1. Introduction
Wetting phenomena are ubiquitous in nature and play a crucial role in numerous industrial applications, including oil recovery (Tangparitkul et al. Reference Tangparitkul, Charpentier, Pradilla and Harbottle2018), inkjet printing (Park & Moon Reference Park and Moon2006), microfluidics manipulation (Stone, Stroock & Ajdari Reference Stone, Stroock and Ajdari2004) and coating processes (Kajiya et al. Reference Kajiya, Brunet, Royon, Daerr, Receveur and Limat2014). A fundamental problem in wetting dynamics is the motion of the contact line, where the interface between two immiscible fluids meets a solid substrate, and one fluid displaces the other (De Gennes Reference De Gennes1985; Bonn et al. Reference Bonn, Eggers, Indekeu, Meunier and Rolley2009). A well-known paradox arises in hydrodynamic models: imposing the conventional no-slip condition at the solid surface leads to an unphysical, non-integrable singularity in the viscous dissipation at the moving contact line (Huh & Scriven Reference Huh and Scriven1971; Snoeijer & Andreotti Reference Snoeijer and Andreotti2013). To resolve this issue, various models have been proposed over the past decades, including molecular kinetic models, molecular dynamic models, phase field models, and hydrodynamic models with modified boundary conditions; see the literature reviews in Ren & E (Reference Ren and E.2007) and Ren, Trinh & E (Reference Ren, Trinh and E.2015). Among these models, a particularly simple and widely used approach is the Navier slip model, which introduces a small region of length
$\lambda _0$
where slip of the fluids occurs. This regularises the singularity and allows for a consistent description of contact line motion.
In their theoretical studies, Cox (Reference Cox1986) and Voinov (Reference Voinov1977) conducted a three-region matched asymptotic analysis in the limit of small slip length
$\lambda = \lambda _0/L \ll 1$
, where
$\lambda _0$
is the physical slip length, and
$L$
is a characteristic length scale of the flow. For simplicity, we will hereafter refer to the dimensionless parameter
$\lambda$
as the slip length. The three regions are as follows. (i) The outer region, located at a distance of
$\operatorname {O} (1)$
from the contact line, where fluid slip is negligible and the apparent contact angle
$\theta _{\textit{app}}$
is defined. (ii) The inner region, located at a distance
$\operatorname {O} (\lambda)$
from the contact line, where slip occurs and the microscopic contact angle
$\theta _{{m}}$
is defined; the microscopic contact angle typically differs from the apparent one at a moving contact line. (iii) The intermediate region, which spans distances between
$\operatorname {O} (\lambda)$
and
$\operatorname {O} (\epsilon)$
from the contact line. This transition zone connects the inner and outer regions, and captures the bending of the fluid interface. Here, the intermediate length scale
satisfies
$\lambda \ll \epsilon \ll 1$
and tends to zero as
$\lambda \rightarrow 0$
. We refer to Sibley, Nold & Kalliadasis (Reference Sibley, Nold and Kalliadasis2015) for a review of the Cox–Voinov theory and its underlying asymptotic structure. The main result of the asymptotic analysis is a relation between the contact line speed and the apparent contact angle. This relation has been successfully applied to many wetting problems (Hocking Reference Hocking1977; Hocking & Rivers Reference Hocking and Rivers1982; Cox Reference Cox1986). For thin liquid drops, the lubrication approximation has been used to derive simplified models (Hocking Reference Hocking1983; Savva & Kalliadasis Reference Savva and Kalliadasis2009; Ren et al. Reference Ren, Trinh and E.2015).
The Cox–Voinov theory has been widely used to describe contact line motion. Its application requires proper identification of the relevant length scales. The slip length is not determined within the theory itself; rather, it must be prescribed, for example, through calibration against experimental measurements or molecular-scale models. In addition, the apparent contact angle defined in the outer region can be problem-dependent. For spreading drops, for instance, it is commonly defined by fitting a circular arc to the fluid interface, then calculating the angle that it makes with the substrate.
In contrast to the asymptotic matching between different scales employed by Voinov (Reference Voinov1977) and Cox (Reference Cox1986), Snoeijer (Reference Snoeijer2006) and Chan et al. (Reference Chan, Kamal, Snoeijer, Sprittles and Eggers2020) adopted an alternative approach in which the free surface is treated as a perturbation around a straight wedge. Based on the wedge flow solution (Huh & Scriven Reference Huh and Scriven1971), they derived the ‘generalised lubrication’ equation, which provides a description of the interface profile at small capillary numbers. This method avoids the need to distinguish between different asymptotic regions, and the resulting generalised lubrication equation remains valid even for large interface slopes.
