Skip to main content
Research logoLink to Research
. 2020 May 21;2020:9739231. doi: 10.34133/2020/9739231

Active Matter, Microreversibility, and Thermodynamics

Pierre Gaspard 1,, Raymond Kapral 2
PMCID: PMC7260603  PMID: 32524094

Abstract

Active matter, comprising many active agents interacting and moving in fluids or more complex environments, is a commonly occurring state of matter in biological and physical systems. By its very nature, active matter systems exist in nonequilibrium states. In this paper, the active agents are small Janus colloidal particles that use chemical energy provided by chemical reactions occurring on their surfaces for propulsion through a diffusiophoretic mechanism. As a result of interactions among these colloids, either directly or through fluid velocity and concentration fields, they may act collectively to form structures such as dynamic clusters. A general nonequilibrium thermodynamics framework for the description of such systems is presented that accounts for both self-diffusiophoresis and diffusiophoresis due to external concentration gradients, and is consistent with microreversibility. It predicts the existence of a reciprocal effect of diffusiophoresis back onto the reaction rate for the entire collection of colloids in the system, as well as the existence of a clustering instability that leads to nonequilibrium inhomogeneous system states.

1. Introduction

Active matter is composed of motile entities or agents interacting with each other either directly or through the velocity and concentration fields of the medium in which they move. Such interactions lead to collective dynamics giving rise to states of matter that may differ from those in equilibrium systems. The study of such collective behavior presents challenges and is currently a topic of considerable scientific interest. Systems with many complex agents can be investigated in different ways. One way is to describe collective dynamics at the macroscale in terms of fields representing the distribution of the agents across the system. These fields are ruled by partial differential equations that are established using general symmetries and experimental observations. Another approach is to model active matter as being composed of active particles moving in space according to specific rules that are postulated on the basis of empirical considerations.

Both of these approaches have been used to explore the origins and types of collective dynamics that can be found in active matter systems, and research on this topic ranges from studies of simple active particle models, often satisfying minimal rules, to suspensions of more complex active synthetic or biological agents [111]. The collective behavior in systems where the active agents are chemically propelled colloids, the subject of this paper, has also been the topic of experimental and theoretical research [1225].

Systems containing colloidal particles are governed by physicochemical laws, so that their time evolution can be understood from first principles using statistical-mechanical methods. This approach was pioneered by Einstein [26] and Smoluchowski [2729] at the beginning of the twentieth century and systematically developed since then for passive colloidal particles [3033]. In active matter, the colloidal particles are propelled with energy supplied by the surrounding solution, so that the description should be extended to include the molecular concentrations of fuel and product powering their motion, in addition to the velocity field of the fluid. Through such an approach, active matter can be described from the scale of a single colloidal motor moving in the surrounding fluid, up to the macroscale where many colloidal motors generate collective motion by interaction. At the macroscale, collective dynamics is described in terms of the distribution function giving the orientation as well as the position of the colloidal motors. This statistical-mechanical approach has the advantage that the parameters characterizing active matter at the macroscale can be deduced from the microscopic level of description. The knowledge of these parameters in terms of the properties of materials composing the colloidal motors and the surrounding solution is fundamental for engineering active systems.

The present paper contributes to the statistical-mechanical and nonequilibrium thermodynamic approaches for active matter systems [3442], and considers systems whose active agents are Janus colloids with catalytic and noncatalytic faces moving by diffusiophoresis generated by chemical reactions taking place on their catalytic faces or caps [43, 44, 40]. Because of diffusiophoresis, the velocity and concentration fields are coupled together in the fluid around the Janus particle [45]. We start from the calculation of the diffusiophoretic force and torque on a single Janus particle moving in a fluid in the presence of molecular species corresponding to the fuel and the product of the reaction taking place on its catalytic surface. The concentrations of these molecular species may develop gradients on large scales under nonequilibrium conditions, and these gradients should be included in the calculation of the force and torque. The resulting diffusiophoretic force and torque enter the coupled Langevin equations ruling the displacement, rotation, and overall reaction of a single active particle.

Next, the evolution equation is established for the distribution function of the ensemble of active particles in a dilute colloidal solution. In order to be consistent with microreversibility, the principles of nonequilibrium thermodynamics are used to relate the thermodynamic forces or affinities to the current densities with linear response coefficients satisfying Onsager's reciprocal relations [4653]. This method allows us to obtain all the possible couplings compatible with microreversibility, including a priori unexpected reciprocal effects. Moreover, this method provides an expression for the entropy production rate density for active matter in agreement with the second law of thermodynamics and including the contribution of the reaction powering activity. Through this procedure, macroscopic evolution equations are obtained that govern the collective dynamics of colloidal motors coupled to the molecular concentrations of fuel and product. These equations can be shown to generate the reciprocal effect of diffusiophoresis back onto the reaction rate that has been obtained previously for a single particle [39, 40], but now at the macroscale. Furthermore, pattern formation due to a clustering instability manifests itself under nonequilibrium conditions induced by a bulk reaction replenishing the solution with fuel.

The paper is organized as follows. Section 2 is devoted to the dynamics of a single colloidal motor. The force and torque due to diffusiophoresis are deduced by solving the diffusion equations for the molecular concentrations coupled to the Navier-Stokes equations for the fluid velocity, including the contributions of concentration gradients at large distances from the particle. These contributions were neglected previously [39, 40] and are calculated in detail here. In Section 3, the diffusiophoretic force and torque obtained in Section 2 are incorporated into the evolution equation for the distribution function describing the ensemble of colloidal motors, and the entropy production rate density is explicitly obtained. Two implications of these results are presented in Sections 4 and 5. First, the reciprocal effect due to the diffusiophoretic coupling of an external force and torque back onto the reaction rate is recovered, now at the level of the collective dynamics. Second, a clustering instability leading to pattern formation is shown to manifest itself. The conclusions of the research are given in Section 6. The appendices provide additional details of the calculations.

2. Diffusiophoresis and Colloidal Motors

This section describes the motion of a single spherical Janus colloidal motor of radius R that is propelled by self-diffusiophoresis generated by a reversible reaction,

A+Cκκ+B+C, (1)

with rate constants κ± taking place on its catalytic surface, as depicted in Figure 1. In this reaction, A is the fuel and B the product, which are present in the solution surrounding the particle. Moreover, the concentrations of the A and B molecular species are assumed to have gradients gk with k = A, B at large distances from the particle that also contribute to motion by diffusiophoresis; thus, the motion of the particle is determined by processes in the fluid surrounding the particle.

Figure 1.

Figure 1

Schematic representation of a Janus particle with its catalytic (C) and noncatalytic (N) hemispheres where the surface reaction (1) takes place between fuel A and product B supplied by the solution surrounding the particle. The particle is also subjected to some external force Fext and torque Text. The position of its center of mass is R, and u is the unit vector giving its orientation and pointing in the direction of the catalytic hemisphere.

2.1. Chemohydrodynamics around a Colloidal Motor

In order to determine the force and the torque due to diffusiophoresis, as well as the overall reaction rate, the velocity of the fluid and the concentrations of fuel A and product B should be obtained by solving the Navier-Stokes equations for the fluid velocity v = vfluid coupled to the advection-diffusion equations for the molecular concentrations ck with k = A, B:

ρtv+v·v=p+η2v, (2)
·v=0, (3)
tck+v·ck=Dk2ck, (4)

where ρ is the constant mass density (the fluid being assumed to be incompressible), p the hydrostatic pressure, η the shear viscosity, and Dk the molecular diffusivity of species k.

The coupling between the velocity and concentration fields is established with the boundary conditions [40, 54]

n·vvsolidR=0, (5)
1·vvsolidR=1·bvSkbkckR, (6)
Dkn·ckR=νkκ+cAκcBR, (7)

where n is the unit vector normal to the solid surface, 1 ≡ 1 − nn, b is the slip length, (∇v)S = (∇v+∇vT), T denotes the transpose, bk is the diffusiophoretic coefficient of species k coupling the velocity field to the corresponding concentration field because of different interactions between the solid surface with the molecules of different species. The velocity field inside the solid particle is given by vsolid = V + Ω × (rR) in terms of the translational and angular velocities of the particle, respectively, denoted by V and Ω, and position R of the center of mass of the particle. Equations (5) and (6) are the boundary conditions on the components of the velocity field that are, respectively, normal and tangential to the interface Σ(t), which is located on the sphere ||rR|| = R. The last equation, i.e., equation (7) is the boundary condition for the two reacting species k = A, B, where νk is the stoichiometric coefficient of species k in the reaction (νA = −1 and νB = +1), and κ± are the forward and reverse surface rate constants per unit area.

The velocity field is assumed to vanish at large distances from the particle, so that the fluid is at rest except in the vicinity of the colloid. With the aim of obtaining mean-field equations for a dilute suspension of active particles, we also assume that the concentration fields can have nonvanishing gradients on large spatial scales. Accordingly, the concentration gradients (∇ck) = gk are taken to exist at large distances from the colloidal particle.

