This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Soft-lubrication interactions between a rigid sphere and an elastic wall

Vincent Bertin Univ. Bordeaux, CNRS, LOMA, UMR 5798, 33405 Talence, France. UMR CNRS Gulliver 7083, ESPCI Paris, PSL Research University, 75005 Paris, France.    Yacine Amarouchene Univ. Bordeaux, CNRS, LOMA, UMR 5798, 33405 Talence, France.    Elie Raphaël UMR CNRS Gulliver 7083, ESPCI Paris, PSL Research University, 75005 Paris, France.    Thomas Salez thomas.salez@u-bordeaux.fr Univ. Bordeaux, CNRS, LOMA, UMR 5798, 33405 Talence, France. Global Station for Soft Matter, Global Institution for Collaborative Research and Education, Hokkaido University, Sapporo, Hokkaido 060-0808, Japan.
(February 10, 2025)
Abstract

The motion of an object within a viscous fluid and in the vicinity of a soft surface induces a hydrodynamic stress field that deforms the latter, thus modifying the boundary conditions of the flow. This results in elastohydrodynamic (EHD) interactions experienced by the particle. Here, we derive a soft-lubrication model, in order to compute all the forces and torque applied on a rigid sphere that is free to translate and rotate near an elastic wall. We focus on the limit of small deformations of the surface with respect to the fluid-gap thickness, and perform a perturbation analysis in dimensionless compliance. The response is computed in the framework of linear elasticity, for planar elastic substrates in the limiting cases of thick and thin layers. The EHD forces are also obtained analytically using the Lorentz reciprocal theorem.

I Introduction

The fluid-structure interaction between flows and boundaries is a central situation in continuum mechanics, encountered at many length and velocity scales. A classical example is lubrication, where the addition of a liquid film, a lubricant, between two contacting objects, allows for a drastic reduction of the friction between them. Such a process occurs in a large variety of contexts with hard materials, such as roller bearings, pistons and gears in industry (Dowson and Higginson, 2014), or faults (Brodsky and Kanamori, 2001) and landslides (Campbell, 1989) in geological settings. At large velocity, or moderate loading, the liquid film is continuous with no direct contact between the solids. When the solids are deformable, the friction force can be described using elastohydrodynamic (EHD) models within the soft-lubrication approximation (Dowson and Higginson, 2014).

The previous EHD coupling is also widely encountered in soft condensed matter, but at very different pressure and velocity scales (Karan et al., 2018). Examples encompass the remarkable frictional properties of eyelids (Jones et al., 2008) and cartilaginous joints (Mow et al., 1984; Jahn et al., 2016), as well as biomimetic gels (Gong, 2006) and rubbers (Sekimoto and Leibler, 1993; Moyle et al., 2020; Wu et al., 2020; Hui et al., 2021). Of interest as well are the collisions and rebounds of spheres in viscous environments (Davis et al., 1986; Gondret et al., 2002; Tan et al., 2019), the rheological properties of soft suspensions and pastes (Sekimoto and Leibler, 1993; Meeker et al., 2004), and the self-similar properties of the contact (Snoeijer et al., 2013).

In the last decade, EHD interactions have been of great interest in the material-science community with the emergence of contactless rheological methods to measure the mechanical properties of confined liquids and soft surfaces (Chan et al., 2009; Vakarelski et al., 2010; Leroy and Charlaix, 2011; Leroy et al., 2012; Villey et al., 2013; Wang et al., 2015, 2017a, 2017b; Guan et al., 2017; Wang et al., 2018; Lainé et al., 2019; Bertin et al., 2020). Typically, in such experimental systems, a spherical colloidal probe is immersed in a fluid and driven to oscillate, with a nanometric amplitude, near a surface of interest. The force exerted on the probe is measured by an atomic force microscope, a surface force apparatus or a tuning-fork microscope, and depends on the properties of both the fluid and the solid boundary.

Generally, an object that moves in a confined fluid environment experiences an enhanced drag force with respect to the bulk Stokes law, as a result of the boundary-induced flow modification (Happel and Brenner, 2012). Furthermore, near a soft wall, the hydrodynamic interactions are modified by the deformation of the boundary that they generate, yielding a nonlinear coupling. Perturbation methods, assuming a small deformation of the interface, have been employed in order to calculate the soft-lubrication interactions exerted on a free infinite cylinder immersed in a viscous fluid and near a thin compressible elastic material (Salez and Mahadevan, 2015). In particular, interesting inertial-like features have been predicted despite the low-Re-number aspect of the flow.

Perhaps the most emblematic example of soft-lubrication interaction is the non-inertial lift force predicted for a particle sliding near a soft boundary (Sekimoto and Leibler, 1993; Beaucourt et al., 2004; Skotheim and Mahadevan, 2004, 2005; Urzay et al., 2007; Urzay, 2010). It might have important implications for advected biological entities, such as red blood cells (Grandchamp et al., 2013) and vesicles (Abkarian et al., 2002). Only recently, the associated dynamical repulsion from an immersed soft interface has been studied experimentally. A preliminary qualitative observation was reported in the context of smart lubricants and elastic polyelectrolytes (Bouchet et al., 2015). Then, a study involving the sliding of an immersed macroscopic cylinder along an inclined plane pre-coated with a thin layer of gel, showed quantitatively an effective reduction of friction induced by the EHD lift force (Saintyves et al., 2016). Subsequently, the same effect was observed in the trajectories of micrometric spherical beads within a microfluidic channel coated with a biomimetic polymer layer (Davies et al., 2018), and through the sedimentation of a macroscopic sphere near a pre-tensed suspended elastic membrane (Rallabandi et al., 2018). Finally, direct measurements of the EHD lift force for two types of elastic materials have been performed at small scales, using surface force apparatus and atomic force microscopy, respectively (Vialar et al., 2019; Zhang et al., 2020).

Despite the increasing number of EHD studies involving spherical probes, the soft-lubrication interactions of a free spherical object immersed in a viscous fluid and moving near an elastic substrate still have to be calculated. In the present article, we aim at filling this gap by deriving a soft-lubrication perturbation theory, in order to compute all the forces and torque for this problem, at first order in dimensionless compliance.

The article is organized as follows. First, we introduce the soft-lubrication framework for a sphere translating near a soft planar surface, both in the normal and tangential directions. The substrate deformation is assumed to follow the constitutive response of a linear elastic semi-infinite material. Then, we perform a perturbative approach, assuming the substrate deformation to be small with respect to the fluid-gap thickness, which allows us to find the normal and tangential forces as well as the torque experienced by the sphere, at first order in dimensionless compliance. Finally, we discuss the rotation of the sphere, before providing concluding remarks. Besides, in the appendix, the EHD forces are computed analytically using the Lorentz reciprocal theorem, while the procedure introduced in the main text is reproduced for the compressible and incompressible responses of a thin material.

II Model

Refer to caption
Figure 1: Schematic of the system. A rigid sphere of surface 𝒮0\mathcal{S}_{0} is freely moving in a viscous fluid, near a soft wall of surface 𝒮w\mathcal{S}_{\textrm{w}} in the flat undeformed state. The lubrication pressure field deforms the latter which induces an elastohydrodynamic coupling, with forces and torque exerted on the sphere. Note that the surface deformation is magnified for clarity, but that we restrict the analysis to the δd\delta\ll d case.

The system is depicted in Fig. 1. We consider a sphere of radius aa, immersed in a Newtonian fluid of dynamic shear viscosity η\eta and density ρ\rho. The sphere is moving with a tangential velocity 𝒖(t)=u(t)𝒆x\bm{u}(t)=u(t)\,\bm{e}_{x} directed along the xx-axis (by definition of the latter axis), where 𝒆j\bm{e}_{j} denotes the unit vector along jj. In this first part, we assume that the sphere does not rotate, i.e. the angular velocity reads 𝛀=𝟎\bm{\Omega}=\bm{0}. The sphere is placed at a time-dependent distance d(t)d(t) (thus a d˙𝒆z\dot{d}\,\bm{e}_{z} normal velocity of the sphere) of an isotropic and homogeneous linear elastic substrate of Lamé coefficients λ\lambda and μ\mu, with a reference undeformed flat surface in the xyxy plan at z=0z=0. We suppose that the sphere-wall distance is small with respect to the sphere radius, such that the lubrication approximation is valid. The fluid inertia is neglected here. Specifically, we assume Re(d/a)1\textrm{Re}\,(d/a)\ll 1, with the Reynolds number Re=ρua/η\textrm{Re}=\rho ua/\eta. Furthermore, we suppose that the typical time scale of variation of the sphere velocity is much larger than the diffusion time scale of vorticity that scales as d2/(η/ρ)d^{2}/(\eta/\rho), such that the flow is described by the steady Stokes equations. This amounts to assuming that |u˙/u|η/(ρd2)\lvert\dot{u}/u\rvert\ll\eta/(\rho d^{2}) and |d¨/d˙|η/(ρd2)\lvert\ddot{d}/\dot{d}\rvert\ll\eta/(\rho d^{2}). No-slip boundary conditions are assumed at both the sphere and wall surfaces. Finally, the system is equivalent to a sphere at rest near a wall translating with a (d(t)𝒆z+𝒖(t))-(d(t)\bm{e}_{z}+\bm{u}(t)) velocity. In such a framework, the fluid velocity field can be written as:

𝒗(𝒓,z,t)=p(𝒓,t)2η(zh0(r,t))(zδ(𝒓,t))𝒖(t)h0(r,t)zh0(r,t)δ(𝒓,t),\bm{v}(\bm{r},z,t)=\frac{\bm{\nabla}p(\bm{r},t)}{2\eta}(z-h_{0}(r,t))(z-\delta(\bm{r},t))-\bm{u}(t)\frac{h_{0}(r,t)-z}{h_{0}(r,t)-\delta(\bm{r},t)}, (1)

where 𝒓=(r,θ)\bm{r}=(r,\theta) is the position in the tangential plane xyxy, \bm{\nabla} is the 2D gradient operator on xyxy, δ(𝒓,t)\delta(\bm{r},t) is the substrate deformation, and z=h0(r,t)z=h_{0}(r,t) is the sphere surface. Near contact, the latter can be approximated by its parabolic expansion h0(r,t)d(t)+r2/(2a)h_{0}(r,t)\simeq d(t)+r^{2}/(2a). Volume conservation further leads to the Reynolds equation:

th(𝒓,t)=(h3(𝒓,t)12ηp(𝒓,t)+h(𝒓,t)2𝒖(t)),\partial_{t}h(\bm{r},t)=\bm{\nabla}\cdot\bigg{(}\frac{h^{3}(\bm{r},t)}{12\eta}\bm{\nabla}p(\bm{r},t)+\frac{h(\bm{r},t)}{2}\bm{u}(t)\bigg{)}, (2)

where h(𝒓,t)=h0(r,t)δ(𝒓,t)h(\bm{r},t)=h_{0}(r,t)-\delta(\bm{r},t) is the fluid-gap thickness. In this first part, we assume that the constitutive elastic response is linear and instantaneous, and that the substrate is a semi-infinite medium, such that the deformation reads Davis et al. (1986):

δ(𝒓,t)=(λ+2μ)4πμ(λ+μ)2d2𝒙p(𝒙,t)|𝒓𝒙|.\delta(\bm{r},t)=-\frac{(\lambda+2\mu)}{4\pi\mu(\lambda+\mu)}\int_{\mathbb{R}^{2}}\textrm{d}^{2}\bm{x}\frac{p(\bm{x},t)}{\lvert\bm{r}-\bm{x}\rvert}. (3)

We non-dimensionalize the problem through:

h(𝒓,t)=dH(𝑹,T),𝒓=2ad𝑹,d(t)=dD(T),δ(𝒓,t)=dΔ(𝑹,T),h(\bm{r},t)=d^{*}H(\bm{R},T),\quad\bm{r}=\sqrt{2ad^{*}}\,\bm{R},\quad d(t)=d^{*}D(T),\quad\delta(\bm{r},t)=d^{*}\Delta(\bm{R},T), (4)
p(𝒓,t)=ηu2add2P(𝑹,T),𝒖(t)=uU(T)𝒆x,𝒗=u𝑽,t=2aduT.p(\bm{r},t)=\frac{\eta u^{*}\sqrt{2ad^{*}}}{d^{*2}}\,P(\bm{R},T),\quad\bm{u}(t)=u^{*}\,U(T)\,\bm{e}_{x},\quad\bm{v}=u^{*}\bm{V},\quad t=\frac{\sqrt{2ad^{*}}}{u^{*}}\,T. (5)

where dd^{*} and uu^{*} are characteristic fluid-gap distance and tangential velocity, respectively. The governing equations are then:

12TH(𝑹,T)=(H3(𝑹,T)P(𝑹,T)+6H(𝑹,T)𝑼(T)),12\partial_{T}H(\bm{R},T)=\bm{\nabla}\cdot\bigg{(}H^{3}(\bm{R},T)\bm{\nabla}P(\bm{R},T)+6H(\bm{R},T)\bm{U}(T)\bigg{)}, (6)
H(𝑹,T)=D(T)+R2Δ(𝑹,T),H(\bm{R},T)=D(T)+R^{2}-\Delta(\bm{R},T), (7)

and:

Δ(𝑹,T)=κ2d2𝑿P(𝑿,T)4π|𝑹𝑿|,\Delta(\bm{R},T)=-\kappa\int_{\mathbb{R}^{2}}\textrm{d}^{2}\bm{X}\frac{P(\bm{X},T)}{4\pi\lvert\bm{R}-\bm{X}\rvert}, (8)

where we introduced the dimensionless compliance:

κ=2ηua(λ+2μ)d2μ(λ+μ).\kappa=\frac{2\eta u^{*}a(\lambda+2\mu)}{d^{*2}\mu(\lambda+\mu)}. (9)

The latter is the only dimensionless parameter in the problem. When κ\kappa is small with respect to unity, it corresponds to the ratio between two length scales: the typical substrate deformation δ2ηua(λ+2μ)dμ(λ+μ)\delta\sim\frac{2\eta u^{*}a(\lambda+2\mu)}{d^{*}\mu(\lambda+\mu)} induced by a tangential velocity uu^{*}, and the typical fluid-gap thickness dd^{*}. All along the article, we focus on the small-deformation regime of soft-lubrication where κ1\kappa\ll 1 (Essink et al., 2021).

III Perturbation theory

We perform a perturbation analysis at small κ\kappa (Sekimoto and Leibler, 1993; Beaucourt et al., 2004; Skotheim and Mahadevan, 2004, 2005; Urzay et al., 2007; Urzay, 2010; Salez and Mahadevan, 2015; Pandey et al., 2016; Rallabandi et al., 2017; Saintyves et al., 2020; Zhang et al., 2020), as follows:

H(𝑹,T)=H0(𝑹,T)+κH1(𝑹,T)+O(κ2),H(\bm{R},T)=H_{0}(\bm{R},T)+\kappa\,H_{1}(\bm{R},T)+O(\kappa^{2}), (10)
P(𝑹,T)=P0(𝑹,T)+κP1(𝑹,T)+O(κ2),P(\bm{R},T)=P_{0}(\bm{R},T)+\kappa\,P_{1}(\bm{R},T)+O(\kappa^{2}), (11)

where the subscript 0 corresponds to the solution for a rigid wall, with H0(𝑹,T)=D(T)+R2H_{0}(\bm{R},T)=D(T)+R^{2}.

III.1 Zeroth-order solution: rigid wall

Equation (6) reads at zeroth order O(κ0)O(\kappa^{0}):

12D˙=(H03P0+6H0𝑼).12\dot{D}=\bm{\nabla}\cdot\bigg{(}H_{0}^{3}\bm{\nabla}P_{0}+6H_{0}\bm{U}\bigg{)}. (12)

In polar coordinates, Eq. (12) can be rewritten as:

.P0=R2R2P0+(R+6R3D+R2)RP0+θ2P0=R2(D+R2)3(12D˙12RcosθU),\mathcal{L}.P_{0}=R^{2}\partial_{R}^{2}P_{0}+\bigg{(}R+\frac{6R^{3}}{D+R^{2}}\bigg{)}\partial_{R}P_{0}+\partial_{\theta}^{2}P_{0}=\frac{R^{2}}{(D+R^{2})^{3}}\,\bigg{(}12\dot{D}-12R\cos\theta\,U\bigg{)}, (13)

where \mathcal{L} is a linear operator. We solve this equation, using an angular-mode decomposition:

P0(𝑹,T)=P0(0)(R,T)+P0(1)(R,T)cosθ,P_{0}(\bm{R},T)=P_{0}^{(0)}(R,T)+P_{0}^{(1)}(R,T)\cos\theta, (14)

where the two coefficients are solutions of the ordinary differential equations:

R2d2P0(0)dR2+(R+6R3D+R2)dP0(0)dR=12R2D˙(D+R2)3,R^{2}\frac{\textrm{d}^{2}P_{0}^{(0)}}{\textrm{d}R^{2}}+\bigg{(}R+\frac{6R^{3}}{D+R^{2}}\bigg{)}\frac{\textrm{d}P_{0}^{(0)}}{\textrm{d}R}=12\frac{R^{2}\dot{D}}{(D+R^{2})^{3}}, (15a)
R2d2P0(1)dR2+(R+6R3D+R2)dP0(1)dRP0(1)=12R3U(D+R2)3.R^{2}\frac{\textrm{d}^{2}P_{0}^{(1)}}{\textrm{d}R^{2}}+\bigg{(}R+\frac{6R^{3}}{D+R^{2}}\bigg{)}\frac{\textrm{d}P_{0}^{(1)}}{\textrm{d}R}-P_{0}^{(1)}=-12\frac{R^{3}U}{(D+R^{2})^{3}}. (16a)

In accordance with the boundary conditions, P(R)=0P(R\rightarrow\infty)=0 and P(R=0)<P(R=0)<\infty, the solution is thus:

P0(𝑹,T)=3D˙2(D+R2)2+6RUcosθ5(D+R2)2.P_{0}(\bm{R},T)=-\frac{3\dot{D}}{2(D+R^{2})^{2}}+\frac{6RU\cos\theta}{5(D+R^{2})^{2}}. (17)

The first-order substrate deformation H1H_{1} can then be computed from Eq. (8) at order O(κ)O(\kappa):

H1(𝑹,T)=2d2𝑿P0(𝑿,T)4π|𝑹𝑿|.H_{1}(\bm{R},T)=\int_{\mathbb{R}^{2}}\textrm{d}^{2}\bm{X}\frac{P_{0}(\bm{X},T)}{4\pi\lvert\bm{R}-\bm{X}\rvert}. (18)

Using e.g. the spatial Fourier transform H~1(𝑲)=2H1(𝑹)ei𝑹𝑲d2𝑹\tilde{H}_{1}(\bm{K})=\int_{\mathbb{R}^{2}}H_{1}(\bm{R})e^{-i\bm{R}\cdot\bm{K}}\textrm{d}^{2}\bm{R}, we find:

H1(𝑹,T)=3D˙8D(R2/D)D+R2+3U10RD(D(R2/D)D+R2+𝒦(R2/D))cosθ,H_{1}(\bm{R},T)=-\frac{3\dot{D}}{8\sqrt{D}}\,\frac{\mathcal{E}(-R^{2}/D)}{D+R^{2}}+\frac{3U}{10R\sqrt{D}}\,\bigg{(}-\frac{D\,\mathcal{E}(-R^{2}/D)}{D+R^{2}}+\mathcal{K}(-R^{2}/D)\bigg{)}\cos\theta, (19)

where 𝒦\mathcal{K} and \mathcal{E} are the complete elliptic integrals of the first and second kinds (Abramowitz and Stegun, 1964). The dimensionless substrate deformations are plotted in Fig. 2. In Fig. 2a), the sphere is moving tangentially to the substrate with a unit velocity U=1U=1. The deformation exhibits a dipolar symmetry, with a negative sign (i.e. the substrate is compressed) at the front. Besides, the isotropic term generated by a sphere moving normally to the substrate is shown in Fig. 2b). In particular, for a sphere approaching the substrate, the latter is compressed.

Refer to caption
Figure 2: Dimensionless deformation fields at the free surface of the soft substrate, for a sphere placed at a unit distance D=1D=1 and for two modes of motion: a) the sphere velocity is directed tangentially to the substrate, along the xx-axis (see black arrow), and is fixed to a unit value U=1U=1; b) the sphere is approaching the substrate normally (see black cross) with a unit velocity D˙=1\dot{D}=-1.

III.2 First-order solution

We can now compute the first-order pressure field P1P_{1}, from Eq. (6) at order O(κ)O(\kappa):

12TH1=(H03P1+3H02H1P0+6H1𝑼).12\partial_{T}H_{1}=\bm{\nabla}\cdot\bigg{(}H_{0}^{3}\bm{\nabla}P_{1}+3H_{0}^{2}H_{1}\bm{\nabla}P_{0}+6H_{1}\bm{U}\bigg{)}. (20)

Invoking the same linear operator \mathcal{L} as in Eq. (13), we can rewrite Eq. (20) as:

.P1=R2H03(12TH1[3H02H1P0+6H1𝑼]).\mathcal{L}.P_{1}=\frac{R^{2}}{H_{0}^{3}}\bigg{(}12\partial_{T}H_{1}-\bm{\nabla}\cdot\bigg{[}3H_{0}^{2}H_{1}\bm{\nabla}P_{0}+6H_{1}\bm{U}\bigg{]}\bigg{)}. (21)

We then expand all the terms in the right-hand side of Eq. (21), and we perform once again the angular-mode decomposition:

.P1=F0(R,T)+F1(R,T)cosθ+F2(R,T)cos2θ,\mathcal{L}.P_{1}=F_{0}(R,T)+F_{1}(R,T)\cos\theta+F_{2}(R,T)\cos 2\theta, (22)

where we have introduced the auxiliary functions:

F0(R,T)=18R2U225D1/2(D+R2)6[(10D2+2DR2)(R2D)+(8D2+7DR2R4)𝒦(R2D)]+9R2D˙24D3/2(D+R2)6[(13D2+3R2D+2R4)(R2D)+(4D25R2DR4)𝒦(R2D)]9R2D¨(R2D)2D1/2(D+R2)4,\begin{split}F_{0}(R,T)&=\frac{18R^{2}U^{2}}{25D^{1/2}(D+R^{2})^{6}}\bigg{[}(-10D^{2}+2DR^{2})\,\mathcal{E}\left(-\frac{R^{2}}{D}\right)+(8D^{2}+7DR^{2}-R^{4})\,\mathcal{K}\left(-\frac{R^{2}}{D}\right)\bigg{]}\\ &\,\,+\frac{9R^{2}\dot{D}^{2}}{4D^{3/2}(D+R^{2})^{6}}\bigg{[}(13D^{2}+3R^{2}D+2R^{4})\,\mathcal{E}\left(-\frac{R^{2}}{D}\right)+(-4D^{2}-5R^{2}D-R^{4})\,\mathcal{K}\left(-\frac{R^{2}}{D}\right)\bigg{]}\\ &\,\,-\frac{9R^{2}\ddot{D}\,\mathcal{E}\left(-\frac{R^{2}}{D}\right)}{2D^{1/2}(D+R^{2})^{4}},\end{split} (23)

and:

F1(R,T)=27RUD˙5D1/2(D+R2)6[(2D2+7DR2+R4)(R2D)+2(D+R2)(DR2)𝒦(R2D)]18RU˙5D1/2(D+R2)4[D(R2D)+(D+R2)𝒦(R2D)].\begin{split}F_{1}(R,T)&=-\frac{27RU\dot{D}}{5D^{1/2}(D+R^{2})^{6}}\bigg{[}(-2D^{2}+7DR^{2}+R^{4})\,\mathcal{E}\left(-\frac{R^{2}}{D}\right)+2(D+R^{2})(D-R^{2})\,\mathcal{K}\left(-\frac{R^{2}}{D}\right)\bigg{]}\\ &\,\,-\frac{18R\dot{U}}{5D^{1/2}(D+R^{2})^{4}}\bigg{[}-D\,\mathcal{E}\left(-\frac{R^{2}}{D}\right)+(D+R^{2})\,\mathcal{K}\left(-\frac{R^{2}}{D}\right)\bigg{]}.\end{split} (24)

We note that we have not provided F2F_{2} as it does not contribute in the forces and torque. We also note that, by setting D(T)=1D(T)=1 in the latter expressions, we self-consistently recover the expression of Zhang et al. (2020). Invoking the angular-mode decomposition P1(𝑹,T)=P1(0)(R,T)+P1(1)(R,T)cosθ+P1(2)(R,T)cos2θP_{1}(\bm{R},T)=P_{1}^{(0)}(R,T)+P_{1}^{(1)}(R,T)\cos\theta+P_{1}^{(2)}(R,T)\cos 2\theta, we get in particular:

R2d2P1(0)dR2+(R+6R3D+R2)dP1(0)dR=F0(R,T),R^{2}\frac{\textrm{d}^{2}P_{1}^{(0)}}{\textrm{d}R^{2}}+\bigg{(}R+\frac{6R^{3}}{D+R^{2}}\bigg{)}\frac{\textrm{d}P_{1}^{(0)}}{\textrm{d}R}=F_{0}(R,T), (25)
R2d2P1(1)dR2+(R+6R3D+R2)dP1(1)dRP1(1)=F1(R,T).R^{2}\frac{\textrm{d}^{2}P_{1}^{(1)}}{\textrm{d}R^{2}}+\bigg{(}R+\frac{6R^{3}}{D+R^{2}}\bigg{)}\frac{\textrm{d}P_{1}^{(1)}}{\textrm{d}R}-P_{1}^{(1)}=F_{1}(R,T). (26)

Using scaling arguments, we can write the two relevant first-order pressure components P1(i)P_{1}^{(i)} as:

P1(0)=U2D7/2ϕU2(R/D)+D˙2D9/2ϕD˙2(R/D)+D¨D7/2ϕD¨(R/D),P_{1}^{(0)}=\frac{U^{2}}{D^{7/2}}\phi_{U^{2}}(R/\sqrt{D})+\frac{\dot{D}^{2}}{D^{9/2}}\phi_{\dot{D}^{2}}(R/\sqrt{D})+\frac{\ddot{D}}{D^{7/2}}\phi_{\ddot{D}}(R/\sqrt{D}), (27)

and:

P1(1)=UD˙D4ϕUD˙(R/D)+U˙D3ϕU˙(R/D),P_{1}^{(1)}=\frac{U\dot{D}}{D^{4}}\phi_{U\dot{D}}(R/\sqrt{D})+\frac{\dot{U}}{D^{3}}\phi_{\dot{U}}(R/\sqrt{D}), (28)

where the ϕi\phi_{i} are five dimensionless scaling functions that depend on the self-similar variable R/DR/\sqrt{D} only. Equations (27) and (28) can be solved numerically with a Runge-Kutta algorithm, and a shooting parameter in order to ensure the boundary condition P1(R,θ,T)=0P_{1}(R\rightarrow\infty,\theta,T)=0. All the scaling functions are plotted in Figs. 3 and 4.

As a remark, we recall that the substrate deformation is induced by the flow, and that at first order it is linear in the velocity field (see Eq. (19)). Moreover, the volume-conservation equation involves the time derivative of the fluid-layer thickness, and thus in particular the time derivative of the substrate deformation. As a consequence, when calculating the first-order EHD pressure field, we find terms (and thus forces and torques) that are proportional to the acceleration components D¨\ddot{D} and U˙\dot{U} of the sphere. At first sight, these original inertial-like features might seem inconsistent with steady Stokes flows, but are in fact independent of the fluid density and solely induced by the intimate EHD coupling.

Refer to caption
Figure 3: Scaling functions for P1(0)P_{1}^{(0)} (see Eq. (27)), obtained from numerical integration of Eq. (25), with the boundary conditions RP1(0)(R=0,T)=0\partial_{R}P_{1}^{(0)}(R=0,T)=0 and P1(0)(R,T)=0P_{1}^{(0)}(R\rightarrow\infty,T)=0.

III.3 Forces and torque

The force 𝑭\bm{F} exerted by the fluid on the sphere is given by:

𝑭=𝒮0𝒏𝝈ds,\bm{F}=\int_{\mathcal{S}_{0}}\bm{n}\cdot\bm{\sigma}\,\mathrm{d}s, (29)

where 𝝈=p𝐈+η(𝒗+𝒗T)\bm{\sigma}=-p\mathbf{I}+\eta(\bm{\nabla}\bm{v}+\bm{\nabla}\bm{v}^{\text{T}}) is the fluid stress tensor, 𝒏\bm{n} is the unit vector normal to the sphere surface and pointing towards the fluid, and 𝐈\mathbf{I} is the identity tensor. Within the lubrication approximation, the fluid stress tensor reads 𝝈p𝐈+η𝒆zz𝒗\bm{\sigma}\simeq-p\mathbf{I}+\eta\bm{e}_{z}\partial_{z}\bm{v}. One can then evaluate the normal force, as:

Fz=2p(𝒓)d2𝒓=6πηa2d˙d+0.41623η2u2(λ+2μ)μ(λ+μ)(ad)5/241.912η2d˙2(λ+2μ)μ(λ+μ)(ad)7/2+18.499η2d¨a(λ+2μ)μ(λ+μ)(ad)5/2,\begin{split}F_{z}=\int_{\mathbb{R}^{2}}p(\bm{r})\,\textrm{d}^{2}\bm{r}=&-\frac{6\pi\eta a^{2}\dot{d}}{d}+0.41623\frac{\eta^{2}u^{2}(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{5/2}\\ &-41.912\frac{\eta^{2}\dot{d}^{2}(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{7/2}+18.499\frac{\eta^{2}\ddot{d}a(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{5/2},\end{split} (30)

where the prefactors have been numerically estimated using Eq. (27). We recover in particular the classical Reynolds force 6πηa2d˙/d-6\pi\eta a^{2}\dot{d}/d at zeroth order, i.e. near a rigid wall. We stress that tangential motions do not induce any normal force at zeroth order in κ\kappa, as the corresponding pressure field is antisymmetric in xx (see Eq. (13)). In contrast, such motions do induce a lift force at first order in κ\kappa, due to the symmetry breaking of the contact geometry associated with the elastic deformation. Interestingly as well, normal motions generate a viscous adhesive force at first order in compliance (Wang et al., 2020). Besides, an original EHD force proportional to the sphere’s normal acceleration is also found, as discussed previously. Finally, in the appendix, and following previous works (Rallabandi et al., 2017, 2018; Daddi-Moussa-Ider et al., 2018; Masoud and Stone, 2019), we use the Lorentz reciprocal theorem in order to recover the prefactors of Eq. (30) analytically, which gives respectively: 243π31280020.41623\frac{243\pi^{3}}{12800\sqrt{2}}\approx 0.41623, 3915π32048241.912\frac{3915\pi^{3}}{2048\sqrt{2}}\approx 41.912 and 27π332218.499\frac{27\pi^{3}}{32\sqrt{2}}\approx 18.499. We note that the latter is in agreement with the result of the linear-response theory derived in Leroy and Charlaix (2011). Furthermore, we recover the lift prefactor (0.416) obtained previously numerically (Zhang et al., 2020), as well as analytically in a recently-published work (Kargar-Estahbanati and Rallabandi, 2021).

Refer to caption
Figure 4: Scaling functions for P1(1)P_{1}^{(1)} (see Eq. (28)), obtained from numerical integration of Eq. (26), with the boundary conditions P1(1)(R=0,T)=0P_{1}^{(1)}(R=0,T)=0 and P1(1)(R,T)=0P_{1}^{(1)}(R\rightarrow\infty,T)=0.

Similarly, the tangential force reads:

𝑭=2(p(𝒓,t)𝒓aηz𝒗)z=h0(r,t)d2𝒓.\bm{F}_{\parallel}=\int_{\mathbb{R}^{2}}\,\bigg{(}-p(\bm{r},t)\frac{\bm{r}}{a}-\eta\partial_{z}\bm{v}\bigg{)}_{z=h_{0}(r,t)}\textrm{d}^{2}\bm{r}. (31)

Using symmetry arguments, we can show that the tangential force is directed along xx, i.e. 𝑭=Fx𝒆x\bm{F}_{\parallel}=F_{x}\,\bm{e}_{x}. At small κ\kappa, we further expand it as FxFx,0+κFx,1F_{x}\simeq F_{x,0}+\kappa F_{x,1}, where Fx,0F_{x,0} is the viscous drag force applied on a sphere near a rigid plane wall, and κFx,1\kappa F_{x,1} is the first-order EHD correction. The zeroth-order term cannot be evaluated using the lubrication model introduced in the previous section, because the integral in Eq. (31) diverges, as the shear term ηz𝒗\eta\partial_{z}\bm{v} scales as r2\sim r^{-2} at large rr. An exact calculation has been performed using bispherical coordinates and provides a solution in the form of a series expansion (O’Neill, 1964). Asymptotic-matching methods have also been employed in order to get the asymptotic behavior at small d/ad/a (O’Neill and Stewartson, 1967; Goldman et al., 1967), which reads Fx,06πηau(815log(da)0.95429)F_{x,0}\approx 6\pi\eta au\bigg{(}\frac{8}{15}\log\left(\frac{d}{a}\right)-0.95429\bigg{)} (see (Chaoui and Feuillebois, 2003) for a high-precision expansion). We note that the sphere’s normal velocity does not contribute to the zeroth-order tangential force, as expected by symmetry.

The first-order EHD correction can be computed with the present model, as the correction pressure field and shear stress scale as r5\sim r^{-5}, at large rr. It reads:

Fx,1=2πηua0[2RP1(1)H02(RP1(1)+P1(1)R)H1(1)2RP0(0)H1(0)2(RP0(1)+P0(1)R)+2UH1(0)H02]RdR,\begin{split}F_{x,1}=2\pi\eta u^{*}a\int_{0}^{\infty}\bigg{[}&-2RP_{1}^{(1)}-\frac{H_{0}}{2}\bigg{(}\partial_{R}P_{1}^{(1)}+\frac{P_{1}^{(1)}}{R}\bigg{)}\\ &-\frac{H_{1}^{(1)}}{2}\partial_{R}P_{0}^{(0)}-\frac{H_{1}^{(0)}}{2}\bigg{(}\partial_{R}P_{0}^{(1)}+\frac{P_{0}^{(1)}}{R}\bigg{)}+2\frac{UH_{1}^{(0)}}{H_{0}^{2}}\bigg{]}R\,\textrm{d}R,\end{split} (32)

where H1(i)H_{1}^{(i)} is the amplitude of the ithi^{\textrm{th}} mode in the angular-mode decomposition of H1H_{1}. Evaluating the latter integral numerically, we find:

Fx6πηau(815log(da)0.95429)10.884η2ud˙(λ+2μ)μ(λ+μ)(ad)5/2+0.98661η2u˙a(λ+2μ)μ(λ+μ)(ad)3/2.F_{x}\approx 6\pi\eta au\bigg{(}\frac{8}{15}\log\left(\frac{d}{a}\right)-0.95429\bigg{)}-10.884\frac{\eta^{2}u\dot{d}(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{5/2}+0.98661\frac{\eta^{2}\dot{u}a(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{3/2}. (33)

In the appendix, we use again the Lorentz reciprocal theorem in order to compute the first-order EHD force, and we obtain the following analytical expressions for the coefficients of Eq. (33): 3177π36400210.884-\frac{3177\pi^{3}}{6400\sqrt{2}}\simeq-10.884 and 9π320020.98661\frac{9\pi^{3}}{200\sqrt{2}}\simeq 0.98661, respectively.

The torque exerted by the fluid on the sphere, with respect to its center of mass, is given by:

𝑻=𝒮0a𝒏×(𝒏𝝈)ds.\bm{T}=\int_{\mathcal{S}_{0}}a\bm{n}\times(\bm{n}\cdot\bm{\sigma})\,\mathrm{d}s. (34)

The latter is directed along the yy direction for symmetry reasons, i.e. 𝑻=Ty𝒆y\bm{T}=T_{y}\,\bm{e}_{y}. At small κ\kappa, we further expand it as TyTy,0+κTy,1T_{y}\simeq T_{y,0}+\kappa T_{y,1}. For the same reason as with the the viscous drag force near a rigid wall, the viscous torque near a rigid wall cannot be computed within the lubrication model. Using asymptotic-matching methods (O’Neill and Stewartson, 1967; Chaoui and Feuillebois, 2003), it is found to be Ty,08πηua2(110log(da)0.19296)T_{y,0}\approx 8\pi\eta ua^{2}\bigg{(}-\frac{1}{10}\log\left(\frac{d}{a}\right)-0.19296\bigg{)}. In contrast, the first-order EHD correction can be computed with the present model, and reads:

Ty,1=2ηua2π0[H02(RP1(1)+P1(1)R)+H1(1)2RP0(0)+H1(0)2(RP0(1)+P0(1)R)+2UH1(0)H02]RdR.T_{y,1}=-2\eta u^{*}a^{2}\pi\int_{0}^{\infty}\bigg{[}\frac{H_{0}}{2}\bigg{(}\partial_{R}P_{1}^{(1)}+\frac{P_{1}^{(1)}}{R}\bigg{)}+\frac{H_{1}^{(1)}}{2}\partial_{R}P_{0}^{(0)}+\frac{H_{1}^{(0)}}{2}\bigg{(}\partial_{R}P_{0}^{(1)}+\frac{P_{0}^{(1)}}{R}\bigg{)}+2\frac{UH_{1}^{(0)}}{H_{0}^{2}}\bigg{]}R\,\textrm{d}R. (35)

Evaluating the latter integral numerically, we find:

Ty8πηua2(110log(da)0.19296)+10.884η2uad˙(λ+2μ)μ(λ+μ)(ad)5/20.98661η2u˙a2(λ+2μ)μ(λ+μ)(ad)3/2.T_{y}\approx 8\pi\eta ua^{2}\bigg{(}-\frac{1}{10}\log\left(\frac{d}{a}\right)-0.19296\bigg{)}+10.884\frac{\eta^{2}ua\dot{d}(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{5/2}-0.98661\frac{\eta^{2}\dot{u}a^{2}(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{3/2}. (36)

So far, we focused on the particular case of a semi-infinite elastic material. In appendices C and D, we apply the same soft-lubrication approach to other elastic models describing thin substrates, which are also widespread in practice. We find similar expressions, but with different numerical prefactors and scalings with the sphere-wall distance.

IV Rotation

We now add the rotation of the sphere, with angular velocity 𝛀(t)\bm{\Omega}(t) in the xyxy plane (see Fig. 1), to the previous translational motion. We define β\beta as the angle between 𝛀\bm{\Omega} and the xx-axis. We stress that 𝛀\bm{\Omega} is not necessarily orthogonal (i.e. β=π/2\beta=\pi/2) to the translation velocity. We discard the rotation along the zz-axis (e.g. for a spinner), because it does not induce any soft-lubrication coupling. Finally, the system is equivalent to a purely rotating sphere with angular velocity 𝛀(t)\bm{\Omega}(t), near a wall translating with a 𝒖(t)-\bm{u}(t) velocity. In such a framework, the fluid velocity field at the sphere surface is 𝒗=𝛀×a𝒏\bm{v}=-\bm{\Omega}\times a\bm{n}, and thus 𝒗𝛀×a𝒆z\bm{v}\simeq-\bm{\Omega}\times a\bm{e}_{z}. All together, the fluid velocity field is modified as:

𝒗(𝒓,z,t)=p(𝒓,t)2η(zh0(r,t))(zδ(𝒓,t))𝒖(t)h0(r,t)zh0(r,t)δ(𝒓,t)+a𝛀(t)×𝒆zzδ(𝒓,t)h0(r,t)δ(𝒓,t),\bm{v}(\bm{r},z,t)=\frac{\bm{\nabla}p(\bm{r},t)}{2\eta}(z-h_{0}(r,t))(z-\delta(\bm{r},t))-\bm{u}(t)\frac{h_{0}(r,t)-z}{h_{0}(r,t)-\delta(\bm{r},t)}+a\bm{\Omega}(t)\times\bm{e}_{z}\,\frac{z-\delta(\bm{r},t)}{h_{0}(r,t)-\delta(\bm{r},t)}, (37)

and the Reynolds equation becomes:

th(𝒓,t)=(h3(𝒓,t)12ηp(𝒓,t)+h(𝒓,t)2[𝒖(t)a𝛀(t)×𝒆z𝒖~]).\partial_{t}h(\bm{r},t)=\bm{\nabla}\cdot\bigg{(}\frac{h^{3}(\bm{r},t)}{12\eta}\bm{\nabla}p(\bm{r},t)+\frac{h(\bm{r},t)}{2}\bigg{[}\underbrace{\bm{u}(t)-a\bm{\Omega}(t)\times\bm{e}_{z}}_{\tilde{\bm{u}}}\bigg{]}\bigg{)}. (38)

The problem is thus formally equivalent to the one of a sphere that is purely translating with effective velocity 𝒖~(t)=𝒖(t)a𝛀(t)×𝒆z\bm{\tilde{u}}(t)=\bm{u}(t)-a\bm{\Omega}(t)\times\bm{e}_{z}. Therefore, we can directly apply the results from the previous sections, and write all the forces and torque exerted on the sphere, as:

Fz=6πηa2d˙d+243π3128002η2|𝒖a𝛀×𝒆z|2μ(ad)5/23915π320482η2d˙2μ(ad)7/2+27π3322η2d¨aμ(ad)5/2,F_{z}=-\frac{6\pi\eta a^{2}\dot{d}}{d}+\frac{243\pi^{3}}{12800\sqrt{2}}\frac{\eta^{2}\lvert\bm{u}-a\bm{\Omega}\times\bm{e}_{z}\rvert^{2}}{\mu}\bigg{(}\frac{a}{d}\bigg{)}^{5/2}-\frac{3915\pi^{3}}{2048\sqrt{2}}\frac{\eta^{2}\dot{d}^{2}}{\mu}\bigg{(}\frac{a}{d}\bigg{)}^{7/2}+\frac{27\pi^{3}}{32\sqrt{2}}\frac{\eta^{2}\ddot{d}a}{\mu}\bigg{(}\frac{a}{d}\bigg{)}^{5/2}, (39)
𝑭=6πηa𝒖[815log(da)0.95429]+6πηa2𝒆z×𝛀[215log(da)+0.25725]3177π364002η2(𝒖a𝛀×𝒆z)d˙μ(ad)5/2+9π32002η2(𝒖˙a𝛀˙×𝒆z)aμ(ad)3/2,\begin{split}\bm{F}_{\parallel}&=6\pi\eta a\bm{u}\bigg{[}\frac{8}{15}\log\bigg{(}\frac{d}{a}\bigg{)}-0.95429\bigg{]}+6\pi\eta a^{2}\bm{e}_{z}\times\bm{\Omega}\bigg{[}\frac{2}{15}\log\bigg{(}\frac{d}{a}\bigg{)}+0.25725\bigg{]}\\ &-\frac{3177\pi^{3}}{6400\sqrt{2}}\frac{\eta^{2}(\bm{u}-a\bm{\Omega}\times\bm{e}_{z})\dot{d}}{\mu}\bigg{(}\frac{a}{d}\bigg{)}^{5/2}+\frac{9\pi^{3}}{200\sqrt{2}}\frac{\eta^{2}(\bm{\dot{u}}-a\bm{\dot{\Omega}}\times\bm{e}_{z})a}{\mu}\bigg{(}\frac{a}{d}\bigg{)}^{3/2},\end{split} (40)

and:

𝑻=8πηa2𝒆z×𝒖[110log(da)0.19296]+8πηa3𝛀[25log(da)0.37085]+3177π364002η2(𝒖a𝛀×𝒆z)ad˙μ(ad)5/29π32002η2(𝒖˙a𝛀˙×𝒆z)a2μ(ad)3/2,\begin{split}\bm{T}_{\parallel}&=8\pi\eta a^{2}\bm{e}_{z}\times\bm{u}\bigg{[}-\frac{1}{10}\log\bigg{(}\frac{d}{a}\bigg{)}-0.19296\bigg{]}+8\pi\eta a^{3}\bm{\Omega}\bigg{[}\frac{2}{5}\log\bigg{(}\frac{d}{a}\bigg{)}-0.37085\bigg{]}\\ &+\frac{3177\pi^{3}}{6400\sqrt{2}}\frac{\eta^{2}(\bm{u}-a\bm{\Omega}\times\bm{e}_{z})a\dot{d}}{\mu}\bigg{(}\frac{a}{d}\bigg{)}^{5/2}-\frac{9\pi^{3}}{200\sqrt{2}}\frac{\eta^{2}(\bm{\dot{u}}-a\bm{\dot{\Omega}}\times\bm{e}_{z})a^{2}}{\mu}\bigg{(}\frac{a}{d}\bigg{)}^{3/2},\end{split} (41)

where we have invoked the force and torque induced by the rotation of a sphere near a rigid wall (Goldman et al., 1967; Urzay, 2010) and where the analytical prefactors are computed in the appendix. We stress that the expressions of the EHD forces and torque for a sphere purely translating near thin elastic substrates, as derived in appendices C and D, can be generalized to further include the sphere’s rotation by similarly following the transformation 𝒖(t)𝒖(t)a𝛀(t)×𝒆z\bm{u}(t)\rightarrow\bm{u}(t)-a\bm{\Omega}(t)\times\bm{e}_{z}.

V Conclusion

We developed a soft-lubrication model in order to compute the EHD interactions exerted on an immersed sphere undergoing both translational and rotational motions near various types of elastic walls. The deformation of the surface was assumed to be small, which allowed us to employ a perturbation analysis in order to obtain the leading-order EHD forces and torque. The obtained interaction matrix exhibits a qualitatively similar form as the one found for a two-dimensional cylinder moving near a thin compressible substrate (Salez and Mahadevan, 2015). In both cases, the EHD coupling is nonlinear and generates quadratic terms in the sphere velocity, thus breaking the time-reversal symmetry of the Stokes equations. In addition, original inertial-like terms proportional to the acceleration of the sphere are found – despite the assumption of steady flows. Therefore, while the quantitative details such as numerical prefactors and exponents differ in 3D and when using more realistic constitutive elastic responses, we expect that the typical zoology of trajectories identified previously (Salez and Mahadevan, 2015) will also hold for spherical objects – and will even be extended with the added degree of freedom. As such, the asymptotic predictions obtained here may open new perspectives in colloidal science and biophysics, through the understanding and control of the emerging interactions within soft confinement or assemblies.

Acknowledgments

We thank Abelhamid Maali, Zaicheng Zhang, Howard Stone and Bhargav Rallabandi for stimulating discussions.

Fundings

The authors acknowledge funding from the Agence Nationale de la Recherche (ANR-21-ERCC-0010-01 EMetBrown) and from the Jean Langlois foundation.

Declaration of Interests

The authors report no conflict of interest.

Appendix A Lorentz reciprocal theorem: normal force

In this appendix, we compute the first-order normal EHD force using the Lorentz reciprocal theorem for Stokes flows (Rallabandi et al., 2017, 2018; Daddi-Moussa-Ider et al., 2018; Masoud and Stone, 2019), in order to recover analytically the numerical prefactors obtained in the main text. To do so, we introduce the model problem of a sphere moving in a viscous fluid and towards an immobile, rigid, planar surface. We note 𝑽^=V^𝒆z\bm{\hat{V}}_{\perp}=-\hat{V}_{\perp}\bm{e}_{z} the velocity at the particle surface 𝒮0\mathcal{S}_{0}, and we assume a no-slip boundary condition at the undeformed wall surface 𝒮w\mathcal{S}_{\text{w}} located at z=0z=0 (see Fig. 1). The viscous stress and velocity fields of the model problem follow the steady, incompressible Stokes equations 𝝈^=𝟎\bm{\nabla}\cdot\bm{\hat{\sigma}}_{\perp}=\bm{0} and 𝒗^=0\bm{\nabla}\cdot\bm{\hat{v}}_{\perp}=0, and we use the lubrication approximation. In this framework, the stress tensor is 𝝈^p^𝐈+η𝒆zz𝒗^\bm{\hat{\sigma}}_{\perp}\simeq-\hat{p}_{\perp}\mathbf{I}+\eta\bm{e}_{z}\partial_{z}\bm{\hat{v}}_{\perp}, with:

p^(𝒓)=3ηV^ah^2(𝒓),𝒗^(𝒓,z)=p^(𝒓)2ηz(zh^(𝒓)),h^(𝒓)=d+𝒓22a.\hat{p}_{\perp}(\bm{r})=\frac{3\eta\hat{V}_{\perp}a}{\hat{h}^{2}(\bm{r})},\quad\quad\bm{\hat{v}}_{\perp}(\bm{r},z)=\frac{\bm{\nabla}\hat{p}_{\perp}(\bm{r})}{2\eta}z(z-\hat{h}(\bm{r})),\quad\quad\hat{h}(\bm{r})=d+\frac{\bm{r}^{2}}{2a}. (42)

The Lorentz reciprocal theorem states that:

𝒮𝒏𝝈𝒗^ds=𝒮𝒏𝝈^𝒗ds,\int_{\mathcal{S}}\bm{n}\cdot\bm{\sigma}\cdot\bm{\hat{v}}_{\perp}\,\mathrm{d}s=\int_{\mathcal{S}}\bm{n}\cdot\bm{\hat{\sigma}}_{\perp}\cdot\bm{v}\,\mathrm{d}s, (43)

where 𝒮=𝒮0+𝒮w+𝒮\mathcal{S}=\mathcal{S}_{0}+\mathcal{S}_{\text{w}}+\mathcal{S}_{\infty} is the total surface bounding the flow, and 𝒮\mathcal{S}_{\infty} is the surface located at 𝒓\bm{r}\rightarrow\infty. The latter does not contribute here. Using the boundary conditions for the model problem, we get:

𝑽^𝑭=V^Fz=𝒮𝒏𝝈^𝒗ds.\bm{\hat{V}}_{\perp}\cdot\bm{F}=-\hat{V}_{\perp}F_{z}=\int_{\mathcal{S}}\bm{n}\cdot\bm{\hat{\sigma}}_{\perp}\cdot\bm{v}\,\mathrm{d}s. (44)

To find the force exerted on the sphere in the real problem, one needs to specify the boundary conditions for the real velocity field. Here, we assume that the sphere does not rotate, and we describe the flow in the translating reference frame of the particle. The no-slip boundary condition thus reads 𝒗=𝟎\bm{v}=\bm{0} on 𝒮0\mathcal{S}_{0}. We further assume a small deformation of the wall, so that the velocity field at the undeformed wall surface can be obtained using the Taylor expansion:

𝒗|z=0=𝒗|z=δδz𝒗0|z=0=u𝒆xd˙𝒆z+(tux)δ𝒆zδz𝒗0|z=0,\begin{split}\bm{v}|_{z=0}&=\bm{v}|_{z=\delta}-\delta\partial_{z}\bm{v}_{0}|_{z=0}\\ &=-u\bm{e}_{x}-\dot{d}\bm{e}_{z}+(\partial_{t}-u\partial_{x})\delta\bm{e}_{z}-\delta\partial_{z}\bm{v}_{0}|_{z=0},\end{split} (45)

where 𝒗0\bm{v}_{0} is the zeroth-order velocity field near a rigid surface. Using results from the main text, we find:

z𝒗0|z=0=3d˙r(d+r22a)2𝒆r+2u5(d+r22a)((76dd+r22a)cosθ𝒆rsinθ𝒆θ).\partial_{z}\bm{v}_{0}|_{z=0}=-\frac{3\dot{d}r}{(d+\frac{r^{2}}{2a})^{2}}\bm{e}_{r}+\frac{2u}{5(d+\frac{r^{2}}{2a})}\left(\left(7-\frac{6d}{d+\frac{r^{2}}{2a}}\right)\cos\theta\bm{e}_{r}-\sin\theta\bm{e}_{\theta}\right). (46)

Combining Eqs. (42) and (44), we find the normal force:

Fz=1V^2(p^(d˙+tδuxδ)+ηz𝒗^|z=0z𝒗0|z=0δ)d𝒓.F_{z}=\frac{1}{\hat{V}_{\perp}}\int_{\mathbb{R}^{2}}\bigg{(}\hat{p}_{\perp}(-\dot{d}+\partial_{t}\delta-u\partial_{x}\delta)+\eta\partial_{z}\bm{\hat{v}}_{\perp}|_{z=0}\cdot\partial_{z}\bm{v}_{0}|_{z=0}\delta\bigg{)}\,\textrm{d}\bm{r}. (47)

After some algebra, and computing the integral in Fourier space, we recover the same expression as in Eq. (30), that reads:

Fz=6πηa2d˙d+Aη2u2(λ+2μ)μ(λ+μ)(ad)5/2Bη2d˙2(λ+2μ)μ(λ+μ)(ad)7/2+Cη2d¨a(λ+2μ)μ(λ+μ)(ad)5/2,F_{z}=-\frac{6\pi\eta a^{2}\dot{d}}{d}+A\frac{\eta^{2}u^{2}(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{5/2}-B\frac{\eta^{2}\dot{d}^{2}(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{7/2}+C\frac{\eta^{2}\ddot{d}a(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{5/2}, (48)

where the numerical coefficients A,B,CA,B,C can be found analytically as:

A=9π2520k2K0(k)(2K1(k)+kK2(k))kdk=243π3128002,A=\frac{9\pi}{25\sqrt{2}}\int_{0}^{\infty}k^{2}K_{0}(k)\bigg{(}-2K_{1}(k)+kK_{2}(k)\bigg{)}\,k\textrm{d}k=\frac{243\pi^{3}}{12800\sqrt{2}}, (49)
B=9π20k2K1(k)(K2(k)kK3(k)8)kdk=3915π320482,B=9\pi\sqrt{2}\int_{0}^{\infty}k^{2}K_{1}(k)\bigg{(}K_{2}(k)-\frac{kK_{3}(k)}{8}\bigg{)}\,k\textrm{d}k=\frac{3915\pi^{3}}{2048\sqrt{2}}, (50)
C=92π20k2K12(k)dk=27π3322,C=\frac{9\sqrt{2}\pi}{2}\int_{0}^{\infty}k^{2}K_{1}^{2}(k)\,\textrm{d}k=\frac{27\pi^{3}}{32\sqrt{2}}, (51)

and where KiK_{i} is the modified Bessel function of the second kind of order ii (Abramowitz and Stegun, 1964).

Appendix B Lorentz reciprocal theorem: tangential force

In order to compute the tangential force acting on the particle, we apply the Lorentz reciprocal theorem, but we introduce a different model problem with respect to the previous section. We consider a sphere translating parallel to a rigid immobile substrate, with a velocity V^\hat{V}_{\parallel} along the xx-axis, and no-slip boundary conditions at both the sphere and substrate surfaces. The velocity and stress fields are denoted 𝝈^\bm{\hat{\sigma}}_{\parallel} and 𝒗^\bm{\hat{v}}_{\parallel}, respectively, and are solutions of the Stokes equations. The lubrication approximation is used here. The solution reads:

p^(𝒓)=6ηV^rcosθ5h^2(𝒓),𝒗^(𝒓,z)=p^(𝒓)2ηz(zh^(𝒓))+𝑽^zh^(𝒓),\hat{p}_{\parallel}(\bm{r})=\frac{6\eta\hat{V}_{\parallel}r\cos\theta}{5\hat{h}^{2}(\bm{r})},\quad\quad\bm{\hat{v}}_{\parallel}(\bm{r},z)=\frac{\bm{\nabla}\hat{p}_{\parallel}(\bm{r})}{2\eta}z(z-\hat{h}(\bm{r}))+\bm{\hat{V}}_{\parallel}\,\frac{z}{\hat{h}(\bm{r})}, (52)

as shown in the main text. The Lorentz reciprocal theorem leads to:

𝑽^𝑭=V^Fx=𝒮𝒏𝝈^𝒗ds.\bm{\hat{V}}_{\parallel}\cdot\bm{F}=\hat{V}_{\parallel}F_{x}=\int_{\mathcal{S}}\bm{n}\cdot\bm{\hat{\sigma}}_{\parallel}\cdot\bm{v}\,\mathrm{d}s. (53)

Using the lubrication expression of the stress tensor of the model problem, 𝝈^p^𝐈+η𝒆zz𝒗^\bm{\hat{\sigma}}_{\parallel}\simeq-\hat{p}_{\parallel}\mathbf{I}+\eta\bm{e}_{z}\partial_{z}\bm{\hat{v}}_{\parallel}, we get an expression for the tangential force as:

Fx=1V^2[ηz𝒗^u(t)𝒆xp^(tu(t)x)δη(z𝒗^z𝒗0)δ]d𝒓.F_{x}=\frac{1}{\hat{V}_{\parallel}}\int_{\mathbb{R}^{2}}\bigg{[}-\eta\partial_{z}\bm{\hat{v}}_{\parallel}\cdot u(t)\bm{e}_{x}-\hat{p}_{\parallel}(\partial_{t}-u(t)\partial_{x})\delta-\eta(\partial_{z}\bm{\hat{v}}_{\parallel}\cdot\partial_{z}\bm{v}_{0})\delta\bigg{]}\textrm{d}\bm{r}. (54)

For the same reason as the one invoked in the main text, the zeroth-order tangential drag force (i.e. the integral of ηz𝒗^u(t)𝒆x-\eta\partial_{z}\bm{\hat{v}}_{\parallel}\cdot u(t)\bm{e}_{x} in Eq. (54)) cannot be computed here as the integral diverges within the lubrication approximation. In contrast, the first-order EHD force is well defined in the lubrication framework and can be computed in Fourier space using Parseval’s theorem, leading to:

κFx,1=3177π364002η2ud˙(λ+2μ)μ(λ+μ)(ad)5/2+9π32002η2u˙a(λ+2μ)μ(λ+μ)(ad)3/2.\kappa F_{x,1}=-\frac{3177\pi^{3}}{6400\sqrt{2}}\frac{\eta^{2}u\dot{d}(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{5/2}+\frac{9\pi^{3}}{200\sqrt{2}}\frac{\eta^{2}\dot{u}a(\lambda+2\mu)}{\mu(\lambda+\mu)}\bigg{(}\frac{a}{d}\bigg{)}^{3/2}. (55)

Appendix C Thin compressible substrate

Refer to caption
Figure 5: Dimensionless deformation fields at the free surface of two soft substrates, for a sphere placed at a unit distance D=1D=1 and for two modes of motion. In a) and b), the substrate’s mechanical response follows the Winkler foundation (see Eq. (56)). In c) and d), the substrate’s mechanical response is the one of a thin incompressible layer (see Eq. (67)).

In this appendix, we derive the EHD interactions exerted on a sphere immersed in a viscous fluid and near a thin compressible substrate of thickness hsubh_{\textrm{sub}}. The deformation field follows the Winkler foundation:

δ(𝒓,t)=hsub(2μ+λ)p(𝒓,t),\delta(\bm{r},t)=-\frac{h_{\textrm{sub}}}{(2\mu+\lambda)}p(\bm{r},t), (56)

which is valid for substrates of thickness smaller than the typical extent of the pressure field, namely the hydrodynamic radius 2ad\sqrt{2ad} (Leroy and Charlaix, 2011; Chandler and Vella, 2020; Kargar-Estahbanati and Rallabandi, 2021). We perform the same asymptotic expansion as the one in the main text, defining the Winkler dimensionless compliance as (Salez and Mahadevan, 2015):

κW=2hsubηua1/2d5/2(2μ+λ).\kappa^{\textrm{W}}=\frac{\sqrt{2}h_{\textrm{sub}}\eta u^{*}a^{1/2}}{d^{*5/2}(2\mu+\lambda)}. (57)

The first-order substrate deformation, or equivalently here the zeroth-order pressure, reads:

H1W(𝑹,T)=P0(𝑹,T)=3D˙2(D+R2)2+6RUcosθ5(D+R2)2.H_{1}^{\text{W}}(\bm{R},T)=P_{0}(\bm{R},T)=\frac{3\dot{D}}{2(D+R^{2})^{2}}+\frac{6RU\cos\theta}{5(D+R^{2})^{2}}. (58)

The first-order deformation fields are plotted in Fig. 5a) and b) for tangential and normal motions of the sphere, respectively. The deformation exhibits the same structure as the one in Fig. 2 for semi-infinite substrates, but the lateral extent of the deformation is narrower. This is expected as the deformation response induced by a given applied pressure is local for a thin compressible layer (see Eq. (56)), while semi-infinite substrates display a non-local response due to the convolution of the pressure with their Green’s function (see Eq. (3)). The first-order pressure correction follows the same type of equation as in the main text:

.P1W=F0W(R,T)+F1W(R,T)cosθ+F2W(R,T)cos2θ,\mathcal{L}.P_{1}^{\textrm{W}}=F_{0}^{\textrm{W}}(R,T)+F_{1}^{\textrm{W}}(R,T)\cos\theta+F_{2}^{\textrm{W}}(R,T)\cos 2\theta, (59)

with:

F0W(R,T)=144R2U225(D+R2)7[D26DR2+R4]+18R2D˙2(D+R2)7[5D4R2]18R2D¨(D+R2)5,F_{0}^{\textrm{W}}(R,T)=-\frac{144R^{2}U^{2}}{25(D+R^{2})^{7}}\bigg{[}D^{2}-6DR^{2}+R^{4}\bigg{]}+\frac{18R^{2}\dot{D}^{2}}{(D+R^{2})^{7}}\bigg{[}5D-4R^{2}\bigg{]}-\frac{18R^{2}\ddot{D}}{(D+R^{2})^{5}}, (60)

and:

F1W(R,T)=216R3UD˙5(D+R2)7[5D+R2]+72RU˙5(D+R2)5.F_{1}^{\textrm{W}}(R,T)=\frac{216R^{3}U\dot{D}}{5(D+R^{2})^{7}}\bigg{[}-5D+R^{2}\bigg{]}+\frac{72R\dot{U}}{5(D+R^{2})^{5}}. (61)

We note that F2WF_{2}^{\textrm{W}} does not contribute for the forces and torque. The isotropic component of the pressure can be found analytically, using polynomial fractions, as:

P1W,(0)(R,T)=912575Y2(1+Y2)5U2D434071+55Y2+30Y4(1+Y2)5D˙2D5+321(1+Y2)3D¨D4,P_{1}^{\textrm{W},(0)}(R,T)=\frac{9}{125}\frac{7-5Y^{2}}{(1+Y^{2})^{5}}\frac{U^{2}}{D^{4}}-\frac{3}{40}\frac{71+55Y^{2}+30Y^{4}}{(1+Y^{2})^{5}}\frac{\dot{D}^{2}}{D^{5}}+\frac{3}{2}\frac{1}{(1+Y^{2})^{3}}\frac{\ddot{D}}{D^{4}}, (62)

where Y=R/D1/2Y=R/D^{1/2} is the self-similar variable. However, the first angular component of the pressure does not exhibit such an analytical solution, and is thus found by numerical integration of two scaling functions. Its general expression reads:

P1W,(1)(R,T)=UD˙D9/2ϕUD˙W(RD1/2)+U˙D7/2ϕU˙W(RD1/2).P_{1}^{\textrm{W},(1)}(R,T)=\frac{U\dot{D}}{D^{9/2}}\phi_{U\dot{D}}^{\textrm{W}}\bigg{(}\frac{R}{D^{1/2}}\bigg{)}+\frac{\dot{U}}{D^{7/2}}\phi_{\dot{U}}^{\textrm{W}}\bigg{(}\frac{R}{D^{1/2}}\bigg{)}. (63)

Following the same calculation as in the main text, we find the normal force as:

FzW=6πηa2d˙d+48π125η2u2hsuba(2μ+λ)(ad)372π5η2d˙2hsuba(2μ+λ)(ad)4+6πη2d¨hsub(2μ+λ)(ad)3.F_{z}^{\textrm{W}}=-\frac{6\pi\eta a^{2}\dot{d}}{d}+\frac{48\pi}{125}\frac{\eta^{2}u^{2}h_{\textrm{sub}}}{a(2\mu+\lambda)}\bigg{(}\frac{a}{d}\bigg{)}^{3}-\frac{72\pi}{5}\frac{\eta^{2}\dot{d}^{2}h_{\textrm{sub}}}{a(2\mu+\lambda)}\bigg{(}\frac{a}{d}\bigg{)}^{4}+\frac{6\pi\eta^{2}\ddot{d}h_{\textrm{sub}}}{(2\mu+\lambda)}\bigg{(}\frac{a}{d}\bigg{)}^{3}. (64)

We stress that the prefactors 48π/12548\pi/125 and 6π6\pi are in agreement with the results in Urzay et al. (2007) and Leroy and Charlaix (2011), respectively. Similarly, the force along xx reads:

FxW=6πηau(815log(da)0.95429)23.9η2ud˙hsuba(2μ+λ)(ad)3+4.520η2u˙hsub(2μ+λ)(ad)2.F_{x}^{\textrm{W}}=6\pi\eta au\bigg{(}\frac{8}{15}\log\left(\frac{d}{a}\right)-0.95429\bigg{)}-23.9\frac{\eta^{2}u\dot{d}h_{\textrm{sub}}}{a(2\mu+\lambda)}\bigg{(}\frac{a}{d}\bigg{)}^{3}+4.520\frac{\eta^{2}\dot{u}h_{\textrm{sub}}}{(2\mu+\lambda)}\bigg{(}\frac{a}{d}\bigg{)}^{2}. (65)

The torque can be evaluated as well, and reads:

TyW=8πηua2(110log(da)0.19296)+12.2η2ud˙hsub(2μ+λ)(ad)31.51η2u˙ahsub(2μ+λ)(ad)2.T_{y}^{\textrm{W}}=8\pi\eta ua^{2}\bigg{(}-\frac{1}{10}\log\left(\frac{d}{a}\right)-0.19296\bigg{)}+12.2\frac{\eta^{2}u\dot{d}h_{\textrm{sub}}}{(2\mu+\lambda)}\bigg{(}\frac{a}{d}\bigg{)}^{3}-1.51\frac{\eta^{2}\dot{u}ah_{\textrm{sub}}}{(2\mu+\lambda)}\bigg{(}\frac{a}{d}\bigg{)}^{2}. (66)

All the prefactors for the EHD corrections of the tangential force and torque have been found numerically. Finally, following the approach in the main text, it is straightforward to generalize Eqs. (64), (65) and (66) in order to incorporate rotation.

Appendix D Thin incompressible substrate

In this appendix, we suppose that the substrate of thickness hsubh_{\textrm{sub}} is incompressible, i.e. of Poisson ratio ν=1/2\nu=1/2, which means that the first Lamé coefficient λ\lambda is infinite. In this situation, the Winkler foundation is not valid. Instead, the mechanical response of a thin substrate follows the relation (Leroy and Charlaix (2011); Chandler and Vella (2020)):

δ(𝒓,t)=hsub33μ2p(𝒓,t),\delta(\bm{r},t)=\frac{h^{3}_{\textrm{sub}}}{3\mu}\bm{\nabla}^{2}p(\bm{r},t), (67)

where 2\bm{\nabla}^{2} denotes the 2D Laplacian operator in the (x,y)(x,y) plan. We perform the same asymptotic expansion as in the main text, defining the thin-incompressible dimensionless compliance as:

κt-i=ηuhsub332μd7/2a1/2.\kappa^{\textrm{t-i}}=\frac{\eta u^{*}h_{\textrm{sub}}^{3}}{3\sqrt{2}\mu d^{*7/2}a^{1/2}}. (68)

The first-order substrate deformation reads:

H1t-i(𝑹,T)=2P0(𝑹,T)=12D˙(D2R2)(D+R2)4+48RU(2DR2)cosθ5(D+R2)4.H_{1}^{\text{t-i}}(\bm{R},T)=-\bm{\nabla}^{2}P_{0}(\bm{R},T)=-\frac{12\dot{D}(D-2R^{2})}{(D+R^{2})^{4}}+\frac{48RU(2D-R^{2})\cos\theta}{5(D+R^{2})^{4}}. (69)

The deformation fields are plotted in Figs. 5 c) and d), for tangential and normal motions of the sphere, respectively. The first-order pressure correction follows the same type of equation as in the main text:

.P1t-i=F0t-i(R,T)+F1t-i(R,T)cosθ+F2t-i(R,T)cos2θ,\mathcal{L}.P_{1}^{\textrm{t-i}}=F_{0}^{\textrm{t-i}}(R,T)+F_{1}^{\textrm{t-i}}(R,T)\cos\theta+F_{2}^{\textrm{t-i}}(R,T)\cos 2\theta, (70)

with:

F0t-i(R,T)=1152R2U2(R22D)(+2D211R2D+2R4)25(D+R2)9+432R2(2D(D5R2)+3R4)D˙2(D+R2)9+144R2(2R2D)D¨(D+R2)7,\begin{split}F_{0}^{\textrm{t-i}}(R,T)&=\frac{1152R^{2}U^{2}\left(R^{2}-2D\right)\left(+2D^{2}-11R^{2}D+2R^{4}\right)}{25\left(D+R^{2}\right)^{9}}\\ &+\frac{432R^{2}\left(2D\left(D-5R^{2}\right)+3R^{4}\right)\dot{D}^{2}}{\left(D+R^{2}\right)^{9}}+\frac{144R^{2}\left(2R^{2}-D\right)\ddot{D}}{\left(D+R^{2}\right)^{7}},\end{split} (71)

and:

F1t-i(R,T)=2592R3D˙U(7D212R2D+R4)5(D+R2)9576R3U˙(2D+R2)5(D+R2)7.F_{1}^{\textrm{t-i}}(R,T)=-\frac{2592R^{3}\dot{D}U\left(7D^{2}-12R^{2}D+R^{4}\right)}{5\left(D+R^{2}\right)^{9}}-\frac{576R^{3}\dot{U}\left(-2D+R^{2}\right)}{5\left(D+R^{2}\right)^{7}}. (72)

We note that F2t-iF_{2}^{\textrm{t-i}} does not contribute for the forces and torque. The isotropic component of the pressure can be found analytically, using polynomial fractions, as:

P1t-i,(0)(R,T)=288(7Y421Y2+17)875(1+Y2)7U2D5+126Y21987(1+Y2)7D˙2D6+365(1+Y2)5D¨D5,P_{1}^{\textrm{t-i},(0)}(R,T)=\frac{288\left(7Y^{4}-21Y^{2}+17\right)}{875\left(1+Y^{2}\right)^{7}}\frac{U^{2}}{D^{5}}+\frac{126Y^{2}-198}{7\left(1+Y^{2}\right)^{7}}\frac{\dot{D}^{2}}{D^{6}}+\frac{36}{5\left(1+Y^{2}\right)^{5}}\frac{\ddot{D}}{D^{5}}, (73)

where Y=R/D1/2Y=R/D^{1/2} is the self-similar variable. However, the first angular component of the pressure does not exhibit such an analytical solution, and is thus found by numerical integration of two scaling functions. Its general expression reads:

P1t-i,(1)(R,T)=UD˙D11/2ϕUD˙t-i(RD1/2)+U˙D9/2ϕU˙t-i(RD1/2).P_{1}^{\textrm{t-i},(1)}(R,T)=\frac{U\dot{D}}{D^{11/2}}\phi_{U\dot{D}}^{\textrm{t-i}}\bigg{(}\frac{R}{D^{1/2}}\bigg{)}+\frac{\dot{U}}{D^{9/2}}\phi_{\dot{U}}^{\textrm{t-i}}\bigg{(}\frac{R}{D^{1/2}}\bigg{)}. (74)

Following the same calculation as in the main text, we find the normal force as:

Fzt-i=6πηa2d˙d+432π875η2u2hsub3a3μ(ad)4192π35η2d˙2hsub3a3μ(ad)5+12π5η2d¨hsub3a2μ(ad)4.F_{z}^{\textrm{t-i}}=-\frac{6\pi\eta a^{2}\dot{d}}{d}+\frac{432\pi}{875}\frac{\eta^{2}u^{2}h_{\textrm{sub}}^{3}}{a^{3}\mu}\bigg{(}\frac{a}{d}\bigg{)}^{4}-\frac{192\pi}{35}\frac{\eta^{2}\dot{d}^{2}h_{\textrm{sub}}^{3}}{a^{3}\mu}\bigg{(}\frac{a}{d}\bigg{)}^{5}+\frac{12\pi}{5}\frac{\eta^{2}\ddot{d}h_{\textrm{sub}}^{3}}{a^{2}\mu}\bigg{(}\frac{a}{d}\bigg{)}^{4}. (75)

We stress that the prefactor 12π/512\pi/5 is consistent with the linear-response theory in Leroy and Charlaix (2011). Similarly, the force along xx reads:

Fxt-i=6πηau(815log(da)0.95429)12.2η2ud˙hsub3a3μ(ad)4+2.41η2u˙hsub3a2μ(ad)3.F_{x}^{\textrm{t-i}}=6\pi\eta au\bigg{(}\frac{8}{15}\log\left(\frac{d}{a}\right)-0.95429\bigg{)}-12.2\frac{\eta^{2}u\dot{d}h_{\textrm{sub}}^{3}}{a^{3}\mu}\bigg{(}\frac{a}{d}\bigg{)}^{4}+2.41\frac{\eta^{2}\dot{u}h_{\textrm{sub}}^{3}}{a^{2}\mu}\bigg{(}\frac{a}{d}\bigg{)}^{3}. (76)

The torque can be evaluated as well, and reads:

Tyt-i=8πηua2(110log(da)0.19296)+7.75η2ud˙hsub3a2μ(ad)40.804η2u˙hsub3aμ(ad)3.T_{y}^{\textrm{t-i}}=8\pi\eta ua^{2}\bigg{(}-\frac{1}{10}\log\left(\frac{d}{a}\right)-0.19296\bigg{)}+7.75\frac{\eta^{2}u\dot{d}h_{\textrm{sub}}^{3}}{a^{2}\mu}\bigg{(}\frac{a}{d}\bigg{)}^{4}-0.804\frac{\eta^{2}\dot{u}h_{\textrm{sub}}^{3}}{a\mu}\bigg{(}\frac{a}{d}\bigg{)}^{3}. (77)

Here again, the prefactors of the transverse force and torque are found using the Lorentz reciprocal theorem, as discussed above in the appendix. We stress that the thin-incompressible limit is mathematically valid for strictly incompressible substrates, but that its range of application is limited in practice. Indeed, usual elastomers and gels, that are considered as almost incompressible, have a Poisson ratio close to ν0.49\nu\simeq 0.49, and thus a tiny but finite compressibility. In recent studies Saintyves et al. (2016); Rallabandi et al. (2017); Saintyves et al. (2020), it has been observed that the mechanical response of very thin incompressible elastic substrates is better described by the Winkler foundation (i.e. thin and compressible) than the thin-incompressible limit discussed here. This observation has then been established on solid theoretical grounds for the EHD lift Chandler and Vella (2020), and is intimately rooted in the structure of the elastic Green’s function. An empirical scaling, based on the numerical calculation of the EHD lift coefficient, has been derived subsequently Kargar-Estahbanati and Rallabandi (2021) and suggests that the thin-incompressible model is valid for thicknesses comprised in the range 73(1/2ν)1/2hsub/2ad0.12\frac{\sqrt{7}}{3}(1/2-\nu)^{1/2}\ll h_{\text{sub}}/\sqrt{2ad}\leq 0.12. For ν=0.49\nu=0.49, the lower bound of the latter range is 0.0880.088, which confirms that the validity window of the thin-incompressible model is limited.

References

  • Dowson and Higginson (2014) D. Dowson and G. R. Higginson, Elasto-hydrodynamic lubrication: international series on materials science and technology (Elsevier, 2014).
  • Brodsky and Kanamori (2001) E. E. Brodsky and H. Kanamori, Journal of Geophysical Research: Solid Earth 106, 16357 (2001).
  • Campbell (1989) C. S. Campbell, The Journal of Geology 97, 653 (1989).
  • Karan et al. (2018) P. Karan, J. Chakraborty, and S. Chakraborty, Journal of the Indian Institute of Science 98, 159 (2018).
  • Jones et al. (2008) M. B. Jones, G. Fulford, C. P. Please, D. McElwain, and M. J. Collins, Bulletin of mathematical biology 70, 323 (2008).
  • Mow et al. (1984) V. C. Mow, M. H. Holmes, and W. M. Lai, Journal of biomechanics 17, 377 (1984).
  • Jahn et al. (2016) S. Jahn, J. Seror, and J. Klein, Annual Review of Biomedical Engineering 18, 235 (2016).
  • Gong (2006) J. P. Gong, Soft matter 2, 544 (2006).
  • Sekimoto and Leibler (1993) K. Sekimoto and L. Leibler, EPL (Europhysics Letters) 23, 113 (1993).
  • Moyle et al. (2020) N. Moyle, H. Wu, C. Khripin, F. Bremond, C.-Y. Hui, and A. Jagota, Soft matter 16, 1627 (2020).
  • Wu et al. (2020) H. Wu, N. Moyle, A. Jagota, and C.-Y. Hui, Soft Matter 16, 2760 (2020).
  • Hui et al. (2021) C.-Y. Hui, H. Wu, A. Jagota, and C. Khripin, Tribology Letters 69, 1 (2021).
  • Davis et al. (1986) R. H. Davis, J.-M. Serayssol, and E. Hinch, Journal of Fluid Mechanics 163, 479 (1986).
  • Gondret et al. (2002) P. Gondret, M. Lance, and L. Petit, Physics of fluids 14, 643 (2002).
  • Tan et al. (2019) M. R. Tan, Y. Wang, and J. Frechette, Physical Review Fluids 4, 084305 (2019).
  • Meeker et al. (2004) S. P. Meeker, R. T. Bonnecaze, and M. Cloitre, Physical Review Letters 92, 198302 (2004).
  • Snoeijer et al. (2013) J. H. Snoeijer, J. Eggers, and C. H. Venner, Physics of Fluids 25, 101705 (2013).
  • Chan et al. (2009) D. Y. Chan, E. Klaseboer, and R. Manica, Soft Matter 5, 2858 (2009).
  • Vakarelski et al. (2010) I. U. Vakarelski, R. Manica, X. Tang, S. J. O’Shea, G. W. Stevens, F. Grieser, R. R. Dagastine, and D. Y. Chan, Proceedings of the National Academy of Sciences 107, 11177 (2010).
  • Leroy and Charlaix (2011) S. Leroy and E. Charlaix, Journal of Fluid Mechanics 674, 389 (2011).
  • Leroy et al. (2012) S. Leroy, A. Steinberger, C. Cottin-Bizonne, F. Restagno, L. Léger, and É. Charlaix, Physical Review Letters 108, 264501 (2012).
  • Villey et al. (2013) R. Villey, E. Martinot, C. Cottin-Bizonne, M. Phaner-Goutorbe, L. Léger, F. Restagno, and E. Charlaix, Physical Review Letters 111, 215701 (2013).
  • Wang et al. (2015) Y. Wang, C. Dhong, and J. Frechette, Physical Review Letters 115, 248302 (2015).
  • Wang et al. (2017a) Y. Wang, G. A. Pilkington, C. Dhong, and J. Frechette, Current Opinion in Colloid & Interface Science 27, 43 (2017a).
  • Wang et al. (2017b) Y. Wang, M. R. Tan, and J. Frechette, Soft Matter 13, 6718 (2017b).
  • Guan et al. (2017) D. Guan, E. Charlaix, R. Z. Qi, and P. Tong, Physical Review Applied 8, 044010 (2017).
  • Wang et al. (2018) Y. Wang, B. Zeng, H. T. Alem, Z. Zhang, E. Charlaix, and A. Maali, Langmuir 34, 1371 (2018).
  • Lainé et al. (2019) A. Lainé, L. Jubin, L. Canale, L. Bocquet, A. Siria, S. H. Donaldson, and A. Niguès, Nanotechnology 30, 195502 (2019).
  • Bertin et al. (2020) V. Bertin, Z. Zhang, R. Boisgard, C. Grauby-Heywang, E. Raphael, T. Salez, and A. Maali, ArXiv:2009.14531 (2020).
  • Happel and Brenner (2012) J. Happel and H. Brenner, Low Reynolds number hydrodynamics: with special applications to particulate media, vol. 1 (Springer Science & Business Media, 2012).
  • Salez and Mahadevan (2015) T. Salez and L. Mahadevan, Journal of Fluid Mechanics 779, 181 (2015).
  • Beaucourt et al. (2004) J. Beaucourt, T. Biben, and C. Misbah, EPL (Europhysics Letters) 67, 676 (2004).
  • Skotheim and Mahadevan (2004) J. Skotheim and L. Mahadevan, Physical Review Letters 92, 245509 (2004).
  • Skotheim and Mahadevan (2005) J. Skotheim and L. Mahadevan, Physics of Fluids 17, 092101 (2005).
  • Urzay et al. (2007) J. Urzay, S. G. Llewellyn Smith, and B. J. Glover, Physics of Fluids 19, 103106 (2007).
  • Urzay (2010) J. Urzay, Journal of fluid mechanics 653, 391 (2010).
  • Grandchamp et al. (2013) X. Grandchamp, G. Coupier, A. Srivastav, C. Minetti, and T. Podgorski, Physical Review Letters 110, 108101 (2013).
  • Abkarian et al. (2002) M. Abkarian, C. Lartigue, and A. Viallat, Physical Review Letters 88, 068103 (2002).
  • Bouchet et al. (2015) A.-S. Bouchet, C. Cazeneuve, N. Baghdadli, G. S. Luengo, and C. Drummond, Macromolecules 48, 2244 (2015).
  • Saintyves et al. (2016) B. Saintyves, T. Jules, T. Salez, and L. Mahadevan, Proceedings of the National Academy of Sciences 113, 5847 (2016), ISSN 0027-8424.
  • Davies et al. (2018) H. S. Davies, D. Débarre, N. El Amri, C. Verdier, R. P. Richter, and L. Bureau, Physical Review Letters 120, 198001 (2018).
  • Rallabandi et al. (2018) B. Rallabandi, N. Oppenheimer, M. Y. B. Zion, and H. A. Stone, Nature Physics 14, 1211 (2018).
  • Vialar et al. (2019) P. Vialar, P. Merzeau, S. Giasson, and C. Drummond, Langmuir 35, 15605 (2019).
  • Zhang et al. (2020) Z. Zhang, V. Bertin, M. Arshad, E. Raphael, T. Salez, and A. Maali, Physical Review Letters 124, 054502 (2020).
  • Essink et al. (2021) M. H. Essink, A. Pandey, S. Karpitschka, C. H. Venner, and J. H. Snoeijer, Journal of Fluid Mechanics 915 (2021).
  • Pandey et al. (2016) A. Pandey, S. Karpitschka, C. H. Venner, and J. H. Snoeijer, Journal of fluid mechanics 799, 433 (2016).
  • Rallabandi et al. (2017) B. Rallabandi, B. Saintyves, T. Jules, T. Salez, C. Schönecker, L. Mahadevan, and H. A. Stone, Physical Review Fluids 2, 074102 (2017).
  • Saintyves et al. (2020) B. Saintyves, B. Rallabandi, T. Jules, J. Ault, T. Salez, C. Schönecker, H. A. Stone, and L. Mahadevan, Soft Matter 16, 4000 (2020).
  • Abramowitz and Stegun (1964) M. Abramowitz and I. A. Stegun, Handbook of mathematical functions with formulas, graphs, and mathematical tables, vol. 55 (US Government printing office, 1964).
  • Wang et al. (2020) Y. Wang, Z. Feng, and J. Frechette, Current Opinion in Colloid & Interface Science p. 101397 (2020).
  • Daddi-Moussa-Ider et al. (2018) A. Daddi-Moussa-Ider, B. Rallabandi, S. Gekle, and H. A. Stone, Physical Review Fluids 3, 084101 (2018).
  • Masoud and Stone (2019) H. Masoud and H. A. Stone, Journal of Fluid Mechanics 879 (2019).
  • Kargar-Estahbanati and Rallabandi (2021) A. Kargar-Estahbanati and B. Rallabandi, Physical Review Fluids 6, 034003 (2021).
  • O’Neill (1964) M. E. O’Neill, Mathematika 11, 67 (1964).
  • O’Neill and Stewartson (1967) M. O’Neill and K. Stewartson, Journal of Fluid Mechanics 27, 705 (1967).
  • Goldman et al. (1967) A. Goldman, R. Cox, and H. Brenner, Chemical engineering science 22, 653 (1967).
  • Chaoui and Feuillebois (2003) M. Chaoui and F. Feuillebois, Quarterly Journal of Mechanics and Applied Mathematics 56, 381 (2003).
  • Chandler and Vella (2020) T. G. Chandler and D. Vella, Proceedings of the Royal Society A 476, 20200551 (2020).