In recent years, the field of elastocapillarity – the interplay between elasticity and capillarity – has attracted much research attention. Wetting problems in this context, often referred to as soft wetting, have been considered in a large body of works; see Bico, Reyssat & Roman (Reference Bico, Reyssat and Roman2018) and Andreotti & Snoeijer (Reference Andreotti and Snoeijer2020). Soft wetting differs from classical wetting due to the deformability of the substrate. In this setting, the substrate can deform in response to fluid pressure, viscous stress, and capillary forces exerted at the contact line. Notably, a wetting ridge forms at the contact line on a (visco)elastic solid (Jerison et al. Reference Jerison, Xu, Wilen and Dufresne2011; Style & Dufresne Reference Style and Dufresne2012; Pandey et al. Reference Pandey, Andreotti, Karpitschka, van Zwieten, Van Brummelen and Snoeijer2020), and is associated with interesting phenomena such as viscoelastic braking (Shanahan Reference Shanahan1988) and stick-slip motion of the contact line (Kajiya et al. Reference Kajiya, Daerr, Narita, Royon, Lequeux and Limat2013; Karpitschka et al. Reference Karpitschka, Das, Van Gorcum, Perrin, Andreotti and Snoeijer2015). More recently, the dynamics of thin liquid films on (visco)elastic substrates has been studied (Charitatos & Kumar Reference Charitatos and Kumar2020; Henkel, Snoeijer & Thiele Reference Henkel, Snoeijer and Thiele2021; Tamim & Bostwick 2021, 2023). These works typically employ the lubrication approximation to derive reduced models, and use either a sharp interface or a precursor film approach to model the moving contact line. Various factors affecting the spreading dynamics have been investigated, including fluid evaporation, Marangoni effects and substrate properties.
Among soft wetting problems, the wetting of thin elastic sheets has been an active subject of research. For example, a drop deposited on an ultrathin sheet can induce wrinkling, a phenomenon first reported by Huang et al. (Reference Huang, Juszkiewicz, De Jeu, Cerda, Emrick, Menon and Russell2007). The equilibrium properties of this drop–sheet system, including the wrinkle extent and wavelength, have been studied in detail (Davidovitch et al. Reference Davidovitch, Schroll, Vella, Adda-Bedia and Cerda2011; King et al. Reference King, Schroll, Davidovitch and Menon2012; Schroll et al. Reference Schroll, Adda-Bedia, Cerda, Huang, Menon, Russell, Toga, Vella and Davidovitch2013; Davidovitch & Vella Reference Davidovitch and Vella2018). Following the work of Py et al. (Reference Py, Reverdy, Doppler, Bico, Roman and Baroud2007) on capillary folding, subsequent studies have been carried out to establish the criteria for the reopening of an elastic sheet upon drop evaporation (Péraud & Lauga Reference Péraud and Lauga2014; Brubaker & Lega Reference Brubaker and Lega2015; Li & Ren Reference Li and Ren2024a ,Reference Li and Ren b ). Static interface profiles for contact lines on elastic membranes were studied in Zhang, Yao & Ren (Reference Zhang, Yao and Ren2020). In addition, Bradley investigated the spontaneous transport of drops on elastic sheets through a mechanism called ‘bendotaxis’ (Bradley et al. Reference Bradley, Box, Hewitt and Vella2019). Onsager’s variational principle (Zhang & Qian Reference Zhang and Qian2022) or non-equilibrium thermodynamics (Yao, Zhang & Ren Reference Yao, Zhang and Ren2023) has been used to develop models to further investigate the mechanisms underlying this spontaneous transport.