We suppose that the diffusiophoretic coefficients take the values bkc and bkn on the catalytic and noncatalytic hemispheres, respectively, while the surface rate constants per unit area take positive values κ±c on the catalytic hemisphere and vanish on the noncatalytic hemisphere, κ±n = 0. Using spherical coordinates (θ, φ) with polar angle θ defined with respect to the axis of the cylindrical symmetry of the Janus particle, we have

bkθ,φ=h=c,nbkhHhθ,κ±θ,φ=h=c,nκ±hHhθ, (8)

where Hh(θ) denotes the Heaviside function such that Hh(θ) = 1 on hemisphere h and is zero otherwise. The catalytic hemisphere is taken as 0 ≤ θ ≤ (π/2), and the noncatalytic hemisphere as (π/2) < θπ.

Solving equations (2)–(4) with the boundary conditions (5)–(7), the velocity and concentration fields can be obtained in the vicinity of the colloidal motor [45, 40, 41]. Accordingly, the force and the torque exerted by the fluid on the motor, as well as the overall reaction rate at its catalytic surface, are given by the following surface integrals at the fluid-colloid interface Σ(t),

F=t P·ndΣ,T=trR×P·ndΣ,W=tκ+cAκcBdΣ, (9)

where P = p1 − η(∇v)S is the pressure tensor of the fluid. The fluctuating contributions from thermal noise can also be included [40, 41].

2.2. Coupled Langevin Equations for the Motor

The orientation of the Janus particle is described by the unit vector u attached to the axis of the cylindrical symmetry of the Janus particle and pointing towards the catalytic hemisphere. Accordingly, the displacement and rotation of the particle are ruled by

dRdt=V,dudt=Ω×u, (10)

in terms of the translational and rotational velocities. These velocities, as well as the number N of reactive events taking place on the particle, are governed by the following coupled Langevin equations [39, 40, 41]:

MdVdt=γtV+Fd+Fext+Fflt, (11)
I·dΩdt=γrΩ+Td+Text+Tflt, (12)
dNdt=Wrxn+Wd+Wflt, (13)

where M and I denote the mass and inertia tensor of the motor, γt = 6πηR(1 + (2b/R))/(1 + (3b/R)) is the translational friction coefficient, γr = 8πηR3/(1 + (3b/R)) the rotational friction coefficient, Fd and Td the diffusiophoretic force and torque, Fext and Text the external force and torque exerted on the particle, while Ffl(t) and Tfl(t) are the contributions to the force and torque due to thermal fluctuations. The overall net reaction rate is Wrxn, Wd is the reciprocal contribution of diffusiophoresis back onto the reaction rate, and Wfl(t) is the fluctuating reaction rate. If the Janus particle has a magnetic dipole μ and is subjected to an external magnetic field B, then the external torque would be given by Text = μu × B. In the overdamped regime, the coupled Langevin equations are obtained by neglecting the inertial terms in equations (11) and (12).

Solving the Navier-Stokes equations (2) and (3) coupled to equation (4) for molecular concentrations with the boundary conditions (5)–(7), the force and the torque exerted on a spherical particle of radius R in a fluid with shear viscosity η and the overall net reaction rate are given by [40]

Fd=6πηR1+3b/Rkbk1·ck¯s, (14)
Td=12πηR1+3b/Rkbkr×ck¯s, (15)
Wrxn=4πR2κ+cAκcB¯s, (16)

expressed in terms of the surface average

·¯s=14π·r=Rdcosθdφ. (17)

The expressions (14) and (15) find their origin in the generalization of Faxén's theorem to a sphere moving in a time-dependent velocity field [55, 56]. When writing these equations, we have taken into account the possibility that the diffusiophoretic coefficients bk and the surface rate constants κ± may be nonuniform on the particle surface.

2.3. Motion in Molecular Concentration Gradients

If molecular diffusion is fast enough so that the concentration fields adopt stationary profiles around the catalytic particle in the concentration gradients gk, the diffusiophoretic translational and rotational velocities can be written as follows (see Appendix A):

Vd=Fdγt=Vsdu+kξk1+εkQu·gk, (18)
Ωd=Tdγr=kλku×gk, (19)

where the parameters ξk, εk, and λk are given in equations (A.35)–(A.40) in terms of the diffusiophoretic coefficients bkh, the rate constants per unit area κ±c, the slip length b, the molecular diffusivities Dk, and the geometry of the Janus particle. The 3 × 3 identity matrix is 1, while

Quuu131. (20)

The self-diffusiophoretic velocity, expressed in terms of the molecular concentrations c¯k extrapolated to the center of the particle, is

Vsd=kζkc¯k=ςκ+cc¯Aκcc¯B, (21)

since the parameters ζk may be written in the forms ζA = ςκ+c and ζB = −ςκc (see Appendix A, equations (A.33)-(A.34).

In the absence of a reaction, we recover the diffusiophoretic velocities given in Refs. [57, 58]:

Vd=kξk0gkwith  ξk0=bkc+bkn21+2b/R,Ωd=kλk0u×gkwith  λk0=916Rbkcbkn. (22)

Moreover, if the diffusiophoretic coefficients are the same on both hemispheres bkc = bkn, the angular velocity is equal to zero, Ωd = 0.

In the presence of a reaction, but without gradients (gk = 0), we have κ+cc¯Aκcc¯B and the linear velocity reduces to the contribution of self-diffusiophoresis, Vd = Vsdu, characterizing the activity of the Janus particle.

The overall reaction rate can be written as follows:

Wrxn=k+c¯Akc¯B+ϖk+gAkgB·u, (23)

in terms of rate constants k± = Γκ±c and a parameter ϖ = O(R) given in equation (A.45). In the absence of the concentration gradients, we recover the expression obtained in Ref. [40]. In the presence of the concentration gradients gk, there is an extra contribution depending on the direction u of the Janus particle. However, this last term is normally negligible because we typically have R||gk||c¯k for micrometric particles and macroscopic gradients of molecular concentrations.

We note that both the self-diffusiophoretic velocity (21) and the leading term of the reaction rate (23) are proportional to each other. Their ratio defines the self-diffusiophoretic parameter χ which was introduced in Refs. [39, 40],

χVsdk+c¯Akc¯B=ςΓ, (24)

where the last equality was obtained using k± = Γκ±c.

3. Active Suspension of Colloidal Motors

3.1. Onsager's Reciprocal Relations

We now show that Onsager's principle of nonequilibrium thermodynamics [4653] can be used to establish coupled diffusion-reaction equations of motion for active matter that are consistent with microreversibility. According to Onsager's principle, currents are related to thermodynamic forces (or affinities) by

Jα=βLαβAβ, (25)

where the linear response coefficients satisfy the Onsager reciprocal relations,

Lαβ=Lβα, (26)

if the affinities are even under time reversal. The thermodynamic entropy production rate density is given by

σs=kBαJαAα=kBαβLαβAαAβ0, (27)

where kB is Boltzmann's constant.

3.2. Mean-Field Equations for the Active Suspension

The system we consider is a dilute solution containing the reactive molecular A and B species together with colloidal motors C in an inert solvent S. The motors are spherical Janus particles and, as described in Section 2, have hemispherical catalytic surfaces where the reaction AB takes place. Moreover, we suppose that the solution is globally at rest, so that the velocity field is equal to zero on scales larger than the size of colloids. The solution is described at the macroscale in terms of the molecular densities nA(r, t) and nB(r, t), as well as the distribution function of the colloidal motors, f(r, u, t), where r = (x, y, z) is the position and u = (sinθcosφ, sinθsinφ, cosθ) is the unit vector giving the orientation of the Janus particles (expressed in spherical coordinates in the laboratory frame). The distribution function is defined as

fr,u,ti=1NCδ3rritδ2uuit, (28)

where {ri, ui}i=1NC are the positions and orientational unit vectors of the colloidal motors. For a dilute suspension, the evolution equation of this distribution function can be deduced from the Fokker-Planck equation for the probability that a single colloidal motor is located at the position r with the orientation u [39, 40, 41]. Once, this distribution function is known, we can obtain the successive moments of u:

nCr,tfr,u,td2u, (29)
pr,tufr,u,td2u, (30)
qr,tQufr,u,td2u, (31)

where d2u = dcosθdφ, nC is the density or concentration of colloidal motors, p is the polar order parameter or polarization of the colloidal motors, and q is the traceless order parameter analogous to that for apolar nematic liquid crystals expressed in terms of the tensor (20) and, thus, satisfies tr(q) = 0.

At the macroscale, the reaction is

A+Ckk+B+C, (32)

with the rate constants k±. For the colloidal suspension treated here, the reaction should be described by a reaction rate density w that is proportional to the distribution function of colloidal motors and determined by the surface reaction taken into account with the boundary conditions (7) in Section 2.

