Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2015 Jun 29;112(28):E3639–E3644. doi: 10.1073/pnas.1506825112

Hydrodynamic collective effects of active protein machines in solution and lipid bilayers

Alexander S Mikhailov a,b, Raymond Kapral c,d,1
PMCID: PMC4507222  PMID: 26124140

Significance

Biological cells contain large numbers of active proteins that repeatedly change their conformations and need a supply of ATP or other substrates to maintain their cyclic operation. Whereas these protein machines have a variety of specific functions, acting as motors, ion pumps, or enzymes, they also induce fluctuating hydrodynamic flows in the cytoplasm. Because such fluctuating flows are nonthermal, energy can be extracted from them and work can be performed. We show that these flows can substantially enhance diffusive motions of passive particles. Furthermore, when gradients in concentrations of active proteins or substrate (ATP) are present, a chemotaxis-like drift should take place. Such nonequilibrium effects are universal: They hold for all passive particles and also for the protein machines themselves.

Keywords: active proteins, collective hydrodynamic effects, nonthermal fluctuation effects, enhanced passive particle diffusion

Abstract

The cytoplasm and biomembranes in biological cells contain large numbers of proteins that cyclically change their shapes. They are molecular machines that can function as molecular motors or carry out various other tasks in the cell. Many enzymes also undergo conformational changes within their turnover cycles. We analyze the advection effects that nonthermal fluctuating hydrodynamic flows induced by active proteins have on other passive molecules in solution or membranes. We show that the diffusion constants of passive particles are enhanced substantially. Furthermore, when gradients of active proteins are present, a chemotaxis-like drift of passive particles takes place. In lipid bilayers, the effects are strongly nonlocal, so that active inclusions in the entire membrane contribute to local diffusion enhancement and the drift. All active proteins in a biological cell or in a membrane contribute to such effects and all passive particles, and the proteins themselves, will be subject to them.


Protein machines play a fundamental role in biological cells (1, 2). Operating as motors, they are responsible for intracellular transport and force generation. As manipulators, they perform various operations involving other biomolecules, including RNA and DNA. As pumps, they transfer ions across biomembranes. A common feature of protein machines is that they undergo cyclic conformational changes that are induced by ligand binding and product release. Thus, all protein machines are enzymes where substrate binding, catalytic conversion to products, and product release are accompanied by internal mechanochemical motions. Conformational changes within turnover cycles are also characteristic of many other enzymes, which need not function as molecular machines. The results we present in this paper are also applicable to these enzymes.

When a macromolecule cyclically changes its shape, it induces hydrodynamic flows in the surrounding fluid or biomembrane in which it resides. Such pulsating flows can act on any passive particles in solution or lipid bilayers. The aim of the present study is to analyze the collective hydrodynamic effects that active macromolecules have on passive particles in the medium. We shall show that these effects lead to substantial modifications of the diffusion constants of passive particles. Furthermore, directed drift of passive particles can be induced when there are spatial gradients of active macromolecules, a phenomenon that is reminiscent of chemotaxis.

The investigation of hydrodynamic effects in active fluids is an important field of current research (3, 4). Although the hydrodynamics of bacterial motion has been studied often, the active elements may be of inorganic origin and operate through various flow-generation mechanisms (513). There has been a considerable amount of work on swimmers that can propel themselves by cyclically changing their shapes (14). Interactions between such swimmers and their collective flows have been analyzed (1521). Also, interactions between active hydrodynamic dipoles have been investigated theoretically (22, 23) and experimentally (24, 25). Effects of active dipoles on chromatin dynamics in a two-fluid model have been considered (26).

Hydrodynamic effects on individual protein machines have been studied; for example, investigations of simple models of such machines have shown how they propel themselves and behave under a load (27). Also, the effects of hydrodynamic interactions on the internal dynamics have been analyzed (28). Active protein inclusions in lipid bilayers can act as hydrodynamic dipoles (29) and, under certain conditions, such inclusions can behave as active membrane swimmers (30).

The focus of the present study differs in several respects from the work recounted above. We are not interested in the effects of hydrodynamics on the operation of a single machine. Instead, we concentrate on the advection effects that protein machines can have on passive particles in the medium. Although some proteins can indeed behave as motors, we only require that such machines act as cyclic hydrodynamic force dipoles. Consequently, our analysis concerns the statistical effects that populations of incoherently oscillating dipoles can have on passive particles in the system.

In bulk solution, the laws of 3D hydrodynamics need to be applied; however, as already pointed out by Saffmann and Delbrück (31), biological membranes should behave as 2D fluids when lipid flows in a membrane that occur on scales shorter than a micrometer are considered (3133). Recently, 2D lipid flows were directly observed in mesoscopic simulations of lipid bilayers (34); additionally, 2D diffusion in biomembranes was experimentally investigated (35). It is well known that 2D hydrodynamics is characterized by the presence of ultra-long-ranged logarithmic interactions that make it qualitatively different from the 3D case (33). We study the hydrodynamic effects of active machines in both 3D and 2D systems.

The outline of the paper is as follows. First, we give a brief discussion of how protein machines undergoing random cyclic changes in response to substrate binding and product release under nonequilibrium conditions can act as force dipoles. A simple model for an active protein is used to illustrate how force dipole effects arise but our general results do not rely on the specific structure of the model. We then show that a cyclically fluctuating hydrodynamic force dipole will induce diffusive motion and directed drift of a passive particle located at some distance from it. When a population of cyclic hydrodynamic force dipoles is randomly distributed in the medium, they will enhance the diffusion of all passive particles in the medium. Moreover, if such dipoles are nonuniformly distributed and concentration gradients in these species are present, directed flows of all passive particles will be induced. Numerical estimates of the magnitudes of the effects are given, and a discussion of the results is presented.