Despite the extensive literature on soft wetting, a theoretical analysis of contact line motion on elastic sheets remains lacking. In this work, we consider the motion of a thin drop on an elastic sheet. The sheet is thin, stretched, highly bendable, and simply supported in the far field. The liquid height and the sheet deformation are assumed to be small, allowing us to apply the lubrication approximation; see e.g. Hocking (Reference Hocking1983), Ehrhard & Davis (Reference Ehrhard and Davis1991) and Oron, Davis & Bankoff (Reference Oron, Davis and Bankoff1997). Our main objective is to derive a relation – analogous to the Cox–Voinov relation – between the contact line speed and measurable quantities such as the drop radius and the apparent contact angles. Compared with classical wetting, our problem is complicated by the substrate deformability and its coupling with the fluid. The elastic sheet bends moderately beneath the drop but exhibits sharp bending near the contact line. This localised bending occurs on the dimensionless length scale
where
$C_b$
is the bending modulus, and
$\gamma$
is the sheet tension. This scale arises from the balance between the bending resistance force and the tension force. We expect that the sheet bending on this scale influences the contact line dynamics. As a consequence of the sharp bending, we introduce two apparent contact angles: the angle
$\alpha _{\textit{app}}$
that the fluid interface makes with the horizontal plane, and the angle
$\beta _{\textit{app}}$
that the sheet makes with the horizontal plane, both measured at a distance
$\operatorname {O} (1)$
from the contact line; see figure 1. We expect that both angles, rather than
$\theta _{\textit{app}} = \alpha _{\textit{app}} - \beta _{\textit{app}}$
, the angle between the interface and the sheet alone, contribute to determining the contact line dynamics.

Figure 1. Spreading of a thin liquid drop on an elastic sheet. The fluid interface is represented by
$z=h(x,t)$
, the sheet by
$z=g(x,t)$
, and the contact line position by
$x=a(t)$
. The apparent contact angles
$\alpha _{\textit{app}}$
and
$\beta _{\textit{app}}$
are defined relative to the horizontal plane for the fluid interface and the sheet, respectively. In the configuration shown,
$\alpha _{\textit{app}}\gt 0$
and
$\beta _{\textit{app}}\lt 0$
.
We extend the classical Cox–Voinov analysis to account for the sheet bending. We consider the distinguished limit in which the capillary number is
${\textit{Ca}} = \operatorname {O} (\epsilon)$
and the bending length is
$l = \operatorname {O} (\epsilon)$
, where
$\epsilon$
is defined in (1.1). A bending region of size
$\operatorname {O} (\epsilon)$
is introduced where the sheet deformation plays a dominant role. We systematically solve the interface profile and the sheet deformation in the outer, bending and inner regions, and use the intermediate region to connect the solutions in the bending and inner regions. Through this matched asymptotic analysis, we derive a relation that expresses the contact line speed in terms of the apparent contact angles
$\alpha _{\textit{app}}$
and
$\beta _{\textit{app}}$
:
where
$\dot {a}$
is the contact line speed,
$\theta _Y$
is the equilibrium contact angle, and
$\gamma _1$
,
$\gamma _2$
are the effective tensions in the wet and dry parts of the sheet, respectively. This relation reveals that the contact line dynamics is significantly influenced by the substrate rigidity: softer or less stretched sheets lead to retarded spreading and enhanced receding, as confirmed by numerical simulations. Furthermore, despite the assumption
$l = \operatorname {O} (\epsilon)$
in the analysis, the relation (1.3) is independent of the bending modulus. Our numerical simulations further confirm that the contact line dynamics is insensitive to the bending length, and that the relation holds even for extremely bendable sheets with
$l$
as small as
$\operatorname {O} (\epsilon ^2)$
.
The rest of this paper is organised as follows. In § 2, we derive a system of fourth-order equations, together with appropriate boundary and contact line conditions, that govern the evolution of the fluid interface and the elastic sheet. In § 3, we perform a four-region matched asymptotic analysis of the model, and derive a relation describing the contact line motion on an elastic sheet. In § 4, we validate the theoretical results through numerical simulations, and examine retarded spreading, enhanced receding and long-time dynamics. Finally, we conclude the paper in § 5.
