3 indicate the times when the center of the active stretch wave enters and exits the system. The time argument for the quench solution is multiplied by ηw, m so that it can be compared to the other model solutions. For the quench model, the quench occurs at t = 2. (B) Same as (A) but with ηw, m = 2 and ηw, d = 20. (C and D) Scaling of λ˙max with ηw, m (C) and ηw, d (D) for the three models. The colors are the same as in (A). For panel (C), ηw, d is fixed at 8, while for panel (D), ηw, m is fixed at 2. (E) Heatmap of χ, the peak elastic energy stored in the system as a fraction of the maximum possible energy, as ηw, m and ηw, d are varied for the full model. Contours are drawn and labeled at selected values of χ.
A stark feature of the full model compared to the quench model is the asymmetry of contraction: In the quench model, all springs instantly have new rest lengths, so all springs relax simultaneously and symmetrically about the center s = 0.5. This is not the case for the full model, where the wave first arrives at s = 0, causing the springs to begin relaxing before those at s = 1. This asymmetry of contraction can be observed in the experimental images of Spirostomum and Vorticella in Fig. 1 A and B, and we further illustrate this in Fig. 4AC. In Fig. 4DG we show x(s, t) at several values of t for the full and quench models corresponding to the length trajectories shown in Fig. 3 A and B.
Fig. 4.
Asymmetric contraction of Spirostomum. (A) A sequence of contours of Spirostomum obtained from 8,000-Hz video imaging. Colors range from yellow to purple as time ranges from 5.5 ms to 8 ms. The initial and final midlines, whose arc lengths determine the measured length of the organism, are drawn in yellow and purple. (B) The trajectory of the length during this contraction event. The dashed curved line shows the trajectory of the simulated model, which was fit to match the experimentally measured solid curve. (C) The displacement trajectories of the top (blue) and bottom (red) ends of the midlines and the corresponding quantities in the fitted model are shown as dashed curves. These displacements were measured by projecting the endpoints to the initial midline curve and computing the arc length from the end to the projected point. Asymmetric contraction can be observed as the top end begins contracting before the bottom end (by a time difference labeled ΔT), a feature that the fitted model reproduces. (D and E) Corresponding to Fig. 3 A and B, the entire x(s, t) curves for the full and quench models. The colors in these panels range from yellow to blue as t increases from 2 to 4 in steps of 0.1. Panels (F) and (G) show the same curves as a function of t, rather than s.
From the fits to the experimental data of Vorticella and Spirostomum contracting in water (see SI Appendix, Supplementary Results, section B), neither organism is expected to obey the quench-like dynamics, being closer to the slow-wave limit. This implies that the organisms are relaxing appreciably while the Ca2+ wave traverses their lengths.