Protein Machines As Force Dipoles

Molecular machines are biomolecules, most often proteins, that undergo structural changes in shape during their operation cycles. These cyclic shape changes, induced by ligand binding and product release, take place under nonequilibrium conditions; therefore, they differ from thermally induced shape fluctuations for which microscopic reversibility holds and the fluctuation-dissipation theorem applies. These molecular machines operate in a viscous environment and their dynamics takes place under low Reynolds number conditions so that inertia does not play a significant role. As a result, if a force is applied to a particle in the fluid, the same force acts on the fluid.

Such protein machines act as stochastic oscillating force dipoles that can influence the motions of other particles in the system. For example, consider a protein with two domains that operates as an enzyme converting substrate into product molecules. The protein domains close in response to binding of ATP or other substrate molecules and open after the reaction and release of a product. We assume that the substrate is continuously supplied and the products are instantaneously removed from the system. When the substrate binds to the protein new bonds are formed and thus the chemical energy, needed to induce conformational changes and cause the domains to close, is supplied. When the product is released, the additional chemical bonds are broken, leading to domain opening, and the protein returns to its initial state. Within one cycle, an active protein consumes the chemical energy whose net value is determined by the difference in internal energies of the substrate and product molecules. If reverse conversion of a product into the substrate is allowed, an active protein can also operate in the opposite direction. Generally, its cycles are driven by the difference in Gibbs potentials of substrate and product; the sign of this difference determines the operation direction of the machine. Because the net force on the protein is zero, the forces that act on the domains are equal in magnitude and opposite in direction, so that a force dipole is created. This oscillating force dipole will act on the surrounding viscous fluid to generate hydrodynamic flows that can induce motions of passive particles in the fluid. A simple dimer model of such an active protein, where the domains in a bidomain protein are represented by beads, is formulated in SI Text.

SI Text

Here, we formulate and analyze a simple model of a catalytically active bidomain protein. In this model, originally used to study synchronization phenomena in coupled protein machines (45), the two protein domains are represented by beads connected by an elastic spring. Ligand binding is treated by assuming that a short additional link connecting the two beads is created. Thus, substrate binding causes dimer contraction. As the distance between the two domains shortens, a conformation favorable for a catalytic reaction is reached. When substrate conversion into a product occurs, the ligand is released. In the model, this corresponds to disappearance of a short additional link between the protein domains. In the absence of such link, the dimer elongates and returns to its equilibrium conformation where the next substrate molecule can bind, so that the turnover cycle is repeated. Under these conditions, the machine goes through cycles of expansion and contraction, similar to the cyclic conformational changes in a real protein machine.

Specifically, we let the internal state s=0 correspond to the free protein and the state with s=1 correspond to the ligand–protein complex. The free protein has the elastic energy

E0=12k0(x0)2, [S1]

where k0 is the spring constant, 0 is the natural length of the spring and x=x1x2 is the distance between the beads 1 and 2. When the ligand is present, it establishes an additional short elastic link connecting the domains, so that

E1=12k0(x0)2+12kc(xc)2, [S2]

where kc is the spring constant and 0 is the length of the additional link (c<0). This energy can be also written as

E1=C+12k1(x1)2, [S3]

where

k1=k0+kc,  1=k00+kcck0+kc,C=12(k002+kcc2)12(k00+kcc)2(k0+kc)2, [S4]

so that 1 is the equilibrium length of the dimer with the ligand bound. Assuming equal mobility γ for both beads, the internal dynamics of the protein in the two different internal states s=0 and s=1 is described by the stochastic differential equation

x˙=γE0,1x+η(t). [S5]

Here, Gaussian white noise η(t), which satisfies the fluctuation-dissipation relation,

η(t)η(t)=2γkBTδ(tt), [S6]

with T the temperature, is included to take into account thermal fluctuations. Effects of hydrodynamic interactions on the motion of the beads are neglected.

Binding of a ligand is modeled as an instantaneous stochastic transition from the state s=0 to the state s=1. It is characterized by the rate proportional to the substrate concentration cS and depends on the distance x between the domains. We assume that binding is only possible in the equilibrium state of the free protein, so that the binding rate is ν0(x)=cSu0δ(x0). When the substrate binds, it brings the energy, ΔE01=12kc(c0)2, assumed to be much higher than thermal energy kBT. After binding, the ligand–protein complex undergoes relaxation to its own equilibrium state with the distance x=1 between the domains. When this state is reached, a reaction converting the ligand into the product particle occurs and, as we assume, the product is immediately released (and removed, so that it cannot bind again). This step corresponds to an instantaneous reverse transition from s=1 to s=0 taking place at rate ν1(x)=u1δ(x1). The product carries away energy ΔE10=12kc(c1)2. Following product release, the free protein undergoes relaxation to its equilibrium state with x=0 at which the next substrate particle may bind, repeating the entire cycle.

The energy supplied with the ligand to the protein (and dissipated within each its cycle) is

ΔE01ΔE10=12kc(0c)212kc(1c)2=kc2(2k0+kc)2(k0+kc)(0c)2. [S7]

Introducing probability densities for the two states, the model can be equivalently described by a system of two coupled Fokker–Planck equations,

p0t=γx(E1x+kBTx)p0ν0(x)p0+ν1(x)p1p1t=γx(E1x+kBTx)p1+ν0(x)p0ν1(x)p1. [S8]

Note that detailed balance is not satisfied because the condition

ν1(x)ν0(x)e(E0(x)E1(x))/kBT [S9]

cannot hold if ν0(x)=cSu0δ(x0) and ν1(x)=u1δ(x1).

When such an oscillating dimer is immersed in a fluid, it behaves as a fluctuating force dipole,

m(t)=Es(t)xx(t). [S10]

Its mean value is zero and the mean-square intensity is

m(t)2=s=0,1dx(Esx)2x2p¯s(x), [S11]

where p¯0,1(x) is the stationary probability distribution determined from Eq. S8. Moreover, the time correlation function of the force dipole is given by

m(t)m(0)=s,s0=0,1dxdx0(Esx)2×(Es0x0)2x2x02p¯s(x,s,t;x0,s0,0), [S12]

where p¯s(x,s,t;x0,s0,0) is the joint probability to find the system in the state with x and s at time t and in the state with x0 and s0 at the initial time t=0. Note that, at t we have m(t)m(0)=m(t)m(0)=0. Thus, the correlation function S(t) decays on some characteristic correlation time τc. Therefore, its time integral S=0S(t)dt can also be defined.