2. A thin-film model
We begin with a hydrodynamics model for the drop–sheet system. Consider a two-dimensional liquid drop surrounded by air and sitting on a one-dimensional elastic sheet; see figure 1. The interfacial tension coefficient for the liquid–air interface is
$\gamma _3$
. For simplicity, we assume that the drop is symmetric with respect to
$x=0$
, and that the contact line is located at
$x=\pm a(t)$
. We represent the liquid–air interface by the height function
$z=h(x,t)$
,
$-a(t) \leqslant x \leqslant a(t)$
, and the sheet profile by the height function
$z = g(x,t)$
,
$-\infty \lt x \lt \infty$
. The liquid is assumed to be incompressible and Newtonian, with dynamical viscosity
$\eta$
. Its dynamics is modelled by the stationary Stokes equations. The viscosity of the surrounding air is assumed to be negligible, and its influence on the liquid is neglected. The elastic sheet is modelled using the Willmore bending energy with bending modulus
$C_b$
. Motivated by the experimental set-ups in Huang et al. (Reference Huang, Juszkiewicz, De Jeu, Cerda, Emrick, Menon and Russell2007) and King et al. (Reference King, Schroll, Davidovitch and Menon2012), we assume that the sheet has a very large stretching modulus, is nearly inextensible, and is uniformly stretched with a pre-stress
$\sigma _{\textit{pre}} \gt 0$
. We then define the effective tensions as
where
$\gamma _{i,o},\ i=1,2$
are the interfacial tension coefficients of the liquid–sheet and air–sheet interfaces, respectively. We refer to Yao et al. (Reference Yao, Zhang and Ren2023) for a complete description and the derivation of the hydrodynamics model.
In the following, we consider a thin liquid film that is symmetric about
$x=0$
, and restrict our attention to the half-domain
$x \geqslant 0$
. Let
$L$
denote the characteristic film radius, and
$U$
the characteristic contact line velocity. We assume that the film height is of order
$\delta L$
, where
$\delta \ll 1$
is a small parameter. Furthermore, we assume that the slope of the film height is
$\partial h/ \partial x = \operatorname {O} (\delta)$
. Similarly, we assume
$g = \operatorname {O} (\delta L)$
and
$\partial g / \partial x = \operatorname {O} (\delta)$
. We introduce the rescalings
with the dimensionless variables marked by uppercase letters. We also introduce the dimensionless parameters
Following the standard procedure (see e.g. Oron et al. Reference Oron, Davis and Bankoff1997), we derive from the hydrodynamics model the evolution equations for the interface height
$Z = H(X, T)$
and the sheet profile
$Z = G(X, T)$
:
\begin{gather} \tilde {C}_b\, \partial _X^4 G = \begin{cases} \tilde {\gamma }_1\, \partial _{XX}G + {\textit{Ca}}^{-1}\, \partial _{XX}H, & 0 \lt X \lt A(T), \\[3pt] \tilde {\gamma }_2\, \partial _{XX}G, & A(T) \lt X. \end{cases} \end{gather}
The boundary conditions are as follows. At the origin, symmetry conditions are imposed:
\begin{equation} \left \{ \begin{aligned} & \partial _X H = \partial _{X}^3 H = 0, \\ & \partial _X G = \partial _{X}^3 G = 0, \end{aligned} \right . \quad \text{at }X=0. \end{equation}
At the far field, we assume simply supported conditions:
Additionally, the kinematic compatibility condition requires that the liquid–air interface and the elastic sheet meet at the moving contact line:
We note that (2.4) has the same form as the usual thin-film equation with slip, except that the liquid height is now given by
$H-G$
. Meanwhile, (2.4c
) is a fourth-order linear beam equation with an additional second-order tension term.
2.1. Thermodynamics and contact line conditions
Based on (2.4)–(2.7), we now derive appropriate contact line conditions following the principles of non-equilibrium thermodynamics (Ren, Hu & E Reference Ren, Hu and E.2010; Ren & E 2011). We start with the system total energy under the lubrication approximation:
\begin{equation} \begin{aligned} \mathcal{E} & = \mathcal{E}_s + \mathcal{E}_b, \\ \mathcal{E}_s & = \frac {1}{{\textit{Ca}}} \int _0^{A(t)}\left (1 + \frac {\delta ^2}{2} \left |\partial _XH\right |^2\right) \mathrm {d} X \\ &\quad {}+ \tilde {\gamma }_1 \int _0^{A(t)}\left (1 + \frac {\delta ^2}{2} \left |\partial _XG\right |^2 \right)\mathrm {d} X + \tilde {\gamma }_2 \int _{A(t)}^{X_r} \left (1 + \frac {\delta ^2}{2} \left |\partial _XG\right |^2\right) \mathrm {d} X, \\ \mathcal{E}_b & = \frac {\tilde {C}_b}{2} \int _0^{X_r} \delta ^2 \left |\partial _{XX} G\right |^2 \mathrm {d} X, \end{aligned} \end{equation}
where
$\mathcal{E}_s$
denotes the total interfacial energy, and
$\mathcal{E}_b$
is the bending energy of the elastic sheet. To avoid an infinite interfacial energy, here we apply the far-field condition (2.6) at the fixed boundary
$X=X_r$
(instead of
$X=+\infty$
) to suppress any energy dissipation at this point. (It is more rigorous to deal with a finite energy, although the derivation in this subsection remains valid, at least formally, with
$X_r = \infty$
.)
