Fact-checked by Grok 2 weeks ago

Stokes flow

Stokes flow, also known as creeping flow, describes the motion of a viscous, incompressible at very low Reynolds numbers, where viscous forces overwhelmingly dominate inertial forces, rendering advective terms negligible in the governing equations. This regime is characterized by the linearity of the Stokes equations, \nabla p = \mu \nabla^2 \mathbf{u} and \nabla \cdot \mathbf{u} = 0, which represent a simplification of the full Navier-Stokes equations by omitting the nonlinear convective term. Such flows exhibit key properties including kinematic reversibility—meaning that reversing the of all velocities reverses the flow pattern—and the minimization of viscous dissipation energy among all divergence-free velocity fields satisfying the boundary conditions. The concept originates from the work of Irish mathematician and physicist George Gabriel Stokes, who in 1845 derived the general form of the Navier-Stokes equations in his paper "On the Theories of the Internal of Fluids in Motion, and of the and Motion of ." Stokes further advanced the field in 1851 with his seminal publication "On the Effect of the Internal of Fluids on the Motion of Pendulums," where he analyzed the experienced by a moving slowly through a viscous , yielding : the F_d = 6\pi \mu a U, with \mu as the dynamic viscosity, a the radius, and U the velocity. This law provides an exact solution to the Stokes equations for a translating in an unbounded and forms the foundation for understanding low-Reynolds-number hydrodynamics. Due to their linear nature, solutions to Stokes flow problems can be superposed, facilitating analytical and numerical treatments for complex geometries via methods like boundary integrals or multipole expansions. Stokes flow is prevalent in natural and engineered systems where characteristic lengths or velocities yield Reynolds numbers much less than unity, such as the sedimentation of colloidal particles, the locomotion of microorganisms like and , and flows in microfluidic devices for technologies. In , it governs the fluid dynamics around flexible structures like flagella, enabling models of microbial and nutrient transport in cellular environments. Engineering applications include , polymer processing, and the design of (microelectromechanical systems), where precise control of viscous-dominated flows is essential.

Fundamentals

Definition and physical context

Stokes flow, also known as creeping flow, describes the motion of a viscous, incompressible at very low s, where advective inertial forces are negligible compared to viscous forces. This regime occurs when the dimensionless Re = \rho U L / \mu \ll 1, with \rho denoting fluid density, U the characteristic velocity, L the characteristic length scale, and \mu the dynamic viscosity; the small Re signifies dominance of diffusion-like viscous effects over inertia. In Stokes flow, the quasi-steady approximation is typically invoked, neglecting unsteady acceleration terms in the momentum balance when temporal changes are slow relative to viscous timescales. This simplification captures the essential physics of slowly varying flows without transient inertial contributions. Stokes flow applies to diverse physical contexts, including microscale phenomena in microfluidic devices where channel dimensions yield inherently low Re. It also governs the of small particles, such as grains settling in air, which motivated early derivations of laws for spherical objects. Additionally, it models slow viscous motions in lubricants and biological systems, like the swimming of microorganisms in aqueous environments.

Historical background

The foundational concepts underlying Stokes flow emerged from early 19th-century efforts to incorporate into the equations of fluid motion. In 1822, derived the first form of the Navier-Stokes equations by adding a viscous term to Euler's inviscid equations, motivated by molecular interactions in fluids. This work laid the groundwork for modeling viscous effects, though it was initially overlooked. independently re-derived a similar set of equations in 1829, emphasizing the role of internal friction in and contributing to the theoretical framework for low-speed flows. The term "Stokes flow" originates from George Gabriel Stokes' seminal 1851 paper, "On the Effect of the Internal Friction of Fluids on the Motion of Pendulums," where he analyzed the of pendulums in viscous . In this work, Stokes derived the drag force on a moving at low Reynolds numbers, given by F_d = 6\pi \mu a U, where \mu is the dynamic , a is the radius, and U is its velocity; this formula has since become central to understanding and colloidal suspensions. Stokes' analysis approximated the full Navier-Stokes equations by neglecting inertial terms, valid for creeping flows where viscous forces dominate. A key limitation of this approximation, known as the , was first noted by Stokes himself in , when he observed the absence of a steady-state for past a under the same low-Reynolds-number assumptions that succeeded for three-dimensional spheres. This paradox arises because the inertial terms, though small locally, cannot be uniformly neglected in unbounded domains, leading to inconsistencies in the far field. Carl Wilhelm Oseen addressed this in 1910 with his approximation, which partially retains convective inertia to yield a valid for cylindrical flows, though it extends slightly beyond the strict Stokes regime. In the 20th century, Stokes flow theory expanded through tools like the Stokeslet, the fundamental for point forces in viscous fluids, introduced by William Hancock in to model the propulsion of microscopic organisms such as via flagella. Though derived earlier by Oseen in 1927, Hancock's application popularized the Stokeslet for low-Reynolds-number hydrodynamics, enabling simulations of slender-body motions and biological swimming where inertia is negligible. These developments facilitated broader applications in and science, emphasizing the regime's relevance to systems with characteristic Reynolds numbers much less than unity.

Mathematical Formulation

Derivation from Navier-Stokes equations

The incompressible Navier-Stokes equations for a Newtonian fluid, in the absence of body forces, govern the motion of viscous fluids and serve as the starting point for deriving the Stokes equations. These equations consist of the momentum balance \rho \left( \frac{\partial \mathbf{u}}{\partial t} + (\mathbf{u} \cdot \nabla) \mathbf{u} \right) = -\nabla p + \mu \nabla^2 \mathbf{u}, coupled with the incompressibility condition \nabla \cdot \mathbf{u} = 0, where \mathbf{u} is the velocity field, p is the pressure, \rho is the fluid density, and \mu is the dynamic viscosity. To analyze the relative importance of terms, the equations are non-dimensionalized using characteristic velocity scale U and length scale L. Define dimensionless variables as \mathbf{u}^* = \mathbf{u}/U, \nabla^* = L \nabla, t^* = t U / L, and p^* = p / (\rho U^2). Substituting these into the momentum equation yields \frac{\partial \mathbf{u}^*}{\partial t^*} + (\mathbf{u}^* \cdot \nabla^*) \mathbf{u}^* = -\nabla^* p^* + \frac{1}{\mathrm{Re}} \nabla^{*2} \mathbf{u}^*, with \nabla^* \cdot \mathbf{u}^* = 0, where the Reynolds number is \mathrm{Re} = \rho U L / \mu. This non-dimensional form highlights the scaling: the inertial terms (left-hand side) are O(1), the pressure gradient is O(1), and the viscous term is O(1/\mathrm{Re}). In the low Reynolds number limit, \mathrm{Re} \to 0, the inertial terms become negligible compared to the viscous terms, which then balance the pressure gradient. An alternative pressure scaling, p^* = p L / (\mu U), adjusts the equation to \mathrm{Re} \left( \frac{\partial \mathbf{u}^*}{\partial t^*} + (\mathbf{u}^* \cdot \nabla^*) \mathbf{u}^* \right) = -\nabla^* p^* + \nabla^{*2} \mathbf{u}^*, confirming that both the unsteady and convective inertial terms scale with \mathrm{Re} and can be neglected as \mathrm{Re} \ll 1. The pressure term scales as O(1/\mathrm{Re}) in the original form but balances viscous diffusion after rescaling. This yields the dimensionless Stokes equations \nabla^* p^* = \nabla^{*2} \mathbf{u}^* and \nabla^* \cdot \mathbf{u}^* = 0, or in dimensional form, -\nabla p + \mu \nabla^2 \mathbf{u} = 0, \quad \nabla \cdot \mathbf{u} = 0. The approximation holds under the assumptions of incompressibility (constant \rho), Newtonian rheology (linear stress-strain relation with constant \mu), and neglect of body forces (which can be added as a modification, e.g., \mathbf{f}/\rho on the right-hand side).