Note that the force dipole intensity S depends on substrate concentration cS, that is, S=S(cS). When substrate is not supplied, only thermal fluctuations are present. They give rise to a thermal force dipole with intensity ST=ST(cS=0). Generally, the force dipole intensity can be decomposed into two parts, S(cS)=ST+SA(cS). The substrate concentration dependence can be determined by solving Eq. S8 using Eq. S12. When a change in the diffusion coefficient of passive particles due to the hydrodynamical effects of an oscillating force dipole is calculated, only the nonequilibrium contribution SA=S(cS)ST should be considered, because the contribution from thermal force dipole fluctuations is already always included into the equilibrium diffusion coefficient, in the absence of the substrate supply. The function SA(cS) will be proportional to cS at low substrate concentrations and reach saturation at high concentrations cS.

In the above version, all reverse processes were neglected (which is possible assuming high activation energy and immediate removal of the products). The model can, however, also be modified to include such processes and thus make it completely reversible. In this case, Eq. S8 should be replaced by

p0t=γx(E1x+kBTx)p0(cSu0+p0u0p1)δ(x0)(cPu1+p0u1p1)δ(x1)p1t=γx(E1x+kBTx)p1+(cSu0+p0u0p1)δ(x0)+(cPu1+p0u1p1)δ(x1), [S13]

where cS and cP are substrate and product concentrations, respectively. In the extended model, the substrate can bind and dissociate without a reaction in the equilibrium state of the free protein with rate constants cSu0+ and u0. It can be converted into a product and immediately released in the equilibrium state of the ligand–protein complex with the rate constant u1. Furthermore, product also can bind back to the free protein if, owing to a thermal fluctuation, it currently finds itself in a state with the distance 1 between the domains, characteristic for the equilibrium state of the ligand–protein complex. This occurs at the rate constant cPu1+. In our model, the product is immediately converted into the substrate once it became bound back to the enzyme. Depending on the difference in substrate and product concentrations, the reversible model enzyme can cycle in any direction, predominantly converting substrate into product or product into substrate. When operating in the reverse direction, it needs, however, to be assisted by thermal fluctuations, so that the differences in the elastic energies, 1/2k0(10)2 and 1/2k1(10)2, between open and closed conformations of the free enzyme and of the ligand–enzyme complex cannot then be large compared with the thermal activation energy kBT.

Hydrodynamic Effects

When a force F is applied to the fluid at a point r, it produces a fluid flow field at R that advects a particle at this location with velocity

R˙α=Gαβ(Rr)Fβ(r), [1]

where Gαβ is the mobility tensor which, for sufficiently large distances, can be evaluated in the Oseen approximation. (The Einstein summation convention over repeated indices will be used throughout this paper.) For an oscillating dimer of length x with orientation given by the unit vector e and interaction force magnitude F, we have R˙α=[Gαβ(Rrxe)Gαβ(Rr)]eβF, because the forces on the dimer beads have equal magnitude but are opposite in direction. If the dimer length x is small compared with the distance |Rr|, we can approximately write

R˙α=Gαβrγeβeγm, [2]

where m(t)=x(t)F(t) denotes the magnitude of the force dipole. Although we derived Eq. 2 for a specific two-bead model, it is general. If an object immersed in the fluid changes its shape, it generates a hydrodynamic flow that, at large separations from this object, can be described as being produced by an active force dipole, unless special symmetries are present. Below, we shall treat any active protein machine as a nonequilibrium stochastic force dipole.

Consider a collection of such active force dipoles, located at positions {Ri} with orientations {ei} and subject to fluctuations arising from thermal and active nonthermal effects. The oscillating dipoles collectively create a fluctuating flow field that induces stochastic advection of a passive particle. At low Reynolds numbers, the passive particle will exactly follow the local flow velocity field. The equation of motion of a passive particle at point Rα may then be written as

R˙α=iGαβ(RRi)Riγeiβeiγmi(t)+fα(t). [3]

The passive particle is subject to fluctuations from two sources. The random force fα(t) has zero mean and satisfies the fluctuation-dissipation relation, fα(t)fα(t)n=2kBTγδααδ(tt), where γ is the mobility coefficient of the passive particle. The particle is also subject to thermal and active nonthermal fluctuations of the force dipoles mi(t) with zero mean, m(t)n=0, and correlation function mi(t)mj(0)n=m(t)m(0)nδij=S(t)δij, which defines the force dipole correlation function S(t)=m(t)m(0)n. For simplicity we have assumed that the force dipole correlations of different proteins are independent. We have further assumed that their orientations are randomly distributed. The force dipole tensor eiβeiγmi(t) of a protein is invariant under inversion of the orientation vector ei and has properties like the tensor order parameter in nematic liquid crystals. When correlations among active proteins are taken into account it may be possible that nematically ordered active protein states could be found (26, 36). We shall not consider such effects here.

The force dipole correlation function S(t) plays a central role in our study. This correlation function contains effects that arise from both thermal and nonthermal noise. This correlation function may be written as S(t)=2STδ(t)+SA(t), where the thermal component ST(t) is determined by the temperature, whereas the active component SA(t) results from nonthermal colored noise. In the absence of substrate, or when the system operates under equilibrium conditions, only the delta correlated thermal contribution remains. The integral intensity S=0dtS(t)ST+SA may also be defined. This quantity will enter in the expressions for the diffusion and drift derived below.