Next, we compute the energy dissipation rate. For the bending energy, we have
\begin{equation} \begin{aligned} \frac {\mathrm {d}\mathcal{E}_b}{\mathrm {d} T} ={} &\delta ^2 \tilde {C}_b \left ( - [\![ \partial _{XX}G ]\!] {\cdot }\partial _{XT}G + [\![ \partial _{X}^3G ]\!] {\cdot }\partial _TG \right)_{X=A(T)} \\ &{}+ \delta ^2 \frac {\tilde {C}_b}{2} [\![ \left |\partial _{XX}G\right |^2 ]\!]\, \dot {A}(T) \\ &{}+ \delta ^2 \tilde {C}_b \left \{\int _0^{A(T)} \partial _X^4G {\cdot }\partial _TG \ \mathrm {d} X + \int _{A(T)}^{X_r} \partial _X^4G {\cdot }\partial _TG \ \mathrm {d} X\right \}\!, \end{aligned} \end{equation}
where we have used the symmetry condition (2.5) at
$X=0$
, and the far-field condition (2.6) at
$X=X_r$
. The symbol
$[\![ {\cdot }]\!]$
denotes the jump of a quantity across the contact line from
$X=A(T)^-$
to
$X=A(T)^+$
.
The dissipation rate of the interfacial energy can be decomposed into three contributions:
\begin{align} \frac {\mathrm {d} \mathcal{E}_s}{\mathrm {d} T} ={}& \dot {\mathcal{E}}_{s,1} + \dot {\mathcal{E}}_{s,2} + \dot {\mathcal{E}}_{s,3}, \notag \\ \dot {\mathcal{E}}_{s,1} ={}& \delta ^2 \left \{ \frac {1}{{\textit{Ca}}}\, \partial _XH {\cdot }\partial _TH + (\tilde {\gamma }_1 - \tilde {\gamma }_2)\, \partial _XG {\cdot }\partial _TG \right \}_{X=A(T)}, \notag \\ \dot {\mathcal{E}}_{s,2} ={}& \dot {A}(T) \left \{ \frac {1}{{\textit{Ca}}} \left (1 + \frac {\delta ^2}{2}\left \vert \partial _XH\right \vert ^2\right) + \left ( \tilde {\gamma }_1 - \tilde {\gamma }_2 \right) \left ( 1 + \frac {\delta ^2}{2}\left \vert \partial _XG\right \vert ^2 \right) \right \}_{X=A(T)},\notag \\ \dot {\mathcal{E}}_{s,3} ={}& -\delta ^2 \left \{ \frac {1}{{\textit{Ca}}}\int _0^{A(T)} \partial _{XX}H {\cdot }\partial _TH \ \mathrm {d} X \right . \notag \\ &\left . {}+ \tilde {\gamma }_1 \int _0^{A(T)} \partial _{XX}G {\cdot }\partial _TG \ \mathrm {d} X + \tilde {\gamma }_2 \int _{A(T)}^{X_r} \partial _{XX}G {\cdot }\partial _TG \ \mathrm {d} X \right \}\!. \end{align}
We take the time derivative of the kinematic compatibility condition (2.7) and use the resulting relation to simplify
$\dot {\mathcal{E}}_{s,1}$
and
$\dot {\mathcal{E}}_{s,2}$
. After some manipulation, we obtain
\begin{equation} \begin{aligned} \dot {\mathcal{E}}_{s,1} + \dot {\mathcal{E}}_{s,2} ={} & \delta ^2 \left \{ \left ( \frac {1}{{\textit{Ca}}}\,\partial _XH + (\tilde {\gamma }_1 - \tilde {\gamma }_2)\, \partial _XG \right)\partial _TG \right \}_{X=A(T)} \\ &{} + \frac {\delta ^2}{{\textit{Ca}}}\, \dot {A}(T) \left \{ \frac {1}{2}\varTheta _Y^2 - \frac {1}{2}\left \vert \partial _XH - \partial _XG\right \vert ^2 \right \}_{X=A(T)} + \operatorname {O}\big(\delta ^4\big), \end{aligned} \end{equation}
where
$\varTheta _Y$
is the rescaled Young’s angle defined via Young’s relation (Young Reference Young1805)
the right-hand side being the small-angle expansion of the cosine function. To handle
$\dot {\mathcal{E}}_{s,3}$
, we use the thin film (2.4), which leads to the identity
\begin{equation} \begin{aligned} &\int _0^{A(T)} \partial _{XX}H {\cdot }\partial _TH \ \mathrm {d} X \\ &= \int _0^{A(T)} \partial _{XX}H {\cdot }\partial _TG \ \mathrm {d} X + \int _0^{A(T)} \frac {1}{{\textit{Ca}}} \left (H-G\right)^2\left (H-G+\lambda \right)\big \vert \partial _{X}^3 H\big \vert ^2 \mathrm {d} X, \end{aligned} \end{equation}
where we have used integration by parts and the fact that
$F=0$
at
$X=0$
and
$X = A(T)$
.