Stokes equations and boundary conditions

The Stokes equations describe the motion of a viscous, incompressible at very low Reynolds numbers, where inertial effects are negligible. In form, they consist of the for incompressibility and the balance of viscous and forces: \nabla \cdot \mathbf{u} = 0, \nabla p = \mu \nabla^2 \mathbf{u}, where \mathbf{u} is the velocity field, p is the , and \mu is the dynamic viscosity of the . In Cartesian coordinates (x_1, x_2, x_3), the incompressibility condition becomes \frac{\partial u_i}{\partial x_i} = 0 \quad (\text{sum on } i), while the momentum equations take the component form \frac{\partial p}{\partial x_i} = \mu \frac{\partial^2 u_i}{\partial x_j \partial x_j} \quad (\text{sum on } j=1,2,3; \quad i=1,2,3). These equations arise from the divergence-free nature of the velocity field and the of forces in the absence of . Typical boundary conditions for Stokes flow problems include the on solid surfaces, where \mathbf{u} = \mathbf{0} (or \mathbf{u} = \mathbf{U}_S if the surface moves with \mathbf{U}_S). For external flows, such as flow past an isolated object, the far-field condition is \mathbf{u} \to \mathbf{U}_\infty as |\mathbf{x}| \to \infty, ensuring uniformity at large distances. In confined geometries like channels, may be applied to model repeating flow structures. The in Stokes flow, which is essential for computing forces on , is the for a : \sigma_{ij} = -p \delta_{ij} + \mu \left( \frac{\partial u_i}{\partial x_j} + \frac{\partial u_j}{\partial x_i} \right). Traction boundary conditions are then specified as \sigma_{ij} n_j = t_i, where \mathbf{n} is the and \mathbf{t} is the applied surface traction . This formulation allows direct calculation of hydrodynamic forces, such as , by integrating the over the surface. The Stokes equations form an elliptic system of partial differential equations, with the velocity satisfying a Laplace-like equation (\nabla^2 \mathbf{u} = (1/\mu) \nabla p) under the constraint of incompressibility. When reformulated using a stream function \psi in two dimensions (where u = \partial \psi / \partial y, v = -\partial \psi / \partial x), the governing equation becomes the biharmonic equation \nabla^4 \psi = 0, highlighting the diffusive, elliptic character of the flow.

Key Properties

Linearity and superposition principle

The Stokes equations governing low-Reynolds-number flows are linear in the velocity field \mathbf{u} and p, taking the form -\nabla p + \mu \nabla^2 \mathbf{u} = 0 and \nabla \cdot \mathbf{u} = 0, where \mu is the dynamic ; these equations are homogeneous in the interior , with inhomogeneities confined to the conditions. This arises from the neglect of inertial terms in the Navier-Stokes equations, eliminating nonlinear convective acceleration (\mathbf{u} \cdot \nabla) \mathbf{u}. The linearity permits a straightforward of the . Suppose (\mathbf{u}_1, p_1) and (\mathbf{u}_2, p_2) are solutions satisfying the Stokes equations and compatible boundary conditions. Then, for arbitrary scalars \alpha and \beta, the combined fields \mathbf{u} = \alpha \mathbf{u}_1 + \beta \mathbf{u}_2 and p = \alpha p_1 + \beta p_2 also satisfy the equations, as the is linear: substituting yields -\nabla (\alpha p_1 + \beta p_2) + \mu \nabla^2 (\alpha \mathbf{u}_1 + \beta \mathbf{u}_2) = \alpha (-\nabla p_1 + \mu \nabla^2 \mathbf{u}_1) + \beta (-\nabla p_2 + \mu \nabla^2 \mathbf{u}_2) = 0, and \nabla \cdot \mathbf{u} = \alpha \nabla \cdot \mathbf{u}_1 + \beta \nabla \cdot \mathbf{u}_2 = 0. Superposition of boundary conditions holds similarly if they are linear. A key implication of this is the behavior of hydrodynamic quantities: forces and stresses in Stokes flow are proportional to \mu U L, where U is a and L a , reflecting viscous dominance; notably, the \rho does not appear in the equations, rendering solutions independent of inertial effects. This scalability simplifies analysis for varying viscosities or speeds while preserving flow patterns. The enables the construction of solutions for complex configurations by adding elementary solutions, such as for flow past multiple bodies or distributed forces, exemplified by arrays of particles where individual disturbances are summed to approximate collective motion. In low-Reynolds-number hydrodynamics, this facilitates multipole expansions, where the far-field flow around a particle is represented as a series of point-force singularities (Stokeslets) and higher-order terms, allowing efficient of interactions in dilute suspensions. In biological applications, superposition aids modeling of multi-particle systems at low Reynolds numbers, such as cell aggregation in viscous , where hydrodynamic interactions between cells are approximated by superposing flows induced by individual motile elements like flagella or cilia.

Time-reversibility

Stokes flow exhibits time-reversibility due to the neglect of inertial effects in the Navier-Stokes equations at low Reynolds numbers, where the nonlinear convective term (\mathbf{u} \cdot \nabla)\mathbf{u} is absent. This term, which changes sign under time reversal in full Navier-Stokes dynamics, is responsible for the irreversibility observed in higher-Reynolds-number flows, such as , where fluid particle paths do not retrace themselves upon reversal. In contrast, the linear Stokes equations balance gradients and viscous forces without such asymmetry, preserving under temporal inversion. The time-reversal invariance can be demonstrated directly from the Stokes equations. Consider the steady incompressible form: \nabla p = \mu \nabla^2 \mathbf{u}, \quad \nabla \cdot \mathbf{u} = 0, with suitable boundary conditions. If \mathbf{u}(\mathbf{x}) and p(\mathbf{x}) satisfy these equations, then substituting \mathbf{u}'(\mathbf{x}) = -\mathbf{u}(\mathbf{x}) and p'(\mathbf{x}) = -p(\mathbf{x}) yields \nabla p' = \mu \nabla^2 \mathbf{u}' and \nabla \cdot \mathbf{u}' = 0, confirming that the transformed fields also solve the system. For time-dependent cases under the quasi-steady Stokes approximation with time-varying boundaries, reversing the temporal sequence of boundary conditions produces the negated velocity field, restoring the initial configuration. This property stems from the nature of the Stokes operator in appropriate function spaces, which ensures the linearity supports such symmetries without dissipative memory effects beyond . Physically, time-reversibility implies that particle trajectories in Stokes flow are reversible, meaning particles retrace their paths exactly when time is inverted, in stark contrast to the chaotic, non-retracing paths in turbulent regimes. A representative example is the of particle clusters in a viscous , where, absent , particles follow closed or periodic paths—such as rotating equilateral triangles for three particles—due to the reversible interactions mediated by the Stokeslet fundamental solution. This reversibility underscores challenges in microscopic locomotion, as symmetric motions (e.g., a scallop-like opening and closing) yield no net displacement, a central to understanding low-Reynolds-number . However, this reversibility holds strictly for deterministic, steady or quasi-steady flows; in unsteady scenarios with significant temporal variation, higher-order inertial corrections may emerge. Moreover, at microscopic scales, introduces stochastic fluctuations that break time-reversal , rendering paths irreversible through , an aspect often overlooked in classical deterministic analyses of Stokes flow.

Stokes paradox

The Stokes paradox arises in the context of two-dimensional steady creeping past an infinite , where no exists that satisfies both the no-slip boundary on the and a at . This paradox manifests because the in such flows exhibits a logarithmic with distance from the , leading to a that prevents the from approaching a far-field . The mathematical origin of the paradox lies in the fundamental of the two-dimensional Stokes equations, known as the Stokeslet, which possesses a logarithmic singularity at large distances. This slow decay, characterized by terms proportional to \log r where r is the radial distance, is incompatible with the requirement of flow as r \to \infty, as the perturbation from the obstacle does not diminish sufficiently to match the prescribed boundary condition at . In contrast, three-dimensional Stokes flow past a admits a well-posed , with the velocity disturbance decaying as $1/r, allowing the far-field to recover the without issue. The paradox was first noted by George Gabriel Stokes in 1851 during his analysis of viscous drag on cylinders. A key resolution came from Carl Wilhelm Oseen in 1910, who addressed the issue by incorporating a perturbation of the inertial terms in the Navier-Stokes equations, yielding a drag force correction of order \log(\mathrm{Re}) where \mathrm{Re} is the , though this approximation extends beyond the pure Stokes regime. These historical developments highlighted the limitations of the Stokes approximation for unbounded two-dimensional domains. The implications of the Stokes paradox restrict the applicability of pure Stokes flow to finite domains or three-dimensional geometries, necessitating approximations like Oseen's linearization for practical two-dimensional problems at low but finite Reynolds numbers. In modern computational approaches, such as boundary element methods or Boltzmann simulations, the paradox is circumvented by imposing artificial finite boundaries, which effectively resolve the flow while approximating infinite-domain conditions through extrapolation.