Because the conformational transitions that produce the force dipole depend on substrate binding, the dependence on substrate concentration enters the description through the force dipole correlation function S(t) and, thus, the integral intensity SA of active force dipoles should depend on the substrate concentration cS. Although the precise form of the dependence SA(cS) can only be determined by considering the kinetics of a particular enzymatic molecular system, general comments on its structure may nonetheless be made. The contribution SA of active force dipoles must vanish in the absence of substrate because only thermal fluctuations are then present and they are accounted for in ST. At high enough substrate concentrations the enzymes will operate at their maximum conversion rates and nonthermal effects will saturate. Consequently, SA(cS) should display a similar behavior: a linear proportionality at small substrate concentration cS and saturation at large concentrations. For example, a functional form like that of Michaelis–Menten kinetics, SA(cS)S0cS/(K+cS), where S0 is the saturation value of SA and K is a constant, satisfies these criteria, but other forms for SA(cS) with similar limiting regimes may well result from a calculation of SA for a specific molecular system.

If, within the time interval being considered, the displacements in the position R(t) of a passive particle are small, we can write Rα(t)=R0,α+ρα(t) and retain only terms that are of the first order in the displacements ρα(t), so that

ρ˙α=i[Gαβ(R0Ri)Ri,γ+2Gαβ(R0Ri)Ri,γR0,δρδ]×ei,βei,γmi(t)+fα(t). [4]

It is convenient to write the first term on the right side of this equation in field-point notation,

ρ˙α=dr[Gαβ(r)rγ2Gαβ(r)rγrδρδ]×iei,βei,γmi(t)δ(RiR0r)+fα(t), [5]

where the term involving the sum over the proteins represents the microscopic density of force dipoles, ei,βei,γmi(t), at a point r with origin at the position R0 of the passive particle. Note that the first and the second derivatives of the Green function correspond to dipole and quadrupole contributions. This equation shows that the instantaneous position of a passive particle evolves with time according to a stochastic differential equation with nonthermal multiplicative noise arising from the collective operation of active force dipoles, in addition to the additive thermal noise.

The diffusion tensor Dαα and the mean drift velocity V¯ of a passive particle can be determined from

Dαα=0dtδVα(t)δVα(0),V¯α=Vα, [6]

where δVα=VαV¯α and the angle bracket denotes an average over the stochastic fluctuations, both thermal and nonthermal, as well as the orientations and positions of active force dipoles.

The diffusion tensor and mean drift velocity of the passive particle may be obtained by substituting the expression in Eq. 5 for the velocity Vα=ρ˙α of a particle, retaining only leading terms, into Eq. 6. As discussed earlier, when computing the average values in Eq. 6, we assume that the orientations of active force dipoles are not correlated with their positions so that ieβeβeγeγδ(RiR0r)=eβeβeγeγc(r)Ωββγγc(r), where Ωββγγ=Cd[δββδγγ+δβγδβγ+δβγδβγ], with C2=1/8 and C3=1/15 for two and three dimensions, respectively, and c(r)=iδ(Rir) is the local concentration of active force dipoles at a point r in the fluid. We find

Dαα(R0)=DααT(R0)+SAΩββγγdrGαβ(r)rγGαβ(r)rγc(R0+r)DααT(R0)+DααA(R0) [7]
V¯α(R0)=SAΩββγγdr2Gαβ(r)rγrδGδβ(r)rγc(R0+r), [8]

where DααT(R0) is the equilibrium diffusion tensor of the passive particle averaged over protein configurations. It contains effects arising from the thermal contribution ST as well as the mobility of the individual passive particle. The last line in Eq. 7 defines the contribution of active force dipoles, DααA(R0), to the total diffusion tensor.

We shall now analyze Eqs. 7 and 8 separately for 3D and 2D systems.

3D Systems.

For applications to protein machines in bulk solution, for example in the cytoplasm of biological cells, the 3D Green function in the Oseen approximation is Gαβ(r)=(8πηr)1(δαβ+r^αr^β), where η is the fluid viscosity and r^ is the unit vector specifying the direction of r. Suppose that the machines are uniformly distributed in space with constant concentration c0. In this case, diffusion is isotropic, Dαα=Dδαα, and Eq. 7 yields

DA=SAc03ΩββγγdrGαβ(r)rγGαβ(r)rγ, [9]

for the active force dipole contribution to the diffusion coefficient. Note that whereas the integral (Eq. 9) diverges as 1/r for short distances r, volume exclusion between an enzyme and the passive particle restricts the domain of integration to distances exceeding some cutoff length c. The prime on the integral in Eq. 9 indicates this restriction. Introducing dimensionless coordinates z=r/c, the nonthermal diffusion coefficient DA can be estimated as

DAζ3SAc0cη2, [10]

where the dimensionless factor ζ3 is given by

ζ3=Ωββγγ192π2dzgαβ(z)zγgαβ(z)zγ=160π, [11]

and the dimensionless function gαβ(z)=z1(δαβ+z^αz^β).

Turning to Eq. 8, we notice that the drift velocity vanishes for a uniform distribution of active dipoles. Suppose instead that a constant concentration gradient in the direction n^, c=n^(n^c)=n^(c), is present and c(r)=c0+(rn^)c. Now, the integration in Eq. 8 yields

V¯α=SAΩββγγdr2Gαβ(r)rγrδGδβ(r)rγ(rn^)c. [12]

Again, introducing dimensionless coordinates z=r/c, we find

V¯ξ3SAcη2c, [13]

where the positive dimensionless factor ξ3 is

ξ3=Ωββγγ64π2dzn^α2gαβ(z)zγzδgδβ(z)zγ(zn^)=130π. [14]

Because ξ3 is positive the drift is in the direction of the enzyme concentration gradient. We further note for 3D that the dominant contributions to the integrals for the general expressions for the diffusion and drift given in Eqs. 7 and 8 come from short distance separations, so Eqs. 10 and 13 also hold for arbitrary concentration distributions; in this case local concentration values and gradients should be taken to determine the effects at a given point.

2D Systems.

