Paper The following article is Open access

Efficiencies of a molecular motor: a generic hybrid model applied to the F1-ATPase

and

Published 16 October 2012 © IOP Publishing and Deutsche Physikalische Gesellschaft
, , Citation Eva Zimmermann and Udo Seifert 2012 New J. Phys. 14 103023 DOI 10.1088/1367-2630/14/10/103023

1367-2630/14/10/103023

Abstract

In a single-molecule assay, the motion of a molecular motor is often inferred by measuring the stochastic trajectory of a large probe particle attached to it. We discuss a simple model for this generic setup taking into account explicitly the elastic coupling between the probe and the motor. The combined dynamics consists of discrete steps of the motor and the continuous Brownian motion of the probe. Motivated by recent experiments on the F1-ATPase, we investigated three types of efficiencies both in simulations and in a Gaussian approximation. Overall, we obtained good quantitative agreement with the experimental data. In particular, we clarify the conditions under which one of these efficiencies becomes larger than 1.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Molecular motors are protein complexes of the size of nanometers that convert chemical energy into mechanical motion [1, 2]. Operating in an aqueous solution, they exhibit stochastic dynamics and energetics due to the influence of thermal fluctuations. Unbalanced concentrations of the molecules providing chemical energy as the input cause the motor proteins to operate under non-equilibrium conditions, which induces a rectified motion with non-zero average velocity. Consequently, molecular motors are often modelled using the Langevin, Fokker–Planck or master equations. The so-called ratchet models combine continuous diffusive spatial motion with a stochastic switching between different potentials corresponding to different chemical states [3, 4]. Alternatively, transitions among a discrete state space governed by master equations provide another way to model molecular motors [59].

A quantity of general interest is the efficiency of such stochastic machines [1014]. For motor proteins, different kinds of efficiencies can be defined, depending on whether one focuses on the work against an external force, i.e. thermodynamic efficiency, or whether the work against viscous friction is also taken into account like in the Stokes or generalized efficiency [1520].

Experimentally, the properties of motor proteins can be investigated in single-molecule experiments by attaching probe particles of the size of micrometers to the motor protein and observing the trajectories of the probes. Additionally, such probes allow us to exert forces on these motor proteins [2124]. Literally speaking, in these assays one cannot observe the motion of the motor directly, but rather one has to infer its properties by analyzing the trajectory of the probe particle. Generically, some elastic linker couples these two elements. Inferring the properties of the motor protein requires one to consider the interaction effects that depend on the linkage between the motor protein and the probe [2532].

In this paper, we discuss a minimal hybrid model for such a motor protein assay that includes this elastic link explicitly. The motor protein and the probe will be modelled as two degrees of freedom moving along a spatial coordinate. In particular, we investigate the different kinds of efficiencies used previously for describing the energetics of molecular motors, and compare our results quantitatively with recent experiments on the rotary motor protein F1-ATPase [3336]. Previous theoretical modelling of F1-ATPase using a discrete state model as well as a ratchet model assuming the probe to stick directly at the motor has focused in particular on the dependence of the rotational behavior on friction, external forces, nucleotide concentrations and temperature as well as on chemical and thermodynamic efficiency and the fluctuation theorem [37, 38]. A detailed modelling of the rotary mechanism and the involved subunits can be found in [39, 40].

This particularly well-studied molecular motor consists of three α and three β subunits arranged around a central γ shaft [41]. The binding and hydrolysis of an ATP molecule at a β subunit causes a rotation of the γ shaft of 120° [21], which has been observed to consist of two substeps of 90° and 30° [42]. An external torque exerted on the γ shaft (as experimentally done in [23] or by the Fo part within the cell) induces ATP synthesis. Coupled to the membrane embedded Fo part, F1-ATPase provides ATP for further hydrolysis reactions, thereby playing an important role in the energy transfer of cells. Experimental observations of the F1-ATPase in the hydrolysis direction include measurements of different kinds of efficiencies. The Stokes efficiency, a Stokes efficiency confined to single jumping events and the thermodynamic efficiency, especially in stall conditions, have been investigated [34, 43, 44]. These experiments led to values for the Stokes efficiency and thermodynamic efficiency of almost 1, suggesting that the F1-ATPase can use almost the complete chemical energy either to drive the probe through a viscous medium or to perform work against an external force. Recently, a measure of the efficiency that explicitly takes care of fluctuations was introduced [33]. The definition of efficiency used there also provided values close to 1 for the examined parameters. Our analysis will show that the latter efficiency can easily reach values larger than 1.

