Viviana
Mancuso
ab,
Mihail N.
Popescu
c and
William E.
Uspal
*ab
aDepartment of Mechanical Engineering, University of Hawai’i at Mānoa, 2540 Dole St., Holmes Hall 302, Honolulu, HI 96822, USA. E-mail: uspal@hawaii.edu
bInternational Institute for Sustainability with Knotted Chiral Meta Matter (WPI-SKCM2), Hiroshima University, 1-3-1 Kagamiyama, Higashi-Hiroshima, Hiroshima 739-8526, Japan
cDepartment of Atomic, Molecular, and Nuclear Physics, University of Seville, 41080 Seville, Spain. E-mail: mpopescu@us.es
First published on 8th October 2024
Many biological microswimmers are capable of chemotaxis, i.e., they can sense an ambient chemical gradient and adjust their mechanism of motility to move towards or away from the source of the gradient. Synthetic active colloids endowed with chemotactic behavior hold considerable promise for targeted drug delivery and the realization of programmable and reconfigurable materials. Here, we study the chemotactic behavior of a Janus particle, which converts “fuel” molecules, released at an axisymmetric chemical patch located on a planar wall, into “product” molecules at its catalytic cap and moves by self-phoresis induced by the product. The chemotactic behavior is characterized as a function of the interplay between the rates of release (at the patch) and the consumption (at the particle) of fuel, as well as of details of the phoretic response of the particle (i.e., its phoretic mobility). Among other results, we find that, under certain conditions, the particle is attracted to a stable “hovering state” in which it aligns its axis normal to the wall and rests (positions itself) at an activity-dependent distance above the center of the patch.
For catalytic Janus particles, the reaction rate, and therefore the interfacial swimming actuation, depends on the local concentration of molecular fuel.8 This observation raises intriguing possibilities for controlling particle behavior. In particular, an understanding of how these particles respond to external gradients of fuel concentration may allow for realization of artificial chemotaxis, inspired by chemotaxis in biological micro-organisms. Here, one can make a suggestive analogy: in order to move towards regions with a higher concentration of nutrients or other chemoattractants, micro-organisms sense local concentrations and adjust their locomotion mechanisms accordingly. A classical example is E. coli, which can migrate towards a food source in liquid solution by temporally sampling the local nutrient concentration, and, acting on these measurements, suitably modulating a random sequence of straight line “runs” and uncontrolled “tumbling” events that randomize the bacterium's swimming direction.12 While E. coli performs temporal sampling, eukaryotic cells are large enough to directly sense chemical gradients across their body length using surface receptors.13–15 For instance, for the amoeba Dictyostelium, an asymmetric distribution of bound and unbound surface receptors can trigger polarization of internal biochemical pathways, allowing this organism to steer towards a food source.14 Neutrophils can chemotax towards an inflammation site through similar mechanisms.16,17 Although Dictyostelium and neutrophil motility usually involves contact with a solid surface, chemotaxis has also been observed for these cells when freely suspended and swimming in solution.18
Motivated by the swimming-squirming analogy between the motility of catalytic Janus particles and that of micro-organisms,19–21 many studies have considered whether such analogy extends towards an equivalent of chemotaxis of synthetic swimmers exposed to spatial gradients in the concentration of reactant. Here, we highlight a few of these studies; comprehensive reviews are provided by ref. 22 and 23. In an early work, Hong et al. studied the behavior of Pt/Au bimetallic nano-rods in the vicinity of a gel soaked with hydrogen peroxide “fuel.” The rods were observed to accumulate at the gel.24 Subsequently, Byun et al. pointed out the possibility of hydrodynamic flows sourced by the fuel patch in this set-up, and developed a framework for distinguishing advective and self-propulsive contributions to particle motion.25 Baraban et al. studied chemotactic behavior of tubular micro-jets and Pt@SiO2 Janus spheres in a microfluidic device.26 A solution containing hydrogen peroxide was continuously injected in one inlet port of the device, leading to a gradient of chemical “fuel” perpendicular to the direction of flow. The microswimmers exhibited a tendency to orient towards the region of high fuel concentration and accumulate in that region. In a similar microfluidic set-up, catalase-coated and urease-coated liposomes were observed to migrate towards and away from, respectively, regions of high substrate concentration.27 More recently, Xiao et al. used stop-flow microfluidic gradient generation to study chemotactic behavior of Cu@SiO2 Janus spheres, finding a tendency of these particles to align their cap towards higher fuel concentration.28 By stopping the flow, they could eliminate any potential contribution of surface-assisted cross-stream migration to particle motion.29 In an effort to realize chemotactic active colloids that use biocompatible fuels, Mou et al. fabricated ZnO-based Janus spheres fuelled by dissolved carbon dioxide.30 They observed these Janus particles to chemotax towards a carbon dioxide source through reorientation. Additionally, Zhou et al. studied bottle-shaped particles powered by internalized enzymatic decomposition of glucose.31 Similarly, they found that “opening” of a bottle will rotate towards a glucose source, leading to chemotactic migration.
To shed some light on mechanisms underlying experimental results, Popescu et al. presented a framework for understanding how the microscopic coupling between the orientation of a self-phoretic Janus particle and an ambient chemical gradient can lead to chemotaxis, defined to be active reorientation of the swimming direction with respect to the gradient.32 Active reorientation distinguishes chemotaxis from chemokinesis, defined as a variation of particle speed due to ambient chemical gradients.32,33 As a microscopic mechanism that induces reorientation, Popescu et al. pinpointed a crucial role for a quantitative difference in chemi-osmotic response of the catalytic and inert “faces” of the particle to a chemical gradient,32 which is the equivalent for the active particles of the classic result of Anderson concerning the electrophoresis of colloids with non-uniform zeta potential.10 The importance of a contrast in phoretic mobilities of the two “faces” for inducing alignment of the particle with an externally maintained gradient had also been noted in the context of thermophoretic Janus particles.34 Building upon the analysis by Saha et al.,35 Tătulea-Codrean and Lauga presented a comprehensive theoretical model for the motion of a self-phoretic Janus particle in a linear background reactant gradient in unconfined solution.36 They assumed that the reaction rate is proportional to the reactant concentration (i.e., first-order kinetics), considered the phoretic response to both the reactant and product gradients, and developed comprehensive analytical expressions for the translational and angular velocity of a Janus particle with arbitrary orientation with respect to the background gradient. Using these results, they considered the dispersion of a dilute suspension of such Janus particles exposed to a linear gradient in fuel. Finally, we mention that chemotaxis of chemically-powered micromotors in a fuel gradient has also been studied by means of mesoscopic particle-based simulations.37,38
Moving from the single-particle behavior to the context of suspensions, chemotaxis has been noted to play a role in phase separation, clustering, and other collective phenomena in systems of self-phoretic active colloids.39 When these particles interact through self-generated chemical gradients, the sign of chemotactic alignment, i.e., the attractive or repulsive character, determines whether these systems form finite-size clusters, or collapse into a single large cluster.40 Saha et al. distinguished four modes of response of an active colloid to an ambient chemical gradient, including chemotactic alignment, and mapped out a phase diagram of collective behaviors induced by chemotaxis.35 Chemotactic alignment can cooperate or compete with other interparticle interactions, especially activity-sourced hydrodynamic interactions.41–43 For instance, it has been shown that activity-induced fluid stirring can disrupt chemotactic pattern formation.43
Preceding theoretical studies have laid the groundwork for understanding chemotaxis of catalytic Janus particles. However, important aspects of the problem still remain to be explored. For instance, most experiments involve a spatially localized and finite-sized source of fuel (i.e., a “patch”). A patch will generate a fuel concentration field that has nonlinear dependence on spatial position. It is only far away from the patch, i.e., for distances from the patch that are much larger than the characteristic length scale of the patch, that a linearly varying field is a good approximation. However, nonlinear gradients may play a significant role in determining particle behavior, especially as the Janus particle approaches the patch. Secondly, the interplay of the characteristic sizes of the patch and particle may be important, especially as the particle approaches the patch. Thirdly, in experimental realizations, the patch is usually associated with a confining boundary, such as when the patch is the surface of another colloidal particle, or when the patch is imprinted on a planar substrate. A solid confining boundary will modify the chemical and hydrodynamic fields sourced by patch and the Janus particle, affecting particle motion. Here, we recall that even an inert planar wall can induce wall-bounded “sliding” and “hovering” states of a catalytic Janus particle.44
In this study, we investigate the motion of a Janus particle activated by a finite-sized, spatially localized patch source of “fuel” in a confining geometry. Specifically, we consider the motion of a catalytic Janus particle near a chemical patch with axisymmetric shape that continuously emits chemical fuel. The model assumes first-order chemical kinetics and therefore resolves the dependence of the catalytic activity of the particle on the local concentration of reactant and motion by self-phoresis induced solely by the product. Using first-order kinetics, the model can probe the relative significance of reaction rate (rate of fuel consumption by the particle) and fuel diffusion.45,46 Moreover, the model also explicitly considers the sizes of the particle and the patch, thus going beyond a far-field point particle and point source analysis (which is recovered in the limiting case of large distances). In large part, this study will focus on an axisymmetric configuration of the particle and patch, in which the particle center is located above the patch center, and the particle axis is aligned with the patch normal, with the general aim of understanding the conditions under which a Janus particle will navigate towards or away from the patch. Although this focus will limit our direct consideration of chemotactic reorientation, it is the logical starting point for understanding general three-dimensional motion.
In summary, our study provides foundations for understanding how a Janus particle responds to a localized, not point-like, source of fuel. The distinctive features include the consideration of a non-uniform reactant concentration field within the solution, sourced by a chemically active patch, as well as the role of confining geometry. Our main finding is that, under certain conditions, these features induce a novel “hovering” state in which the particle remains motionless at a steady height above the patch on the wall. The precise value of the hovering height depends on the dimensionless reaction rate (Damköhler number), as well as the material properties of the particle (phoretic surface mobility) and the details of the geometry (the shape of the particle, the shape of the patch, the size ratio, etc.) Depending on these parameters, the hovering state can be stable or unstable against vertical perturbations of the particle position. We also briefly investigate the stability of the hovering state against general three-dimensional (3D) motions. We show that the hovering state can indeed be stable in 3D, and attracts a particle from an arbitrary initial position and orientation that is not an axisymmetric configuration. Therefore, our study provides proof of concept that a fuel-releasing chemical patch can attract a Janus particle through chemotaxis.
From a mathematical standpoint, the model described above means determining a chemical field, which consists of the spatial distributions of reactant and product molecules, and a hydrodynamic field induced by the active actuation at the surface of the particle, as well as the overdamped motion of the particle. These points will be addressed in the next subsections. Before proceeding, here we discuss several approximations, already alluded to in the previous paragraph, which are suggested by the typical experimental studies of micrometer sized colloids in aqueous suspensions (these seem to be also the most suitable set-ups for realizations of the model system considered by us). These will allow significant simplifications of the mathematical description.
First, for systems involving suspensions with density and viscosity similar to that of water, particles with sizes typically in the order of micrometers, and flow and particle velocities typically in the order of micrometers per second, viscous forces dominate over inertia. This is quantified by the Reynolds number , where ρ is the density of the fluid, |U| a characteristic velocity (e.g., the translational velocity of the particle), and μ is the dynamic viscosity. Using Rp ∼ 1 μm, |U| ∼ 1 μm s−1, ρ ∼ 1000 kg m−3, and μ ∼ 10−3 Pa s renders Re ∼ 10−6 ≪ 1. Accordingly, the hydrodynamics is governed by the Stokes equation, and, consistently, the motion of the particle is in the overdamped regime.
Second, by the Stokes–Einstein relation, diffusion of the molecularly sized reactant and product species in aqueous solutions is fast, implying that the transport of the reactant and product species by diffusion dominates over that by advection. This assumption is substantiated by the value of the Péclet number (Pe), which compares the time scales of diffusion and advection over a lengthscale comparable to that of the particle size: the same numbers as above render D ∼ 10−9 m2 s−1 for Å-sized molecules and thus Pe = |U|Rp/D ≈ 10−3 ≪ 1. Consequently, the advection term can be neglected in the continuity equation for the product. This is a major simplification because it effectively decouples the equations governing the chemical field from the hydrodynamics of the solution.
Given these considerations, the reactant chemical field cr is modeled by the Laplace equation,
∇2cr = 0, | (1a) |
– at infinity
(1b) |
– release of reactant molecules at the patch at the wall (the plane z = 0):
−Dr[∇cr·ẑ]|z=0 = QK(r)|z=0, | (1c) |
– a sink at the active side, SA of the JP surface S where the conversion of reactant into product is promoted, and a reflective surface at the inert side SI of the JP surface, i.e.,
−Dr[∇cr·]|S = −F[f(r)cr]|S. | (1d) |
As discussed, we consider in this study only the case of patches with axisymmetric shapes, and thus we write K(r) = K(r) (z = 0, implicitly) and define the origin of the coordinate system at the center of the patch. The solution to the diffusion problem outlined above is more conveniently expressed as the superposition of a background field produced by the patch in the absence of the JP, and a “disturbance” field, cr = cbr + cdr. The background field obeys the Laplace equation, the boundary condition at infinity, eqn (1b), and accounts for the boundary condition at the wall eqn (1c). This problem can be solved analytically using the separation of variables and the Fourier–Bessel transform,47,48 yielding the solution:
(2) |
(3) |
(4) |
This leads to:
(5) |
Anticipating an extension of the type of geometries analysed as a test for the theoretical far-field, point-particle predictions (see the following section), we note here that from the solution for the background reactant field sourced by a circular patch, we can easily obtain the solution for the background field sourced by, e.g., a ring-shaped patch. Specifically, the ring-shaped patch has K(r) = 1 for Ri < r < Ro, and K(r) = 0 elsewhere. Due to the linearity of Laplace's equation, we can obtain the solution for a ring-shaped patch by subtracting the background field for a circular patch of radius Rd = Ri from the background field for a circular patch of radius Rd = Ro.
In terms of the – now known – background density field cbr, the diffusion problem for the disturbance field is formulated as follows. The field cdr obeys the Laplace equation, the boundary condition at infinity, eqn (1b), and – owing to cbr already fulfilling eqn (1c) – a zero normal current (reflective) boundary condition at the wall. The concentration gradient along the direction normal to the JP surface assumes the form:
[∇cr·]|S = [∇cbr· + ∇cdr·]|S, |
(6) |
(7) |
Da := FRp/Dr ≡ FC0/Q. | (8) |
Turning now to the distribution of product, we recall that at the JP surface the R-molecules are converted into P-molecules. At steady state, the product chemical distribution is the solution of the Laplace equation,
∇2cp = 0, | (9a) |
(9b) |
(9c) |
(9d) |
For simplicity, in the followings we assume equal diffusion coefficients, Dp = Dr.
vs(r) = −b(r)∇‖cp, | (10) |
The latter is a material-dependent parameter that describes the effective intermolecular interaction between the product and the colloid surface. The value of this parameter quantifies the strength of intermolecular forces, and the sign indicates whether these interactions are attractive (positive sign) or repulsive (negative sign). Moreover, the phoretic mobility can be either uniform or non-uniform over the particle surface. Typically, as the two sides of the Janus particle are composed of different materials, their distinct chemical properties result in separate and independent interactions with the product molecules. Consequently, the parameter “b” assumes distinct values bcb0 and bib0, on the cap and inert side, respectively. Here, bc and bi are dimensionless quantities, while b0 > 0 is a characteristic surface mobility. When considering the materials commonly utilized in the design of self-phoretic Janus particles, such as Pt and SiO2, experimental evidence suggests that the ratio bi/bc typically falls within the range of 0 to 0.3.53 This parameter holds particular significance in this problem as it can be adjusted to achieve specific motion dynamics.
Finally, we note that from the form of the slip velocity, we can define a characteristic velocity V0 ≡ b0C0/Rp. Substituting our expression for C0, we obtain
V0 = b0Q/Dr. | (11) |
∇·u = 0, | (12) |
∇·σ = 0; | (13a) |
u(|r| → ∞) = 0, | (13b) |
u(r)|S = [V + Ω × (r − rp) + vs(r)]S, | (13c) |
u(r)|z=0 = 0. | (13d) |
To close the system of equations (recall that V and Ω are unknown at this point), we consider the net force and the net torque acting on the colloid, defined as Fnet and the Tnet respectively. Since Re ≪ 1, these instantaneously vanish: Fnet ≈ 0 and Tnet ≈ 0. Both can be seen as the sum of hydrodynamic and external contributions so that: Fnet = Fh + Fext and Tnet = Th + Text. The hydrodynamic terms, Fh and Th, are exerted by the fluid on the particle. On the other hand, Fext and Text refer to external forces and torques acting on the system. In this study, we assume that there are no external forces or torques on the system. Consequently, the net force becomes:
(14a) |
Similarly, the net torque becomes:
(14b) |
At this stage, the hydrodynamic problem has been fully characterized and the analysis can be simplified by introducing the Lorentz reciprocal theorem, which connects the slip velocity and the translational and rotational velocities, leading to:54
(15) |
In the preceding, we did not assume any particular shape of the Janus particle. In the following, we will mainly consider spherical particles of radius Rp that are half-covered by catalyst. We will also consider prolate spheroidal particles with semi-major axis length Rp. The semi-minor axis length is quantified by the aspect ratio re as reRp, where 0 < re ≤ 1. Both spherical and prolate spheroidal particles are assumed to have axisymmetric catalyst coverage, and we will also consider the dependence of hovering on the particle shape, i.e., the value of re.
We briefly discuss the reactant background field, shown in Fig. 2 for a circular patch with Rd = Rp. In the bulk liquid near the patch (0.1 < z/Rp < 1 and |y|/Rp < 1.5), there is strong spatial variation of the concentration. In very close vicinity of the patch, z ≈ 0 and |y|/Rp < 1, the reactant field is approximately uniform in the lateral direction. At this close distance, the patch resembles a planar wall that uniformly releases reactant. Since the only length scale in determination of the background field for a circular patch is Rd, the values of the dimensionless concentration for other values of Rd can be easily inferred by rescaling of the coordinate axes and the dimensionless concentration by Rd/Rp (see the prefactor in eqn (5).)
Fig. 2 Background reactant concentration br sourced by a circular patch with radius Rd = Rp. The patch is shown as a thick black line. |
Fig. 3 shows the reactant concentration (on the left) and product concentration (on the right) for h/Rp = 1.2 and Da = 0.001, 1, 100. For the reactant, the insets (top-right corner) provide a zoomed-in view of the region between the patch and the JP, for a better visualization of the gradient. For the low value of Da, diffusion is the predominant process. A particular concentration, cr/C0 = 1 (yellow tone) reaches a distance on the wall (z = 0) equal to y/Rp = 0.8. In the vicinity of the catalytic “pole” of the particle, located at z/Rp = 2.2, the reactant concentration cr/C0 ≈ 0.15 (bluish color). At Da = 1, the two processes (reaction and diffusion) assume equal importance. At high Da, the reactant is quickly consumed by the catalytic cap, and therefore is less able to diffuse into the surrounding solution. Here, cr/C0 = 1 reaches only y/Rp = 0.5 on the wall. Moreover, the reactant concentration is nearly exhausted in the vicinity of the catalytic pole (cr/C0 ≈ 0.01, purple color.) In all cases, the results obtained numerically, with the BEM, and semi-analytically, using bipolar coordinates, show excellent quantitative agreement (see Appendix B).
In addition, Fig. 3 shows (on the right) the product concentration for the same height and values of Da. For the smallest value of Da, Da = 0.001, the concentration is highest near the catalytic pole. The maximum of the product concentration being located at the catalytic pole is a typical scenario for catalytic Janus spheres.46 Over the cap, the spatial gradient of concentration is directed from the “equator” to the pole, with a positive component in the z-direction. Over the inert region of the particle, the spatial gradient is from the pole to the equator; again, the z-component is positive. Recalling that vs(r) = −b(r)∇‖cp, we expect that the slip velocity will have a positive z-component where b is negative. Therefore, for bi < 0, the inert region would drive swimming towards the wall. (Recall from Fig. 1 that slip velocity induces translational motion in the opposite direction.) Likewise, for bc < 0, the catalytic region would also drive swimming towards the wall.
For the largest value of Da, however, we see an “inversion” effect. The concentration is highest near the “equator” of the particle. Thus, on the catalytic side, the spatial gradient of the concentration is directed from the catalytic “pole” to the “equator,” with a negative component in the z-direction. Therefore, if bc < 0, the slip velocity on the cap will be directed from the “pole” to the “equator,” and the catalytic cap would drive swimming away from the wall. On the inert side, we still have a spatial gradient from the “pole” to the “equator,” and a slip velocity in the same direction for bi < 0. Therefore, the inert side would still drive swimming towards the wall when bi < 0, as in the low Da case.
From the foregoing, it is clear that, for a given configuration of the particle and patch, the Damköhler number Da can have a profound effect on the product concentration field. Increasing Da for a particle at a given position can lead to a “concentration-inversion” effect. We can also show that, at a given value of Da, the inversion effect depends on proximity to the patch. In Fig. 4, we show the product concentration field for a sphere with Da = 100 that is far away from the patch (h/Rp = 100). The product concentration is mostly uniform over the cap, with a sharp decrease in the vicinity of the equator as one approaches the equator from the catalytic pole. Therefore, there is no inversion effect, and for bc < 0, the catalytic cap would drive swimming towards the wall.
Fig. 4 Concentration of the chemical product for a cap-up sphere located far away from the wall (h/Rp = 100) with Da = 100. The concentration is mostly uniform over the cap, with a sharp drop as one approaches the “equator” from the catalytic pole (note that the yellow region is thinner near the equator.) This variation is in contrast with the “inversion” effect seen for Da = 100 in Fig. 3. The patch size is Rd = Rp, and the concentration was computed with the BEM. |
Now we are in a position to qualitatively infer the existence of a “hovering” state. To simplify the following argument, suppose bi = 0, such that the inert side of the particle does not contribute to the particle velocity. For Da = 100 and bc < 0, the catalytic cap drives swimming away from the wall when the particle is at a height h/Rp = 1.2. For the same value of Da and the same bc < 0, the catalytic cap drives swimming towards the wall when h/Rp = 100. Thus, we expect that at some intermediate height 1.2 < h/Rp < 100, the velocity of the particle will be zero. The particle will exhibit motionless “hovering” at this height (although it will continuously pump the surrounding fluid.) The following section will confirm this reasoning quantitatively.
Appendix B provides a quantitative comparison of reactant concentration fields obtained with the BEM and via semi-analytical solution in bipolar coordinates (Fig. 17). We note that the agreement shown in Fig. 17 between the two methodologies provides important validation of our implementation of the boundary element method for first-order chemical kinetics. Beyond the parameters shown in the figures, the BEM can easily be extended to calculate chemical fields for very large Da numbers and/or very large distances h, as well as non-axisymmetric configurations. Moreover, the BEM straightforwardly handles other particle shapes, and is computationally significantly faster.
The power law behavior can be recovered in the framework of a far-field, “point-particle” theory. We model the reactant background concentration as being due to a point source located at the origin:
(16) |
cbr(r) = cbr(rp) + ∇cbr|r=rp·(r − rp) + …, | (17) |
From Fig. 5, it is apparent that for low values of Da, each Vzvs. h curve is associated with its own power law. Therefore, in the far-field, point-particle framework, we first consider the scenario that Da ≪ 1. In this scenario, the consumption of the fuel by the particle is negligible, and we approximate cr(r) ≈ cbr(r). Moreover, in the limit Da → 0, we expect to recover results from zeroth order kinetics, i.e., results for a Janus particle with a boundary condition −Dp[∇cp·] = κ on the catalytic cap, where κ = Fcbr(rp). For this boundary condition, it is known that a Janus particle in free space (unconfined solution), with its symmetry axis aligned with the z-axis, will have velocity , where Ṽfs is a dimensionless constant. Therefore, we obtain the scaling
(18) |
(19) |
The predicted scaling, plotted as dashed-dotted lines in Fig. 5, shows good agreement with the Da ≪ 1 data. In order to emphasize the universality of this scaling for Da ≪ 1, in Fig. 6 we show data for Da = 0.001 and Da = 0.01 for cap-up and cap-down spheres activated by circular patches of different sizes Rd, a cap-up prolate particle activated by a circular patch, and a cap-up sphere activated by a ring-shaped patch. We exploit our scaling to collapse the data onto a universal curve.
Fig. 6 In the limit Da → 0, the velocity Vz as a function of height h can be collapsed onto a universal curve, given by eqn (19), for various particle shapes, patch sizes, and patch shapes. Except where indicated in the legend, the shape of the particle is assumed to be spherical, the configuration is assumed to be cap-up, and the shape of the patch is assumed to be circular. The surface mobility is assumed to be uniform, with the sign chosen to give a positive velocity. For the prolate particle, re = 1/3, and Ṽfs was determined numerically as Ṽfs = 0.1247. For the ring, the outer radius is Ro/Rp = 2.5 and the inner radius is Ri/Rp = 1.5. |
Now we turn to the opposite limit, Da → ∞. From our data in Fig. 5, it is apparent that the function Vzvs. h/Rp for different values of Da collapses onto a universal curve as Da → ∞. In other words, the far-field scaling for Vzvs. h/Rp loses dependence on Da as Da → ∞. We give arguments in Appendix A rationalizing this loss of dependence on Da. Moreover, in Appendix A, we argue that the prefactor of the scaling Vz/V0 ∼ ÃdRp/h should be the dimensionless number 0.15 for a sphere. This predicted scaling is shown on Fig. 5, and has good agreement with the numerical data.
In Fig. 7, we show the results for a cap-up sphere for the cases of bc = 0, bi = 1 and bc = 1, bi = 0. In the first case (phoretic slip only on the inert side), we obtain a similar result as in the case of uniform mobility. Again, the low Da curves are described by the scaling law given in eqn (19). In this case, the value of Ṽfs appearing in the low Da scaling is Ṽfs = 1/8. We also obtain collapse for the Da ≫ 1 curves, although for higher values of Da than in the case of uniform mobility.
Fig. 7 Dependence of the vertical velocity Vz on the height h for a cap-up sphere with non-uniform surface mobility near a circular patch with Rd = Rp. |
Turning to the second case (phoretic slip only on the catalytic side), we obtain some intriguing behavior. Specifically, for Da ≫ 1, we find that the curves have negative Vz at low values of h/Rp, but switch over to positive Vz at high values of h/Rp. Accordingly, each of these curves intersect Vz = 0 at some crossover height hc(Da). These crossover points can be identified as hovering states. Since the slope of Vzvs. h at the crossover height in Fig. 7 is positive, the states shown in that figure are unstable against perturbations of the particle height. However, changing bc = 1 to bc = −1 would invert the sign of the velocity Vz, and therefore change the sign of the slope, which would make the hovering states stable against vertical perturbations.
Another intriguing aspect of bi = 0, bc = 1 can be observed for the Da ≫ 1 curves. Specifically, they follow a Vz ∼ (Rp/h)2 power law at intermediate heights, and cross over to a Vz ∼ (Rp/h) power law at large heights. The crossover behavior is especially apparent for Da = 350, shown in Fig. 8. The observed crossover between power laws can rationalized by the far-field perturbation theory developed for Da ≫ 1 in Appendix A. As shown there, in the limit Da → ∞, and for a uniform background concentration of reactant, the product concentration on the cap would be uniform: (∞)p = br(rp) on the cap, where the superscript (∞) indicates the Da → ∞ limit. Therefore, for bc = 1 and bi = 0, there would not be a surface gradient of product concentration on the region of the particle surface with a non-zero phoretic mobility. The particle velocity would therefore be zero. However, in this limit (Da → ∞), the particle can still self-propel if there is a background reactant gradient. Since the background reactant field is sourced by a patch at the origin, the gradient in the vicinity of the particle, ∇cbr|r=rp, decays as ∇cbr|r=rp ∼ (Rp/h)2. Therefore, the contribution to velocity that is leading order in inverse powers of Da is V(∞)z ∼ (Rp/h)2.
Fig. 8 Dependence of Vz on h for a cap-up sphere with bc = 1, bi = 0, and Da = 350. Crossover between two power law scalings is clearly visible. |
To complete the picture, we must explain the Vz ∼ Rp/h scaling at h/Rp ≫ 1 and the crossover between the two power laws. Here, we note that there is a subleading (Da−1) contribution to particle velocity for Da ≫ 1. As detailed in Appendix A, the subleading contribution does give particle motion in a uniform background concentration of reactant. Therefore, the subleading contribution to velocity goes as . Here, the superscript (1) indicates that this contribution to the velocity Vz is (Da−1). For a fixed, finite value of Da, this scaling will eventually dominate the leading order V(∞)z ∼ (Rp/h)2 scaling for sufficiently large h. The crossover height hc can be estimated from , or hc/Rp ∼ Da. In order to test this prediction, we show the crossover height as a function of Da in Fig. 9. We find that a linear relationship does indeed capture the dependence of hc/Rp on Da.
Fig. 9 Crossover height hc as a function of Da for a spherical, cap-up Janus particle with bc = 1, bi = 0. The particle is near a circular patch with radius Rd = Rp. We approximate the crossover height by the value where the particle velocity Vz is zero, as seen in the bottom panel of Fig. 7. The red line in the figure is a linear fit hc/Rp = 0.172 Da + 2.253. As discussed in the text, we predict a linear dependence of hc/Rp on Da using perturbation theory. |
Our analysis explained the crossover phenomenon mathematically. A physical explanation can be inferred from the “inversion” of the product concentration observed for Da = 100 in Fig. 3. When a particle is close to the patch, it experiences a strong reactant concentration gradient. Therefore, for the diffusion-limited regime (Da ≫ 1), the region of the catalyst near the “equator” has a higher rate of reaction than the region near the “pole,” due to the much greater availability of fuel closer to the patch. This effect explains the importance of the reactant gradient ∇cbr in determining the particle velocity for Da ≫ 1 and h < hc. However, when the particle is far away from the patch, as in Fig. 4, the reactant gradient is much weaker. Therefore, the product concentration field is qualitatively the same as for a particle in a uniform background reactant field, i.e., the maximum is at the pole.
Now we consider a cap-down particle. For bi = −1, bc = 0, shown in Fig. 10, top panel, we obtain a similar behavior with that exhibited by a cap-up configuration with bi = 1 and bc = 0. For bi = 0, bc = −1, shown in Fig. 10, bottom panel, we do not obtain any change in the sign of Vz. Accordingly, there are no hovering states for these parameters. However, for the high Da curves, we still observe a transition from Vz ∼ (Rp/h)2 behavior at intermediate heights h, to Vz ∼ (Rp/h) behavior at large heights h. The Vz ∼ (Rp/h)2 scaling is most evident for the largest value of Da, Da = 5000. This curve follows a Vz ∼ (Rp/h)2 scaling for a broad range of heights, with only a slight deviation for the largest height considered. The transition to Vz ∼ (Rp/h) behavior is clearly evident for curves with 10 ≤ Da ≤ 200.
Fig. 10 Dependence of the vertical velocity Vz on the height h for a cap-down sphere with non-uniform surface mobility near a circular patch with Rd = Rp. |
Having characterized the cases (bi = ±1, bc = 0) and (bi = 0, bc = ±1), we now turn to scenarios in which both parameters are non-zero. As a specific example, Fig. 11 shows Vzvs. h for a cap-up sphere with bi = −0.3 and bc = −1. It can be seen that there is a crossover for all Da ≥ 50. Moreover, the crossover heights are close to the wall: hc/Rp < 10 for each curve with crossover. Thirdly, at crossover, the slope of Vzvs. h is negative, indicating that the hovering state is stable against vertical perturbations. Given these observations, we conclude that bi = −0.3 and bc = −1 are particularly favorable values of the surface mobility for realizing stable hovering states near the wall. Fig. 11 also shows the scaling predicted by eqn (19) for Da ≪ 1.
Fig. 11 Dependence of the vertical velocity Vz on the height h for a cap-up sphere with bc = −1 and bi = −0.3 near a circular patch with Rd = Rp. |
By varying bi and bc and identifying the existence and location of hovering states, we determine a hovering “phase diagram” in the parameter space bi/bc and Da, as shown in Fig. 12 for a cap-up sphere near a circular patch with Rd = Rp. The ratio bi/bc is chosen as the horizontal axis because the sign of bc only affects the vertical stability of the hovering state (i.e., the slope of Vzvs. h at the crossover point), and not the existence of a hovering state. Additionally, the phase diagram is restricted to hovering states that occur for 1 ≤ h/Rp ≤ 10. In comparison with the previously considered case of bi = 0 and bc = 1, for which the hovering state occurred for high values of Da, there are also hovering states for Da < 1 for −1.0 < bi/bc < −0.3. More generally, “low” and “high” Da regimes can be distinguished from Fig. 12 on the basis of two features. The first feature concerns the dependence of the hovering height and phase boundaries on Da. In the low Da regime, there is no such dependence, and the left and right borders of the hovering region are vertical lines. In the high Da regime, hovering height does depend on Da, and the left and right borders are curved. The second feature concerns the dependence on bi/bc. The hovering states in the low Da regime occur at significantly lower bi/bc than the hovering states in the high Da regime. In consideration of these two features, we identify Da ≈ 1 as the border between the low and high Da regimes.
Considering the product concentration field shown in Fig. 3 for low Da (Da = 0.001), it is apparent that the low Da hovering states are not due to the “inversion” effect. Instead, it is likely that they have the same origin as the hovering states demonstrated in ref. 44, i.e., hydrodynamic and chemical interactions with the wall. Note that ref. 44 assumed zeroth-order chemical kinetics, which can be obtained from first-order kinetics in the limit Da → 0. Additionally, ref. 44 considered an inert wall and uniform background concentration of “fuel.” In contrast with the present work, the problem considered in ref. 44 gives hovering states in the range 1.02 ≤ h/Rp ≤ 7 for −1.4 ≲ bi/bc ≲ −0.95 for a Janus sphere that is half covered by catalyst. Therefore, although the Da ≪ 1 hovering states obtained here may have the same physical origin as in ref. 44, we find that sourcing the reactant field with a finite-sized patch can have a significant effect on the range of parameters that give a hovering state. Moreover, we note that the hovering states in ref. 44 are “degenerate” in that they can occur at any position in the plane of the wall (i.e., any x − y position), due to translational symmetry. In the present work, hovering states are localized directly above the patch (x = 0 and y = 0).
We also consider the phase diagrams for other patch sizes: Rd/Rp = 1/3 (Fig. 13, top) and Rd/Rp = 3 (Fig. 13, bottom). The dependencies of the phase behavior on Da and bi/bc described for Rd = Rp remain consistent when varying the patch size. However, there are quantitative differences between all three phase diagrams. Notably, for the smaller patch, there is a larger range of bi/bc that gives a hovering state, for both small and large Da. For the larger patch, the range of bi/bc giving a hovering state at low Da is more narrow than for Rd/Rp = 1. The range at high Da is similar, although shifted to the left.
The “edge case” bi = −bc merits some additional consideration, as a particle with these mobility parameters in a uniform background concentration would be motionless in the limit Da → 0. In other words, Ṽfs = 0. However, recalling that the patch creates a concentration gradient ∇cbr, we expect a scaling law for Da → 0. We indeed recover this scaling for Da ≪ 1, as shown in Fig. 14. (The numerical prefactor listed in the legend is determined from Tătulea-Codrean and Lauga.36) For high Da, there is no reason to expect these mobility parameters to give zero velocity in a uniform reactant field, and we indeed obtain a scaling for Da ≫ 1. For intermediate values of Da, there is crossover between the two scaling laws: for small h/Rp, and for large h/Rp. This crossover is evident even for Da as low as Da = 0.001. Given previous results on crossover, we expect this crossover to result from a competition between leading and subleading terms, although in this case for an expansion in powers of Da, instead of inverse powers of Da. We defer a more detailed examination to later work.
As a first step towards this direction of investigation, we show in Fig. 15 an example of a three-dimensional trajectory, computed using the BEM and a quaternion-based rigid-body dynamics method57 (see Appendix C for a brief discussion of this method). This demonstrates that a prolate spheroid with phoretic mobilities bi = −0.4 and bc = −1, Damköhler number Da = 100, and aspect ratio re = 1/3 is attracted to the hovering state from an arbitrarily chosen initial position and orientation. The initial position of the particle (xp/Rp = 4, yp/Rp = 0, and zp/Rp = 7) is located off the patch axis. Additionally, the initial particle orientation is neither aligned with the patch axis, nor with the plane defined by the patch axis and the patch-to-particle vector. Specifically, the initial orientation of the spheroid is θ = 60° and φ = 5°. Here, θ is the angle between the particle orientation vector and ẑ, where is defined to point from the inert pole (black) to catalytic pole (red) of the particle. The quantity φ is the angle between and . Accordingly, all potential symmetries of the patch-and-particle configuration are broken. Despite this fact, the particle is able to steer to the patch, align its axis with the patch normal, and remain in the hovering state for the rest of the simulation. Fig. 15 shows the rod trajectory and the time evolution of its configuration. Since the azimuthal angle φ is not well-defined when θ = 0, we use p⊥ = ·(ẑ × p) to characterize the component of the orientation vector that is out of the plane defined by the patch normal ẑ and the particle position rp. Here, p = rp/|rp|.
Interestingly, for spheres, we did not observe this long-time stability for any of the parameters tested. For some parameters, we found that the spheres could initially steer themselves towards the hovering state, but subsequently exhibited a slow drift away from the patch axis over a longer timescale. An example trajectory, illustrating such a behavior, is shown in Fig. 16. This observation suggests that shape may have a crucial role in promoting stable “docking,” most likely by a tendency for elongated particles to align with nonlinear gradients.58 In future work, we will undertake a systematic examination of the three-dimensional linear stability of the hovering state as a function of particle aspect ratio, Da, and surface mobilities. This examination will settle whether slender shape and a high value of Da are indeed necessary conditions for three-dimensional stability.
Throughout this manuscript, we have neglected the effect of thermal fluctuations. The stiffness of the hovering states with respect to perturbations in vertical position can be quantified by evaluating the slope of the Vz/V0vs. h/Rp curve at the hovering height. Specifically, denoting the slope by m, the characteristic relaxation time for vertical perturbations is τ = Rp/|m|V0. For Da = 50 and surface mobility parameters bi = −0.3 and bc = −1, we obtain |m| ≈ 0.0017 (Fig. 11). For comparison with experiments, we consider a particle with Rp = 2.5 μm that typically moves with speed V = 9 μm s−1 in a uniform background concentration of molecular “fuel.”3,59 Numerically, we obtain that V ≈ 0.1V0 for Da = 50, and thus τ ≈ 16 s. This relaxation time can be compared against the characteristic timescale for translational diffusion, τt = Rp2/D = 6πμRp3/kBT, where D is the diffusion coefficient of the particle, and the second equality assumes a spherical particle. In ref. 44, τt was calculated as τt ≈ 70 s. Since τ < τt, we expect the hovering state to be robust against thermal fluctuations in the vertical direction. Similar arguments can be made for a hovering state with a small value of Da. For example, for Da = 0.1, bi = 0.5, and bc = −1, and using the same dimensional particle radius and speed as before, we obtain V ≈ 0.006V0 and |m| ≈ 0.0003, giving τ ≈ 6 s (i.e., this state is even more robust against thermal fluctuation, as the relaxation back to hovering happens on even shorter timescales than in the large Da example.) Future work on three-dimensional motion can address stability against translational diffusion in a horizontal direction, as well as stability against rotational diffusion.
Although we observed hovering states over a wide range of Da, from Da = 0.0001 to Da = 5000, our detailed analysis (Appendix A) largely focused on the high Da regime. This focus was driven by the novelty of the patch-induced “concentration inversion” mechanism for inducing near-wall hovering states. In comparison, the low Da hovering states are most likely due to chemical and hydrodynamic interactions with the wall. Comparatively, this latter mechanism is rather well-known, and can be analyzed using singular solutions to the Laplace and Stokes equations (including image singularities to account for the presence of the wall).44,60 Future work could take this approach, building on our eqn (19), to obtain mechanistic insight into low Da hovering.
Our study focused on an axisymmetric configuration of the particle and patch. However, we laid the groundwork for consideration of general three-dimensional motion, and demonstrated that the hovering state can be stable against general three-dimensional perturbations. Specifically, we demonstrated that a particle with arbitrary initial orientation and position can steer to a chemical patch, i.e., exhibit chemotaxis. Three-dimensional motion would therefore be a natural direction for continuation of our work. This direction could address the phase diagram for three-dimensional stability of the hovering states, as well as how the system parameters can be tuned to enlarge the basin of stability, i.e., enlarge the range of initial orientations and positions from which a particle will be attracted to the hovering state. Continuation of our study could also probe the possible existence of other dynamical fixed points, such as continuous “sliding” along the edge of a circular patch. This sliding would be analogous to the orbiting observed for active colloids near geometric obstacles.59,61,62 One can also imagine hovering states that are shifted away from the center of the patch. In this scenario, axisymmetry would imply that there is a continuous ring of degenerate hovering states. For a non-axisymmetric patch, such as a square or star-shaped patch, there could be hovering states at a discrete set of spatial positions. Three-dimensional Brownian motion of the colloid can also be straightforwardly included in the framework of our model.29,63
Our results for a particle near a patch-decorated planar wall may have implications for bound pair formation or “docking” of two freely suspended colloidal particles.64–66 We recall that a planar wall can be regarded as a sphere with infinite radius of curvature. To our knowledge, a situation in which a reactant source is localized to one particle, and a catalytic region to a second particle, has not previously been considered in the literature, especially for first-order kinetics. Temporal variation of the rate of release for a particle/particle system could be exploited for temporally sequenced cargo pick-up and release.67 Incorporation of additional chemical species and/or nonlinear chemical kinetics in our framework could allow for realization of more complex particle behaviors.68
Finally, our findings may provide conceptual and technical guidance for advanced applications involving active colloids. In lab-on-a-chip systems, a chemical patch with a time-varying rate of release could be used for programmed “trap-and-release” of an active colloid. A planar surface decorated with many patches could template assembly of an active colloidal crystal. In materials science applications, a material that is impregnated with reactant, and which releases the reactant upon damage, could recruit active colloids to the damage site.69 (Alternatively, the scenario can be that of locating corrosion spots on metal surfaces.) Additionally, for drug delivery applications, the chemical patch in this study could model an in vivo biological source of reactant.6 A stably hovering particle that continuously releases a drug would provide a highly concentrated and localized dose. For biomedical applications, future consideration of aspects such as those of a non-Newtonian carrier fluid70 or the effect of ambient flow29 would enhance the realism of our model. One could also consider a patch embedded in a fluid–fluid interface or a bilayer membrane as a model for binding sites at biological interfaces.71
(20) |
The governing equation becomes
(21) |
(22) |
On the inert side of the particle, we have
(23) |
The boundary conditions for (∞)r on the wall and at infinity are the same as given for r in Section 3. Therefore, the inhomogeneity in the boundary condition on the wall for r is accounted for in the leading order problem. For all i ≥ 1, the boundary condition on the wall is , and the boundary condition at infinity is (i)r = 0.
Thus, we have a well-defined boundary value problem for (∞)r, with so-called mixed boundary conditions on the JP51 (i.e., Dirichlet on the cap and Neumann on the inert face.)
Now we turn to the product concentration. Again, we expand as follows:
(24) |
(25) |
On the inert side of the particle, we have
(26) |
Thus, we have obtained a boundary value problem for (∞)p with Neumann conditions on both sides of the JP, with the prescribed product flux on the catalytic cap obtained from the solution to the problem for (∞)r.
The product concentration gradient determines the particle velocity. We define
Vz = V(∞)z + V(1)z + V(2)z + … | (27) |
At this point, we have established that, in the limit Da → ∞, the rate of production of product molecules on the cap no longer depends on Da. This result entails that, for a given set of mobility parameters, the Vzvs. h/Rp curves for different values of Da should collapse onto a universal curve as Da → ∞. This is observed for Da ≫ 1 curves in various figures (e.g., Fig. 5), although the values of Da for which collapse is observed depends on the mobility parameters. Additionally, for some mobility parameters, collapse of a Da curve may occur over a limited range of heights h. For instance, in Fig. 10, collapse is observed at lower values of h/Rp. The curves with higher Da exhibit collapse for a larger range of h/Rp.
(28) |
Davis and Yariv also obtain the Janus particle velocity for the case of a sphere with uniform phoretic mobility.50 However, they assume – in contrast with the present manuscript – that the slip velocity is driven by the surface gradient of the reactant concentration, and not by the gradient of a product concentration. Since the product concentration field obeys Neumann conditions over the whole surface of the JP, we cannot immediately use this result. However, if we write the product concentration field as , we obtain that An = −Bn for all n. Therefore, their result that V(∞)z/V0 = ±0.3br(p) applies here. (The sign of the velocity is determined by the cap-up or cap-down character of the particle.) We therefore obtain that as Da → ∞. Aside from the sign, this result is independent of the cap-up or cap-down character of the particle, since it was obtained using the approximation of a uniform background concentration of reactant.
In order to obtain this crossover, we argue that, at leading order in Da as Da → ∞, the particle velocity is zero for a uniform background concentration of reactant. The reason is the following. We recall that in the leading order problem, i.e., the problem for (∞)r, the boundary condition on the catalytic cap is (∞)r = 0. Additionally, assuming a uniform background concentration of reactant br(p), and neglecting the confining wall, we recall that (∞)r and (∞)p can be written as harmonic expansions, with the coefficients related by An = −Bn. Since (∞)r = 0 on the cap, the coefficients Bn are such as to precisely cancel the background concentration br(p) on the cap. In other words, on the catalytic cap, . From An = −Bn, it follows that, on the catalytic cap, .
At this point, we have obtained that the product concentration on the cap is uniform as Da → ∞, for a uniform background concentration of reactant. In the case bc = 1, bi = 0, only the cap is osmotically responsive, i.e., we can only have slip on the cap. However, since the slip is proportional to the surface gradient of product concentration, we obtain that the slip is zero over the entire surface of the particle. Therefore, to leading order in Da as Da → ∞, the particle does not move in a uniform background concentration of reactant.
However, we can still obtain motion in the leading order problem by considering the role of the second term in eqn (17), which represents a linear gradient in the background reactant concentration. At leading order, the reactant concentration can be written as . We still have the boundary conditions (∞)r = 0 on the catalytic cap and on the inert face. Therefore, the values of Cn must be such as to cancel the linearly varying background reactant concentration on the catalytic face. Concerning the product concentration, we can write this field as , with the same boundary conditions as before. By the same arguments, we obtain Dn = −Cn. Therefore, the product concentration varies with position on the catalytic face. This spatial variation is sufficient to give motion of the particle when bc = 1 and bi = 0. Since the background gradient decays as (Rp/h)2, we obtain that V(∞)z/V0 ∼ (Rp/h)2. The sign will depend on the cap-up or cap-down character of the particle, since the linear gradient in background reactant concentration is not isotropic.
We briefly argue that the subleading problem will give a contribution to velocity V(1)z/V0 ∼ Da−1(Rp/h). Approximating the JP as being in unconfined solution, we have the boundary condition (1)r() → 0 as | − p| → ∞, and we write . We recall that the boundary conditions on the JP are on the catalytic cap, and on the inert face. This is a mixed boundary value problem, giving the dual series equations on the catalytic face, and on the inert face, where the Bn are known from solution of the leading order problem. Now, for the particle to remain motionless, the reactant field must take a constant value on the catalytic face. We recall that the coefficients Bn give a constant value of the reactant on the catalytic face and no-flux on the inert face. By uniqueness of solutions to linear equations, we deduce that En = αBn. (The constant α is included to allow for the scenario in which (∞)r and (1)r take different uniform values on the cap.) However, from the dual series equations, it is clear that En cannot be a constant multiple α of Bn: En ≠ αBn. Therefore, the field (1)r cannot take a uniform value on the cap and also satisfy the no-flux boundary condition on the inert face. Since the field , where Fn = −En, it follows that (1)p is non-uniform on the cap. Therefore, in the subleading problem, the particle can move in a uniform background reactant concentration. We thus obtain V(1)z/V0 ∼ Da−1(Rp/h).
(29) |
(30) |
ξ0 = arccosh(h/Rp). | (31) |
The metric factors, which enter in the calculation of gradients and area elements (e.g., ∇ = eξgξ−1∂ξ + eηgη−1∂η + eϕgϕ−1∂ϕ), are given by
(32) |
In terms of the bipolar coordinates system introduced above, the plane z = 0 of the wall corresponds to ξ = 0, while the surface at infinity is parametrized by ξ = 0, η = 0 (ω = 1). The activity distribution K(s) ↦ K(ω) is calculated by noting that ξ = 0, s = 0 (the center of the patch) corresponds to η = π (ω = −1) while ξ = 0, s = Rd (the circumference of the patch) corresponds to
(33) |
(34) |
Similarly, the activity distribution f(rJ) ↦ f(ω), where rJ denotes positions on the surface of the Janus particle, is calculated by noting that: (i) the “south” pole (s = 0, z = h − R) corresponds to (ξ = ξ0, ω = −1), (ii) the “north” pole (s = 0, z = h + R) corresponds to ξ = ξ0, ω = +1; and (iii) the “equator” {s = R, z = h} corresponds to ξ = ξ0,
ωe = (coshξ0)−1 = Rp/h. | (35) |
(36) |
(37) |
(38) |
An = 0, n ≥ 0. | (39) |
(40) |
(41) |
(42) |
(43) |
For the results presented in the manuscript we have used Nmax = 10 and numerically solved the system using the software Mathematica (version 14.0); this was sufficient to pass the cross-check against the corresponding BEM results (see the good agreement between the two shown by the results in Fig. 17.) If very accurate calculations of the concentrations are needed, additional analytical manipulations of the equations derived above may be necessary, as discussed by ref. 75.
(44) |
:= Dr/Dp, | (45) |
(46) |
(47) |
Owing to the symmetry of the system, only motion along the z-axis occurs; accordingly, only one auxiliary problem is needed in eqn (15), and this is chosen as the motion of a chemically inert spherical particle, of same radius Rp and located at the same height h with velocity V′ = V′ez. This was studied by ref. 73, which provides the force Fa in the form of the well-known wall-correction of the Stokes formula and outlines the solution for the hydrodynamic flow in terms of a stream function Ψa(ξ,ω := cosη) = V′Rp2ψa(ξ,ω), where 0 ≤ ξ ≤ ξ0 while ψa(ξ,ω) is dimensionless, which admits the series representation in bipolar coordinates as:73,74
(48) |
(49) |
The phoretic slip velocity in the rhs of eqn (15) is given by
(50) |
Since the unit vector along the direction normal to the surface of the particle is n = −eξ, the integrand in the rhs of eqn (15) will involve only the contraction eξ·σa·eη of the shear stress tensor of the auxiliary problem with the unit vectors eξ,η, evaluated at ξ = ξ0. This contraction is given by (see also ref. 78),
(51) |
(52) |
After performing the derivatives analytically, the integral in the rhs of eqn (15) is computed numerically and the integrand is approximated by truncating the series representations corresponding to the phoretic slip and of the stress-tensor contraction to the first Nmax = 30 terms, which seems sufficient as long as h/Rp ≳ 1.1.
In Fig. 18, we show the velocity of a cap-up sphere with non-uniform surface mobility near a circular patch with radius Rd = Rp, computed with the boundary element method and using bipolar coordinates. The two methods show close quantitative agreement for the range of h/Rp and values of Da shown. The system defined by truncating eqn (47) to n equations is observed to have slow convergence with n for larger h/Rp and higher Da. Accordingly, we use the BEM for velocity data shown in Section 3.
The trajectory of a particle is obtained as follows. At each simulation timestep, we use the BEM to compute the particle's translational velocity V(t) and angular velocity Ω(t), which depend on the particle's instantaneous position rp(t) and orientation q(t). The position and orientational quaternion are then integrated forward in time using a predictor-corrector Euler method with timestep Δt:
rp(t + Δt) = rp(t) + V(t)Δt, | (53) |
(54) |
q(t + Δt) = q* + λqq(t). | (55) |
(56) |
(57) |
This journal is © The Royal Society of Chemistry 2024 |