As noted by Saffman and Delbrück (31), biological membranes should behave as 2D lipid fluids on submicrometer length scales. Therefore, the effects of an ensemble of active protein machines on the motion of a passive particle in a lipid bilayer provides an example where a 2D description is appropriate. In the Oseen approximation, the 2D Green function of lipid bilayers is (32)

Gαβ(r)=(4πηm)1((1+ln(κr))δαβ+r^αr^β). [15]

Here, ηm is the 2D viscosity of the lipid bilayer, which is related to its 3D viscosity ηL by ηm=ηLh, where h is the thickness of the bilayer. In contrast to the 3D case, hydrodynamic interactions in 2D are ultra-long-ranged, owing to the logarithmic dependence on the distance r. For biomembranes, one can use the estimate (32) κ1=ηLh/(2η), where η is the viscosity of the surrounding aqueous medium. Typically, κ1 is of the order of a micrometer. At separations larger than this distance, 3D effects become important for biomembranes. Therefore, our analysis applies only for relatively small membranes of micrometer size; for larger membranes the crossover to 3D hydrodynamics at long length scales needs to be taken into account.

The general expressions in Eqs. 7 and 8 hold in the 2D case as well; however, because of the logarithmic distance dependence in the Green function, hydrodynamic effects are nonlocal. Therefore, it is not possible to obtain precise estimates similar to those in Eqs. 10 and 13 for such systems. Still, some estimates can be made.

Consider a passive particle at the center of a circular membrane with radius 0 so that 2D behavior applies. If we again consider a uniform distribution of active enzymes, the diffusion coefficient given by Eq. 7 is isotropic at the center of the membrane. Introducing the rescaled dimensionless coordinates z=r/0, we obtain

DAζ2SAc0ηm2, [16]

where the dimensionless factor ζ2 is given by

ζ2=Ωββγγ32π2dzqαβ(z)zγqαβ(z)zγ=132πlnoc. [17]

Here qαβ(z)=ln(z)δαβ+z^αz^β. The integral (Eq. 17) diverges logarithmically at z=0 and a cutoff at z=c/0 has been introduced.

In a similar manner, the drift velocity in 2D systems can be estimated. Taking c(r)=c0+(rn^)c in Eq. 8, and changing variables as indicated above, we find

V¯ζ2SAηm2c, [18]

where

ξ2=Ωββγγ16π2dzn^α2qαβ(z)zγzδqδβ(z)zγ(zn^)=132πlnoc, [19]

so that ξ2=ζ2. Note that because the integrals in Eqs. 17 and 19 diverge logarithmically at large distances, the finite size of the membrane is important. Moreover, as follows from Eq. 19 and similar to the 3D case, passive particles drift toward higher concentrations of active proteins.

If the passive particle is not at the center of the domain or concentration distributions of active protein inclusions are more general than the constant and linear-gradient distributions considered above, the diffusion will no longer be isotropic and the diffusion and drift will depend on the concentration distribution c(r) over the entire membrane, in contrast to the 3D case where only the local structure of the enzyme concentration field in the neighborhood of the passive particle is of importance.

Numerical Estimates

The magnitude of a force dipole m of a protein machine can be roughly estimated as mFdP, where F is the force generated by the machine and dP is the linear size of the protein. Molecular motors, such as myosin or kinesin, typically generate forces about 1 pN and this can be chosen as the characteristic value for F. Taking the size of a protein to be about 10 nm, the force dipole can be estimated to be about m=1020 N⋅m. The correlation time for force-dipole fluctuations can be taken to be the duration tc of the cycle time in a chemical machine. Although enzyme cycle times vary widely, we choose a time of about tc1 ms. The parameter SA can then be evaluated to give SAm2tc=1043 N2m2⋅s. Note that this estimate corresponds to substrate (typically ATP) saturation conditions: The machine binds a new substrate molecule and enters into a new cycle immediately once the previous cycle finishes. If this condition is not satisfied, the protein machine must wait for a new substrate molecule to arrive. During this waiting period, the machine does not act as a force dipole and this will decrease the value of SA. Obviously, the effects disappear when the substrate is not supplied.

Concentrations of active proteins inside a biological cell can vary over a large range. The highest concentrations of the order of 104 M are characteristic for the enzymes involved in glycolysis. As a rough estimate, a value of 106 M can be chosen, corresponding to c0=1021m3, so that the mean distance between the proteins being considered is about 100 nm. Given the protein size, we choose a cutoff length of c=10 nm. The viscosity of water is about 103 Pa⋅s. For the dimensionless numerical factors in Eqs. 11 and 14 we take their order-of-magnitude values ζ3=102 and ξ3=102.

With these values, the contribution (Eq. 10) to the diffusion coefficient owing to hydrodynamic effects arising from protein machines in bulk 3D solutions is about DA106cm2/s. This result should be compared with typical diffusion constants in water that can vary from about 105cm2/s for small molecules to 107cm2/s for small proteins in water. If one takes the viscosity of the cytoplasm to be two to four times that of water, the active contribution DA will decrease by approximately an order of magnitude. However, it is still approximately the same as the thermal Brownian contribution DTkBTγ=kBT/(6πηRp), where Rp is the radius of the passive particle, under the same viscosity conditions using a Stokes law estimate. To estimate the magnitude of the drift velocity from Eq. 13, we can take c=Δc/L, where Δc is the concentration difference across the cell and L=10μm is the typical eukaryotic cell size. If we choose Δc0.1c0, where again c0=1021m3, we obtain a drift velocity magnitude of about V¯1μm/s.

Proceeding to lipid bilayers, we observe that their 3D viscosity ηL1 Pa⋅s is typically a factor of 103 higher than that of water. The 2D viscosity of such bilayers is ηm=ηLh, where h1 nm has been taken as the thickness of the bilayer. The magnitude S1043N2m2s may again be used for the force dipoles. Taking the mean distance between proteins to be ∼100 nm, the 2D concentration is about c0=1014m2. As rough estimates of these factors we again choose their order-of-magnitude values, ζ2=ξ2=102. A membrane of micrometer size is considered and a cut-off distance of 10 nm is introduced.