2. The hybrid model

2.1. Single-molecule dynamics

The one-dimensional model we will use to describe a molecular motor with an attached probe particle consists of two degrees of freedom, representing the motor protein at position n(t) and the probe at position x(t), respectively, see figure 1. For a rotary motor like the F1-ATPase, the rotary motion is mapped to a linear one for simplicity. Both constituents are linked via a harmonic potential

Equation (1)

with spring constant κ, where we have included a possible rest length of the linker in the definition of x.

Figure 1.

Figure 1. Schematic representation of the motor protein (blue) with the attached probe (red). The instantaneous distance between the motor protein and the probe is denoted by y(t). The probe moves along a continuous spatial coordinate x(t) and is subject to an external force fex. With transition rates k±(nj,x) the motor protein jumps at times τj between discrete states nj separated a distance d. The load sharing factors θ+ and θ indicate the position of an underlying unresolved potential barrier relative to the potential minimum.

Standard image

The motion of the probe particle is described by an overdamped Langevin equation with friction coefficient γ and constant external force fex,

Equation (2)

including the random force γζ(t) that the solvent exerts on the probe. The thermal fluctuations are assumed to be Gaussian white noise with zero mean and correlations 〈ζ(t1)ζ(t2)〉 = 2kBTδ(t1 − t2)/γ, where kB is Boltzmann's constant and T is the temperature of the solvent.

The motor protein jumps at times τj from nj to nj ± d with transition rates k±(nj,x), hydrolyzing (or synthesizing) one ATP molecule per jump, which corresponds to tight mechanochemical coupling. In this minimal model, we take into account only one chemical state. The transition rates have to fulfill a local detailed balance condition of the form

Equation (3)

where we assume that the jumps of the motor protein take place instantaneously. The free energy change of the solvent

Equation (4)

is associated with ATP turnover. Implementing mass action law kinetics and the concept of a barrier in the potential of mean force for the unresolved chemical steps, the individual rates become

Equation (5)

and

Equation (6)

with

Equation (7)

and ci the concentrations of the nucleotides (i = ATP, ADP, Pi). Here, the transition rate keq applies to equilibrium concentrations of nucleotides. The load sharing factors θ+ and θ, with θ+ + θ = 1, depend on the unresolved shape of the free-energy landscape of the motor protein [5, 45].

2.2. The Fokker–Planck equation

The transition rates, as well as the force the motor protein exerts on the probe, depend on the distance

Equation (8)

between the motor and the probe. The corresponding probability density p(y,t) obeys the Fokker–Planck-type equation

Equation (9)

which contains in the first line the contributions of the transitions of the motor protein and in the second line the drift and diffusion of the probe particle.

For large t and constant nucleotide concentrations, the system reaches a stationary state with time-independent ps(y) and constant mean velocity $\langle \dot {n} \rangle =\langle \dot {x} \rangle \equiv v$ with

Equation (10)

and

Equation (11)

where 〈...〉 denotes the average using the stationary distribution ps(y), which, however, cannot be determined analytically.

3. Efficiencies

3.1. The first law on a single trajectory

Following the concept of stochastic thermodynamics [46, 47], one can assign a first law at the level of a single trajectory. If the probe moves a small distance Δx, the first law becomes

Equation (12)

where ΔqP is the heat dissipated by the probe, fexΔx is the work against the external force and (∂xVx is the change in the internal energy of the spring due to the motion of only the probe. A jump of the motor protein gives rise to a first law in the form of [48]

Equation (13)

without a contribution of the internal energy of the motor, as its internal energy does not change in the one-state model. The change in the internal energy of the spring is given by

Equation (14)

where the sign depends on the direction of the jump. Due to ATP turnover, the internal energy of the solution changes by ΔESol = −Δμ + TΔSSol, where ΔSSol is the change in the entropy of the solution. The heat dissipated by the motor protein in this transition is denoted by ΔqM.