The mean concentrations of molecular species are defined by nk=1ϕc¯k, where ϕ = 4πR3nC/3 is the volume fraction of the suspension. Their corresponding gradients are related to those considered in Section 2 by ∇nk = (1 − ϕ)gk for a dilute enough suspension. The coupled diffusion-reaction equations for the different species take the following forms:

tnk+·jk=νkwk=A,B, (33)
tf+·VfDtf=DrLrf, (34)

where jk are the molecular current densities, V is the total drift velocity obtained by adding the drift velocity due to the external force Vext = Fext/γt to the diffusiophoretic velocity (18) giving

V=Vsdu+kξk1+εkQu·nk+βDtFext, (35)

with the self-diffusiophoretic velocity (equation (21))

Vsd=kζknk=ςκ+cnAκcnB, (36)

now expressed in terms of the mean concentrations nk, and the inverse temperature β = (kBT)−1. In equation (34), Dt is an effective translational diffusion coefficient related to the effective translational friction coefficient by Einstein's formula Dt = kBT/γt and Dr is an effective rotational diffusion coefficient related to the effective rotational friction coefficient by Dr = kBT/γr. Since the shear viscosity increases as ηη(0)(1 + 2.5ϕ) with the volume fraction ϕ of the suspension [26, 31], both friction coefficients γt and γr also increase, and the diffusion coefficients decrease. In particular, it is known that DtDt(0)(1 − 2.1ϕ) [31]. A similar dependence on the volume fraction ϕ is expected for the parameters ς, ξk, εk, and λk given in Appendix A, since these parameters are proportional to the diffusiophoretic coefficients bkh that are known to be inversely proportional to shear viscosity, bkhη−1 [54, 57, 58]. The effects of this dependence would manifest themselves if the colloidal suspension became dense enough. Here, such effects are assumed to play a negligible role.

The Janus particles have a spherical shape so that their random rotational and translational motions are decoupled. In this case, the rotational diffusion operator is given by

Lrf=1sinθθsinθeβUrθeβUrf+1sin2θφeβUrφeβUrf, (37)

expressed in terms of the rotational energy associated with the torque exerted by an external magnetic field B on some magnetic dipole μ of the particle [52] and that due to the diffusiophoretic effect, we have the following:

Ur=μB·uγrkλknk·u. (38)

3.3. Translational and Rotational Current Densities

Before proceeding with nonequilibrium thermodynamics, we need to identify in equation (34) the current densities associated with the translational and rotational movements of the colloidal motors. The distribution function f(r, u) for colloidal Janus particles is defined in the five-dimensional space (x, y, z, θ, φ), where (x, y, z) are the Cartesian coordinates for the position r and (θ, φ) the spherical coordinates for the orientation u. Vector calculus is used in these coordinates to obtain the corresponding gradients and divergences [59].

For the rotational degrees of freedom we have

du2=dθ2+sin2θdφ2=gijdqidqjwith gij=100sin2θ. (39)

The scalar product between a pair of rotational vectors ar, br ∈ ℝ2 is given by ar · br = ∑i,j=θ,φgijaribrj, and the scalar product of such a vector with itself is denoted ar2 = ar · ar. In spherical coordinates, the rotational gradient and divergence are given, respectively, by [59]

gradrX=θX1sin2θφX, (40)
divrXr=1sinθθXrθsinθ+φXrφ. (41)

In the five-dimensional space, the gradient is given by

gradX=XgradrXwith X=xXyXzX, (42)

and the divergence of a five-dimensional vector, X = (Xt, Xr)T, is

divX=·Xt+divrXr. (43)

Using these notations, equation (34) can be written in the form of a local conservation law involving the five-dimensional current density, JC = (jt, jr)T, as

tf+divJC=0,or   tf+·jt+divrjr=0, (44)

with translational current density

jt=VfDtf=fVsdu+fkξk1+εkQu·nkDtf+fβUt, (45)

where Ut(r) = −Fext · r is the translational potential energy due to the external force Fext, rotational current density

jr=DreβUrgradreβUrf=fkλknk·gradruDrgradrffβμB·gradru, (46)

and their translational and rotational divergences, ∇·jt and divrjr = −DrLrf, where Lr is the operator (37).

3.4. Nonequilibrium Thermodynamics of the Active Suspension

Local thermodynamic equilibrium is assumed on scales larger than the size of the colloidal motors where the description by the mean-field equations (33) and (34) is valid and the fluid is at rest. According to this assumption, thermodynamic quantities can be locally expressed in the active suspension in terms of the molecular densities nk(r, t) and the distribution function f(r, u, t). Furthermore, we suppose that the system is isothermal and isobaric and the solution is dilute in the species A, B, and C. The appropriate thermodynamic potential is thus Gibbs' free energy given by the following volume integral of the corresponding density:

G=d3rnSψS+k=A,Bnkψk+nkkBTlnnkenS+d2ufψC+fkBTlnf4πenS+fUtrfμB·u, (47)

where the first term is the contribution from the solvent S of density nS, the next terms from fuel k = A and product k = B dilutely dispersed in the solvent [52], and the last terms from all the orientations u of the colloidal motors moving in the mechanical potential energies due to the external force Fext and the external torque exerted by the magnetic field B on the magnetic dipoles of the colloids. We can thus deduce the following chemical potentials:

μS=δGδnS=ψSkBTnSnA+nB+nC, (48)
μk=δGδnk=ψk+kBTlnnknSk=A,B, (49)
μC=δGδf=ψC+kBTlnf4πnS+UtrμB·u. (50)

Here, ψk = μk0 + kBTln(nS/n0), where μk0 is the standard chemical potential of species k and n0 = 1 mole/liter is the standard concentration. Since the solution is dilute, we have taken the solvent density nS to be essentially uniform in space and constant in time.

Next, we use the principles of nonequilibrium thermodynamics in order to express the current densities in terms of the affinities or thermodynamic forces given in Table 1. For the reaction (32), the affinity is given by

Arxn=1kBTμAμB=lnk+nAknB, (51)

and the corresponding current density is the rate density w introduced in equation (33). At chemical equilibrium, we have Arxn = 0, w = 0, and k+nA,eq = knB,eq. In the linear regime close to equilibrium where δnk = nknk,eq, the chemical affinity (51) can be approximated as

Arxn=δnAnA,eqδnBnB,eq=1Drxnk+δnAkδnB, (52)

where

Drxn12k+nA+knBeq, (53)

is the diffusivity of the reaction taking place on the colloidal motors [39, 40]. For the diffusion processes of species k, the affinity associated with the current density jk is given by Ak = −grad(μk/kBT) in terms of the chemical potential μk. For molecular species, the gradient is tridimensional in Euclidian space, so that Ak = −∇(μk/kBT) = −nk−1nk. For the colloid with chemical potential (50), the affinity is given by the five-dimensional gradient (42) as

AC=gradμCkBT=1ff+fβUtgradrffβμB·gradru, (54)

if the magnetic field B is uniform. In this five-dimensional space, the associated current density JC = (jt, jr)T given by equations (45) and (46) can thus be written in the following form:

JC=fVsdu0+fk=A,Bξk1+εkQuλkgradru·nkfDt100Dr1r·gradμCkBT, (55)

where 1r is the 2 × 2 identity matrix. In this form, we see that the first term is related to the reaction affinity since the self-diffusiophoretic velocity can be written as Vsd = χDrxnArxn. The next two terms can be related to the affinities of molecular species, and the last term to the affinity of the colloidal species.

Table 1.

Current densities and corresponding affinities or thermodynamic forces in the active suspension: w is the reaction rate density introduced in equation (33) corresponding to the affinity (51), jA and jB are the molecular current densities of fuel A and product B, jt is the translational current density of colloids given by equation (45), jr is the rotational current density of colloids given by equation (46). The translational and rotational current densities of colloids form the five-dimensional current density (55) according to JC = (jt, jr)T. Similarly, the translational and rotational affinities of colloids form the five-dimensional affinity (54). β = (kBT)−1 denotes the inverse temperature, μk the chemical potentials, ∇ = (x, y, z) the gradient in Euclidean space, and gradr the rotational gradient (40). We note that colloidal motors with the given orientation u are considered as so many independent species in the free energy (47), which is expressed by equation (56).

Process Current Affinity Dimension
Reaction w β(μAμB) 1
Molecular diffusion of fuel j A −∇(βμA) 3
Molecular diffusion of product j B −∇(βμB) 3
Translational diffusion of colloids j t −∇(βμC) 3
Rotational diffusion of colloids j r −gradr(βμC) 2

According to the Curie principle, there is no coupling between processes with different tensorial characters. However, the Janus particles have a director given by the unit vector u and we have adopted a description in terms of the distribution function f(r, u, t) for the Janus particles. Consequently, it is possible that a vectorial process such as diffusion may be coupled to a scalar process such as reaction if it is polarized by the unit vector u. If we introduce the densities 𝒩C = fΔ2u for Janus particles having their orientation u in cells with a size of Δ2u, along with the associated current densities,

JC=JCΔ2u, (56)

we may write a general coupling (25) of the following form:

wjAjBJC=LrrLrALrBLrCLArLAALABLACLBrLBALBBLBCLCrLCALCBLCC·ArxnμAkBTμBkBTgradμCkBT, (57)

up to possible nonlinear contributions that may be required in order for the reaction rate to obey the mass-action law. In equation (57), we have that Lrr is 1 × 1, Lrk1 × 3, LrC1 × 5, Lkr3 × 1, Lkl3 × 3, LkC3 × 5, LCr5 × 1, LCk5 × 3, and LCC5 × 5 (for k, l = A, B).

According to Onsager's reciprocal relations (26), the linear response coefficients should obey

Lrk=LkrT,LrC=LCrT,Lkl=LlkT,LCC=LCCT,LkC=LCkT, (58)

for k = A, B and where T again denotes the transpose.