Given these numerical values, the hydrodynamic effects of active protein inclusions are predicted to increase the diffusion of passive particles within the membrane by about DA109cm2/s. For comparison, Brownian diffusion constants for proteins in lipid bilayers are of the order of 1010cm2/s and diffusion constants for lipids are about 108cm2/s. The characteristic magnitude of the drift velocity of passive particles in lipid bilayers is estimated to be about V¯102μm/s, assuming that Δc=0.1c0 and the characteristic length for concentration variation in a membrane is about 1 μm.

These numerical estimates should be used only as rough guide to the possible magnitudes of the effects because many of the parameters may vary significantly from one system to another, or are known only poorly. For example, forces have only been measured for some molecular motors and, for protein machines that are not motors or for enzymes, they may be smaller than 1 pN. However, the concentrations of some proteins that behave as force dipoles may be significantly higher than the value we have assumed. For example, the enzyme phosphoglycerate kinase involved in glycolysis is present in the living cell in the concentrations up to 104 M, two orders of magnitude larger than our assumed value. Proteins typically contribute about 40% of the mass in biomembranes and their concentrations may well be significantly higher than our 2D values. In addition, the cytoplasm is a very complicated medium that is crowded by a variety of macromolecules, filaments, organelles, and other structures (37). Diffusion in such a crowded environment differs from that in a simple solution. Crowding can also influence the hydrodynamic effects owing to active force dipoles discussed in this paper. Such effects will have to be considered in descriptions of transport in the cell that model the detailed structure of the cytoplasm. Consequently, the uncertainty in the estimates of hydrodynamic effects is high and deviations of up to two orders of magnitude from the numerical estimates given above may well be possible.

The effects considered here depend on the concentration of substrate through SA(cS) as discussed earlier. If the substrate concentration varies in a cell or a membrane, the coordinate-dependent dipole intensity SA should be retained within integrals in Eqs. 7 and 8. In three dimensions, diffusion is determined by the local dipole force intensity according to Eq. 10, whereas the drift velocity is determined by the local gradient of SA(r)c(r), replacing the gradient of concentration in Eq. 13. In two dimensions, local diffusion and drift are generally dependent on the concentration distribution of active inclusions and substrates over the entire membrane.

Discussion and Conclusions

In a biological cell there are large populations of active proteins, both molecular machines and enzymes, that change their conformations within catalytic cycles. In this paper we showed that when active proteins are present, either in solution or in lipid bilayers, they can substantially modify the diffusion constants of passive particles in the system. These modifications affect all passive particles, and all active proteins, even of different kinds, contribute to the effect provided they are supplied with substrate and remain active. The magnitude of the effect can be comparable to the value of Brownian diffusion constants under physiological conditions.

Furthermore, if protein machines are nonuniformly distributed in a cell or in a biomembrane, directed drift of passive particles, analogous to chemotaxis, can occur. However, the mechanism is completely different: All active proteins contribute toward it and all passive particles experience the drift. Drift velocities of the order of micrometers per second can be realized. The enhancement of diffusion and chemotaxis-like drift should take place for the protein machines (enzymes) themselves as well. Note that the drift velocity is in the same direction as the concentration gradient and therefore the hydrodynamic attraction of incoherent active proteins should occur. Generally, the same proteins would exhibit different interactions depending on whether they are catalytically active or inactive (no ATP is supplied). However, thus far we have not considered collective effects due to hydrodynamic interactions on the populations of active proteins. It may be that orientation alignment leading to nematic order (26, 36) and cycle synchronization also arise.

In three dimensions, hydrodynamic interactions are already long-ranged, with power-law dependence. In two dimensions they become ultra-long-ranged with a logarithmic dependence on distance. Thus, the effects predicted to exist in 3D and 2D systems differ substantially. In solution, the change in the diffusion constant is determined by the local protein machine concentration and the drift velocity is controlled by its local spatial gradient. In contrast, in 2D systems such as lipid bilayers, the effects are essentially nonlocal: The change in the diffusion constant and the drift of passive particles at a given location are determined by the distribution of active inclusions over the entire membrane. Note, however, that only relatively small membranes with micrometer sizes were considered here. In general, diffusion in biomembranes should be anisotropic, reflecting the asymmetry of protein distribution and the membrane shape.

Our description of how active molecular machines, through hydrodynamic interactions, influence the dynamics of passive particles was based on the equations of continuum hydrodynamics. One might question the use of such a continuum description for molecular systems. It is well established (38) that hydrodynamic effects are observable on even very small scales of tens of solvent particle diameters. They persist despite strong fluctuations and their presence is signaled in the long-time tails of velocity correlation functions or even in the transport properties of polymers. Our use of continuum equations is restricted to rather long scales so that the main conclusions of our study should be robust.

Sen and coworkers (39, 40) have shown that catalytically active enzymes have larger diffusion coefficients than their inactive counterparts in the absence of substrate. Recently, chemotaxis-like drift of enzymes in the presence of substrate gradients has been observed and used for sorting of the enzymes (41). Although additional analysis of the experimental data is needed, it may be that such observations can be explained by the effects considered above. Furthermore, in vivo studies have revealed that the diffusion of particles decreases in ATP-depleted biological cells (42). In addition, several studies have proposed specific explanations for the importance of nonthermal random motions in living cells, whose origin lies in the forces generated by the uncorrelated activity of protein machines. (26, 4244) Such nonthermal fluctuations may also be a consequence of the universal hydrodynamic effects, described here, that arise from active conformational changes in molecular motors and other protein machines powered by ATP.

