2016
DOI: 10.1017/jfm.2016.41
|View full text |Cite
|
Sign up to set email alerts
|

Drag and diffusion coefficients of a spherical particle attached to a fluid–fluid interface

Abstract: Explicit analytical expressions for the drag and diffusion coefficients of a spherical particle attached to the flat interface between two immiscible fluids are constructed for the case of a vanishing viscosity ratio between the fluid phases. The model is designed to account explicitly for the dependence on the contact angle between the two fluids and the solid surface. The Lorentz reciprocal theorem is applied in the context of geometric perturbations from the limiting cases of 90 • and 180 • contact angles. … Show more

Help me understand this report
View preprint versions

Search citation statements

Order By: Relevance

Paper Sections

Select...
1
1

Citation Types

3
66
0

Year Published

2017
2017
2022
2022

Publication Types

Select...
7
1

Relationship

0
8

Authors

Journals

citations
Cited by 66 publications
(69 citation statements)
references
References 34 publications
3
66
0
Order By: Relevance
“…3b) and this restriction again does not lead to an anomalously large drag. The Stokes continuum drag of a nonrotating smooth particle moving along a gas/liquid interface [3][4][5][6] is shown in Fig. 3b as ξ c and is larger than the drag computed from the MD simulations.…”
mentioning
confidence: 76%
See 1 more Smart Citation
“…3b) and this restriction again does not lead to an anomalously large drag. The Stokes continuum drag of a nonrotating smooth particle moving along a gas/liquid interface [3][4][5][6] is shown in Fig. 3b as ξ c and is larger than the drag computed from the MD simulations.…”
mentioning
confidence: 76%
“…Alternatively, if an external force is applied to the particle to yield a steady velocity, ξ is calculated directly as their ratio. Using the latter method, continuum calculations of the surface drag exerted on a spherical, smooth, nonrotating particle in the limit of zero inertia and a flat, zero-thickness interface separating two immiscible liquids have been undertaken [3][4][5][6]. The continuum drag increases with immersion into the more viscous phase, and approaches the Stokes bulk drag coefficient 6πµR, where µ is the fluid viscosity far from the surface.…”
mentioning
confidence: 99%
“…On other hand, small particles dynamically enter and detach from the aqueous–aqueous interface, which is driven by Brownian motion, resulting in reversible and unstable interfacial adsorption. Particularly, when the kinetic energy of Brownian motion is relative higher than the thermal energy to trap the surfactant at the aqueous–aqueous interface, surfactants cannot be irreversibly trapped at the interface, and the neighboring droplets will eventually coalesce . For instance, latex particles with a radius no less than 0.1 µm can be irreversibly trapped at the interfaces of the polyethylene oxide (PEO)/dextran ATPS with interfacial tensions down to 10 −6 N m −1 , as shown by the confocal laser scanning microscope (CLSM) images in a,b .…”
Section: All‐aqueous Interface Templated Biomaterials: Assembly Of Armentioning
confidence: 99%
“…Past studies using either boundary integral 15,16 , finite element 17,18 or integral transform methods 12,19 have focused on the translation, and have assumed that the colloid does not rotate due to the presence of surface roughness or chemical heterogeneities that pin the three-phase contact line (contact angle hysteresis). However recently, evaluating the drag exerted on non-smooth colloids 14 with heterogeneous surfaces as they diffuse along an interface while specifically examining the role of contact line pinning and de-pinning have been important given the extensive interest in particle-laden interfaces.…”
Section: Introductionmentioning
confidence: 99%