3.2. The first law: ensemble average

On average, the chemical energy gained from ATP consumption that involves changes in the entropy of the solvent will be dissipated as heat Q in the environment and/or is delivered as work against the external force. In the stationary state, the internal energy of the spring is constant on average. Taking the average rates of (12) and (13) and summing the two contributions, this first-law condition can be expressed as

Equation (15)

where the dot denotes a rate and

Equation (16)

is the rate of free energy consumption. The rate of dissipated heat $\dot {Q}=\dot {Q}_{\mathrm {P}}+\dot {Q}_{\mathrm {M}}$ has two contributions. First, the heat flow through the motor protein is given by

Equation (17)

representing the fact that while jumping, the motor protein uses free energy from the hydrolysis to load the spring which corresponds to a change of the internal energy of the spring $\dot {V}_n$ with

Equation (18)

Equation (19)

The energy thus stored in the spring is then dissipated by the probe whose heat flow is given by

Equation (20)

where

Equation (21)

is the local mean velocity of the probe for a given y [49, 50] which corresponds to the current arising from the motion of only the probe in (9).

3.3. Three different efficiencies

We will now focus on three different definitions of efficiency that have been proposed for motor proteins.

In the absence of an external force (fex = 0), one can compare the energy that the motor protein transfers to the spring, $\dot {V}_n$ , with its available chemical energy $\dot {\Delta \mu }$ . From (15) and (17) it follows that $\dot {V}_n=\dot {Q}_{\mathrm {P}}$ . The ratio of the on average dissipated heat through the probe and the available free energy

Equation (22)

was proposed as the definition of efficiency [33]. We will see below that ηQ is not bounded by 1, as has been anticipated earlier [16, 51], and therefore we will call it a pseudo efficiency. A second type of efficiency is the Stokes efficiency,

Equation (23)

which compares the mean drag force γv the probe feels with the available chemical force. In contrast to ηQ, ηS is bounded by 1 [16]. If the motor protein exerted a constant force on the probe, the Stokes efficiency would be equal to the pseudo efficiency ηQ, because in this case the average heat dissipated by the probe is the mean drag force times d.

Finally, in the presence of an external force acting on the probe, the thermodynamic efficiency of the system is the ratio between the mechanical work delivered to the external force and the available free energy [10]

Equation (24)

For fex ≠ 0, the pseudo efficiency ηQ can be defined as

Equation (25)

4. Gaussian approximation

4.1. Derivation

For a comparison with the simulations and in order to gain more analytical insights, it will be convenient to have a simple approximation for the stationary distribution ps(y). For a Gaussian probability distribution

Equation (26)

the free parameters $\bar {y}$ for the mean and σ2 for the variance can be determined by requiring that the time derivative of these quantities as calculated with the Fokker–Planck equation (9) vanishes in the steady state. These conditions result in the following two equations for $\bar {y}$ and σ2:

Equation (27)

and

Equation (28)

where we have introduced the average jump rates

Equation (29)

Equation (30)

These equations can easily be solved numerically.

4.2. The limits Δμ → 0 and Δμ → 

Close to chemical equilibrium, i.e. Δμ = 0, and for fex = 0, we expand $\bar {y}$ and κσ2 − kBT up to first order in Δμ and find that

Equation (31)

and

Equation (32)

The coefficients A, $\tilde {A}$ and B obtained by solving the first order of (27) and (28) are too long to be shown here.

In the limit Δμ →  and fex = 0, we obtain for $\bar {y}$ and κσ2 − kBT

Equation (33)

and

Equation (34)

as long as θ > 0. The coefficients

Equation (35)

and

Equation (36)

are obtained by solving (27) and (28) to first and second order in Δμ.

4.3. Efficiencies

Within this Gaussian approximation, the average heat flow through the probe as given by (20) is calculated using the local mean velocity (21)

Equation (37)

The average over y can now be performed leading to

Equation (38)

This expression is used to determine ηQ in this approximation as

Equation (39)

with $\bar {y}$ and σ2 being the solution of (27) and (28) for given Δμ and keq.

For small Δμ, using (31) and (32), ηQ takes the form

Equation (40)

If θ+ ≠ θ, ηQ diverges for vanishing Δμ. For θ+ > θ, ηQ can become negative due to those jumps of the motor protein that occur when the previous diffusion of the probe has resulted in y < −0.5d. Then, the energy stored in the spring is dissipated by the motor protein during jumping.

In the limit of large Δμ, we use (34) and (35) to obtain

Equation (41)

which approaches 1.

The Stokes efficiency in the Gaussian approximation without an external force is simply given by

Equation (42)

For κσ2 > kBT, which is the case for θ+ < 0.5, the Stokes efficiency is always smaller than ηQ. For vanishing Δμ, ηS approaches a finite value, $\eta _{\mathrm {S}}\approx d\kappa A+d\kappa \tilde {A}(\theta _{+}-\theta _{-})^2$ , while for Δμ →  it also converges to 1.

5. Results

In this section, we study the three efficiencies for our hybrid model as functions of the chemical energy Δμ, the absolute concentrations of the nucleodides, i.e. keq, the external force fex and the load sharing factor θ+. The data have been obtained from simulations, using a Gillespie algorithm [52] similar to that in [37] with the motion of the probe being spatially discretized in steps of Δx = d/1000, and compared with the Gaussian approximation.

We use the model parameters as given in table 1, which are motivated by experimental results for the F1-ATPase as described in section 6. The load sharing factor θ+ remains a free parameter.

Table 1. Values of the model parameters used for the simulation and the Gaussian approximation.

γ (kBT s/d2) κ (kBT/d2) keq/ceqATP (M−1s−1)
0.407 40 3×107

Simulated trajectories with the same nucleotide concentrations as those used in the experiment [33] are shown in figure 2. In the presence of low nucleotide concentrations, only a few backward jumps of the motor protein take place and the trajectory of the probe shows an almost staircase-like form. For high nucleotide concentrations, following a forward step the motor often performs a backward jump. Such a sequence of two jumps is not necessarily visible in the trajectory of the probe, which remains almost linear.

Figure 2.

Figure 2. Trajectories of the molecular motor (black) and the attached probe particle (green) obtained from the simulation (without external force). In the presence of low nucleotide concentrations (I) the trajectory of the probe exhibits a more stepwise form, while it becomes almost linear for high nucleotide concentrations (III). The parameters are θ+ = 0.1, (I) cATP = 0.4 μM, cADP = 0.4 μM, cPi = 1 mM, (II) cATP = 2 μM, cADP = 2 μM, cPi = 1 mM, (III) cATP = 100 μM, cADP = 100 μM, cPi = 1 mM. For all parameter sets, we have Δμ = 19.14kBT. The nucleotide concentrations are the same as those used in the experiment [33] shown in figure 8 with the same labelling (I–III) below.

Standard image

5.1. Pseudo efficiency ηQ

We will first investigate the pseudo efficiency ηQ as a function of Δμ, keq and θ+. We extract $\dot {Q}_{\mathrm {P}}$ from the numerical data by averaging over one sufficiently long trajectory. The results are shown in figure 3. The most striking fact regarding these data is the observation that ηQ is larger than 1 for small enough Δμ and θ+, which shows up in the Gaussian approximation as well. This effect can be understood as follows. In a jump, the motor protein can take heat from the solution in order to change the internal energy of the spring by an amount larger than Δμ. If, subsequently, the probe dissipates this internal energy of the spring as heat back into the environment, $\dot {Q}_{\mathrm {P}}$ can indeed become larger than $\dot {\Delta \mu }$ without any violation of the second law. Using the obtained parameter for the spring constant κ, the motor protein transfers 20 kBT to the spring if it starts the jump from the minimum of the harmonic potential. For small values of θ+, the forward jump rate of the motor protein depends only weakly on the current position of the probe as shown in figure 4. Therefore, jumps will occur even if the associated change in internal energy of the spring, ΔV , is larger than Δμ. For a rather small keq, backward jumps are rare and the probe relaxes to the potential minimum between successive forward jumps (see data set I in figure 2). This leads to $\dot {V}_n>\dot {\Delta \mu }$ on average and hence to ηQ > 1 for Δμ considerably smaller than 20kBT as shown in figure 3. As the value of θ+ increases, ηQ decreases because the forward jumps of the motor protein are suppressed. On average, in this case the motor protein jumps only if the probe has diffused forward and exerts a pulling force on the motor through the spring.