Our analysis focused on general aspects of the phenomena and is not intended to address the full complexity seen in biological systems. Nevertheless, in accord with the above findings, our results suggest a modified physical picture of kinetic processes in the biological cell. When energy is supplied by ATP or other substrates to active proteins in a cell, such as molecular motors, other protein machines, or enzymes, they cyclically change their shapes in the course of carrying out their various specific functions. In addition to their functions, all such proteins act as oscillating active force dipoles and collectively create a fluctuating hydrodynamic field over the entire cell or a biomembrane. This nonequilibrium flow field can be maintained because a fraction of the energy flux arriving with substrates is diverted through the force-dipole activity to hydrodynamic flows in the cytoplasm. Such fluctuating fields arise from nonequilibrium effects; therefore, in contrast to thermal hydrodynamic fluctuations, the fluctuation-dissipation theorem does not apply to them. Because these fluctuating fields arise from nonthermal noise, it is possible that they can be rectified and work or energy can be extracted from them. Thus, active proteins in a cell not only execute their specific functions but, collectively, they supply power in a distributed way to the system. Such power, originating from substrate supply to active proteins, spans the entire cell.

Acknowledgments

R.K. thanks the research training group GRK 1558 funded by Deutsche Forschungsgemeinschaft for financial support through a Mercator Fellowship. This work was supported by a grant from the Natural Sciences and Engineering Research Council of Canada (to R.K.). Financial support (to A.S.M.) through a grant from the Volkswagen Foundation and from the Ministry for Education, Culture, Sports, Science and Technology of Japan through a program on “Dynamic Approaches to Living Systems” are gratefully acknowledged.

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1506825112/-/DCSupplemental.