Analytical Solution Methods

Stream function approach

In two-dimensional incompressible Stokes flow, the velocity field can be represented using a scalar \psi(x, y) defined such that the velocity components are u = \frac{\partial \psi}{\partial y} and v = -\frac{\partial \psi}{\partial x}. This formulation inherently satisfies the \nabla \cdot \mathbf{u} = 0, as the divergence of the curl is zero. Substituting this representation into the Stokes momentum equations and eliminating the pressure by taking the curl yields a single governing equation for \psi: the \nabla^4 \psi = 0. The vorticity \omega = -\nabla^2 \psi then satisfies \nabla^2 \omega = 0, linking the directly to the rotational aspects of the flow. Analytical solutions to the can be obtained using in simple geometries, such as rectangular or circular domains, or via expansions for periodic or channel-like configurations. For numerical implementation, methods discretize the biharmonic operator on structured grids, often decoupling it into coupled equations for \psi and \omega to improve efficiency and . The stream function approach offers key advantages, including the reduction of the vector-valued Stokes problem to a scalar equation, which simplifies both analytical manipulations and numerical schemes by avoiding direct handling of pressure. Additionally, the relation \omega = -\nabla^2 \psi facilitates the incorporation of vorticity transport in extensions to slightly higher Reynolds numbers. However, this method is inherently limited to two-dimensional flows; for axisymmetric three-dimensional cases, an extension known as the Stokes stream function is employed, which satisfies a biharmonic-like equation in spherical or cylindrical coordinates. As an illustrative outline without full derivation, consider plane Poiseuille flow in a channel driven by a pressure gradient: the stream function \psi takes a cubic polynomial form in the cross-channel direction, satisfying no-slip boundaries at the walls, though detailed solutions are addressed in specific geometry sections.

Green's function and the Stokeslet

In Stokes flow, the Green's function represents the velocity field induced by a point force and serves as the fundamental building block for constructing solutions to more complex problems. Known as the Stokeslet, it describes the flow due to a delta-function force \mathbf{F} applied at the origin in an unbounded viscous fluid. The velocity \mathbf{u} at position \mathbf{r} is given by u_i(\mathbf{r}) = \frac{1}{8\pi\mu r} \left( \delta_{ij} + \frac{r_i r_j}{r^2} \right) F_j, where \mu is the dynamic viscosity, r = |\mathbf{r}|, \delta_{ij} is the Kronecker delta, and summation over repeated indices is implied. This expression, also referred to as the Oseen tensor in the limit of zero Reynolds number, was first derived by Oseen in 1927. The term "Stokeslet" was coined by Hancock in 1953 to denote this singular solution. The Stokeslet arises from solving the Stokes equations with a singular forcing term: \mu \nabla^2 \mathbf{u} - \nabla p = -\mathbf{F} \delta(\mathbf{x}), subject to the incompressibility condition \nabla \cdot \mathbf{u} = 0. A standard approach to derive it involves applying the to the linearized equations, solving for the transformed , and inverting the transform to obtain the spatial form, which yields the above tensor after imposing incompressibility. This fundamental solution satisfies the Stokes equations everywhere except at the origin, where the singularity corresponds to the point force. Key properties of the Stokeslet include its decay as $1/r far from the source in three dimensions, ensuring finite energy dissipation despite the singularity. It forms the basis for higher-order singularities, such as the rotlet (associated with a point torque) and the stresslet (related to a force dipole), which are essential for representing multipole expansions in Stokes flows. The Stokeslet enables integral representations of velocity fields by distributing point forces over volumes or surfaces, facilitating solutions in bounded domains via the . For instance, Blake (1971) developed an image system for a Stokeslet near a no-slip plane wall, consisting of an opposing Stokeslet, a stresslet, and a source doublet to enforce the boundary condition. Similarly, Hancock (1953) applied distributions of Stokeslets along slender filaments to model , a technique later extended by Chwang and Wu (1975) to systematize singularity methods for arbitrary low-Reynolds-number configurations. In biological contexts, such as flagellar of microorganisms, Stokeslet distributions capture the hydrodynamic interactions driving , as demonstrated in early models of sea-urchin spermatozoa.

Papkovich–Neuber representation

The Papkovich–Neuber representation offers a general integral-free method to express solutions to the three-dimensional Stokes equations using harmonic potentials, leveraging the analogy between incompressible Stokes flow and linear elasticity at Poisson's ratio ν = 0.5. In elasticity, the displacement field is represented as \mathbf{u} = \mathbf{B} - \frac{1}{4(1-\nu)} \nabla (\mathbf{x} \cdot \mathbf{B} + \phi), where \mathbf{B} and \phi are vector and scalar harmonic functions satisfying \nabla^2 \mathbf{B} = \mathbf{0} and \nabla^2 \phi = 0. For Stokes flow, substituting ν = 0.5 yields the adapted form for the velocity field:
u_i = \chi_i - \frac{\partial}{\partial x_i} \left( \frac{x_j \chi_j}{2} + \frac{\phi}{4} \right),
where \chi is a harmonic vector potential (\nabla^2 \chi = \mathbf{0}) and \phi is a scalar harmonic potential (\nabla^2 \phi = 0). The corresponding pressure is p = -\mu \nabla \cdot \chi. This form ensures the velocity is divergence-free and satisfies the Stokes momentum equation.
The representation decouples the coupled Stokes system into independent Laplace equations for the potentials, simplifying the solution process by reducing it to problems for functions. This decoupling is particularly advantageous for exterior domains, such as flows around immersed bodies, where the potentials decay at infinity and conditions on can be translated to conditions on \chi and \phi. It also connects to the by expressing the solenoidal as a of irrotational and components. Historically, the Papkovich–Neuber solution was introduced independently by Papkovich in 1940 and Neuber in 1940 for three-dimensional elasticity problems. Its adaptation to Stokes flow exploits the structural similarity between the biharmonic equation in elasticity and the vector Laplace equation in creeping flow, as detailed in subsequent works on low-Reynolds-number hydrodynamics. In applications, the representation facilitates analytical solutions for flows around arbitrary bodies by prescribing the potentials' values and normal derivatives on the body surface to match no-slip conditions. For instance, it has been employed to solve exterior problems in spheroidal coordinates, enabling explicit series expansions for translating or rotating particles. Computationally, it integrates well with potential theory-based codes, such as those using fast multipole methods for harmonic solvers, to handle complex geometries without singular kernels like the Stokeslet.

Boundary integral methods

Boundary integral methods for Stokes flow represent the and fields as integrals over the bounding surfaces, leveraging the fundamental solution of the Stokes equations to reduce the dimensionality of the problem from volume to surface . This approach is particularly suited for exterior flows around immersed objects, where the infinite domain is handled naturally without artificial boundaries. The core of the method is the boundary integral representation formula, derived from the applied to the Stokes system. For a point \mathbf{x} in the fluid domain, the \mathbf{u}(\mathbf{x}) is given by \mathbf{u}(\mathbf{x}) = \int_S \mathbf{U}(\mathbf{x}, \mathbf{y}) \cdot \mathbf{t}(\mathbf{y}) \, dS(\mathbf{y}) - \int_S \mathbf{T}(\mathbf{x}, \mathbf{y}) : (\mathbf{u}(\mathbf{y}) \otimes \mathbf{n}(\mathbf{y})) \, dS(\mathbf{y}), where S is the surface enclosing the solid, \mathbf{n} is the outward unit normal, \mathbf{U} is the Stokeslet tensor (the velocity due to a point force), \mathbf{T} is the associated stress tensor, and \mathbf{t} = \boldsymbol{\sigma} \cdot \mathbf{n} is the traction on the surface. The Stokeslet U_{jk}(\mathbf{r}) = \frac{1}{8\pi\mu} \left( \delta_{jk} + \frac{r_j r_k}{r^2} \right) / r and stress tensor T_{kjl}(\mathbf{r}) = -\frac{3}{8\pi\mu} \frac{r_k r_j r_l}{r^3}, with \mathbf{r} = \mathbf{x} - \mathbf{y} and \mu the viscosity, capture the singular behavior at the source point. When \mathbf{x} approaches the boundary, the formula requires a jump relation, leading to a Fredholm integral equation of the second kind for the unknown boundary data (typically velocity for no-slip conditions). The method employs single-layer and double-layer potentials as building blocks. The single-layer potential, with density corresponding to a fictitious traction distribution, generates a velocity field \int_S \mathbf{U}(\mathbf{x}, \mathbf{y}) \cdot \boldsymbol{\mu}(\mathbf{y}) \, dS(\mathbf{y}), which is continuous across the surface but has a discontinuous normal derivative. The double-layer potential, \int_S \mathbf{u}(\mathbf{y}) \cdot \mathbf{T}(\mathbf{x}, \mathbf{y}) \cdot \mathbf{n}(\mathbf{y}) \, dS(\mathbf{y}), produces a jump in the velocity equal to the density difference, making it ideal for representing no-slip boundaries where the interior velocity is zero. Combinations of these potentials form the integral equations solved numerically. For instance, the standard representation for exterior no-slip problems yields \mathbf{u}(\mathbf{x}) = -\int_S \mathbf{u}(\mathbf{y}) \cdot \mathbf{T}(\mathbf{x}, \mathbf{y}) \cdot \mathbf{n}(\mathbf{y}) \, dS(\mathbf{y}) for \mathbf{x} on S (up to a solid angle factor), reducing the problem to solving for surface velocities. Numerically, the surface is meshed into panels (e.g., triangles or quadrilaterals), and the integrals are discretized using basis functions for the unknowns, leading to a dense . Collocation methods evaluate the equation at points matching the nodes, simplifying implementation but potentially less accurate for ill-conditioned kernels; Galerkin methods project onto the same space, improving symmetry and convergence for second-kind equations. in the kernels, arising when \mathbf{x} and \mathbf{y} coincide, are handled through regularization techniques such as singularity subtraction, analytical integration over flat panels, or specialized rules like Gaussian for weakly singular parts and polar for hypersingular ones. These ensure accurate evaluation near the diagonal, with error controlled to O(h^2) or higher, where h is the size. The primary advantages lie in the reduction of computational domain: for exterior problems, only the object surface needs meshing, avoiding volume grids in unbounded regions and facilitating simulations of complex geometries like particles or biological structures. This dimensionality reduction lowers significantly compared to finite element or volume methods, with scaling O(N^2) for N panels, mitigated by fast multipole accelerations that achieve O(N) or O(N \log N) for large systems via hierarchical expansions of the Stokeslet. Historical development began in the late with early boundary element applications to viscous flows, gaining momentum in the through works like those of Tran-Cong and Phan-Thien, who formulated direct methods for multiparticle Stokes problems. Modern implementations often incorporate fast methods, as in the 1990s extensions by Phan-Thien for elasticity analogs, and are available in open-source libraries like Bempp, which supports Stokes formulations via interfaces for applications.

Solutions in Specific Geometries

Uniform flow past a

The uniform flow past a fixed of radius a in an unbounded viscous fluid of dynamic \mu, with undisturbed far-field \mathbf{u} = U_\infty \mathbf{e}_z, and no-slip boundary condition on the surface, represents a canonical solvable problem in Stokes flow. This setup assumes axisymmetry about the z-axis and neglects inertial effects, valid at low Reynolds numbers \mathrm{Re} = \frac{2 a U_\infty \rho}{\mu} \ll 1, where \rho is the fluid density. The analytical solution employs the \psi(r, \theta) in spherical coordinates, where r is the radial distance from the sphere center and \theta is the polar angle from the z-axis. The stream function takes the form of a and is given by \psi = \frac{1}{4} U_\infty \left(2 r^2 - 3 a r + \frac{a^3}{r}\right) \sin^2 \theta. This expression satisfies the \nabla^4 \psi = 0 governing axisymmetric Stokes flow. The corresponding velocity components are u_r = U_\infty \cos \theta \left( 1 - \frac{3a}{2r} + \frac{a^3}{2r^3} \right), u_\theta = -U_\infty \sin \theta \left( 1 - \frac{3a}{4r} - \frac{a^3}{4r^3} \right). These components derive from u_r = \frac{1}{r^2 \sin \theta} \frac{\partial \psi}{\partial \theta} and u_\theta = -\frac{1}{r \sin \theta} \frac{\partial \psi}{\partial r}. At the sphere surface (r = a), u_r = u_\theta = 0, enforcing no-slip; as r \to \infty, \mathbf{u} \to U_\infty \mathbf{e}_z; and the fields satisfy the Stokes equations \nabla p = \mu \nabla^2 \mathbf{u} and \nabla \cdot \mathbf{u} = 0 everywhere outside the sphere. The associated pressure field is p = p_\infty - \frac{3 \mu U_\infty a \cos \theta}{2 r^2}. The total hydrodynamic force on the sphere, obtained by integrating the normal and shear stresses over the surface, yields a drag \mathbf{F}_D = 6 \pi \mu a U_\infty \mathbf{e}_z with no lift due to fore-aft symmetry. This result, known as , arises from the viscous dissipation in the flow. This solution underpins applications such as the sedimentation of small spherical particles, where gravitational settling balances drag to give a terminal velocity U_t = \frac{2 a^2 g (\rho_p - \rho_f)}{9 \mu} for particle density \rho_p > \rho_f. Near bounding walls, Faxén's laws introduce corrections to the force and torque; for instance, the drag on a sphere moving parallel to a plane wall at distance h \gg a increases by a factor $1 + \frac{9a}{16 h} + O\left( (a/h)^3 \right).

Hele-Shaw flow

Hele-Shaw flow describes the slow, viscous motion of an incompressible confined between two closely spaced parallel plates separated by a small fixed gap of height h, where the gap width is much smaller than the lateral dimensions of the flow domain. This configuration approximates under the lubrication approximation, neglecting inertial effects and assuming a low regime. The no-slip boundary conditions on the plates induce a parabolic velocity profile across the gap, with the velocity varying quadratically from zero at the walls to a maximum at the midplane. Depth-averaging the over the yields an effective two-dimensional , where the averaged \mathbf{u} satisfies a Darcy-like relation \mathbf{u} = -\frac{h^2}{12\mu} \nabla p, with \mu the fluid and p the . Incompressibility \nabla \cdot \mathbf{u} = 0 then implies that the is , \nabla^2 p = 0, making the flow mathematically equivalent to two-dimensional for an inviscid, irrotational fluid. This analogy allows streamlines and equipotentials to be interchanged, facilitating analytical solutions via complex potentials. Unlike the full two-dimensional Stokes equations, which suffer from the Stokes paradox—no bounded steady solution exists for uniform flow past a due to logarithmic velocity growth at —the Hele-Shaw approximation resolves this issue through the transverse viscous friction from the finite gap width, effectively introducing a that regularizes the flow. However, this model ignores along the plates' lateral boundaries and assumes the gap is uniformly thin, limiting its validity to scenarios where variations in the flow direction parallel to the plates dominate. A prominent application involves the evolution of free interfaces in Hele-Shaw cells, such as when a less viscous fluid displaces a more viscous one, leading to the Saffman-Taylor viscous fingering instability. Here, perturbations at the interface grow exponentially, forming branched finger-like patterns whose width is often half the channel width in the absence of surface tension, providing an analog to interfacial instabilities in ideal fluid dynamics. For a moving interface, the normal velocity follows from the kinematic condition, with pressure jumps across the interface incorporating surface tension effects. In modern contexts, Hele-Shaw flows in microfluidic devices enable precise control of shear stresses on cells and facilitate patterning applications, such as droplet manipulation and soft lithography.

Slender-body theory

Slender-body theory provides an asymptotic approximation for the slow viscous flow around thin, elongated bodies, such as rods or filaments, where the body length L greatly exceeds its radius a, defining the small parameter \epsilon = a/L \ll 1. This approach exploits the slenderness to simplify the Stokes equations, treating the body as a line distribution of singular forces while matching inner and outer expansions to capture the flow field. At leading order, the local drag force per unit length \mathbf{f}(s) on the slender body is approximated by resistive force theory, which assumes the drag depends linearly on the local velocity \mathbf{U}(s) and decomposes into and perpendicular components relative to the body's . For motion to the local , f_\parallel(s) \approx \frac{2\pi \mu U_\parallel(s)}{\ln(2L/a)}, while for perpendicular motion, f_\perp(s) \approx \frac{4\pi \mu U_\perp(s)}{\ln(2L/a)}, where \mu is the fluid and the logarithmic factor arises from the long-range nature of the Stokeslet interactions. These coefficients stem from early analyses of propulsion and have been refined for arbitrary cross-sections. The velocity field \mathbf{u}(\mathbf{x}) induced by the slender body is then represented as a of Stokeslets along the centerline \mathbf{y}(s): \mathbf{u}(\mathbf{x}) = \frac{1}{8\pi\mu} \int_{-L/2}^{L/2} \mathbf{f}(s) \cdot \mathbf{G}(\mathbf{x} - \mathbf{y}(s)) \, ds, where \mathbf{G} is the Stokeslet tensor, the fundamental solution to the Stokes equations. The force distribution \mathbf{f}(s) is determined by solving an enforcing the on the body surface, approximated to leading order in \epsilon. This formulation yields a resistance matrix relating total force and torque to the body's velocity and , enabling computation of for rigid or flexible filaments. A prominent application is the of bacterial flagella, where applied these approximations to model the helical motion of microscopic organisms through viscous fluids, predicting speeds based on flagellar and drag anisotropy. Higher-order corrections in slender-body theory account for effects like body and end singularities, improving accuracy for non-straight or finite-length filaments by including O(\epsilon) terms in the force expansion.

Flows in cylindrical and spherical coordinates

In axisymmetric Stokes flow, the velocity field can be derived from a \psi(r, \theta) in spherical coordinates (r, \theta), which satisfies the E^4 \psi = 0, where the operator E^2 is given by E^2 = \frac{\partial^2}{\partial r^2} + \frac{\sin \theta}{r^2} \frac{\partial}{\partial \theta} \left( \frac{1}{\sin \theta} \frac{\partial}{\partial \theta} \right). This equation arises from the curl of the Stokes equations, eliminating and yielding a fourth-order PDE for the stream function. assumes \psi(r, \theta) = f(r) g(\theta), leading to ordinary differential equations; the angular part involves Gegenbauer functions C_n^{(3/2)}(\cos \theta) as eigenfunctions of the operator, while the radial part yields solutions of the form r^{n+1} and r^{-n} for n = 1, 2, \dots. These separated solutions form a complete basis for expanding the stream function in bounded or unbounded domains, facilitating analytical or semi-analytical treatments of axisymmetric geometries beyond uniform flows. In cylindrical coordinates (\rho, \phi, z), solutions for Stokes flow in infinite or semi-infinite domains often employ expansions in the axial direction combined with in the radial direction to satisfy boundary conditions at and on surfaces. For example, in an unbounded domain, the velocity components can be expressed as integrals over wave numbers, with radial dependence via modified I_n(k\rho) and K_n(k\rho) to ensure decay at large \rho. This approach is particularly useful for axisymmetric or helical flows, where the general solution decomposes into toroidal and poloidal modes. For flow past an infinite , the Stokes paradox prevents a uniform far-field solution, but approximations via matched asymptotic expansions in the near and far fields resolve this by incorporating weak inertial effects or domain finiteness. In finite cylindrical domains, such as or containers, the is avoided due to the bounded , allowing exact series solutions using sine/cosine series in z and Fourier-Bessel series in \rho, with coefficients determined from conditions on the walls and ends. These expansions capture three-dimensional effects, including secondary flows and corner singularities, and converge rapidly for aspect ratios near unity. Such methods are applied analytically to pressure-driven flows in or orifices, where the series link to approximations for slender geometries by averaging over cross-sections to yield effective one-dimensional models. A representative example is the Jeffery-Hamel flow in a of $2\alpha, which admits an exact similarity for steady radial Stokes flow, with u_r = \gamma g(\theta)/r and u_\theta = 0, where g(\theta) solves a nonlinear reducible to linear in the low-Reynolds-number limit. Solutions exist and are unique for sufficiently small \gamma, with pure outflow stable for acute wedges. This provides a for viscous flows in converging/diverging channels. Another example is the sequence of Moffatt eddies near a sharp corner formed by two rigid walls at $2\alpha < 146^\circ, where an infinite of counter-rotating eddies arises, with successive eddy strengths decaying geometrically by factors m_1 \approx 2000 at $90^\circ and sizes by p_1 \approx 10, driven by weak stirring far from the corner. These eddies highlight singular behavior in Stokes flow at acute s. Applications include modeling blood flow in capillaries, where Stokes equations govern the low-Reynolds-number motion of deformable red blood cells (RBCs) in tubes of diameter 4–10 \mum, with viscosity \mu \approx 0.001 Pa·s. RBCs occupy nearly the full cross-section, leading to nonlinear resistance via the Fåhraeus-Lindqvist effect, where drops with decreasing tube radius due to cell deformation and plasma layering; two-phase models predict entry pressures up to 1000 dyn/cm² for RBC trains. The endothelial glycocalyx adds hydraulic resistance, modeled as a porous layer increasing flow impedance by factors of 10–100.

Associated Theorems

Stokes-Helmholtz decomposition

The Stokes-Helmholtz decomposition adapts the classical Helmholtz theorem from vector calculus to the context of Stokes flow, expressing the velocity field \mathbf{u} as the sum of an irrotational component and a solenoidal component: \mathbf{u} = \nabla \phi + \nabla \times \mathbf{A}, where \phi is a scalar potential and \mathbf{A} is a divergence-free vector potential (\nabla \cdot \mathbf{A} = 0). This decomposition holds for sufficiently smooth vector fields in simply connected domains, with uniqueness ensured under suitable decay conditions at infinity for exterior problems. In the Stokes regime, where the flow is governed by the linear equations -\nabla p + \mu \nabla^2 \mathbf{u} = 0 and \nabla \cdot \mathbf{u} = 0, the incompressibility condition implies \nabla^2 \phi = 0, making \phi a harmonic function. The solenoidal part \nabla \times \mathbf{A} then satisfies a modified equation derived by substituting into the momentum balance, leading to \nabla p = \mu \nabla^2 (\nabla \times \mathbf{A}). Since the irrotational part contributes nothing to the pressure gradient (as \nabla^2 (\nabla \phi) = \nabla (\nabla^2 \phi) = 0), the pressure is fully determined by the rotational component. This decomposition facilitates the solvability of Stokes flow by decoupling the irrotational and rotational contributions, enabling separate solution of problems for the potentials subject to conditions. In exterior Dirichlet problems, where the is prescribed on a bounded obstacle and approaches a uniform flow at infinity, the solution is unique up to additive constants in the potentials, reflecting the gauge freedom in \mathbf{A} and the fact that constants do not affect the gradients. The irrotational part \nabla \phi can incorporate the uniform far-field flow, which is itself both irrotational and solenoidal, while the rotational part adjusts to enforce no-slip conditions on the . This approach links directly to the Papkovich–Neuber representation (see Analytical Solution Methods), where potentials are used to construct biharmonic fields satisfying the Stokes equations, with pressure given by p = \nabla \cdot \boldsymbol{\Phi} for a \boldsymbol{\Phi}. Numerical methods often exploit this decomposition to improve stability and efficiency by projecting onto solenoidal spaces. Historically, George Gabriel Stokes employed an analogous decomposition in his 1851 analysis of uniform flow past a , using scalar and vector potentials to derive the exact solution, marking an early application to low-Reynolds-number hydrodynamics. The method underscores the role of harmonic functions in representing decaying , which works well in three dimensions where multipole expansions converge. However, in two dimensions, the decomposition reveals the Stokes paradox: harmonic functions like the fundamental solution \ln r grow logarithmically at , preventing a uniform far-field flow from being matched with a decaying perturbation that satisfies no-slip on a , as no such solution exists in standard spaces. This dimensionality issue arises because 2D exterior domains lack the rapid decay of 3D harmonics (e.g., $1/r), leading to inconsistencies in the and necessitating higher-order approximations like Oseen's equations.

Lorentz reciprocal theorem

The Lorentz reciprocal theorem, introduced by Hendrik Antoon Lorentz in to analyze the of particles in viscous fluids, is a fundamental identity in Stokes flow that relates the and velocity fields of two distinct solutions to the Stokes equations over the same domain. For two incompressible Stokes flows denoted by velocity fields \mathbf{u}^a, \mathbf{u}^b and corresponding tensors \boldsymbol{\sigma}^a = -p^a \mathbf{I} + 2\mu \mathbf{e}^a, \boldsymbol{\sigma}^b = -p^b \mathbf{I} + 2\mu \mathbf{e}^b (where \mathbf{e} is the rate-of-strain tensor and \mu is the viscosity), the theorem states that the surface integral over a closed boundary S enclosing the fluid domain vanishes in the absence of body forces: \int_S \left( \boldsymbol{\sigma}^a \cdot \mathbf{n} \cdot \mathbf{u}^b - \boldsymbol{\sigma}^b \cdot \mathbf{n} \cdot \mathbf{u}^a \right) \, dS = 0, where \mathbf{n} is the outward unit normal. This reciprocity arises from the self-adjoint nature of the Stokes operator and holds for arbitrary boundary conditions on S, provided both flows satisfy the incompressibility condition \nabla \cdot \mathbf{u} = 0 and the momentum balance \nabla \cdot \boldsymbol{\sigma} = 0. The derivation exploits the linearity of the Stokes equations and the symmetry of the stress tensor. Consider the volume V bounded by S; the difference \boldsymbol{\sigma}^a : \nabla \mathbf{u}^b - \boldsymbol{\sigma}^b : \nabla \mathbf{u}^a is divergence-free due to the momentum equations and incompressibility, as \nabla \cdot (\boldsymbol{\sigma}^a \cdot \mathbf{u}^b) = \mathbf{u}^b \cdot (\nabla \cdot \boldsymbol{\sigma}^a) + \boldsymbol{\sigma}^a : \nabla \mathbf{u}^b = \boldsymbol{\sigma}^a : \mathbf{e}^b (and similarly for the other term), with equality following from the symmetry \boldsymbol{\sigma}^a : \mathbf{e}^b = \boldsymbol{\sigma}^b : \mathbf{e}^a = 2\mu \mathbf{e}^a : \mathbf{e}^b. Applying the divergence theorem then yields the surface integral form directly. If body forces \mathbf{f}^a, \mathbf{f}^b are present, the theorem generalizes to include volume integrals \int_V (\mathbf{u}^b \cdot \mathbf{f}^a - \mathbf{u}^a \cdot \mathbf{f}^b) \, dV = 0 on the left-hand side. A primary application is computing hydrodynamic forces and couples on a body by interchanging the roles of the two flows. For instance, the drag force \mathbf{F}^a on body a translating with velocity \mathbf{U}^a in quiescent fluid can be obtained from the known uniform flow \mathbf{u}^b = \mathbf{U}^b past the same body, yielding \mathbf{F}^a \cdot \mathbf{U}^b = \mathbf{F}^b \cdot \mathbf{U}^a, where \mathbf{F}^b = 6\pi \mu a \mathbf{U}^b for a sphere of radius a. This reciprocity simplifies calculations for sedimentation or drag in non-uniform backgrounds without solving the full boundary-value problem. The theorem extends naturally to couples \mathbf{L}^a \cdot \boldsymbol{\omega}^b = \mathbf{L}^b \cdot \boldsymbol{\omega}^a (with \boldsymbol{\omega} the angular velocity) and higher moments like stresslets, facilitating analysis of rotational and deformational effects. In multiparticle systems, the reciprocal theorem underpins methods for hydrodynamic interactions, such as the method of reflections, where forces on one particle are related to velocities induced on others, enabling efficient computation of collective dynamics in dilute suspensions. It has proven particularly valuable in colloidal assembly, where reciprocity links phoretic flows around active particles to emergent self-propulsion in clusters; for example, in dense colloidal aggregates, the theorem reveals how hydrodynamic coupling switches between phoretic and osmotic propulsion mechanisms, driving in non-equilibrium systems.

Faxén's laws

Faxén's laws provide expressions for the hydrodynamic force and torque exerted on a rigid, no-slip sphere of radius a immersed in an arbitrary, slowly varying Stokes flow, accounting for the effects of flow gradients beyond the uniform incident flow approximation. These laws extend the classical Stokes drag by incorporating corrections due to the spatial variation of the undisturbed incident velocity field \mathbf{u}^\infty. Originally derived by Hilding Faxén in 1922, they are fundamental for analyzing particle dynamics in low-Reynolds-number regimes where inertia is negligible. The force \mathbf{F} on the sphere, assuming it is held fixed at the origin with no , is given by \mathbf{F} = 6\pi \mu a \left( \mathbf{u}^\infty + \frac{a^2}{6} \nabla^2 \mathbf{u}^\infty \right) \bigg|_{\mathbf{x}=0}, where \mu is the fluid viscosity and the terms are evaluated at the sphere's center. The leading term $6\pi \mu a \mathbf{u}^\infty recovers the uniform Stokes drag, while the second term \pi \mu a^3 \nabla^2 \mathbf{u}^\infty represents a correction arising from the flow's Laplacian, which vanishes for uniform flows but becomes significant in non-uniform fields. For the \mathbf{T}, required to prevent rotation of the fixed sphere, \mathbf{T} = 8\pi \mu a^3 \boldsymbol{\omega}^\infty \bigg|_{\mathbf{x}=0}, where \boldsymbol{\omega}^\infty = \frac{1}{2} \nabla \times \mathbf{u}^\infty is the of the incident ; this term ensures zero in irrotational flows. These expressions assume the incident satisfies the Stokes equations far from the . The derivation of Faxén's laws employs the Lorentz reciprocal theorem, which relates the stresses and velocities in two Stokes flows, applied here to an auxiliary problem of uniform flow past the sphere. By integrating the reciprocal identity over the sphere's surface and invoking the mean-value property of functions (since the velocity satisfies biharmonic conditions in Stokes flow), the force and torque emerge as averages of the incident flow and its derivatives over the sphere's volume, reducing to center-point evaluations for spheres. This approach highlights the no-slip boundary condition's role in coupling the particle to higher-order flow moments. The Laplacian correction in the force law captures the sphere's response to flow curvature, effectively representing a Faxén force that adjusts the classical for non-uniformity, with implications for enhanced or reduced drag depending on the flow's convexity. In practice, these laws simplify computations in multiparticle suspensions by avoiding full boundary integral solutions for weakly non-uniform ambient flows. Applications include of charged particles in varying electric fields, where the correction influences mobility, and manipulating microspheres in gradient flows for precise positioning. In modern contexts, such as systems, modified Faxén laws describe ' interactions in vortical or sheared environments, aiding models of bacterial swarms or synthetic microswimmers.

References

  1. [1]
    Stokes Flow - an overview | ScienceDirect Topics
    Stokes flow refers to the flow of viscous fluids at low Reynolds numbers, characterized by the linearity of the Stokes equation, which allows for analytical ...
  2. [2]
    Basic Concepts of Stokes Flows - SpringerLink
    Sep 26, 2019 · In this chapter, we discuss some fundamental properties of Stokes flows, namely: negligibility of inertial forces, reversibility and the minimum energy ...
  3. [3]
    (PDF) Basic Concepts of Stokes Flows - ResearchGate
    Various properties essential to the understanding of Stokes flow are discussed, including reversibility, negligibility of inertial forces and minimum energy ...
  4. [4]
    [PDF] Stokes's Fundamental Contributions to Fluid Dynamics
    Jul 12, 2019 · The fluid motion known as Stokes's Second Problem is the flow around an infinite flat plate that moves sinusoidally in its own plane. There ...
  5. [5]
    [PDF] on the Motion of Pendulums. By G. G. Stokes, M.A., Fellow of ...
    On the Effect of the Internal Friction of Fluids on the Motion of Pendulums. of Mathematics in the University ofCambridge. [Read December 9, 1850.]
  6. [6]
    Viscous-Dominated Flows – Introduction to Aerospace Flight Vehicles
    Viscous-dominated flows, often referred to as “creeping” or “Stokes” flows, occur when viscous forces rather than inertial forces predominantly govern fluid ...
  7. [7]
    [PDF] Stokes flow
    Thus we have solved the Stokes flow problem of a sphere spinning in an infinite expanse of viscous fluid. It is not difficult to see, however, that (7.3) does ...
  8. [8]
    Fluid flows and forces in development: functions, features and ...
    Stokes flow (or creeping flow). A type of fluid flow in which viscous forces dominate the advective inertial forces. It occurs at a Reynolds number less ...
  9. [9]
    [PDF] Stokes Flow Around Beating Microorganisms - VCU Scholars ...
    This model has applications in multiple disciplines, including medicine, biology, technology, and engineering. Specifically, the modeling of Stokes flow ...
  10. [10]
    [PDF] Analysis of the Immersed Boundary Method for Stokes Flow
    It has since proven to be a general and robust method for diverse flow problems arising in biology and engineering. Recently, the immersed boundary method has ...
  11. [11]
    The unsteady motion of solid bodies in creeping flows
    Apr 26, 2006 · In treating unsteady particle motions in creeping flows, a quasi-steady approximation is often used, which assumes that the particle's ...
  12. [12]
    A Versatile Flow-Profile Engineering Method in the Stokes Flow ...
    In microfluidic channels, the flow is usually in the low-Reynolds-number laminar state due to the small dimension and the often highly viscous working fluids.
  13. [13]
    I. Specific gravity of pollen grains - ScienceDirect
    Adequate application of Stokes' law to the problem of volumetric pollen content of the air, even for fairly spherical pollen grains, requires reasonably ...
  14. [14]
    Suspension biomechanics of swimming microbes - PMC
    It can therefore be assumed that the flow field around a micro-organism is a Stokes flow and hence that the inertial force is negligible compared with the ...
  15. [15]
    [PDF] The First Five Births of the Navier-Stokes Equation
    Navier's original proof of 1822 was not influential, and the equation was rediscovered or re-derived at least four times, by Cauchy in 1823, by Poisson in 1829 ...
  16. [16]
    Stokes' law, viscometry, and the Stokes falling sphere clock - Journals
    Aug 3, 2020 · Stokes' paper was 'On the effect of the internal friction of fluids on the motion of pendulums'. In 1851, such research was very important for ...
  17. [17]
    Wilhelm Oseen and Stokes' paradox - MacTutor
    In the 1840s George Gabriel Stokes investigated the motion of a sphere through a viscous fluid. He published On the theories of the internal friction of ...Missing: Rayleigh | Show results with:Rayleigh
  18. [18]
    The self-propulsion of microscopic organisms through liquids
    The best-known problem that neglects all inertial forces is Stokes's solution for the slow steady fluid motion past a sphere, in which the velocity field can be ...
  19. [19]
    [PDF] 7 The Navier-Stokes Equations - MIT Mathematics
    For low Reynolds number it may be possible to ignore the inertial terms in the Navier-Stokes equations and obtain the so-called slow (or creeping) flow ...
  20. [20]
    [PDF] an introduction to - fluid dynamics
    Page 1. AN INTRODUCTION TO. FLUID DYNAMICS. BY. G. K. BATCHELOR, F.R.S.. Prof.,1Or of Applid Mathmatit' in 1M Un;v.,tity of Cambridg• .. :.:. CAMBRIDGE.
  21. [21]
    [PDF] On the theories of the internal friction of fluids in motion
    It appears however from experiments on pendulums that the total friction varies as the first power of the velocity, and consequently we may suppose that ...
  22. [22]
    [PDF] Low Reynolds number hydrodynamics: Stokes' flow
    The above property of Stokes flow is know as instantaneity. From instantaneity, time-reversibility immediately follows.
  23. [23]
    [PDF] Time-Reversibility and Particle Sedimentation
    (time reversibility is due to infinite viscosity and spatial ... particles with gravity acting as a delta function, Stokes flow for individual particles is.
  24. [24]
    [PDF] Microscopic irreversibility and chaos. - NYU Physics department
    Aug 8, 2006 · Irreversibility of fluid motion also oc- curs at a high Reynolds number, but it is not surprising there because the equations are nonlinear and ...
  25. [25]
    [PDF] Fluid Mechanics - DAMTP
    • George Batchelor, An Introduction to Fluid Dynamics. This is considered the ... that the stream function must be biharmonic. ∇4Ψ = 1 r. ∂. ∂r r. ∂. ∂r.
  26. [26]
    [PDF] Stokes Equation
    This proves that the stream function solves the biharmonic equation with homo. BC. ∆2ψ = curlf in Ω φ = 0 on ∂Ω. ∂φ. ∂n. = 0 on ∂Ω. Page 13. 4.4. STREAM ...
  27. [27]
    Stokes flow — MEK4300 Lecture notes - Professor Mikael Mortensen
    Flows at very low Reynolds numbers are often called Stokes flow. If the Navier-Stokes equations are nondimensionalized using the characteristic velocity U and ...
  28. [28]
    A fast finite difference method for biharmonic equations on irregular ...
    Jul 24, 2007 · A fast finite difference method for biharmonic equations on irregular domains and its application to an incompressible Stokes flow · Abstract.
  29. [29]
    Computation of Two-Dimensional Stokes Flows via Lightning and ...
    In this paper, we present the LARS algorithm (lightning-AAA rational Stokes) for computing two-dimensional (2D) Stokes flows in domains with smooth boundaries.2. 2d Stokes Flow And... · 4.1. Stokes Flow In A... · 5.2. Stokes Flow Between Two...
  30. [30]
    [PDF] Low Reynolds number flow
    Low Reynolds number flow is governed by the biharmonic ... The Stokes stream function. The flow is axisymmetric and best posed in spherical polar coordinates.
  31. [31]
    Stokes flow - Wikipedia
    Stokes flow also named creeping flow or creeping motion, is a type of fluid flow where advective inertial forces are small compared with viscous forces.
  32. [32]
    Hydromechanics of low-Reynolds-number flow. Part 2. Singularity ...
    Mar 29, 2006 · Chwang, A. T. & Wu, T. Y. 1974a Hydromechanics of low-Reynolds-number flow. Part 1. Rotation of axisymmetric prolate bodies. J. Fluid Mech ...
  33. [33]
    A note on the image system for a stokeslet in a no-slip boundary
    The image system consists of a stokeslet (opposite sign), a stokes-doublet, and a source-doublet, with displacement axes in the original force direction.
  34. [34]
    The Propulsion of Sea-Urchin Spermatozoa
    Gray, G. J. Hancock; The Propulsion of Sea-Urchin Spermatozoa. J Exp Biol 1 December 1955; 32 (4): 802–814. doi: https://doi.org/10.1242/jeb.32.4.802.ABSTRACT · II. PROPULSIVE EFFECT OF... · APPLICATION OF...
  35. [35]
    Papkovich–Neuber type representations with differential forms
    Jun 18, 2021 · This representation is also applicable for the solutions of the homogeneous steady-state Stokes equations, that is, for stationary Stokes flows.<|control11|><|separator|>
  36. [36]
    [PDF] ON THE PAPKOVICH-NEUBER FORMULATION FOR STOKES ...
    Abstract: We investigate, via the Papkovich-Neuber formulation using pro- late spheroidal coordinates, a fully three-dimensional Stokes flow in the exterior.
  37. [37]
    [PDF] General Solutions of the Stokes' Flow Equations - CORE
    The Papkovich-Neuber solution of the Stokes' flow equations for a viscous incompressible flow at very low Reynolds numbers is derived using three alternative ...
  38. [38]
    [PDF] ON THE PAPKOVICH–FADLE AND OTHER EIGENFUNCTIONS ...
    Before the appearance of Papkovich's (1940), Fadle's (1940), and ... The convergence of biorthogonal se- ries for biharmonic and stokes flow edge problems.
  39. [39]
    General solutions of the Stokes' flow equations - ScienceDirect.com
    The Papkovich-Neuber solution of the Stokes' flow equations for a viscous incompressible flow at very low Reynolds numbers is derived using three alternative ...Missing: formula | Show results with:formula
  40. [40]
    [PDF] Simple and efficient representations for the fundamental solutions of ...
    We have derived very simple image formulas for Stokes flow in a half-space induced by any of the standard fundamental solutions - the Stokeslet, stresslet, ...
  41. [41]
    Boundary Integral and Singularity Methods for Linearized Viscous ...
    This book presents a coherent introduction to boundary integral, boundary element and singularity methods for steady and unsteady flow at zero Reynolds ...
  42. [42]
    [PDF] The boundary integral formulation of Stokes flows includes slender ...
    Jul 16, 2018 · The incompressible Stokes equations can classically be recast in a boundary integral. (BI) representation, which provides a general method ...Missing: Phan- Thien
  43. [43]
    Boundary regularized integral equation formulation of Stokes flow
    Feb 5, 2015 · Single-phase Stokes flow problems with prescribed boundary conditions can be formulated in terms of a boundary regularized integral equation ...Missing: seminal | Show results with:seminal
  44. [44]
    Direct second kind boundary integral formulation for Stokes flow ...
    A direct boundary element method is formulated for the Stokes flow problem based on an integral equation representation for the components of traction.
  45. [45]
    [PDF] Nearly Singular Integrals in 3D Stokes Flow 1 Introduction - Duke Math
    A straightforward method is presented for computing three-dimensional. Stokes flow, due to forces on a surface, with high accuracy at points near the sur- face.<|control11|><|separator|>
  46. [46]
    [PDF] Boundary integral methods for Stokes flow
    To expedite the derivation we consider the fundamental solution of the Laplace equation ... singular kernel given by the fundamental solution of Stokes equations ...
  47. [47]
    Boundary integral equation analysis for suspension of spheres in ...
    Jul 18, 2017 · ... fast multipole method (FMM). We employ this discretization scheme to analyze several boundary integral formulations of interest including ...
  48. [48]
    Boundary element solution for half-space elasticity or stokes ...
    Tran-Cong, T., Phan-Thien, N. Boundary element solution for half-space elasticity or stokes problem with a no-slip boundary. Computational Mechanics 1, 259 ...
  49. [49]
    Stokes problems of multiparticle systems: A numerical method for ...
    Mar 1, 1989 · A boundary element method (BEM) is developed and used to compute the arbitrary Stokes flow of a system of N particles of arbitrary shape and ...
  50. [50]
    [PDF] A Solving Boundary Integral Problems with BEM++ - Bempp
    In this paper we present BEM++, a novel open-source library for the solution of boundary integral equations for Laplace,. Helmholtz and Maxwell problems in ...Missing: flow | Show results with:flow
  51. [51]
    [PDF] Low Reynolds number hydrodynamics
    Stokes equations at low Reynolds numbers. This singular perturbation ... solution of Stokes' equations for a translating disk32,3g,. K I = 136 c(311 t ...
  52. [52]
    The Flow of Water | Nature
    The Flow of Water. H. S. HELE-SHAW. Nature volume 58, pages 34–36 (1898)Cite this article ... A simple numerical method for Hele–Shaw type problems by the method ...Missing: original | Show results with:original
  53. [53]
    [PDF] On pressure-driven Hele-Shaw flow - DiVA portal
    Hele-Shaw experimentally (in 1898) concerning the steady motion of a fluid under pressure between two fixed closely spaced parallel plates. In those experiments ...<|separator|>
  54. [54]
    [PDF] Two-dimensional Stokes and Hele-Shaw flows with free surfaces
    dimensional Stokes flow both with and without surface tension. ... This, together with the relation ∂u/∂t = p, gives a complete description of the interior.
  55. [55]
    Obstructed viscoplastic flow in a Hele-Shaw cell | Phys. Rev. Fluids
    Jan 6, 2020 · Slow viscous flow around an obstacle is a classical problem in fluid mechanics, illustrating the so-called Stokes paradox and its resolution.
  56. [56]
    The penetration of a fluid into a porous medium or Hele-Shaw cell ...
    The penetration of a fluid into a porous medium or Hele-Shaw cell containing a more viscous liquid. Philip Geoffrey Saffman.
  57. [57]
    Hele Shaw microfluidic device: A new tool for systematic ...
    This method describes a novel approach to systematically investigate the effect of the fluid shear stress (FSS) on epithelial cells thanks to a single ...
  58. [58]
    Slender-body theory for particles of arbitrary cross-section in Stokes ...
    Mar 29, 2006 · A rigid body whose length (2l) is large compared with its breadth (represented by R0) is straight but is otherwise of arbitrary shape.
  59. [59]
    On the Spheroidal Semiseparation for Stokes Flow - Dassios - 2008
    The stream function ψ, for axisymmetric Stokes flow, satisfies the well-known equation E4ψ = 0. Despite the fact that in spherical coordinates this equation ...
  60. [60]
    Steady Stokes flow in a finite cylinder | Request PDF - ResearchGate
    Aug 9, 2025 · This article addresses an analytical approach for solving the problem of steady three–dimensional creeping flow in a finit cylindrical ...
  61. [61]
    [PDF] Jeffery-Hamel's Flows in the Plane
    Abstract. We consider a radial steady flow of an incompressible viscous fluid which either converges or diverges in a two dimensional wedge domain.
  62. [62]
    [PDF] Viscous and resistive eddies near a sharp corner - DAMTP
    ... Stokes flow near the leading edge of a wedge has the character of an infinite sequence of eddies if a sufficiently strong line current flows along the ...
  63. [63]
    [PDF] Mechanics of Blood Flow in Capillaries - CORE
    Mechanics of Blood Flow in Capillaries. Timothy W ... fluid, in Stokes flow. A finite element package ... Skalak, R., 1976. Rheology of red blood cell.
  64. [64]
    Helmholtz decomposition coupling rotational to irrotational flow of a ...
    Sep 26, 2006 · The Helmholtz decomposition gives rise to an exact theory of potential flow in the frame of the Navier–Stokes equations in which rotational and ...
  65. [65]
    A 'reciprocal' theorem for the prediction of loads on a body moving in ...
    Oct 20, 2011 · A 'reciprocal' theorem for the prediction of loads on a body moving in an inhomogeneous flow at arbitrary Reynolds number - Volume 689.
  66. [66]
    [PDF] Complex Fluids” dated 17th January, 2023 1 Stokes flow ... - People
    The viscous dissipation in Stokes flow is thus instantaneously equal to the sum of the rate of working by the body force and the rate of working by the surface ...
  67. [67]
    [PDF] A generalized traction integral equation for Stokes flow, with ...
    Dec 29, 2016 · a Papkovich–Neuber potential [104]. In this paper we use formulas obtained following the Lorentz method. The procedure is as follows. Given ...
  68. [68]
    Clustering induces switching between phoretic and osmotic ...
    Jul 6, 2024 · Using the Lorentz reciprocal theorem, we demonstrate that in large clusters self-propulsion emerges by considering the hydrodynamic flow via ...
  69. [69]