We assume that the molecular species A and B undergo Fickian diffusion without cross-diffusion, so that

Lkl=Dknkδkl1, (59)

and that the reaction rate does not depend on the gradients ∇nA or ∇nB, whereupon

LrA=LrB=0. (60)

This last assumption consists in neglecting the terms with the coefficient ϖ in equation (23), which is usually justified as mentioned in Section 2.

The scalar coefficient associated with the reaction can be identified as

Lrr=DrxnnC, (61)

and the linear response coefficients LCr, LCk, and LCC in equation (57) can be determined using the current density (55), as described in Appendix B. As a consequence of Onsager's reciprocal relations, we can conclude that the reaction rate and the current densities should be given by

w=DrxnnCArxnχDrxnu·f+fβUtd2u, (62)
jk=Dknk+nkξk1+εkQu·f+fβUt+λkgradru·gradrffβμB·gradrud2u. (63)

In equation (62), the second term describes the reciprocal effects of diffusiophoresis back onto reaction. The second term in equation (63) is due to cross-diffusion between the molecular and colloidal species due to diffusiophoresis. We see that the linear response coefficients depend on the unit vector u in a manner similar to that already shown in Refs. [39, 40].

With respect to standard expressions, the terms involving the integral ∫d2u in equation (63) are required in order to satisfy Onsager's reciprocal relations and for these quantities to be compatible with microreversibility. However, these extra terms can be shown to be negligible, although the reciprocal terms are not negligible in equations (45) and (46). In order to show that the extra terms are negligible, we suppose that the self-diffusiophoretic and diffusiophoretic velocities take the typical value Vsd ~ Vd ~ 10 μm/s [60]. According to Ref. [58], the molecular gradients used in experiments of diffusiophoresis are of the order of ‖∇nk‖ ~ 105 mol/m4, so that diffusiophoretic parameters have the value ξk, εk ~ 10−10 m5 s−1 mol−1. Moreover, we have λk ~ ξk/R, but since ‖gradrf‖ ~ R‖∇f‖, the effect of the coefficients λk is again of the same order of magnitude as ξk and εk. Molecular diffusivities typically have the value Dk ~ 10−9 m2/s, while the translational diffusion coefficient of a micrometric colloidal particle is of the order of Dt ~ 10−13 m2/s. The molecular concentrations used in experiments on self-diffusiophoresis are about nk ~ 103 mol/m3, while the density of micrometric colloidal particles is approximately nC ~ 1018m−3 ~ 10−6mol/m3, or lower. If we assume that the molecular and colloidal gradients take comparable values ‖∇nk‖/nk ~ ‖∇f‖/f, the ratio between the extra term and the standard molecular diffusion term in equation (63) is given by

nkξkfDk||nk||~ξkfDk~108, (64)

which shows that the second term in equation (63) is negligible. Accordingly, the standard Fickian expressions jk≃−Dknk are very well justified for the molecular current densities. In the presence of colloidal motors, the expressions compatible with microreversibility are nevertheless given by equations (62) and (63). In contrast, the terms associated with the diffusiophoretic parameters in the colloidal current density (55) have effects that are not negligible.

The conclusion from these considerations is that active matter can be described as generalized diffusion-reaction processes in complete compatibility with microreversibility and Onsager's reciprocal relations. In this way, the program of nonequilibrium thermodynamics is complete and application of equation (27) gives the following expression for the thermodynamic entropy production rate density:

kB1σs=DrxnnCArxn2+k=A,BDknk2nk2χDrxnArxnu·f+fβUtd2u2k=A,Bnk·ξk1+εkQu·f+fβUtd2u2k=A,Bλknk·gradru·gradrffβμB·gradrud2u+Dt1ff+fβUt2d2u+Dr1fgradrffβμB·gradru2d2u0. (65)

The second law is satisfied if Dtχ2Drxn > 0, DkDtnCnkξk2 > 0, DkDtnCnkεk2 > 0, and DkDrnCnkλk2 > 0, which is expected.

The results derived in this section provide the basis for the analysis of collective effects in suspensions of active Janus particles. In Sections 4 and 5, we describe two collective phenomena that emerge from this theoretical framework: the effect of an external force and torque on the reaction rate, and a clustering instability.

4. Effect of External Force and Torque

Using a thermodynamic formulation that is consistent with microreversibility, we showed earlier [3941, 61] how the application of an external force and torque on a single colloidal motor can change the reaction rate on its surface and even lead to a net production of fuel rather than product. Now we show how these considerations can be extended to a suspension of such motors.

4.1. Local Evolution Equations

We suppose that the colloidal motors are subjected to an external force Fext = Fext1z and an external torque induced by an external magnetic field B = B1z exerted on the magnetic moment μ of the colloidal particles, both oriented in the z-direction. If βμB is large enough, the distribution function is given by

fr,u,t=nr,tβμB4πsinhβμBexpβμBcosθ, (66)

so that p(r, t) = 1zuznC(r, t) with 〈uz〉 = cothβμB − (βμB)−1. Moreover, the terms with the coefficients ξk, εk, and λk are assumed to be negligible in equation (34). If the concentrations are uniform in the x- and y-directions, the process is ruled by

tnC+zχuzk+nAknBnCDtznCβFextnC=0, (67)

obtained by integrating equation (34) over the orientation u. This equation for nC is coupled to equation (33) with the Fickian molecular current densities jk≃−Dknk and the local reaction rate

w=k+nAknBnCχuzDrxnznCβFextnC, (68)

given by equation (62), as predicted by Onsager's reciprocal relations.

4.2. Global Evolution Equations

Defining the mean value of the z-coordinate for the colloidal motors as

zznCd3rnCd3r, (69)

and using equation (67), we obtain the evolution equation,

dzdt=χuzwrxn+βDtFext, (70)

with mean reaction rate,

wrxnk+nAnCd3rnCd3rknBnCd3rnCd3r. (71)

Furthermore, integrating equation (33) with k = B over the position r with the local rate (68), we get the total reaction rate

dNrxndt=dNBdt=dNAdt=NCwrxn+χuzDrxnβFext, (72)

where NC = ∫nCd3r is the total number of colloidal motors in the suspension. Equations (70) and (72) have precisely the same structure as for a single colloidal motor. However, one should note that the mean reaction rate in equation (71) contains spatial correlations between the solute and colloid concentration fields. Given the structure of the equations, the results obtained in Refs. [39, 40] also apply here. In particular, there exists a regime where the entire ensemble of colloidal motors is propelled and carries out work against the external force by consuming fuel. In addition, there is also a regime where fuel is synthesized if the external force that opposes motion is sufficiently large to reverse the reaction AB. The efficiencies of these processes are given by the same expressions as in Refs. [39, 40].

5. Clustering Instability and Pattern Formation

The equations of motion developed in Section 3 that describe a dilute suspension of colloidal motors moving in a dilute solution of fuel A and product B molecular species will be shown in this section to lead to a clustering instability. This instability can be described by the mean-field equations obtained above for the concentrations of the molecular species and the distribution function of the colloidal motors in the absence of an external force and torque (Fext = 0 and B = 0). A number of other mean-field descriptions that predict instabilities and the formation of various clustering states of collections of diffusiophoretic colloidal particles have appeared in literature [62, 18, 6, 19, 20], and use techniques involving coupled moment equations similar to those adopted in this section.

5.1. Molecular Diffusion and Reaction

The equation for the colloidal motors is coupled to the reaction-diffusion equations for the molecular species A and B, accounting for the fact that the reaction AB occurs both at the surface of the catalytic hemisphere of the colloids and in the bulk:

tnk=Dk2nk+νkwtot, (73)

where the total reaction rate density is given by

wtot=k+nAknBnC+k+2nAk2nB. (74)

The system is driven out of equilibrium if k+/kk+2/k−2 [61].

5.2. Colloidal Density and Polarization

If the second moment (31) as well as higher moments are assumed to be negligible, the evolution equations for the density of colloidal motors (29) and the polarization (30) are given by

tnC+·nCkξknk+Vsdp=Dt2nC, (75)
tp+·pkξknk+13VsdnC+15Δ:kεknkp=Dt2p2Drp+23nCkλknk, (76)

in terms of the fourth-order tensor Δ with the following components: Δijmn = δijδmn + δimδjn − (2/3)δinδjm.

If Dr is large enough so that 2Drp dominates the other terms involving p in equation (76), we can neglect these other terms and this equation can be inverted to obtain

p16DrVsdnC+nCk2λkζknk, (77)

under which circumstances the field p is driven by the gradients of the colloid and species densities. Substituting this result into equation (75) for the density nC of colloidal particles, we find

tnC+·nCkξk+Vsd6Dr2λkζknkDteffnC=0, (78)

with the effective diffusion coefficient

DteffDt+Vsd26Dr, (79)

expressing the enhancement of diffusivity due to the self-diffusiophoretic activity [17].