References

  • 1.Alberts B, Bray D, Lewis J, Raff M, Roberts K, Watson JD. Molecular Biology of the Cell. 3rd Ed Garland Science; New York: 2002. [Google Scholar]
  • 2.Jones RAL. Soft Machines: Nanotechnology and Life. Oxford Univ Press; Oxford: 2004. [Google Scholar]
  • 3.Lauga E, Powers TR. The hydrodynamics of swimming microorganisms. Rep Prog Phys. 2009;72:1–36. [Google Scholar]
  • 4.Cisneros LH, Cortez R, Dombrowski C, Goldstein RE, Kessler JO. Fluid dynamics of self-propelled microorganisms, from individuals to concentrated populations. Exp Fluids. 2007;43:737–753. [Google Scholar]
  • 5.Wang J. Nanomachines: Fundamentals and Applications. Wiley-VCH; Weinheim, Germany: 2013. [Google Scholar]
  • 6.Wang W, Duan W, Ahmed S, Mallouk TE, Sen A. Small power: Autonomous nano- and micromotors propelled by self-generated gradients. Nano Today. 2013;8(5):531–554. [Google Scholar]
  • 7.Kapral R. Perspective: Nanomotors without moving parts that propel themselves in solution. J Chem Phys. 2013;138(2):020901. doi: 10.1063/1.4773981. [DOI] [PubMed] [Google Scholar]
  • 8.Seifert U. Stochastic thermodynamics, fluctuation theorems and molecular machines. Rep Prog Phys. 2012;75(12):126001. doi: 10.1088/0034-4885/75/12/126001. [DOI] [PubMed] [Google Scholar]
  • 9.Mirkovic T, Zacharia NS, Scholes GD, Ozin GA. Nanolocomotion - catalytic nanomotors and nanorotors. Small. 2010;6(2):159–167. doi: 10.1002/smll.200901340. [DOI] [PubMed] [Google Scholar]
  • 10.Golestanian R, Liverpool TB, Ajdari A. Designing phoretic micro- and nano-swimmers. New J Phys. 2007;9:126. [Google Scholar]
  • 11.Mikhailov AS, Calenbuhr V. From Cells to Societies: Models of Complex Coherent Action. Springer; Berlin: 2002. [Google Scholar]
  • 12.Jülicher F, Ajdari A, Prost J. Modeling molecular motors. Rev Mod Phys. 1997;69:1269–1281. [Google Scholar]
  • 13.Aranson IS, Tsimring LS. Patterns and collective behavior in granular media: Theoretical concepts. Rev Mod Phys. 2006;78:641. [Google Scholar]
  • 14.Purcell EM. Life at low Reynolds number. Am J Phys. 1977;45:3–11. [Google Scholar]
  • 15.Saintillan D, Shelley MJ. Emergence of coherent structures and large-scale flows in motile suspensions. J R Soc Interface. 2012;9(68):571–585. doi: 10.1098/rsif.2011.0355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Schwarz-Linek J, et al. Phase separation and rotor self-assembly in active particle suspensions. Proc Natl Acad Sci USA. 2012;109(11):4052–4057. doi: 10.1073/pnas.1116334109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Zöttl A, Stark H. Hydrodynamics determines collective motion and phase behavior of active colloids in quasi-two-dimensional confinement. Phys Rev Lett. 2014;112(11):118101. doi: 10.1103/PhysRevLett.112.118101. [DOI] [PubMed] [Google Scholar]
  • 18.Marchetti MC, et al. Hydrodynamics of soft active matter. Rev Mod Phys. 2013;85:1143–1189. [Google Scholar]
  • 19.Theurkauff I, Cottin-Bizonne C, Palacci J, Ybert C, Bocquet L. Dynamic clustering in active colloidal suspensions with chemical signaling. Phys Rev Lett. 2012;108(26):268303. doi: 10.1103/PhysRevLett.108.268303. [DOI] [PubMed] [Google Scholar]
  • 20.Buttinoni I, et al. Dynamical clustering and phase separation in suspensions of self-propelled colloidal particles. Phys Rev Lett. 2013;110(23):238301. doi: 10.1103/PhysRevLett.110.238301. [DOI] [PubMed] [Google Scholar]
  • 21.Saha S, Golestanian R, Ramaswamy S. Clusters, asters, and collective oscillations in chemotactic colloids. Phys Rev E Stat Nonlin Soft Matter Phys. 2014;89(6):062316. doi: 10.1103/PhysRevE.89.062316. [DOI] [PubMed] [Google Scholar]
  • 22.Alexander GP, Yeomans JM. Dumb-bell swimmers. Europhys Lett. 2008;83:34006. [Google Scholar]
  • 23.Lauga E, Bartolo D. No many-scallop theorem: Collective locomotion of reciprocal swimmers. Phys Rev E Stat Nonlin Soft Matter Phys. 2008;78(3 Pt 1):030901. doi: 10.1103/PhysRevE.78.030901. [DOI] [PubMed] [Google Scholar]
  • 24.Shani I, Beatus T, Bar-Ziv RH, Tlusty T. Long-range orientational order in two-dimensional microfluidic dipoles. Nat Phys. 2014;10:140–144. [Google Scholar]
  • 25.Drescher K, Dunkel J, Cisneros LH, Ganguly S, Goldstein RE. Fluid dynamics and noise in bacterial cell-cell and cell-surface scattering. Proc Natl Acad Sci USA. 2011;108(27):10940–10945. doi: 10.1073/pnas.1019079108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Bruinsma R, Grosberg AY, Rabin Y, Zidovska A. Chromatin hydrodynamics. Biophys J. 2014;106(9):1871–1881. doi: 10.1016/j.bpj.2014.03.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Sakaue T, Kapral R, Mikhailov A. Nanoscale swimmers: Hydrodynamic interactions and propulsion of molecular machines. Eur Phys J B. 2010;75:381–387. [Google Scholar]
  • 28.Cressman A, Togashi Y, Mikhailov AS, Kapral R. Mesoscale modeling of molecular machines: Cyclic dynamics and hydrodynamical fluctuations. Phys Rev E Stat Nonlin Soft Matter Phys. 2008;77(5 Pt 1):050901. doi: 10.1103/PhysRevE.77.050901. [DOI] [PubMed] [Google Scholar]
  • 29.Huang M-J, Kapral R, Mikhailov AS, Chen H-Y. Coarse-grain simulations of active molecular machines in lipid bilayers. J Chem Phys. 2013c;138(19):195101. doi: 10.1063/1.4803507. [DOI] [PubMed] [Google Scholar]
  • 30.Huang M-J, Chen H-Y, Mikhailov AS. Nano-swimmers in biological membranes and propulsion hydrodynamics in two dimensions. Eur Phys J E Soft Matter. 2012a;35(11):119. doi: 10.1140/epje/i2012-12119-5. [DOI] [PubMed] [Google Scholar]
  • 31.Saffman PG, Delbrück M. Brownian motion in biological membranes. Proc Natl Acad Sci USA. 1975;72(8):3111–3113. doi: 10.1073/pnas.72.8.3111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Diamant H. Hydrodynamic interaction in confined geometries. J Phys Soc Jpn. 2009;78(4):041002. [Google Scholar]
  • 33.Di Leonardo R, et al. Hydrodynamic interactions in two dimensions. Phys Rev E Stat Nonlin Soft Matter Phys. 2008;78(3 Pt 1):031406. doi: 10.1103/PhysRevE.78.031406. [DOI] [PubMed] [Google Scholar]
  • 34.Huang M-J, Kapral R, Mikhailov AS, Chen H-Y. Coarse-grain model for lipid bilayer self-assembly and dynamics: Multiparticle collision description of the solvent. J Chem Phys. 2012b;137(5):055101. doi: 10.1063/1.4736414. [DOI] [PubMed] [Google Scholar]
  • 35.Cheung C, Hwang YH, Wu Xl, Choi HJ. Diffusion of particles in free-standing liquid films. Phys Rev Lett. 1996;76(14):2531–2534. doi: 10.1103/PhysRevLett.76.2531. [DOI] [PubMed] [Google Scholar]
  • 36.Woodhouse FG, Goldstein RE. Spontaneous circulation of confined active suspensions. Phys Rev Lett. 2012;109(16):168105. doi: 10.1103/PhysRevLett.109.168105. [DOI] [PubMed] [Google Scholar]
  • 37.Goodsell DS. Inside a living cell. Trends Biochem Sci. 1991;16(6):203–206. doi: 10.1016/0968-0004(91)90083-8. [DOI] [PubMed] [Google Scholar]
  • 38.Alder BJ, Wainwright TE. Decay of the velocity autocorrelation function. Phys Rev A. 1970;1:18–21. [Google Scholar]
  • 39.Sengupta S, et al. Enzyme molecules as nanomotors. J Am Chem Soc. 2013;135(4):1406–1414. doi: 10.1021/ja3091615. [DOI] [PubMed] [Google Scholar]
  • 40.Muddana HS, Sengupta S, Mallouk TE, Sen A, Butler PJ. Substrate catalysis enhances single-enzyme diffusion. J Am Chem Soc. 2010;132(7):2110–2111. doi: 10.1021/ja908773a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Dey KK, et al. Chemotactic separation of enzymes. ACS Nano. 2014;8(12):11941–11949. doi: 10.1021/nn504418u. [DOI] [PubMed] [Google Scholar]
  • 42.Weber SC, Spakowitz AJ, Theriot JA. Nonthermal ATP-dependent fluctuations contribute to the in vivo motion of chromosomal loci. Proc Natl Acad Sci USA. 2012;109(19):7338–7343. doi: 10.1073/pnas.1119505109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.MacKintosh FC, Levine AJ. Nonequilibrium mechanics and dynamics of motor-activated gels. Phys Rev Lett. 2008;100(1):018104. doi: 10.1103/PhysRevLett.100.018104. [DOI] [PubMed] [Google Scholar]
  • 44.Guo M, et al. Probing the stochastic, motor-driven properties of the cytoplasm using force spectrum microscopy. Cell. 2014;158(4):822–832. doi: 10.1016/j.cell.2014.06.051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Kogler F. 2012. Interactions of artificial molecular machines. Diploma thesis (Technical Univ of Berlin, Berlin)

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES