Governing Equation

We consider a two-dimensional domain where a circular solid particle whose radius is \(\dim{a}\) is positioned at the center. Hereafter \(\dim{q}\) denotes that the variable \(q\) is dimensional (i.e., not non-dimensionalized), and vice versa.

The motion of liquid is governed by the incompressibility constraint:

\[\pder{}{\dim{u}_i}{\dim{x}_i} = 0\]

as well as the momentum balance for Newtonian liquids:

\[\dim{\rho} \left( \pder{}{\dim{u}_i}{\dim{t}} + \dim{u}_j \pder{}{\dim{u}_i}{\dim{x}_j} \right) = \pder{}{\dim{\sigma}_{ij}}{\dim{x}_j} = - \pder{}{\dim{p}}{\dim{x}_i} + \dim{\mu} \pder{}{}{\dim{x}_j} \pder{}{\dim{u}_i}{\dim{x}_j}.\]

The particle emits a solute, which is transported into the liquid. In particular, the concentration of the solute (mass per unit area) follows the advection-diffusion equation:

\[\pder{}{\dim{c}}{\dim{t}} + \dim{u}_j \pder{}{\dim{c}}{\dim{x}_j} = \dim{D} \pder{}{}{\dim{x}_j} \pder{}{\dim{c}}{\dim{x}_j},\]

where \(\dim{D}\) is the diffusivity of the solute.

On the particle surface, we request the emission rate of the solute to be constant:

\[\dim{E} \equiv - \dim{D} \vat{ \pder{}{\dim{c}}{\dim{\vr}} }{\dim{r} = \dim{a}} = const.,\]

which serves as a radial boundary condition with respect to the concentration field. In the tangential direction, a slip velocity is prescribed:

\[\vat{ \dim{\ut} }{\dim{r} = \dim{a}} = \dim{M} \vat{ \frac{1}{\dim{\vr}} \pder{}{\dim{c}}{\dim{\vt}} }{\dim{r} = \dim{a}},\]

where \(\dim{M}\) is called mobility.

The particle experiences hydrodynamic force exerted by the surround liquid, which is governed by the Newton-Euler equation:

\[ \begin{align}\begin{aligned}& \dim{\rho} \pi \dim{a}^2 \der{}{\dim{v}_i}{\dim{t}} = \oint \dim{\sigma}_{ij} n_j d\dim{l},\\& \frac{1}{2} \dim{\rho} \pi \dim{a}^4 \der{}{\dim{\omega}_i}{\dim{t}} = \oint \epsilon_{ijk} \dim{r}_j \dim{\sigma}_{kl} n_l d\dim{l},\end{aligned}\end{align} \]

where we assume the particle is neutrally-buoyant. \(\dim{v}_i\) and \(\dim{\omega}_i\) denote the translational and rotational velocity, and the right-hand-side integrals are applied to the interface along the particle.

Normalized equations are

\[\pder{}{u_i}{x_i} = 0,\]
\[\rho \left( \pder{}{u_i}{t} + u_j \pder{}{u_i}{x_j} \right) = \frac{\dim{\mu} \dim{D}}{\dim{\rho} \dim{E} \dim{M} \dim{a}} \left( - \pder{}{p}{x_i} + \pder{}{}{x_j} \pder{}{u_i}{x_j} \right),\]
\[\pder{}{c}{t} + u_j \pder{}{c}{x_j} = \frac{\dim{\rho} \dim{D}}{\dim{E} \dim{a}} \frac{\dim{D}}{\dim{\rho} \dim{M}} \pder{}{}{x_j} \pder{}{c}{x_j},\]

with the corresponding boundary conditions on the particle:

\[\vat{ \pder{}{c}{\vr} }{r = 1} = - \frac{\dim{E} \dim{a}}{\dim{\rho} \dim{D}},\]
\[\vat{ \ut }{r = 1} = \frac{\dim{\rho} \dim{D}}{\dim{E} \dim{a}} \vat{ \frac{1}{\vr} \pder{}{c}{\vt} }{r = 1}.\]

The non-dimensional Newton-Euler equation leads to

\[ \begin{align}\begin{aligned}& \rho \pi a^2 \der{}{v_i}{t} = \frac{\dim{\mu} \dim{D}}{\dim{\rho} \dim{E} \dim{M} \dim{a}} \oint \sigma_{ij} n_j dl,\\& \frac{1}{2} \rho \pi a^4 \der{}{\omega_i}{t} = \frac{\dim{\mu} \dim{D}}{\dim{\rho} \dim{E} \dim{M} \dim{a}} \oint \epsilon_{ijk} r_j \sigma_{kl} n_l dl.\end{aligned}\end{align} \]

Note that we adopt \(\dim{a}\) and \(\dim{E} \dim{M} / \dim{D}\) as reference length and velocity scales to obtain the non-dimensional relations (Hu et al., Phys. Rev. Lett. (123), 2019).

In this project, we assume that the Reynolds number is sufficiently small so that the inertial effects on the momentum transport can be neglected. By fixing the non-dimensional flux to be unity (which does not loss the generality), we obtain a set of equations governing the whole system:

\[\pder{}{u_i}{x_i} = 0,\]
\[0_i = - \pder{}{p}{x_i} + \pder{}{}{x_j} \pder{}{u_i}{x_j},\]
\[\pder{}{c}{t} + u_j \pder{}{c}{x_j} = \frac{1}{Pe} \pder{}{}{x_j} \pder{}{c}{x_j},\]

with the corresponding boundary conditions on the particle:

\[\vat{ \pder{}{c}{\vr} }{r = 1} = - 1,\]
\[\vat{ \ut }{r = 1} = \vat{ \frac{1}{\vr} \pder{}{c}{\vt} }{r = 1}.\]

Note that the Newton-Euler equation in the limit of \(Re \rightarrow 0\) leads to

\[ \begin{align}\begin{aligned}0_i = \oint \sigma_{ij} n_j dl,\\0_i = \oint \epsilon_{ijk} r_j \sigma_{kl} n_l dl,\end{aligned}\end{align} \]

implying that the swimming particle (in the absence of external forces) is force-free and torque-free (Lauga, Cambridge University Press, 2020).

Velocity Field

Here we elaborate the velocity field.

To begin, we set a wall at \(\vr = R\). The wall and the particle satisfy the impermeability condition:

\[\vat{\ur}{\vr = 1} = \vat{\ur}{\vr = R} = 0.\]

On the inner wall, again, a slip velocity is enforced:

\[\vat{\ut}{\vr = 1} = \vat{ \frac{1}{r} \pder{}{c}{\vt} }{\vr = 1},\]

whereas the outer wall follows the no-slip condition:

\[\vat{\ut}{\vr = R} = 0.\]

Since the Stokes equation is a pure boundary-value problem, the entire flow field is uniquely determined once the boundary conditions are given. In particular, the solution can be conveniently expressed using the stream function \(\psi \left( \vr, \vt, t \right)\), given by:

\[\psi \left( \vr, \vt, t \right) = \sum_k \Psi_k \left( \vr, t \right) \expp{I k \vt},\]

where \(I\) is the imaginary unit, and \(\Psi_k\) is the stream function in the frequency space. Because of \(\psi \in \mathbb{R}\), \(\Psi_k\) satisfies the complex-conjugate property, and thus it is sufficient to consider \(k \ge 0\).

As derived in the appendix, this is given by

\[\begin{split}\Psi_k = \begin{cases} k = 0 & \text{unused}, \\ k = 1 & A_1 \vr^{-1} + B_1 \vr + E_1 \vr \log \vr + D_1 \vr^3, \\ \text{otherwise} & A_k \vr^{-k} + B_k \vr^k + C_k \vr^{2 - k} + D_k \vr^{2 + k}, \end{cases}\end{split}\]

where the coefficients \(A_k, B_k, C_k, D_k, E_1\) are determined using the boundary conditions at \(\vr = 1\) and \(\vr = R\).

Assuming that \(R\) is infinity, we request \(B_k, E_1, D_k\) to be zero so that the solution converges, leading to

\[\Psi_k = A_k \vr^{-k} + C_k \vr^{2 - k}.\]

By imposing the boundary conditions at \(\vr = 1\):

\[\begin{split}\begin{pmatrix} 1 & 1 \\ - k & 2 - k \\ \end{pmatrix} \begin{pmatrix} A_k \\ C_k \end{pmatrix} = \begin{pmatrix} 0 \\ - \left( U_{\vt}^s \right)_k, \end{pmatrix}\end{split}\]

we obtain

\[\Psi_k \left( \vr, t \right) = \frac{1 - \vr^2}{2 \vr^{k}} \left( U_{\vt}^s \right)_k.\]

In the above equation, \(U_{\vt}^s\) is the azimuthal velocity along the particle surface, expressed in the frequency domain:

\[\left( U_{\vt}^s \right)_k = I k C_k^s,\]

or in the physical space:

\[\vat{\ut}{r = 1} = \vat{ \frac{1}{\vr} \pder{}{\ut}{\vt} }{ r = 1 }.\]

\(C_k^s\) represents the concentration field on the particle surface (at \(\vr = 1\)) in the frequency domain:

\[c \left( \vr = 1, \vt, t \right) = \sum_k C_k^s \expp{I k \vt}.\]

Concentration Field

In addition to the constant-flux condition on the particle:

\[\vat{\pder{}{c}{\vr}}{\vr = 1} = -1,\]

we impose the Dirichlet condition on the outer boundary:

\[\vat{c}{\vr = R} = 0.\]

In practice, we set \(R\) to a finite value, which mismatches with the obtained stream function derived under the assumption that \(R \to \infty\). This contradiction is originated by the fact that no analytical solution is available due to the non-linear term in the advection-diffusion equation. We assume, nevertheless, that the approximation remains sufficiently accurate for large values of \(R\).