5.3. Coupled Colloidal and Molecular Diffusion-Reaction Equations

In the following, we suppose that the diffusion coefficient is the same for both molecular species: DDA = DB. Consequently, n0 = nA + nB remains uniform during the time evolution if initially so. Therefore, nB = n0nA is known and only nA needs to be determined. Introducing the notations

anA,cnC, (80)

we have the following coupled equations describing the system:

ta=D2aWtot, (81)
Wtot=cKaK0+K2aK20, (82)
tc=·Dt+τrVsd2cξ+σVsdca, (83)
Vsd=ζaV0, (84)

with

Kk++k,K0kn0,K2k+2+k2,K20k2n0,τr6Dr1,ξξAξB,λλAλB,ζζAζB=ςκ+c+κc,V0ζBn0=ςκcn0,στr2λζ. (85)

Moreover, consistency with the existence of equilibrium requires that ζ/K = V0/K0 = ς/Γ = χ is equal to the diffusiophoretic parameter (24) that is the ratio between the self-diffusiophoretic velocity (21) and the leading term of the overall reaction rate (23).

For this system, there exists a uniform nonequilibrium steady state, where c keeps its initial uniform value c0 and the molecular density is also uniform at the value

a0=c0K0+K20c0K+K2, (86)

in order to satisfy the stationary condition Wtot = 0. For this molecular concentration a = a0, we notice that the rate KaK0 of the catalytic reaction on the colloids is not vanishing under the nonequilibrium condition k+/kk+2/k−2.

5.4. Pattern Formation

To analyze the stability of this homogeneous steady state, for simplicity we consider a one-dimensional system where the fields a and c only depend on the variable z. Accordingly, the gradients ∇ can be replaced by partial derivatives z in equations (81) and (83). The set of equations (81)–(84) is then numerically integrated by spatial discretization over the grid z = iΔz with i = 1, 2, ⋯, M with Δz = 0.1 and M = 500. The integration is performed with a Runge-Kutta algorithm of varying order 4-5 over a long enough time interval to reach a steady state. Figure 2 shows numerical results for the parameter values:

c0=1,n0=10,D=1,Dt=1,τr=1,ξ=3,σ=2,V0=0.5,ζ=0.1,K=0.2,K0=1,K2=0.3, (87)

and increasing values of K20. We observe the formation of clusters of colloidal motors in regions where the fuel A is depleted, as expected.

Figure 2.

Figure 2

Nonequilibrium steady state of the one-dimensional system for the parameter values (87): (a) with K20 = 2; (b) with K20 = 2.5.

5.5. Linear Analysis of the Clustering Instability

The threshold of this clustering instability can be found from a linear stability analysis. Linearizing the equations around the uniform steady state, we find that the perturbations obey

tδaδc=Dz2K~wρz2Dteffz2δaδc, (88)

with

K~c0K+K2,wKa0K0,ρc0ξ+σVsd,DteffDt+τrVsd2,Vsdζa0V0. (89)

Supposing that the perturbations behave as δa, δc ~ exp(iqz + st), we obtain the dispersion relations

s±q=12K~+D+Dteffq2±12K~+DDteffq224ρwq2. (90)

These dispersion relations are depicted in Figures 3(a)3(c), respectively, below, at, and beyond the threshold. The leading dispersion relation is associated with the conserved unstable mode of the colloidal motors because s+(0) = 0. The subleading dispersion relation is associated with the reactive mode of the molecular species because s0=K~. We notice that, since K~>0, there is no possibility for a Hopf bifurcation to uniform oscillatory behavior, which would be the case if the eigenvalues s±(0) were complex. We also note that there is no wavelength selection at the level of linear stability analysis in this clustering instability.

Figure 3.

Figure 3

Dispersion relations of linear stability analysis for K20 = 1, K20 = 1.89817, and K20 = 3, respectively, below, at, and beyond the threshold of clustering instability. The dispersion relations are obtained in (a)–(c) with the approximation (90), and in (d)–(f) by truncating equation (C.5) into a 5 × 5 matrix. The other parameter values are given in equation (87) and ε = 1.

Therefore, instability manifests itself if

ρw+K~Dteff<0, (91)

and the threshold is given by the condition

ρw+K~Dteff=0, (92)

which leads to the value K20≃1.89817 for the parameter values (87).

The dispersion relations can also be obtained from the evolution equation (34) for the distribution function. Supposing that f = f(z, θ) and a = a(z), we have

tf+zVdzfDtzf=Drsinθθsinθθf+2λcosθfza, (93)

where

Vdz=Vsdcosθ+ξ+εcos2θ13za, (94)

with εεAεB. Equation (93) is coupled to the diffusion-reaction equation (81) for the concentration field a with c = ∫fd2u.

The linear stability analysis can be carried out for equation (93) coupled to equation (81) with the rate (82) in a similar manner to that for equations (81)–(84). This analysis is presented in Appendix C. The dispersion relations can be computed by truncating the infinite matrix (C.5) in order to obtain the eigenvalues as a function of the wave number q. The result converges to the dispersion relations shown in Figures 3(d)3(f) below, at, and beyond the threshold, for the parameter values (87) and ε = 1. The convergence occurs faster for the leading dispersion relation than for the subleading ones. For the chosen parameter values, we can see that the approximation where we suppose that the vector field p is driven by the gradients (which corresponds to truncating to a 2 × 2 matrix) constitutes a good approximation to describe the instability. Indeed, the leading dispersion relation of Figure 3(b) is already very close to that in Figure 3(e).

The conclusion is that equations (81)–(84) provide a robust description of the clustering instability and of the emerging patterns.

6. Conclusion

Autonomous motion is not possible at equilibrium and active matter relies on the presence of nonequilibrium constraints to drive the system out of equilibrium. As a result the theoretical formulations provided by nonequilibrium thermodynamics and statistical mechanics are a natural starting point for the description of such systems.

Many of the active matter systems currently under study involve active agents such as molecular machines or self-propelled colloidal particles with linear dimensions ranging from tens of nanometers to micrometers. The transition from microscopic to macroscopic description for fluids containing active agents of such sizes takes place in the upper range of this scale. Suspensions of active colloidal particles are interesting in this connection since, as described earlier in this paper, the colloidal particles are large compared to the molecules of the medium in which they reside. The dynamics of the suspension can then be described by considering the equations for the positions, velocities, and orientations of the colloidal particles in the medium, or through field equations that describe the densities of these particles.

Nonequilibrium thermodynamics provides a set of principles that these systems must obey. Most important among these is microreversibility that stems from the basic time reversal character of the microscopic dynamics. On the macroscale, this principle manifests itself in Onsager's reciprocal relations that govern what dynamical processes are coupled and how they are described. For example, for single Janus particles propelled by a self-diffusiophoretic mechanism, microreversibility implies the existence of reciprocal effect where the reaction rate depends on an applied external force [3941, 61].

This paper extended the nonequilibrium thermodynamics formulation to the collective dynamics of ensembles of diffusiophoretic Janus colloids. In particular, we considered Janus colloids driven by both self-diffusiophoresis arising from reactions on the motor catalytic surface as well as motion arising from an external concentration gradient. This latter contribution is essential for the extension of the theory to collective motor dynamics. The resulting formulation is consistent with microreversibility and an expression for the entropy production is provided. From this general formulation of collective dynamics, one can show that if an external force and torque are applied to the system, the overall reaction rate depends on the applied force. In addition, a stability analysis of the equations governing the collective behavior predicts the existence of a clustering instability seen in many experiments of Janus colloids. Such considerations can be extended to ensembles of thermophoretic Janus colloids [63].

Acknowledgments

Research was supported in part by the Natural Sciences and Engineering Research Council of Canada and Compute Canada. Financial support from the Université libre de Bruxelles (ULB) and the Fonds de la Recherche Scientifique - FNRS under Grant PDR T.0094.16 for the project “SYMSTATPHYS” is also acknowledged.

Appendix

A. Force and Torque on a Colloidal Motor

A.1. Stationary Molecular Concentration Fields. We suppose that molecular diffusivity is large enough so that the concentration fields adopt a quasistationary profile in the vicinity of every Janus particle. Accordingly, the concentration fields should obey the following equations:

2ck=0,DkrckR=νkHcθκ+ccAκccBR,ck=gk, (A.1)

for the two reacting species k = A, B, where gk is the concentration gradient of species k at large distances from the center of the catalytic particle. In Ref. [40], we considered the special case where gk = 0, so that the concentration fields are uniform at large distances. The upper hemisphere is catalytic, while the lower one is noncatalytic. The axis of the Janus particle is oriented from the noncatalytic towards the catalytic hemisphere and taken along the z-axis in the frame of the particle.

We introduce the fields

ΦDAcA+DBcB, (A.2)
Ψ2κ+ccAκccB, (A.3)

in terms of the characteristic length of the reaction

κ+cDA+κcDB1. (A.4)

The fields Φ and Ψ have the units of sec−1, and the concentration fields are recovered from them by

cA=1DAκcDBΦ+1Ψ,cB=1DBκ+cDAΦ1Ψ. (A.5)

Similar expressions hold for the concentration gradients at large distances: gA, gB, gΦ, and gΨ.

The fields (A.2) and (A.3) obey the equations

2Φ=0,rΦR=0,Φ=gΦ,2Ψ=0,rΨR=1HcθΨR,Ψ=gΨ, (A.6)

where Hc(θ) is the Heaviside function of the catalytic hemisphere.