D. Scaling Behavior Distinguishes Three Dynamic Regimes.

An additional significant difference between the different models is the scaling of the maximum contraction rate with the viscosity. Notably, this has been studied in previous experimental and computational works on Vorticella and Spirostomum, against which we may validate our model (9, 42, 49).
To explore this scaling behavior, we computed the maximum contraction rate λ˙maxmaxt|tλ(t)| as we varied the nondimensional parameters ηw, m and ηw, d in the full, quench, and slow wave models. The results are displayed in Fig. 3 C and D, again setting w = 0.1 and gmin = 0 for simplicity. The dynamics of the quench model only depend on the dimensionless parameter ηm, d (cf. Eq. 15), and the maximum rate (in units of L/τm=α) asymptotically scales like λ˙maxηm,d1=ηw,m/ηw,d=α/Lμ. This can be qualitatively understood as follows: as shown in SI Appendix, Supplementary Methods, section D, an exact solution for the quench model can be constructed by decomposing the initial condition into spatial Fourier modes and solving for the decay of each mode. Each mode acts like a mass, attached to a wall by a stretched spring with an effective drag constant ηm, d, which is suddenly released (see SI Appendix, Eq. 66). By Stokes’ law, the maximum rate should then scale inversely with ηm, d as observed.
For both the full and slow wave models, λ˙maxηw,m up to a point at which λ˙max plateaus at 1. This results from the contraction speed not significantly exceeding the speed of the traveling active stretch wave, which is defined as 1 in these units. Interestingly, for both models λ˙max (in units of L/τw = V) asymptotically scales like λ˙maxηw,d1/2. To understand this, recall that the slow wave model can be formally mapped to the diffusion equation with a source term (cf. Eqs. 10 and 13). The dynamics are thus characterized by the “diffusion constant” ζ = ηw, m2/ηw, d = α/VμL. For diffusion, it is known that an initial perturbation to the system spreads out over time with a length scale that grows like ζt, which implies that the corresponding rate scales like λ˙maxζ1/2=α/VμL. This fact explains both the observed scaling λ˙maxηw,m up to its saturation at 1, and the asymptotic scaling λ˙maxηw,d1/2. We emphasize that in Fig. 3, we have assumed the wave width W = 0.1 L scales with the length L; as demonstrated in SI Appendix, Fig. S2, different scaling behaviors result if one varies the uncontracted length L while holding the wave width W fixed. In SI Appendix, Fig. S4 we show that this predicted scaling fits our experimental data for Spirostomum.
In refs. 42 and 49, the scaling of λ˙max with the solution viscosity ν for Vorticella is measured to be λ˙maxν1/2 (Fig. 5A). This is consistent with our observed scaling λ˙maxζ1/2 for the full and slow wave models when holding all other parameters constant and only varying μ (which is proportional to ν). Our model thus exhibits the scaling observed in existing experimental data for Vorticella. We note that an alternative description for the ν−1/2 scaling based on power-limited contraction was provided in ref. 42. Our explanation here does not involve a physical argument about power conservation and follows solely from the dynamical equations. It may be possible to show that the energy-based argument is equivalent to the dynamical one presented here, but doing so would likely require a more detailed treatment of thermodynamics of the Ca2+ wave (as in ref. 67), and we leave this to future work.
Fig. 5.
Experimentally measured contraction rate scaling. (A) Published data on the maximum contraction rate of Vorticella in solutions of different viscosities. Each colored line represents a single cell exposed to varying viscosities; these data are from figure 3B of ref. 42. The black points represent results from simulations, reproduced from figure 4A in ref. 49. The circle and square represent two values of the elastic stalk’s Young’s modulus. (B) Published data from figure 2 of ref. 9 on the maximum contraction rate of Spirostomum in different viscosities and confinement methods: Orange represents a setup using plane parallel glass coverslips, and blue represents a setup using an agar channel. Each point represents an average of 10 to 23 measurements, and connecting lines are drawn to guide the eye. (C) Our measurements of the maximum contraction rate of free-swimming Spirostomum in solutions of different viscosities. The published data were extracted from refs. 9, 42, and 49 using WebPlotDigitizer (66).
Scaling data for Spirostomum are available in refs. 8 and 9 (Fig. 5B). In these experiments, to aid imaging, the cells are confined in either a narrow agar channel (blue points) or between a microscope slide and a glass coverslip (orange points). Depending on the confinement method, the scaling appears consistent with either λ˙maxν1, corresponding to the quench regime, or λ˙maxν1/2. Our fits in Fig. 2, together with observation of the asymmetric contraction dynamics (Fig. 5A), suggest that Spirostomum should not be in the quench regime. The quench-like scaling may be an artifact of using pressure on the coverslips to prevent the organisms from swimming out of view; we speculate that this pressure could hinder contraction during the traversal of the wave, effectively quenching the springs. The imaging rate in refs. 8 and 9 (800 Hz) may also be too low to resolve the maximum in the contraction rate profile precisely. We thus performed our own measurements of the contraction of free-swimming Spirostomum using a high-speed camera, recording at a minimum of 2,000 Hz. The results, displayed in Fig. 5C, are consistent with the predicted scaling λ˙maxν1/2, though we cannot conclusively exclude the scaling λ˙maxν1. We note that the low λ˙max at 1 cP is consistent with the predicted roll-off at low viscosity shown in Fig. 3D.
The roll-off corresponds to an approach toward a third dynamic regime, which occurs when the active stretch wave is sufficiently slow, or the springs can sufficiently overcome the drag resistance to follow the active stretch wave very closely, with little deviation from equilibrium. In this “quasi-static” limit, there is no dependence of λ˙max on either the stiffness or drag since the contraction speed is solely determined by the speed of the active stretch wave. This regime occurs in the slow wave model when ζ ≫ 1, explaining the approach toward λ˙max=1 for large ηw, m and small ηw, d in Fig. 3 C and D. The three scaling regimes—the quench regime (λ˙maxν1), the intermediate regime (λ˙maxν1/2), and the quasi-static regime (λ˙maxν0)—fully characterize the qualitative dynamical behavior of myoneme contraction.
To summarize the dynamical behavior of the system as a function of ηw, d and ηw, m, in SI Appendix, Supplementary Methods, section E we introduce the dimensionless quantity χ, which measures the peak elastic energy stored in the system as a fraction of the maximum possible energy which is achieved in the quench limit. In Fig. 3E, we plot χ over ηw, m ∈ [1/32, 32] and ηw, d ∈ [1/4, 256], and in SI Appendix, Fig. S1 we show corresponding trajectories of the energy and length for these conditions. Let us imagine fixing all parameters except the spring stiffness and the viscosity. Fig. 3E then illustrates the system’s ability to respond to the active stretch wave under different mechanical and environmental conditions: for small ηw, m, corresponding to soft springs, the system stores a large fraction of the possible amount of internal energy (χ ≈ 1), but this fraction decreases as the drag, proportional to ηw, d, decreases. By contrast, for large ηw, m, the springs closely follow the active stretch wave and never accumulate much internal energy (χ ≈ 0), and the amount that does accumulate decreases as the drag decreases. This energetic perspective thus provides a unified way to understand the dynamic regimes for organisms using Ca2+-based motion.

3. Discussion

A. Summary.

We present a minimal mathematical model of myoneme contraction that captures the kinematic features of the organisms Spirostomum and Vorticella, which exhibit different shapes, sizes, and myoneme arrangements. This model suggests that a fast-traveling wavefront of active stretch, underpinned by rapid Ca2+ kinetics, synchronizes contraction across the organism’s length. Using an auxiliary reaction–diffusion model for the Ca2+ dynamics, we derive scaling relationships for the speed and width of the active stretch wave that enters the mechanical model. These parameters dictate the rate of chemical driving of the myoneme network. Through nondimensionalization, we determine the minimal independent set of parameters controlling contraction and further examine two limiting cases: a quench limit and a slow wave limit. The latter approximately applies to both Spirostomum and Vorticella in water, as confirmed by fitting the model to experimentally obtained length trajectories. We formally map the dynamics in this limit to the well-studied diffusion equation, with the unitless parameter ζ playing the role of the diffusion constant. The key geometric and mechanical factors that control contractile dynamics include length, stiffness, wave speed, and drag on the organism. These factors manifest through various nondimensional combinations in the different dynamical regimes outlined above. Our analysis reveals three such regimes, each with distinct kinematic signatures such as the symmetry of contraction about the organism’s midpoint and the exponent with which the maximum contraction rate scales with viscosity. We further differentiate these regimes by a quantity representing the peak elastic energy stored in the organism relative to the maximal possible value attained in the quench limit.

B. Biological Implications of Dynamic Regimes.

It is worthwhile to consider the possible biological implications of the three dynamic regimes. In the quasi-static regime, the myoneme contractile apparatus remains at mechanical equilibrium while the Ca2+ wave traverses the organism’s length. In this limit, the contractile speed is fully determined by the wave speed V and does not scale with other parameters. The dissipation of chemical energy is minimized at the expense of the speed of motion. The marine protist Acantharia reportedly has myoneme-based spicules that are used to regulate buoyancy by contracting in a Ca2+-powered process that is two orders of magnitude slower than contraction in Spirostomum or Vorticella (19, 29). In this case, contraction may occur quasi-statically. Conversely, the quench regime attains the highest contractile velocities. These high speeds consequently dissipate the most energy because the myoneme apparatus deviates the furthest from mechanical equilibrium. This regime would be suited to processes in which large speeds are physiologically important. For example, Spirostomum excites contraction of nearby organisms through “hydrodynamic trigger waves,” requiring it to move sufficiently rapidly to disturb the fluid, dissipating energy in the process (10). The intermediate regime, where we place the contraction of Vorticella and Spirostomum, is expected to achieve a balance between thermodynamic efficiency and biological functionality. While myoneme-based systems are predominantly found in ciliated protists, similar ATP-independent contractile protein networks exist in a wide range of other organisms across taxa, facilitating a diverse set of biological functions (810, 1527). If similar mechanochemical dynamics apply in those organisms, then various selection pressures may favor a particular regime for each function.

C. Comparison to Traditional Latch-Mediated Systems.

The theory of latch-mediated spring actuation (LaMSA) provides a general description of motion characterized by a fast unlatching event that releases stored elastic energy (6870). It has been applied to analyze macroscopic biomechanical processes like the striking of a Mantis shrimp’s claw and a human finger snap (71, 72). One may expect that LaMSA also explains myoneme-based contraction, but here energy is stored in the form of a difference in Ca2+ chemical potential between the cytosol and the endoplasmic reticulum, and release of a “chemical latch” corresponds to initiation of a traveling wave of Ca2+ ions. In addition to this difference in the energy storage medium, LaMSA is designed to apply to macroscopic systems, neglecting the strong dissipative forces which are unavoidable at the cellular level. Furthermore, in myonemes, contraction occurs over an approximately continuous spatial domain of many small elastic elements connected together, compared to the discrete mechanical elements considered in LaMSA. These qualitative features of traditional LaMSA systems compared to the myoneme-based “chemical latch” are illustrated in Fig. 6. Outside the quench limit, the Ca2+ wave, which loads mechanical energy into the system, has a finite traversal time. This means that the system can begin relaxing before the energy has been fully loaded. This is another key distinction from traditional LaMSA systems, which are assumed to be fully loaded before relaxing. LaMSA systems also include a constraint force which is lacking in free-swimming organisms like Spirostomum. Exploring the implications of adding constraint forces to myoneme-based systems would be an interesting direction for future research. For example, fixing the ends of Spirostomum and then releasing them after the Ca2+ wave has passed would allow artificially pushing the contraction process into the quench regime. Although how physical latches mediate energy flow has only recently been appreciated in biology (69, 73), our model offers a guiding framework for how a chemical latch dynamically controls and tunes force output in ultrafast single-cell organisms.
Fig. 6.
Schematic illustration of a chemical latch. The Left panel, adapted from ref. 68, illustrates a mechanical latch-and-spring system. A latch, shown in blue, slides frictionally off of a compressed spring, shown in gray, which allows for a fast release of the stored potential energy. The finite time during which the latch slides is highlighted in orange. The potential and kinetic energies of the system are schematically plotted at the bottom. At smaller scales, such as the scale of a single-cell organism like Spirostomum, a different paradigm accounts for the chemical nature of mechanical activation (via a traveling Ca2+ wave). This is illustrated in the Right panel, where analogous features of the dynamics in the macroscopic case are shown in the corresponding colors. The traveling wave is shown as a blue gradient, which controls the local rest length (indicated as the number of ridges) of internal springs, shown in gray.

D. Implications for Synthetic Biology.

Finally, our model offers valuable insights for engineering myoneme-based synthetic systems. Researchers have recently demonstrated the feasibility of reconstituting myonemes in vitro and exogenously controlling them using light-patterned Ca2+ release (74). Myoneme-based systems, therefore, provide a mechanism for exerting mechanical forces at subcellular scales, with orthogonal chemistry to light-controllable actomyosin and kinesin-microtubule systems (7579). Our model uncovers several critical parameters that can modulate the dynamics of myoneme contraction. For example, increasing cross-linking would enhance the effective α, elevating the solution viscosity would raise the effective μ, and increasing the effective light-activated Ca2+ release rate or reducing Ca2+ buffering (thus increasing DC) would adjust V and W of the active stretch wave. By controlling these experimental parameters, researchers can explore the various regimes identified in this work, paving the way for designing synthetic cells equipped with myoneme-based artificial cytoskeletons that harness the full potential of ultrafast contraction dynamics.

Materials and Methods

Ca2+ Dynamics.

The dependence of the active stretch g(S, T) on the local bound Ca2+ concentration may be modeled using the instantaneous linear relation (49)
g(S,T)=1(1gmin)B(S,T)Btot,
[17]
where B(S, T) is the local concentration of Ca2+ bound to myoneme fibers and Btot is the total concentration of binding sites (which is assumed to be uniform over the spatial domain). Finding g(S, T) then requires specifying the dynamics for the bound Ca2+ profile B(S, T). To avoid introducing extra variables, we first analyze the complete dynamics for B(S, T) by extending a tractable model for a traveling Ca2+ wave developed by Kupferman et al. ((58)), and we then use results of this analysis to determine an approximate form for g(S, T) to use in the mechanical model.
The Kupferman model for a traveling Ca2+ wave is
TC=DCS2CΓC+JΘ(CC0).
[18]
Here, C(S, T) is the concentration profile of cytosolic Ca2+ which diffuses with a diffusion constant DC, leaks from the cytosolic compartment with a first-order rate constant Γ, and is released into the cytosol by internal stores at a spatially dependent rate JΘ(C − C0). This last term involves a Heaviside function Θ(C − C0), which approximates the kinetics of calcium-induced calcium release (63). It “turns on” when the local Ca2+ concentration C(S, T) exceeds a spatially uniform threshold C0, at which point Ca2+ is locally released into the cytosol at a rate J. The authors of ref. 58 show that the speed of the traveling wave is given by Eq. 5.
We extend the Kupferman model to allow cytosolic Ca2+ to bind to the myoneme assembly. This introduces the variable B(S, T) for the local concentration of myoneme-bound Ca2+, and the coupled dynamical equations for C and B are
TC=DCS2CΓC+JΘ(CC0)K+C(BtotB)+KB
[19]
TB=K+C(BtotB)KB.
[20]
Here, K+ is the second-order rate constant for binding of cytosolic Ca2+ to open myoneme binding sites, which have concentration Btot − B. K is the first-order rate constant for the subsequent unbinding of myoneme-bound Ca2+.
In SI Appendix, Supplementary Methods, section B, we analyze Eqs. 19 and 20 under mild simplifying assumptions to show that: i) the function B(S, T) is well approximated by a traveling wave with a sigmoidal profile:
B(S,T)Btot=111+e(SVT)/W,
[21]
ii) its speed V is equal to the speed of diffusing Ca2+ wave (cf. Eq. 5), and iii) its width W can be expressed in terms of the chemical parameters according to Eq. 6.
Combing Eqs. 17 and 21 gives the following form for the nonautonomous active stretch:
g(S,T)=g(SVT)=gmin+1gmin1+e(SV(TO))/W.
[22]
The additional parameter O is a time that allows an adjustable offset between T = 0 and the time T = O when the center of the wave arrives at S = 0. This allows shifting time to achieve acceptable numerical agreement between boundary and initial conditions, as described in Materials and Methods, Computational methods.

Experimental Methods.

We obtained Spirostomum organisms from Carolina Biological and cultured them at room temperature. We prepared culturing medium by mixing one part boiled hay medium solution (Ward’s Science, 470177-390), four parts boiled spring water, and two boiled wheat grains (Carolina, Item #132425) as described previously (36). We inoculated organisms into fresh media approximately twice per week. For imaging, we pipetted several organisms at a time into microscopy slides with multiple channels (Ibidi, μ-Slide VI 0.4) or with multiple wells (Ibidi, μ-Slide 8-well). The channels are 400 μm high, allowing the organisms to swim and contract without confinement.
Our Vorticella samples were collected from Morse Pond near the Marine Biological Laboratory. We dissected small pieces of plant material to which individual or small groups of Vorticella were attached. We added these pieces of plant matter, along with enough imaging medium to cover them, into 8-well coverslip-bottomed sample chambers (Ibidi, μ-Slide 8-well).
To alter the viscosity, we prepared stock solutions of Ficoll PM 400 (Sigma F4375) in water and diluted them to concentrations between 3% and 12% (weight per volume). In the case of Spirostomum, we pipetted organisms directly into the Ficoll solution. In the case of Vorticella, we rinsed the samples several times with the solution within the sample chamber to ensure that the appropriate viscosity was reached. In both cases, slow pipetting was important to prevent shearing of the organisms in the high-viscosity solutions. To calculate viscosity ν as a function of percentage p of Ficoll in water, we used the calibration function ν = e0.162p reported in refs. 80 and 81.
We performed brightfield imaging on either a Zeiss inverted microscope equipped with a 0.55 NA condenser, a 0.63× defocusing lens, DIC optics, and a 10× air objective or a Nikon Ti2 equipped with a 2× (MRD70020) or 10× (MRD70170) air objective. In both cases, we collected images using a pco.dimax CS camera equipped with at least 2,000 and up to 10,000 Hz.
For quantifying Spirostomum images, we first used Fiji (82) to threshold the images to identify the pixels associated with the organisms, and then we used Matlab to run Morphometrics (83) to obtain the cell contours. We used the pill-mesh option of Morphometrics to obtain cell shape and length. See SI Appendix, Fig. S3 and Movie S3 for a visualization of this mesh generation procedure.
Movie S3.
Contraction of Spirostomum, showing the original video footage, a thresholded version, and the output of the Morphometrics contour tracking software.
For quantifying Vorticella length over time, we manually tracked the position of the end of the stalk attached to the cell body using home-built Python software.

Computational Methods.

All numerical integration was performed with Mathematica’s NDSolve function, using the method of lines option with a spatial discretization based on tensor product grids. The accuracy and precision goals were set high enough to ensure reasonable numerical convergence. To achieve agreement between the initial conditions Eq. 4 and the boundary conditions Eq. 3, we increased the parameter O in Eq. 22 so that g(S, 0)≈1 across the computational domain. To solve the quench model Eq. 15, we evaluate the exact solution given in SI Appendix, Eq. 60 using 50 terms in the Fourier expansion.

Data, Materials, and Software Availability

All study data are included in the article and/or supporting information. Code and data used for the results in this paper can be found at https://github.com/bhamla-lab/myonememodelPNAS2023 (84).

Acknowledgments

We wish to thank Shane Terrell for help setting up the experiments, Vincent Boudreau for providing samples of Vorticella in fresh pond water, Joe Brzostowski, Tanner Fadero, and Sam Lord for technical assistance with the microscopy, KC Huang and Tristan Ursell for providing the Morphometrics software, and Joe Lannan, Christian Pagán-Medina, Joël Lemière, and Jane Maienschein for helpful discussions. This work was mainly supported by NSF awards EF-1935260, EF-1935261, and EF-1935262. S.V. was supported by the National Institute of General Medical Sciences of the NIH under Award No. R35GM147400. The content is solely the responsibility of the authors and does not necessarily represent the official views of the NIH. C.F. acknowledges support from the University of Chicago through a Chicago Center for Theoretical Chemistry Fellowship. M.S.B. acknowledges funding support from NIH Grant R35GM142588; NSF Grant MCB-1817334; NSF CAREER IOS-1941933; and a gift from the Open Philanthropy Project. F.C. acknowledges funding support from NIH Grant R35GM141796. We acknowledge the University of Chicago’s Research Computing Center for computing resources, the Marine Biological Laboratory Whitman Summer Investigator Program and the Physiology Course for the use of microscopy equipment, and Zeiss and pco. for loans of cameras.

Author contributions

C.F., A.T.M., X.L., J.E.H., F.C., M.W.E., S.V., A.R.D., and M.S.B. designed research; C.F., A.T.M., M.W.E., S.V., A.R.D., and M.S.B. performed research; C.F., A.T.M., M.W.E., S.V., A.R.D., and M.S.B. analyzed data; and C.F., A.T.M., M.W.E., S.V., A.R.D., and M.S.B. wrote the paper.

Competing interests

The authors declare no competing interest.

Supporting Information

Appendix 01 (PDF)
Movie S1.
Contraction of Spirostomum, along with output from the contour-tracking Morphometrics software used to measure the change in length over time.
Movie S2.
Contraction of Vorticella, along with output from the contour-tracking Morphometrics software used to measure the change in length over time.
Movie S3.
Contraction of Spirostomum, showing the original video footage, a thresholded version, and the output of the Morphometrics contour tracking software.

References

1
M. Murrell, P. W. Oakes, M. Lenz, M. L. Gardel, Forcing cells into shape: The mechanics of actomyosin contractility. Nat. Rev. Mol. Cell Biol. 16, 486–498 (2015).
2
W. F. Marshall, Don’t blink: Observing the ultra-fast contraction of spasmonemes. Biophys. J. 94, 4–5 (2008).
3
T. Weis-Fogh, W. Amos, Evidence for a new mechanism of cell motility. Nature 236, 301–304 (1972).
4
J. Cachon, M. Cachon, Movement by non-actin filament mechanisms. Biosystems 14, 313–326 (1981).
5
D. H. Lynn, The Ciliated Protozoa: Characterization, Classification, and Guide to the Literature (Springer, 2008).
6
W. Amos, Reversible mechanochemical cycle in the contraction of Vorticella. Nature 229, 127–128 (1971).
7
S. Ryu, R. E. Pepper, M. Nagai, D. C. France, Vorticella: A protozoan for bio-inspired engineering. Micromachines 8, 4 (2016).
8
R. Hawkes, D. Holberton, Myonemal contraction of Spirostomum. I. Kinetics of contraction and relaxation. J. Cell. Physiol. 84, 225–235 (1974).
9
R. Hawkes, D. Holberton, Myonemal contraction of Spirostomum. II. Some mechanical properties of the contractile apparatus. J. Cell. Physiol. 85, 595–602 (1975).
10
A. J. Mathijssen, J. Culver, M. S. Bhamla, M. Prakash, Collective intercellular communication through ultra-fast hydrodynamic trigger waves. Nature 571, 560–564 (2019).
11
J. Zhang et al., Giant proteins in a giant cell: Molecular basis of ultrafast Ca2+-dependent cell contraction. Sci. Adv. 9, eadd6550 (2023).
12
M. M. Slabodnick, W. F. Marshall, Stentor coeruleus. Curr. Biol. 24, R783–R784 (2014).
13
E. Newman, Contraction in Stentor coeruleus: A cinematic analysis. Science 177, 447–449 (1972).
14
S. M. Coyle et al., Coupled active systems encode an emergent hunting behavior in the unicellular predator Lacrymaria olor. Curr. Biol. 29, 3838–3850 (2019).
15
Y. Moriyama, H. Okamoto, H. Asai, Rubber-like elasticity and volume changes in the isolated spasmoneme of giant Zoothamnium sp. under Ca2+-induced contraction Biophys. J. 76, 993–1000 (1999).
16
R. Moreton, W. Amos, Electrical recording from the contractile ciliate Zoothamnium geniculatum ayrton. J. Exp. Biol. 83, 159–167 (1979).
17
M. Melkonian, P. L. Beech, C. Katsaros, D. Schulze, “Centrin-mediated cell motility in algae” in Algal Cell Motility (Springer, 1992), pp. 179–221.
18
M. Ishida, Y. Shigenaka, K. Taneda, Studies on the mechanism of cell elongation in Blepharisma japonicum. I. A physiological mechanism how light stimulation evokes cell elongation. Eur. J. Protistol. 25, 182–186 (1989).
19
J. Febvre, The myoneme of the Acantharia (Protozoa): A new model of cellular motility. Biosystems 14, 327–336 (1981).
20
M. Ando, Y. Shigenaka, Structure and function of the cytoskeleton in heliozoa: I. Mechanism of rapid axopodial contraction in Echinosphaerium. Cell Motil. Cytoskelet. 14, 288–301 (1989).
21
B. Viguès, R. Damaj, The adhesive disc in the mobilid ciliate Trichodina pediculus: Evidence for centrin-related, calcium-sensitive filaments. Cytoskeleton 69, 113–124 (2012).
22
M. Kawachi, I. Inouye, Ca2+-mediated induction of the coiling of the haptonema in Chrysochromulina hirta (Prymnesiophyta = Haptophyta) Phycologia 33, 53–57 (1994).
23
G. McFadden, D. Schulze, B. Surek, J. L. Salisbury, M. Melkonian, Basal body reorientation mediated by a Ca2+-modulated contractile protein J. Cell Biol. 105, 903–912 (1987).
24
J. Cachon, M. Cachon, An unusual mechanism of cell contraction: Leptodiscinae dinoflagellates. Cell Motil. 4, 41–55 (1984).
25
M. Knoblauch et al., ATP-independent contractile proteins from plants. Nat. Mater. 2, 600–603 (2003).
26
M. Knoblauch, W. S. Peters, Forisomes, a novel type of Ca2+-dependent contractile protein motor Cell Motil. Cytoskelet. 58, 137–142 (2004).
27
W. T. Huck, Responsive polymers for nanoscale actuation. Mater. Today 11, 24–32 (2008).
28
H. Hoffmann-Berling, The mechanism of a new contraction cycle distinct from muscle contraction. Biochim. Biophys. Acta 27, 247–255 (1958).
29
J. Febvre, C. Febvre-Chevalier, Motility processes in Acantharia (Protozoa). III. Calcium regulation of the contraction–relaxation cycles of in vivo myonemes. Biol. Cell 67, 251–261 (1989).
30
K. Katoh, Y. Naitoh, Control of cellular contraction by calcium in Vorticella. J. Exp. Biol. 189, 163–177 (1994).
31
Y. Yokoyama, H. Asai, Contractility of the spasmoneme in glycerinated Vorticella stalk induced by various divalent metal and lanthanide ions. Cell Motil. Cytoskelet. 7, 39–45 (1987).
32
W. Amos, Structure and coiling of the stalk in the peritrich ciliates Vorticella and Carchesium. J. Cell Sci. 10, 95–122 (1972).
33
L. Routledge, W. Amos, B. Gupta, T. Hall, T. Weis-Fogh, Microprobe measurements of calcium binding in the contractile spasmoneme of a vorticellid. J. Cell Sci. 19, 195–201 (1975).
34
A. R. Jones, T. L. Jahn, J. R. Fonseca, Contraction of protoplasm. I. Cinematographic analysis of the anodally stimulated contraction of Spirostomum ambiguum. J. Cell. Physiol. 68, 127–133 (1966).
35
T. Ochiai, H. Asai, K. Fukui, Hysteresis of contraction-extension cycle of glycerinated Vorticella. J. Protozool. 26, 420–425 (1979).
36
L. Xu, M. Bhamla, Biophysical mechanism of ultrafast helical twisting contraction in the giant unicellular ciliate Spirostomum ambiguum. bioRxiv [Preprint] (2019). https://doi.org/10.1101/854836 (Accessed 23 July 2021).
37
Y. Moriyama, S. Hiyama, H. Asai, High-speed video cinematographic demonstration of stalk and zooid contraction of Vorticella convallaria. Biophys. J. 74, 487–491 (1998).
38
S. Ryu, P. Matsudaira, Unsteady motion, finite Reynolds numbers, and wall effect on Vorticella convallaria contribute contraction force greater than the stokes drag. Biophys. J. 98, 2574–2581 (2010).
39
S. Ryu, M. J. Lang, P. Matsudaira, Maximal force characteristics of the Ca2+-powered actuator of Vorticella convallaria Biophys. J. 103, 860–867 (2012).
40
J. Zhou, S. Ryu, D. Admiraal, Flow and transport effect caused by the stalk contraction cycle of Vorticella convallaria. Biomicrofluidics 11, 034119 (2017).
41
E. G. Chung, S. Ryu, Stalk-length-dependence of the contractility of Vorticella convallaria. Phys. Biol. 14, 066002 (2017).
42
A. Upadhyaya et al., Power-limited contraction dynamics of Vorticella convallaria: An ultrafast biological spring. Biophys. J. 94, 265–272 (2008).
43
M. C. Marchetti et al., Hydrodynamics of soft active matter. Rev. Mod. Phys. 85, 1143 (2013).
44
G. Danuser, J. Allard, A. Mogilner, Mathematical modeling of eukaryotic cell migration: Insights beyond experiments. Annu. Rev. Cell Dev. Biol. 29, 501 (2013).
45
J. Prost, F. Jülicher, J. F. Joanny, Active gel physics. Nat. Phys. 11, 111–117 (2015).
46
K. Popov, J. Komianos, G. A. Papoian, MEDYAN: Mechanochemical simulations of contraction and polarity alignment in actomyosin networks. PLoS Comput. Biol. 12, e1004877 (2016).
47
S. L. Freedman, S. Banerjee, G. M. Hocky, A. R. Dinner, A versatile framework for simulating the dynamic mechanical structure of cytoskeletal networks. Biophys. J. 113, 448–460 (2017).
48
F. Nédélec, T. Surrey, A. C. Maggs, S. Leibler, Self-organization of microtubules and motors. Nature 389, 305–308 (1997).
49
G. Misra, R. B. Dickinson, A. J. Ladd, Mechanics of Vorticella contraction. Biophys. J. 98, 2923–2932 (2010).
50
K. Katoh, M. Kikuyama, An all-or-nothing rise in cytosolic. J. Exp. Biol. 200, 35–40 (1997).
51
H. Shiono, Y. Naitoh, Cellular contraction precedes membrane depolarization in Vorticella convallaria. J. Exp. Biol. 200, 2249–2261 (1997).
52
I. Delvendahl et al., Reduced endogenous Ca2+ buffering speeds active zone Ca2+ signaling. Proc. Natl. Acad. Sci. U.S.A. 112, E3075–E3084 (2015).
53
L. Mahadevan, P. Matsudaira, Motility powered by supramolecular springs and ratchets. Science 288, 95–99 (2000).
54
F. Horkay, Polyelectrolyte gels: A unique class of soft materials. Gels 7, 102 (2021).
55
A. Mogilner, G. Oster, Polymer motors: Pushing out the front and pulling up the back. Curr. Biol. 13, R721–R733 (2003).
56
W. S. Slaughter, The Linearized Theory of Elasticity (Springer Science& Business Media, 2012).
57
L. E. Kinsler, A. R. Frey, A. B. Coppens, J. V. Sanders, Fundamentals of Acoustics (John Wiley& Sons, 2000).
58
R. Kupferman, P. P. Mitra, P. Hohenberg, S. Wang, Analytical calculation of intracellular calcium wave characteristics. Biophys. J. 72, 2430–2444 (1997).
59
J. Gore et al., DNA overwinds when stretched. Nature 442, 836–839 (2006).
60
C. Floyd, H. Ni, R. S. Gunaratne, R. Erban, G. A. Papoian, On stretching, bending, shearing, and twisting of actin filaments. I: Variational models. J. Chem. Theory Comput. 18, 4865–4878 (2022).
61
E. Moeendarbary et al., The cytoplasm of living cells behaves as a poroelastic material. Nat. Mater. 12, 253–261 (2013).
62
S. Schuster, M. Marhl, T. Höfer, Modelling of simple and complex calcium oscillations: From single-cell responses to intercellular signalling. Eur. J. Biochem. 269, 1333–1355 (2002).
63
G. Dupont, M. Falcke, V. Kirk, J. Sneyd, Models of Calcium Signalling (Springer, 2016).
64
H. Ishida, T. Suzaki, Y. Shigenaka, Effect of Mg2+ on Ca2+-dependent contraction of a Spirostomum cell model. Eur. J. Protistol. 32, 316–319 (1996).
65
E. M. Ettienne, Control of contractility in Spirostomum by dissociated calcium ions. J. General Physiol. 56, 168–179 (1970).
66
A. Rohatgi, Webplotdigitizer (Version 4.6, Pacifica, California, USA, 2017). https://automeris.io/WebPlotDigitizer (Accessed 25 July 2022).
67
F. Avanzini, G. Falasco, M. Esposito, Thermodynamics of chemical waves. J. Chem. Phys. 151, 234103 (2019).
68
S. Divi et al., Latch-based control of energy output in spring actuated systems. J. R. Soc. Interface 17, 20200070 (2020).
69
M. Ilton et al., The principles of cascading power limits in small, fast biological and engineered systems. Science 360, eaao1082 (2018).
70
S. J. Longo et al., Beyond power amplification: Latch-mediated spring actuation is an emerging framework for the study of diverse elastic systems. J. Exp. Biol. 222, jeb197889 (2019).
71
E. Steinhardt et al., A physical model of mantis shrimp for exploring the dynamics of ultrafast systems. Proc. Natl. Acad. Sci. U.S.A. 118, e2026833118 (2021).
72
R. Acharya, E. J. Challita, M. Ilton, M. Saad Bhamla, The ultrafast snap of a finger is mediated by skin friction. J. R. Soc. Interface 18, 20210672 (2021).
73
N. P. Hyun et al., Spring and latch dynamics can act as control pathways in ultrafast systems. Bioinspir. Biomim. 18, 026002 (2022).
74
X. Lei, T. Chakrabortty, J. Honts, S. Bhamla, In vitro light-controlled patterns and force generation for synthetic cytoskeletal networks. Bull. Am. Phys. Soc. 2022, B07–003 (2022).
75
M. Nakamura et al., Remote control of myosin and kinesin motors using light-activated gearshifting. Nat. Nanotechnol. 9, 693–697 (2014).
76
P. V. Ruijgrok et al., Optical control of fast and processive engineered myosins in vitro and in living cells. Nat. Chem. Biol. 17, 540–548 (2021).
77
R. Zhang et al., Spatiotemporal control of liquid crystal structure and dynamics through activity patterning. Nat. Mater. 20, 875–882 (2021).
78
T. D. Ross et al., Controlling organization and forces in active matter through optically defined boundaries. Nature 572, 224–229 (2019).
79
M. Malik-Garbi et al., Scaling behaviour in steady-state contracting actomyosin networks. Nat. Phys. 15, 509–516 (2019).
80
C. J. Lo, “Sodium energetics of chimeric flagellar motors in Escherichia coli” PhD thesis, University of Oxford, UK (2007).
81
X. Chen, H. C. Berg, Torque–speed relationship of the flagellar rotary motor of Escherichia coli. Biophys. J. 78, 1036–1041 (2000).
82
J. Schindelin et al., Fiji: An open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).
83
T. Ursell et al., Rapid, precise quantification of bacterial cellular dimensions across a genomic-scale knockout library. BMC Biol. 15, 1–15 (2017).
84
C. Floyd et al., Bhamla-lab/myonememodelPNAS2023. GitHub. https://github.com/bhamla-lab/myonememodelPNAS2023. Deposited 18 May 2023.

Information & Authors

Information

Published in

The cover image for PNAS Vol.120; No.25
Proceedings of the National Academy of Sciences
Vol. 120 | No. 25
June 20, 2023
PubMed: 37307463

Classifications

Data, Materials, and Software Availability

All study data are included in the article and/or supporting information. Code and data used for the results in this paper can be found at https://github.com/bhamla-lab/myonememodelPNAS2023 (84).

Submission history

Received: October 17, 2022
Accepted: May 1, 2023
Published online: June 12, 2023
Published in issue: June 20, 2023

Keywords

  1. ultrafast motion
  2. calcium-powered dynamics
  3. protist physiology
  4. mechanochemical modeling
  5. synthetic biology

Acknowledgments

We wish to thank Shane Terrell for help setting up the experiments, Vincent Boudreau for providing samples of Vorticella in fresh pond water, Joe Brzostowski, Tanner Fadero, and Sam Lord for technical assistance with the microscopy, KC Huang and Tristan Ursell for providing the Morphometrics software, and Joe Lannan, Christian Pagán-Medina, Joël Lemière, and Jane Maienschein for helpful discussions. This work was mainly supported by NSF awards EF-1935260, EF-1935261, and EF-1935262. S.V. was supported by the National Institute of General Medical Sciences of the NIH under Award No. R35GM147400. The content is solely the responsibility of the authors and does not necessarily represent the official views of the NIH. C.F. acknowledges support from the University of Chicago through a Chicago Center for Theoretical Chemistry Fellowship. M.S.B. acknowledges funding support from NIH Grant R35GM142588; NSF Grant MCB-1817334; NSF CAREER IOS-1941933; and a gift from the Open Philanthropy Project. F.C. acknowledges funding support from NIH Grant R35GM141796. We acknowledge the University of Chicago’s Research Computing Center for computing resources, the Marine Biological Laboratory Whitman Summer Investigator Program and the Physiology Course for the use of microscopy equipment, and Zeiss and pco. for loans of cameras.
Author contributions
C.F., A.T.M., X.L., J.E.H., F.C., M.W.E., S.V., A.R.D., and M.S.B. designed research; C.F., A.T.M., M.W.E., S.V., A.R.D., and M.S.B. performed research; C.F., A.T.M., M.W.E., S.V., A.R.D., and M.S.B. analyzed data; and C.F., A.T.M., M.W.E., S.V., A.R.D., and M.S.B. wrote the paper.
Competing interests
The authors declare no competing interest.

Notes

This article is a PNAS Direct Submission.

Authors

Affiliations

Department of Chemistry and James Franck Institute, University of Chicago, Chicago, IL 60637
Department of Cell and Tissue Biology, University of California, San Francisco, CA 94143
Xiangting Lei
School of Chemical and Biomolecular Engineering, Georgia Institute of Technology, Atlanta, GA 30318
Jerry E. Honts
Department of Biology, Drake University, Des Moines, IA 50311
Department of Cell and Tissue Biology, University of California, San Francisco, CA 94143
Department of Physics, North Carolina State University, Raleigh, NC 27607
Suriyanarayanan Vaikuntanathan1 [email protected]
Department of Chemistry and James Franck Institute, University of Chicago, Chicago, IL 60637
Department of Chemistry and James Franck Institute, University of Chicago, Chicago, IL 60637
School of Chemical and Biomolecular Engineering, Georgia Institute of Technology, Atlanta, GA 30318

Notes

1
To whom correspondence may be addressed. Email: [email protected], [email protected], [email protected], or [email protected].

Metrics & Citations

Metrics

Note: The article usage is presented with a three- to four-day delay and will update daily once available. Due to ths delay, usage data will not appear immediately following publication. Citation information is sourced from Crossref Cited-by service.


Altmetrics




Citations

Export the article citation data by selecting a format from the list below and clicking Export.

Cited by

    Loading...

    View Options

    View options

    PDF format

    Download this article as a PDF file

    DOWNLOAD PDF

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Personal login Institutional Login

    Recommend to a librarian

    Recommend PNAS to a Librarian

    Purchase options

    Purchase this article to access the full text.

    Single Article Purchase

    A unified model for the dynamics of ATP-independent ultrafast contraction
    Proceedings of the National Academy of Sciences
    • Vol. 120
    • No. 25

    Figures

    Tables

    Media

    Share

    Share

    Share article link

    Share on social media