Figure 3.

Figure 3. Pseudo efficiency ηQ from the simulation (a) and (b) and within the Gaussian approximation (c) and (d). (a) and (c) ηQ as a function of Δμ for different values of the load sharing factor θ+ and fixed keq = 10−5 s−1. (b) and (d) ηQ as a function of Δμ for different values of keq with fixed θ+ = 0.1. The remaining parameters are κ = 40kBT/d2, γ = 0.407kBT s/d2. In the simulation, the error is of the order of the symbol size.

Standard image
Figure 4.

Figure 4. Probability distribution p(y)|jump of the distance y just before a jump of the motor protein for several values of θ+ at Δμ = 13kBT and keq = 10−5 s−1. For small θ+, the forward jumps of the motor protein are almost independent of the position of the probe resulting in a peak at y ≃ 0, whereas for larger θ+ the peak clearly shifts to y < 0, implying that the motor protein prefers to jump when the probe has diffused ahead. The peaks around y = 1 indicate backward jumps which take place more often in the case of small θ+ when the backward rate is more sensitive to the position of the probe.

Standard image

Increasing the absolute concentrations of the nucleotides, i.e. increasing keq, results in more forward but also more backward jumps, which can be seen for data set III in figure 2. For small Δμ, the occasional backward jumps follow especially those forward jumps for which the change in internal energy of the spring has been larger than Δμ, leading to a smaller ηQ.

In the limit of large Δμ, the motor protein jumps even when the spring is previously stretched, which can result in changes of the internal energy of the spring by an amount larger than 20kBT. The coupling between the motor protein and the probe induces a balancing effect between the forward motion of the motor protein and the drag of the probe maintaining a typical $\dot {V}$ that turns out to be approximately $\dot {\Delta \mu }$ , leading to ηQ ≃ 1.

5.2. Stokes efficiency ηS

We also obtain the Stokes efficiency (23) from the simulated trajectories and the Gaussian approximation as shown in figure 5. Characteristically, starting close to 0 for small Δμ, ηS monotonically increases with Δμ reaching 1 for Δμ → . For small Δμ, the trajectory of the probe shows a staircase form with small average velocity leading to small values of the Stokes efficiency in contrast to values of the pseudo efficiency ηQ > 1. For large Δμ, the probe does not relax to the potential minimum between consecutive jumps, resulting in a more linear trajectory of the probe as if it was exposed to an almost constant force. In this limit of an almost linear motion of the probe, the pseudo efficiency becomes the Stokes efficiency. As ηS is bounded by 1, ηQ cannot reach values larger than 1 in this limit either.

Figure 5.

Figure 5. (a) Stokes efficiency ηS from the simulation (a) and (b) and within the Gaussian approximation (c) and (d). The data are obtained from the same trajectories as those used to obtain ηQ in figures 3(a) and (b). (a) and (c) ηS as a function of Δμ for different values of the load sharing factor θ+ and fixed keq = 10−5 s−1. (b) and (d) ηS as a function of Δμ for different values of keq for fixed θ+ = 0.1.

Standard image

Increasing the load sharing factor θ+ results in decreasing average velocities. Therefore, the Stokes efficiency also decreases, which can be seen in figures 5(a) and (c). With increasing absolute concentrations of nucleotides, i.e. with increasing keq, the average velocity and therefore also the Stokes efficiency at fixed Δμ increases as shown in figures 5(b) and (d).

5.3. Thermodynamic efficiency ηT

The thermodynamic efficiency of the system can be studied only if an external force is applied to the probe. As an illustrative example, for fixed Δμ = 19kBT, we examine the thermodynamic efficiency in the presence of external forces smaller than the stall force as shown in figure 6. The thermodynamic efficiency increases linearly with fex and reaches 1 at the stall force which is possible only due to the tight mechanochemical coupling in this model. In figure 6 we also show the pseudo efficiency ηQ as defined in (25) in the presence of external forces. While ηQ is almost independent of the external force, the contribution from the dissipated heat, $\dot {Q}_{\mathrm {P}}/\dot {\Delta \mu }$ , decreases linearly with fex and reaches zero at the stall force.

Figure 6.

Figure 6. Pseudo efficiency ηQ (red and black dots), thermodynamic efficiency ηT (green and blue squares) and dissipated heat ΔQP per Δμ (cyan and purple diamonds) as functions of fex for fixed Δμ = 19kBT, θ+ = 0.1 and keq = 10−5 s−1. The plot contains the data from the simulation (S) as well as from the Gaussian approximation (G).

Standard image

In stall conditions, the work corresponding to the stall force refers to the maximum work the motor protein can convert on average. In the simulation, we find that applying fex = Δμ/d generates a diffusive motion of the motor protein with v ≃ 0 and $\dot {Q}_{\mathrm {P}}\simeq 0$ for various values of θ+ and Δμ, including the ones with ηQ > 1 and ηQ < 1. This observation implies that the motor protein seems, in principle, to be able to convert the full Δμ into extractable work. In our model, where the motor interacts with the external force only via the spring, this result is not trivial, as it would have been if we had applied fex directly to the motor. Under these conditions, ps(y) is Gaussian with 〈y〉 = Δμ/(dk) and the same variance as the Boltzmann distribution in equilibrium, σ2 = kBT/κ.

Within the Gaussian approximation we can insert fex = Δμ/d into (27) and (28). With κσ2 = kBT, $\bar {y}=\Delta \mu /(d\kappa )$ , i.e. v = 0, is a solution for fex = Δμ/d, implying that also in the Gaussian approximation the motor protein is able to convert the full Δμ into extractable work for any values of the load sharing factors given that θ+ + θ = 1.

6. Case study: F1-ATPase

In this section, we apply our hybrid model to the F1-ATPase and compare the simulations with recent experimental data [3335].

6.1. Model parameters

For a quantitative comparison, we have to map the rotary motion of the F1-ATPase to our linear model and determine the model parameters. In our model, one jump of the motor protein covering a distance d corresponds to a rotation of the γ shaft of 120°. Using large probe particles such as polystyrene beads or actin filaments, the substeps in one 120° rotation are not resolved experimentally. Therefore, we will omit the substeps here, too. We assume that the temperature of the solution is T ≃ 24 °C and that the probe consists of two beads of diameter 287 nm [33]. The friction coefficient of the probe can be calculated using the formula for the rotational frictional coefficient Γ from [36, 37] with the viscosity of water (η ≃ 0.001 N s m−2). The frictional torque $N=\Gamma \dot {\varphi }$ acting on the probe with angular velocity $\dot {\varphi }$ corresponds to a frictional force

Equation (43)

acting on the probe at distance r from the γ shaft. Within one 120° rotation, the probe at distance r covers d = 2πr/3. For the linear model, the friction coefficient γ can be calculated as

Equation (44)

leading to γ = 0.407kBT s/d2.

Following the mass action law assumption, the equilibrium transition rate keq is supposed to depend linearly on the concentrations of nucleotides in the solvent. For low ATP concentrations (cATP ≃ 10−6 M), the mean velocity of the motor protein is dominated by the rate of ATP binding. In the one-step model this feature holds for all concentrations. Therefore, we choose keq to be the experimentally determined rate of ATP binding keq ≃ 3 × 107 M−1 s−1 ceqATP [42]. For known non-equilibrium concentrations of nucleotides like in the experiments, the structure of the transition rates (5) and (6) leaves the choice of the equilibrium concentrations arbitrary as long as they obey

Equation (45)

for T = 23 °C and pH 7 [37]. For given keq and Δμ, one possible choice of the non-equilibrium concentrations of nucleotides is cADP = ceqADP, cPi = ceqPi and $c_{\mathrm {ATP}}=c^{\mathrm {eq}}_{\mathrm {ATP}}\exp [\Delta \mu /k_{\mathrm {B}}T]$ which was used for the simulation and the Gaussian approximation.

In order to determine the spring constant κ and the load sharing factor θ+, we use both the experimental data of the mean velocities [33] and the histogram of the angular position of the probe at a jump [35]. While both data sets depend on both parameters, the velocity, especially for large keq, is more sensitive to κ, whereas the peak position of the histogram mainly depends on θ+. Therefore, we primarily use the velocity data to fit κ and determine the load sharing factor θ+ by comparing the peak position of the experimental histogram [35] with the left peak position of the corresponding histograms obtained by our simulation as the ones shown in figure 4.

As a result, we obtain κ = 40 ± 5kBT/d2 and a value of θ+ in the range 0.1 ≲ θ+ ≲ 0.3. In figure 7, we show how for this value of κ changing the load sharing factor affects the mean velocity for which we get the best overall agreement for θ+ = 0.1. For later purposes, we also include data for θ+ = 0.01.

Figure 7.

Figure 7. Comparison of the mean velocities observed experimentally [33] (red squares) and in the simulation for several load sharing factors θ+ (black dots, green diamonds and blue triangles). The labelling I–V refers to the corresponding parameter sets in figure 8.

Standard image

6.2. Comparison of efficiencies with experimental data

Experimentally, the heat flow of the probe is determined using the Harada–Sasa relation [53]. In the appendix, we show that this heat flow is equal to $\dot {Q}_{\mathrm {P}}$ as defined in (20). In figure 8, we plot the average heat released through the probe per step, QP, plus the work against the external force, W, obtained by the simulation for κ = 40kBT/d2 and θ+ = 0.1 and compare it with the experimental results [33]. We found quite good agreement between theory and experiment for the parameter sets I–IV where either the maximum deviation is 15% (II–IV) or our theoretical value is included in the experimental error range (I). As an aside, we note that for the parameter sets I–III (without external force) the simulated mean velocities also coincide well with the experimental values with a maximum deviation of 10% as shown in figure 7. For illustrative purposes, we also plot QP plus W for θ+ = 0.01 which shows better agreement with the experimental data (but is not consistent with the range of θ+ obtained in section 6.1).

Figure 8.

Figure 8. Average heat QP released through the probe (green and cyan) and work W ≡ fexd against the external force (blue) compared to the available free energy per step Δμ (red line). The dissipated heat of the probe is split into two contributions QS and QV according to the two terms of the Harada–Sasa relation (A.3). The contribution from the linear motion with constant mean velocity, QS (cyan), appears in the numerator of the Stokes efficiency while QV (green) is the contribution due to the non-uniform jumping motion of the motor protein. In each of the five parameter sets labelled by I–V, the left and the central bars represent results from the simulation for θ+ = 0.1 and 0.01, respectively, while the right bar shows the experimental results and error bars from [33]. The following parameters were used in the five cases: (I) cATP = 0.4 μM, cADP = 0.4 μM, cPi = 1 mM, i.e. keq = 5.87 × 10−8 s−1 and Δμ = 19.14kBT; (II) cATP = 2 μM, cADP = 2 μM, cPi = 1 mM, i.e. keq = 2.93 × 10−7 s−1 and Δμ = 19.14kBT; (III) cATP = 100 μM, cADP = 100 μM, cPi = 1 mM, i.e. keq = 1.47 × 10−5 s−1 and Δμ = 19.14kBT; (IV) cATP = 2 μM, cADP = 2 μM, cPi = 1 mM, fex = 9.27kBT/d, i.e. keq = 2.93 × 10−7 s−1 and Δμ = 19.14kBT; (V) cATP = 2 μM, cADP = 0.5 μM, cPi = 0.5 μM, i.e. keq = 3.67 × 10−11 s−1 and Δμ = 28.12kBT.

Standard image

Discrepancies between our theory and the experiment are visible in figures 7 and 8 where for the parameter set V both the average velocity and the pseudo efficiency deviate significantly from the experimental values for κ = 40kBT/d2 and θ+ = 0.1. For Δμ = 28.12kBT corresponding to the data set V in figure 8, the probe just reaches the potential minimum between consecutive jumps of the motor protein. Therefore on average at most 20kBT can be transferred to the spring leading to ηQ ≃ 0.7, which is less than the experimental value. This discrepancy is most likely due to the fact that we have omitted substeps in our model. In the simulation, the average velocity then does not show saturation as one would expect it to result from the hydrolysis step [42] which should be experimentally observable at the higher concentrations used in [33].

The confinement of θ+ to the range 0.1 ≲ θ+ ≲ 0.3 implies, on the one hand, that the potential of the mean force of the motor protein should be asymmetric and, on the other hand, that asymmetric potentials with a barrier state close to the initial state seem to enhance the ability of the motor protein to perform work on the spring, in accordance with [54]. If θ+ were larger, ηQ would decrease and the experimentally determined values of ηQ would not be reached in the simulation. If θ+ were smaller, ηQ would approach the experimental values better; however, the distribution of the position of the probe just before a jump as shown in figure 4 would then no longer coincide with the experimentally observed distribution (see [35]).

Information about the thermodynamic efficiency of the motor protein can be gained by applying an external force to the probe. In our simulation, the stall force is found to be fex = Δμ/d, implying that the motor protein is able to convert the full Δμ into extractable work without dissipation in accordance with the experiments performed in [34].

7. Conclusion

In summary, we have discussed a simple generic model which includes the elastic linker between the probe particle and the molecular motor. The properties of the motor become typically accessible only through the observation of the motion of the probe. We have then focused on discussing three types of efficiencies within this model using both simulations and a Gaussian approximation to the stationary distribution for the distance between the motor and the probe. The genuine thermodynamic efficiency is non-zero only if an external force is applied to the probe. The Stokes efficiency deviates from 1 due to the discrete nature of the motor steps which become less relevant with increasing ATP concentration. A pseudo efficiency measuring how much of the free energy of ATP hydrolysis ends up in loading the elastic element can even become larger than 1 close to equilibrium and for a barrier state close to the initial state.

Applying this minimal model to recent experimental data for the F1-ATPase, we find overall good agreement except for those parameters where, in particular, the Pi concentration is very small. In general, one should consider ATP binding and ATP hydrolysis as two separate steps. Such a refinement as well as a splitting of the 120° rotation into two steps of 90° and 30° as experimentally observed using much smaller probe particles does not pose new conceptual challenges to the present framework and will be pursued elsewhere.

Appendix.: Equivalence of heat flow $\dot {Q}_{\mathrm {P}}$ with the one inferred from the Harada–Sasa relation

Experimentally, the heat flow caused by the probe has been inferred from measuring the autocorrelation function $C_{\dot {x}}(\tau )=\langle \dot {x}(t+\tau )\dot {x}(t) \rangle -v^2$ and the linear response function

Equation (A.1)

of the velocity of the probe to a small external perturbation h(t) of the probe within the steady state [33]. The heat flow is then given by an equality derived by Harada and Sasa [53]

Equation (A.2)

Equation (A.3)

with $\tilde {C}_{\dot {x}}$ and $\tilde {R}_{\dot {x}}$ being the Fourier transforms of $C_{\dot {x}}$ and $R_{\dot {x}}$ . Using a path weight approach described in [55] applied to our system, the response function follows as

Equation (A.4)

Inserting $C_{\dot {x}}$ and this $R_{\dot {x}}$ into (A.2), one immediately finds that

Equation (A.5)

which is equal to $\dot {Q}_{\mathrm {P}}$ in (20).

Please wait… references are loading.