The solution of the equations for Φ is given by

Φ=Φg=Φ0+gΦ·r1+12Rr3, (A.7)
Φ0=DAc¯A+DBc¯B, (A.8)

which obeys reflective boundary conditions on the sphere r = R and represents a gradient at large distances. Defining Da ≡ R/ to be the dimensionless Damköhler number of the reaction on the spherical catalytic particle, the field Ψ can be decomposed as

Ψ=ΨgDaΨ0F, (A.9)

in terms of a field similar to equation (A.8)

Ψg=Ψ0+gΨ·r1+12Rr3,Ψ02κ+cc¯Aκcc¯B, (A.10)

and another field F obeying

2F=0,RrFR=HcθDaFRΨgRΨ0,F=0. (A.11)

In the equations above, the concentrations c¯k may be considered as their extrapolations to the center of the Janus particle or the mean concentrations at that position in a dilute suspension of Janus particles. Similarly, gk may also be considered as the mean gradients of concentrations at the location of the Janus particle in a dilute suspension.

A.2. Calculation of the Diffusiophoretic Velocities. The diffusiophoretic force (14) and torque (15) are thus given by the following expressions:

Fd=6πηR1+3b/R32bA1¯s·gA+32bB1¯s·gB+Rκ+cc¯Aκcc¯BbBDBbADA1·F¯s, (A.12)

and

Td=12πηR1+3b/R3R2bAn¯s×gA+3R2bBn¯s×gB+Rκ+cc¯Aκcc¯BbBDBbADAr×F¯s. (A.13)

Next, the field F can be expanded in Legendre polynomials. Since it obeys Laplace's equation and is vanishing at large distances, we find that

Fr,θ,φ=l=0alPlcosθRrl+1+3R2Ψ0gΨzl=0blPlcosθRrl+1+3R2Ψ02gΨxcosφ+gΨysinφ×l=1clPl1cosθll+1Rrl+1, (A.14)

with

alM1·Al, (A.15)
blM1·Bl, (A.16)
clN1·Cl, (A.17)
Al01dξPlξ, (A.18)
Bl01dξP1ξPlξ, (A.19)
Cl01dξP11ξ2Pl1ξll+1, (A.20)
Mll2l+12l+1δll+Da01dξPlξPlξ, (A.21)
Nll2l+12l+1δll+Da01dξPl1ξll+1Pl1ξll+1. (A.22)

First, we calculate the force (A.12) in order to obtain the corresponding velocity Vd. We have that

bknn¯s=16bkc+bkn1,bk1¯s=13bkc+bkn1, (A.23)

and, furthermore,

1x·bk1·F¯s=12γcbkc+γnbkngΨxΨ0,1y·bk1·F¯s=12γcbkc+γnbkngΨyΨ0,1z·bk1·F¯s=12Rαcbkc+αnbkn12βcbkc+βnbkngΨzΨ0, (A.24)

in terms of the integrals

αhdθHhθsin2θdAdθ,βh32dθHhθsin2θdBdθd,γh324dθHhθsinθcosθdCdθ+C, (A.25)

where

Al=0alPlcosθ,Bl=0blPlcosθ,Cl=1clPl1cosθll+1. (A.26)

Then, we calculate the torque (A.13) to get the corresponding angular velocity Ωd. We have that

bkn¯s=14bkcbknu,1x·bkr×F¯s=R2δcbkc+δnbkngΨyΨ0,1y·bkr×F¯s=R2δcbkc+δnbkngΨxΨ0,1z·bkr×F¯s=0, (A.27)

with

δh324dθHhθsinθdCdθ+cosθC, (A.28)

so that

bkr×F¯s=R2Ψ0δcbkc+δnbknu×gΨ. (A.29)

With the following coefficients associated with diffusiophoresis,

YhbAhDAbBhDBfor  h=c,n, (A.30)

the diffusiophoretic linear and angular velocities can be expressed as

Vd=Fdγt=11+2b/R12αcYc+αnYnκ+cc¯Aκcc¯Bu+12bAc+bAngA+12bBc+bBngB+R2γcYc+γnYnκ+cgAκcgB+R2βcγcYc+βnγnYnκ+cgAκcgB·uu, (A.31)
Ωd=Tdγr=916RbAcbAnu×gA+bBcbBnu×gB+34δcYc+δnYnu×κ+cgAκcgB. (A.32)

By defining the parameters,

ζA=ςκ+c, (A.33)
ς=12hαhYh1+2b/R, (A.34)
ξA=12hbAh1+2b/R+Ξκ+c, (A.35)
Ξ=R6hβh+2γhYh1+2b/R, (A.36)
εA=Eκ+c, (A.37)
E=R2hβhγhYh1+2b/R, (A.38)
λA=9bAcbAn16R+Λκ+c, (A.39)
Λ=34hδhYh, (A.40)

the linear and angular velocities in equations (A.31) and (A.32) can be written in the forms given in equations (18) and (19) in the main text.

A.3. Calculation of the Overall Reaction Rate. Moreover, we can also calculate the overall reaction rate (16). According to the boundary condition

DArcAR=Hcθκ+ccAκccBR, (A.41)

the reaction rate on the Janus particle is equivalently given by

Wrxn=r=RDArcARdΣ, (A.42)

with = R2dcosθdφ. The concentration field cA is again decomposed in terms of Φ and Ψ. The contributions from the terms Φ = Φg and Ψg are vanishing, so that there remains

Wrxn=DaΨ0r=RrFRdΣ. (A.43)

Using the expansion (A.14), we obtain the overall reaction rate (23) with the rate constants

k±=4πR2a0κ±, (A.44)

and the parameter

ϖ=3Rb02a0, (A.45)

where the coefficients a0 and b0 are given by equations (A.15) and (A.16) with l = 0.

B. Determination of Onsager's Linear Response Coefficients

B.1. Rotational Gradient and Divergence. In Euclidean space, the contravariant ai and covariant ai components of a vector a ∈ ℝ3 coincide: ai = ai. However, in spherical coordinates, the contravariant ari and covariant ari components of a rotational vector ar ∈ ℝ2 are related to each other according to ari = ∑j=θ,φgijarj in terms of the metric (39). Therefore, the scalar product between two rotational vectors ar, br ∈ ℝ2 has the following equivalent forms, ar · br = ∑i=θ,φaribri = ∑i=θ,φaribri. The inverse of the metric (39) associated with the spherical coordinates is given by

gij=gij1=1001sin2θ, (B.1)

and its determinant by

g=detgij=sin2θ. (B.2)

Hence, the element of angular integration can be written as d2u=gd2q=sinθdθdφ. Using contravariant components, the gradient of some function X is given by [59]

gradrXi=jgijXqj, (B.3)

and the divergence of some vector Xr by

divrXr=i1gqiXrig, (B.4)

which leads to equations (40) and (41) with the metric (39) of spherical coordinates.

B.2. Consequences of Onsager's Principle. Using the chemical potentials (49) and (50) and the assumptions (59) and (60), equation (57) becomes

wjAjBJC=Lrr00LrC0LAA0LAC00LBBLBCLCrLCALCBLCC·ArxnnAnAnBnBgradμCkBT, (B.5)

which implies that

w=LrrArxnuLrC·gradμCkBT, (B.6)
jA=LAA·nAnAuLAC·gradμCkBT, (B.7)
jB=LBB·nBnBuLBC·gradμCkBT, (B.8)
JC=LCrArxnLCA·nAnALCB·nBnBuLCC·gradμCkBT, (B.9)

where the sum extends over the different groups Δ2u of colloidal motors and

gradμCkBT=1ff+fβUtθffβμB·θu1sin2θφffβμsin2θB·φu. (B.10)

Using the expression (55) of the five-dimensional colloidal current density and comparing with the expression (B.9), we find that the linear response coefficients are given by

LCr=fχDrxnu0Δ2uδuu,LCA=fnAξA1+εAQuλAgradruΔ2uδuu,LCB=fnBξB1+εBQuλBgradruΔ2uδuu,LCC=fDt100Dr1rΔ2uδuu. (B.11)

Consequently, we obtain equations (62) and (63).

Remark 1 . —

Interestingly, the assumption (60), according to which the reaction rate does not depend on the gradients of molecular densities, can be relaxed by taking instead

LrA=ϖk+nAp,LrB=+ϖknBp, (B.12)

with the polarization (30). Cross-diffusion between the molecular species A and B may also be included with the coefficients

LAB=LBA=CnAnB1. (B.13)

In this general case where the matrix of linear response coefficients in equation (57) is complete, the reaction rate and the molecular current densities are instead given by

w=k+nAknBnC+ϖk+nAknB·pχDrxnu·f+fβUtd2u, (B.14)
jA=DAnACnAnBϖk+nAArxnp+nAξA1+εAQu·f+fβUt+λAgradru·gradrffβμB·gradrud2u, (B.15)
jB=DBnBCnBnA+ϖknBArxnp+nBξB1+εBQu·f+fβUt+λBgradru·gradrffβμB·gradrud2u. (B.16)

Neglecting the last term in the expression (B.14), we recover a form compatible with the reaction rate (23) obtained in Appendix A by direct calculation using the molecular concentration profiles around a single motor. The scheme has great generality.

C. Linear Stability Analysis Using the Colloidal Distribution Function

Linearizing equation (93) for f around a uniform steady state f0 = c0/(4π), we get

tδf=Dtz2δf+DrsinθθsinθθδfVsdcosθzδffξ+εcos2θ13z2δa+2λζfcosθzδa, (C.1)

where δf = ff0, and δa is ruled by the first line of matricial equation (88). These linear equations can be solved by expanding δf in series of Legendre polynomials as

δf=14πl=0δclPlcosθ. (C.2)

Supposing that δf, δa ~ exp(iqz), we find the following coupled equations

dδcldt=Dtq2+ll+1DrδcliqVsdl2l+1δcl1+l+12l+3δcl+1+c0q2ξδl,0+23εδl,2+iq2λζδl,1δa, (C.3)

for l = 0, 1, 2, ⋯, and

dδadt=Dq2+K~δawδc0. (C.4)

In matrix form, we have

ddtδaδc0δc1δc2δc3δc4δc5=Dq2K~w0000c0ξq2Dtq2iVsd3q000ic02λζqiVsdqDtq22Dri2Vsd5q00c02ε3q20i2Vsd3qDtq26Dri3Vsd7q0000i3Vsd5qDtq212Dri4Vsd9q0000i4Vsd7qDtq220Drδaδc0δc1δc2δc3δc4δc5, (C.5)

which can be solved by truncation to obtain the dispersion relations shown in Figures 3(d)3(f). If the wave number is vanishing (q = 0), the matrix in equation (C.5) has the following successive eigenvalues: s0=K~ for reaction, s+(0) = 0 for translational diffusion, and sl(0) = −l(l + 1)Dr with l = 1, 2, 3, ⋯ for rotational diffusion. In Figure 3(b), they appear in the order s+0=0>s10=2Dr>s0=K~>s20=6Dr> for the parameter values (87). We note that all the dispersion relations satisfy the property qs(0) = 0. At the threshold of clustering instability, the leading dispersion relation s+(q) should moreover satisfy the condition q2s+(0) = 0, which also gives equation (92), thus confirming the validity of the approximation (88) to determine the threshold.

Conflicts of Interest

The authors declare that there are no conflicts of interest.

References

  • 1.Vicsek T., Zafeiris A. Collective motion. Physics Reports. 2012;517(3-4):71–140. doi: 10.1016/j.physrep.2012.03.004. [DOI] [Google Scholar]
  • 2.Romanczuk P., Bär M., Ebeling W., Lindner B., Schimansky-Geier L. Active Brownian particles. The European Physical Journal Special Topics. 2012;202(1):1–162. doi: 10.1140/epjst/e2012-01529-y. [DOI] [Google Scholar]
  • 3.Aronson I. S. Collective behavior in out-of-equilibrium colloidal suspensions. Comptes Rendus Physique. 2013;14(6):518–527. doi: 10.1016/j.crhy.2013.05.002. [DOI] [Google Scholar]
  • 4.Cates M. E., Tailleur J. When are active Brownian particles and run-and-tumble particles equivalent? Consequences for motility-induced phase separation. EPL. 2013;101(2, article 20010) doi: 10.1209/0295-5075/101/20010. [DOI] [Google Scholar]
  • 5.Elgeti J., Winkler R. G., Gompper G. Physics of microswimmers—single particle motion and collective behavior: a review. Reports on Progress in Physics. 2015;78(5, article 056601) doi: 10.1088/0034-4885/78/5/056601. [DOI] [PubMed] [Google Scholar]
  • 6.Speck T., Menzel A. M., Bialké J., Löwen H. Dynamical mean-field theory and weakly non-linear analysis for the phase separation of active Brownian particles. The Journal of Chemical Physics. 2015;142(22, article 224109) doi: 10.1063/1.4922324. [DOI] [PubMed] [Google Scholar]
  • 7.Bechinger C., Di Leonardo R., Löwen H., Reichhardt C., Volpe G., Volpe G. Active particles in complex and crowded environments. Reviews of Modern Physics. 2016;88(4, article 045006) doi: 10.1103/RevModPhys.88.045006. [DOI] [Google Scholar]
  • 8.Marchetti M. C., Fily Y., Henkes S., Patch A., Yllanes D. Minimal model of active colloids highlights the role of mechanical interactions in controlling the emergent behavior of active matter. Current Opinion in Colloid & Interface Science. 2016;21:34–43. doi: 10.1016/j.cocis.2016.01.003. [DOI] [Google Scholar]
  • 9.Zöttl A., Stark H. Emergent behavior in active colloids. Journal of Physics: Condensed Matter. 2016;28(25, article 253001) doi: 10.1088/0953-8984/28/25/253001. [DOI] [Google Scholar]
  • 10.Ramaswamy S. Active matter. Journal of Statistical Mechanics: Theory and Experiment. 2017;2017(5, article 054002) doi: 10.1088/1742-5468/aa6bc5. [DOI] [Google Scholar]
  • 11.Gompper G., Winkler R. G., Speck T., et al. The 2020 motile active matter roadmap. Journal of Physics: Condensed Matter. 2020;32(19, article 193001) doi: 10.1088/1361-648x/ab6348. [DOI] [PubMed] [Google Scholar]
  • 12.Theurkauff I., Cottin-Bizonne C., Palacci J., Ybert C., Bocquet L. Dynamic clustering in active colloidal suspensions with chemical signaling. Physical Review Letters. 2012;108(26, article 268303) doi: 10.1103/PhysRevLett.108.268303. [DOI] [PubMed] [Google Scholar]
  • 13.Gao W., Pei A., Dong R., Wang J. Catalytic iridium-based Janus micromotors powered by ultralow levels of chemical fuels. Journal of the American Chemical Society. 2014;136(6):2276–2279. doi: 10.1021/ja413002e. [DOI] [PubMed] [Google Scholar]
  • 14.Wang W., Duan W., Ahmed S., Sen A., Mallouk T. E. From one to many: dynamic assembly and collective behavior of self-propelled colloidal motors. Accounts of Chemical Research. 2015;48(7):1938–1946. doi: 10.1021/acs.accounts.5b00025. [DOI] [PubMed] [Google Scholar]
  • 15.Ginot F., Theurkauff I., Detcheverry F., Ybert C., Cottin-Bizonne C. Aggregation-fragmentation and individual dynamics of active clusters. Nature Communications. 2018;9(1, article 696) doi: 10.1038/s41467-017-02625-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Thakur S., Kapral R. Collective dynamics of self-propelled sphere-dimer motors. Physical Review E. 2012;85(2, article 026121) doi: 10.1103/PhysRevE.85.026121. [DOI] [PubMed] [Google Scholar]
  • 17.Kapral R. Perspective: nanomotors without moving parts that propel themselves in solution. The Journal of Chemical Physics. 2013;138(2, article 020901) doi: 10.1063/1.4773981. [DOI] [PubMed] [Google Scholar]
  • 18.Saha S., Golestanian R., Ramaswamy S. Clusters, asters, and collective oscillations in chemotactic colloids. Physical Review E. 2014;89(6, article 062316) doi: 10.1103/PhysRevE.89.062316. [DOI] [PubMed] [Google Scholar]
  • 19.Pohl O., Stark H. Self-phoretic active particles interacting by diffusiophoresis: a numerical study of the collapsed state and dynamic clustering. The European Physical Journal E. 2015;38(8, article 93) doi: 10.1140/epje/i2015-15093-4. [DOI] [PubMed] [Google Scholar]
  • 20.Liebchen B., Marenduzzo D., Cates M. E. Phoretic interactions generically induce dynamic clusters and wave patterns in active colloids. Physical Review Letters. 2017;118(26, article 268001) doi: 10.1103/PhysRevLett.118.268001. [DOI] [PubMed] [Google Scholar]
  • 21.Huang M.-J., Schofield J., Kapral R. Chemotactic and hydrodynamic effects on collective dynamics of self-diffusiophoretic Janus motors. New Journal of Physics. 2017;19(12, article 125003) doi: 10.1088/1367-2630/aa958c. [DOI] [Google Scholar]
  • 22.Colberg P. H., Kapral R. Many-body dynamics of chemically-propelled nanomotors. The Journal of Chemical Physics. 2017;147(6, article 064910) doi: 10.1063/1.4997572. [DOI] [PubMed] [Google Scholar]
  • 23.Robertson B., Huang M.-J., Chen J.-X., Kapral R. Synthetic nanomotors: working together through chemistry. Accounts of Chemical Research. 2018;51(10):2355–2364. doi: 10.1021/acs.accounts.8b00239. [DOI] [PubMed] [Google Scholar]
  • 24.Huang M.-J., Schofield J., Gaspard P., Kapral R. From single particle motion to collective dynamics in Janus motor systems. The Journal of Chemical Physics. 2019;150(12, article 124110) doi: 10.1063/1.5081820. [DOI] [PubMed] [Google Scholar]
  • 25.Stark H. Artificial chemotaxis of self-phoretic active colloids: collective behavior. Accounts of Chemical Research. 2018;51(11):2681–2688. doi: 10.1021/acs.accounts.8b00259. [DOI] [PubMed] [Google Scholar]
  • 26.Einstein A. Investigations on the Theory of the Brownian Movement. New York: Dover; 1956. [Google Scholar]
  • 27.Smoluchowski M. V. Über Brownsche Molekularbewegung unter Einwirkung äußerer Kräfte und deren Zusammenhang mit der verallgemeinerten Diffusionsgleichung. Annalen der Physik. 1916;353(24):1103–1112. doi: 10.1002/andp.19163532408. [DOI] [Google Scholar]
  • 28.von Smoluchowski M. Drei Vorträge über Diffusion, Brownsche Molekularbewegung und Koagulation von Kolloidteilchen. Physikalische Zeitschrift. 1916;17:557–571. [Google Scholar]
  • 29.Smoluchowski M. v. Versuch einer mathematischen Theorie der Koagulationskinetik Kolloider Lösungen. Zeitschrift für Physikalische Chemie. 1918;92U(1):129–168. doi: 10.1515/zpch-1918-9209. [DOI] [Google Scholar]
  • 30.Kapral R. Advances in Chemical Physics. Vol. 48. Wiley; 1981. Kinetic theory of chemical reactions in liquids; pp. 71–179. [Google Scholar]
  • 31.Happel J., Brenner H. Low Reynolds Number Hydrodynamics. Martinus Nijhoff Publishers, The Hague; 1983. [Google Scholar]
  • 32.Dhont J. K. G. An Introduction to Dynamics of Colloids. Amsterdam: Elsevier; 1996. [Google Scholar]
  • 33.Nägele G. Colloidal hydrodynamics. In: Bechinger C., Sciortino F., Ziherl P., editors. Physics of Complex Colloids, Proc. of the Int. School of Physics Enrico Fermi. Amsterdam: IOS Press; 2013. pp. 507–601. [Google Scholar]
  • 34.Gerritsma E., Gaspard P. Chemomechanical coupling and stochastic thermodynamics of the F1-ATPase molecular motor with an applied external torque. Biophysical Reviews and Letters. 2011;5(4):163–208. doi: 10.1142/s1793048010001214. [DOI] [Google Scholar]
  • 35.Seifert U. Stochastic thermodynamics, fluctuation theorems and molecular machines. Reports on Progress in Physics. 2012;75(12, article 126001) doi: 10.1088/0034-4885/75/12/126001. [DOI] [PubMed] [Google Scholar]
  • 36.Speck T. Stochastic thermodynamics for active matter. EPL. 2016;114(3, article 30006) doi: 10.1209/0295-5075/114/30006. [DOI] [Google Scholar]
  • 37.Jülicher F., Grill S. W., Salbreux G. Hydrodynamic theory of active matter. Reports on Progress in Physics. 2018;81(7, article 076601) doi: 10.1088/1361-6633/aab6bb. [DOI] [PubMed] [Google Scholar]
  • 38.Fodor É., Cristina Marchetti M. The statistical physics of active matter: from self-catalytic colloids to living cells. Physica A: Statistical Mechanics and its Applications. 2018;504:106–120. doi: 10.1016/j.physa.2017.12.137. [DOI] [Google Scholar]
  • 39.Gaspard P., Kapral R. Communication: mechanochemical fluctuation theorem and thermodynamics of self-phoretic motors. The Journal of Chemical Physics. 2017;147(21, article 211101) doi: 10.1063/1.5008562. [DOI] [PubMed] [Google Scholar]
  • 40.Gaspard P., Kapral R. Fluctuating chemohydrodynamics and the stochastic motion of self-diffusiophoretic particles. The Journal of Chemical Physics. 2018;148(13, article 134104) doi: 10.1063/1.5020442. [DOI] [PubMed] [Google Scholar]
  • 41.Gaspard P., Kapral R. Thermodynamics and statistical mechanics of chemically powered synthetic nanomotors. Advances in Physics: X. 2019;4(1, article 1602480) doi: 10.1080/23746149.2019.1602480. [DOI] [Google Scholar]
  • 42.Speck T. Thermodynamic approach to the self-diffusiophoresis of colloidal Janus particles. Physical Review E. 2019;99(6, article 060602) doi: 10.1103/PhysRevE.99.060602. [DOI] [PubMed] [Google Scholar]
  • 43.Golestanian R., Liverpool T. B., Ajdari A. Designing phoretic micro- and nano-swimmers. New Journal of Physics. 2007;9(5, article 126) doi: 10.1088/1367-2630/9/5/126. [DOI] [Google Scholar]
  • 44.Oshanin G., Popescu M. N., Dietrich S. Active colloids in the context of chemical kinetics. Journal of Physics A: Mathematical and Theoretical. 2017;50(13, article 134001) doi: 10.1088/1751-8121/aa5e91. [DOI] [Google Scholar]
  • 45.Reigh S. Y., Huang M.-J., Schofield J., Kapral R. Microscopic and continuum descriptions of Janus motor fluid flow fields. Philosophical Transactions of the Royal Society A. 2016;374(2080, article 20160140) doi: 10.1098/rsta.2016.0140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Onsager L. Reciprocal relations in irreversible processes I. Physical Review. 1931;37(4):405–426. doi: 10.1103/PhysRev.37.405. [DOI] [Google Scholar]
  • 47.Onsager L. Reciprocal relations in irreversible processes II. Physical Review. 1931;38(12):2265–2279. doi: 10.1103/PhysRev.38.2265. [DOI] [Google Scholar]
  • 48.Prigogine I. Introduction to Thermodynamics of Irreversible Processes. New York: Wiley; 1967. [Google Scholar]
  • 49.Haase R. Thermodynamics of Irreversible Processes. New York: Dover; 1969. [Google Scholar]
  • 50.de Groot S. R., Mazur P. Nonequilibrium Thermodynamics. New York: Dover; 1984. [Google Scholar]
  • 51.Nicolis G. Irreversible thermodynamics. Reports on Progress in Physics. 1979;42(2):225–268. doi: 10.1088/0034-4885/42/2/001. [DOI] [Google Scholar]
  • 52.Landau L. D., Lifshitz E. M. Statistical Physics, Part 1. 3rd. Oxford, UK: Pergamon Press; 1980. [Google Scholar]
  • 53.Callen H. B. Thermodynamics and an Introduction to Thermostatistics. 2nd. New York: Wiley; 1985. [Google Scholar]
  • 54.Gaspard P., Kapral R. Nonequilibrium thermodynamics and boundary conditions for reaction and transport in heterogeneous media. The Journal of Chemical Physics. 2018;148(19, article 194114) doi: 10.1063/1.5031761. [DOI] [PubMed] [Google Scholar]
  • 55.Mazur P., Bedeaux D. A generalization of Faxén's theorem to nonsteady motion of a sphere through an incompressible fluid in arbitrary flow. Physica. 1974;76(2):235–246. doi: 10.1016/0031-8914(74)90197-9. [DOI] [Google Scholar]
  • 56.Albano A. M., Bedeaux D., Mazur P. On the motion of a sphere with arbitrary slip in a viscous incompressible fluid. Physica A: Statistical Mechanics and its Applications. 1975;80(1):89–97. doi: 10.1016/0378-4371(75)90148-X. [DOI] [Google Scholar]
  • 57.Anderson J. L. Colloid transport by interfacial forces. Annual Review of Fluid Mechanics. 1989;21(1):61–99. doi: 10.1146/annurev.fl.21.010189.000425. [DOI] [Google Scholar]
  • 58.Anderson J. L., Prieve D. C. Diffusiophoresis caused by gradients of strongly adsorbing solutes. Langmuir. 1991;7(2):403–406. doi: 10.1021/la00050a035. [DOI] [Google Scholar]
  • 59.Brenner H. Coupling between the translational and rotational brownian motions of rigid particles of arbitrary shape I. Helicoidally isotropic particles. Journal of Colloid Science. 1965;20(2):104–122. doi: 10.1016/0095-8522(65)90002-4. [DOI] [Google Scholar]
  • 60.Wang J. Nanomachines: Fundamentals and Applications. Weinheim, Germany: Wiley-VCH; 2013. [DOI] [Google Scholar]
  • 61.Huang M.-J., Schofield J., Gaspard P., Kapral R. Dynamics of Janus motors with microscopically reversible kinetics. The Journal of Chemical Physics. 2018;149(2, article 024904) doi: 10.1063/1.5029344. [DOI] [PubMed] [Google Scholar]
  • 62.Speck T., Bialké J., Menzel A. M., Löwen H. Effective Cahn-Hilliard equation for the phase separation of active Brownian particles. Physical Review Letters. 2014;112(21, article 218304) doi: 10.1103/PhysRevLett.112.218304. [DOI] [Google Scholar]
  • 63.Gaspard P., Kapral R. The stochastic motion of self-thermophoretic Janus particles. Journal of Statistical Mechanics: Theory and Experiment. 2019;2019(7, article 074001) doi: 10.1088/1742-5468/ab252f. [DOI] [Google Scholar]

Articles from Research are provided here courtesy of American Association for the Advancement of Science (AAAS) and Science and Technology Review Publishing House

RESOURCES