Finally, we combine (2.9), (2.10), (2.11) and (2.13), and apply the sheet (2.4c ) to obtain the dissipation rate for the total energy:
This equation shows that the energy dissipation consists of dominant terms of
$\operatorname {O} (\delta ^2)$
and some higher-order contributions. Among the dominant terms, the term in (2.14a
) arises from the dynamics of the bulk fluid, while the remaining terms (2.14b
)–(2.14d
) are associated with the contact line dynamics. Under the physical constraint
$H-G \geqslant 0$
for
$0 \lt X \lt A(T)$
, the term in (2.14a
) is always non-positive, and the thin-film motion always dissipates energy. To ensure that the total system energy decays, at least to leading order in
$\operatorname {O} (\delta ^2)$
, we will derive and impose appropriate conditions at the moving contact line. For this purpose, we have written the contact-line-related terms in the form of a product of a generalised force and a generalised flux. Next, we examine these terms one by one.
In the first term of (2.14b ), the generalised force is the jump in the curvature across the contact line, while the generalised flux is the rate of sheet rotation about the contact line. We assume that the sheet rotation here does not induce any energy change, so we set the generalised force to zero, and obtain a continuity condition for the curvature:
The second term of (2.14b ) vanishes given the continuity of the sheet curvature established above. Therefore, this term does not contribute to the energy dissipation.
In (2.14c ), the generalised force is the resultant vertical force acting at the contact line, while the generalised flux is the vertical velocity of the sheet at the contact line. (Note that this is not a material velocity, however.) We assume that the sheet motion at the contact line alone does not induce any energy change. This yields the following balance condition at the contact line:
In (2.14d ), the generalised force is an approximation of the unbalanced Young stress in the horizontal direction, while the generalised flux is the contact line speed. Assuming a linear constitutive relation between the two, we obtain
Here,
$\tilde {\mu }_{\varLambda } \gt 0$
is a dimensionless friction coefficient, representing the resistance to contact line motion. It is a physical parameter whose value in general depends on fluid properties, substrate roughness and microscopic characteristics.
To conclude this section, we have derived a model for the motion of a thin film on an elastic sheet based on the lubrication approximation. The primary unknowns of this model are the liquid–air interface profile
$z=h(x,t)$
and the elastic sheet profile
$z=g(x,t)$
, with the contact line located at
$x=a(t)$
. The dimensionless governing equations are summarised below. For better readability, we revert to lowercase variables and omit overhead tildes:
The boundary conditions are
The contact line conditions at
$x=a(t)$
are
In this model, the total energy of the system changes as
where the first and second terms on the right-hand side represent the dissipation rates due to the thin-film motion and the contact line motion, respectively.
3. Asymptotic analysis
The classical Cox–Voinov theory describes the spreading (or receding) rate of a liquid film on a rigid substrate in the limit of small slip length
$\lambda \ll 1$
, and capillary number
${\textit{Ca}} = \operatorname {O} (\epsilon)$
, where
$\epsilon$
is defined in (1.1). The central result can be stated as
where
$\theta _{\textit{app}}$
is the apparent contact angle that the interface makes with the substrate when viewed at a distance of
$\operatorname {O} (1)$
from the contact line. This relation has been verified through previous experimental and theoretical works; see e.g. Hocking & Rivers (Reference Hocking and Rivers1982), Cox (Reference Cox1986) and Ngan & Dussan (Reference Ngan and Dussan1989). The small capillary number assumption
${\textit{Ca}} = \operatorname {O} (\epsilon)$
is essential for the applicability of the Cox–Voinov theory. In this regime, the interface is only weakly curved near the contact line, and the flow can be treated as a perturbation around the wedge solution (Huh & Scriven Reference Huh and Scriven1971), which forms the basis of the Cox–Voinov analysis and its extensions. When
${\textit{Ca}} \sim 1$
, however, the interface may develop a cusp-like structure near the contact line (Kamal et al. Reference Kamal, Sprittles, Snoeijer and Eggers2019). In such cases, the wedge flow approximation is no longer valid, and a more detailed description of the interface is required.
We now set up the formulation to extend the Cox–Voinov analysis to the case where the substrate is an elastic sheet. The governing equations for the dynamics of the fluid interface and the sheet are given in (2.18)–(2.20). We consider the limit of small slip length
$\lambda \ll 1$
, with the scaling
${\textit{Ca}} = {\textit{Ca}}^\ast\, \epsilon$
, where
${\textit{Ca}}^\ast$
is an
$\operatorname {O} (1)$
constant. Since the effective tensions
$\gamma _i\ (i=1,2)$
are of the same order as
$\gamma _3$
, we write
$\gamma _i = \gamma _i^* / \epsilon$
, where the
$\gamma _i^*$
are
$\operatorname {O} (1)$
constants. Additionally, we consider a highly bendable sheet with bending modulus
$C_b = C^*_b \epsilon$
, where
$C^*_b$
is an
$\operatorname {O} (1)$
constant. In this regime, the corresponding (dimensionless) bending length
$l_i = \sqrt {C_b^*/\gamma _i^*}\,\epsilon$
is consistent with typical experimental parameters (e.g. Huang et al. Reference Huang, Juszkiewicz, De Jeu, Cerda, Emrick, Menon and Russell2007) and is therefore of practical interest. (For illustration, consider a slip length on the molecular scale of 1 nm and drop radius 1 mm, giving
$\lambda = 10^{-6}$
and
$\epsilon \approx 0.072$
. For a polystyrene sheet with Young’s modulus 3.4 GPa, thickness 200 nm, Poisson ratio
$1/3$
, and interfacial tension 30
$\text{nN}\ \text{m}^{-1}$
, the resulting dimensionless bending length is
$l \approx 0.0093 \lesssim \epsilon$
.)
We note that the
$\operatorname {O} (1)$
parameter
${\textit{Ca}}^*$
has no fundamental influence on the contact line dynamics, nor is it essential for the mathematical analysis. To keep the exposition simple, we will restrict to the specific case
${\textit{Ca}}^* = 1$
in the rest of this section. The general case can be recovered from this specific one by straightforward rescaling. Below, we summarise the rescaled governing equations, with all asterisks dropped for notational simplicity:
The conditions at the moving contact line
$x=a(t)$
are
The symmetry and far-field boundary conditions remain the same as in (2.19).
The outline of our analysis is illustrated in figures 1 and 2. We perform the analysis in four regions: the outer region (figure 1), the bending region (figure 2 a), the inner region (figure 2 c) and the intermediate region (figure 2 b). These analyses are carried out in §§ 3.1–3.4. In § 3.5, we match the solutions in the different regions, and derive an effective model for the contact line speed, analogous to the Cox–Voinov relation (3.1).

Figure 2. (a) Bending region, viewed at a distance
$\operatorname {O}(\epsilon)$
from the contact line, with coordinates
$(\tilde {y}, \tilde {z})$
; in this region,
$\tilde {\beta }$
denotes the slope of the sheet at the contact line. (b) Intermediate region, with coordinates
$(s, z)$
, where
$0 \lt s = \epsilon \log y + 1 \lt \epsilon \log \epsilon + 1$
. (c) Inner region, viewed at a distance
$\operatorname {O}(\lambda)$
from the contact line, with coordinates
$(\bar {y}, \bar {z})$
; here,
$\theta _{{m}}$
is the microscopic contact angle.
3.1. Outer region
We expand the contact line speed in powers of
$\epsilon$
:
where
$\epsilon \dot {a}_1$
is
$\operatorname {o} (1)$
for the expansion to be well ordered. Similarly, we expand the profiles
Substituting these expansions into the governing equation (3.2), we obtain the leading-order equations for
$h_0$
and
$g_0$
:
The leading-order boundary conditions are given by
\begin{equation} \begin{aligned} &\partial _x h_0 = \partial _{x}^3 h_0 = 0 \quad \text{at } x=0, \\ &\partial _x g_0 = 0 \quad \text{at } x=0, \\ &g_0 = 0 \quad \text{at } x = +\infty , \end{aligned} \end{equation}
and the leading-order contact line conditions are
To solve this system, we first deduce from (3.6c
) and the far-field condition that
$g \equiv 0$
when
$a(t) \leqslant x$
. We then solve (3.6a
) and (3.6b
) with the boundary and contact line conditions. To fully determine the solution, we need to impose the conservation of the liquid volume
$V$
. The leading-order solution is then given by
\begin{equation} \begin{gathered} h_0(x,t) = \frac {p_0}{2} \big(a^2 - x^2\big), \quad 0 \leqslant x \leqslant a(t), \\ g_0(x,t) = \begin{cases} \dfrac {p_0}{2\gamma _1}\left (x^2 - a^2\right)\!, & 0 \leqslant x \leqslant a(t), \\ 0, & a(t) \lt x, \end{cases} \end{gathered} \end{equation}
where we have introduced the pressure
$p_0$
for notational convenience:
We note that this solution automatically satisfies the higher-order boundary conditions
$\partial _x^3g_0 = 0$
at
$x=0$
, and
$\partial _x^2g_0 = 0$
at
$x=+\infty$
. For later use, we compute the leading-order film thickness:
The next-order equations from (3.2) are
The boundary conditions and contact line conditions take the same form as in (3.7) and (3.8), with
$h_0$
,
$g_0$
replaced by
$h_1$
,
$g_1$
, respectively. As before, we have
$g_1 \equiv 0$
when
$a(t) \leqslant x$
. To solve for
$h_1$
, we note that the dependence of the leading-order profiles
$h_0$
and
$g_0$
on time is only through
$a(t)$
. We integrate (3.12a
) once with respect to
$x$
to obtain
We set the integration constant to zero, as it will yield a solution that matches the solutions in the other regions and is consistent with the numerical results. We proceed by further integrations. Using the boundary and contact line conditions as well as volume conservation, we can determine the next-order solutions:
\begin{gather} g_1 = \begin{cases}-\dfrac {1}{\gamma _1}h_1, & 0 \leqslant x \leqslant a(t), \\ 0, & a(t) \lt x. \end{cases} \end{gather}
Before moving on, we introduce a moving coordinate system centred at the contact line. Let
$x = a(t) - y$
, where
$y \leqslant a(t)$
. In this coordinate system, the contact line is located at
$y=0$
, and the governing equation for
$h = h(y,t)$
becomes
The governing equations for
$g = g(y,t)$
remain the same as in (3.2b
)–(3.2c
), but with
$\partial _x$
replaced by
$\partial _y$
. The symmetry conditions are now imposed at
$y=a(t)$
, the far-field conditions at
$y=-\infty$
, and the contact line conditions at
$y=0$
. Their forms remain unchanged except for
$\partial _x$
being replaced by
$\partial _y$
. Near the contact line as
$y\rightarrow 0^+$
, the outer solutions in terms of
$y$
are
\begin{equation} \begin{aligned} h_0 \sim & \alpha _{\textit{app}} y - \frac {\alpha _{\textit{app}}}{2a}y^2, \\ h_1 \sim & \frac {\dot {a}_0}{\theta _{\textit{app}}^2} \left \{y\log {y} + \left (2 - \log {2a} \right) y + \cdots \right \}\!, \end{aligned} \end{equation}
and
\begin{equation} \begin{aligned} g_0 \sim & \beta _{\textit{app}} y - \frac {\beta _{\textit{app}}}{2a} y^2, \\ g_1 \sim & -\frac {1}{\gamma _1} h_1. \end{aligned} \end{equation}
Here,
$\alpha _{\textit{app}}$
,
$\beta _{\textit{app}}$
and
$\theta _{\textit{app}}$
are the apparent contact angles for the interface, for the sheet, and between them, respectively. They are defined as
\begin{equation} \begin{aligned} \alpha _{\textit{app}} =& \frac {\partial h_0(0,{\cdot })}{\partial y} = \frac {3V}{a^2 \big(1 + \gamma _1^{-1}\big)}, \\ \beta _{\textit{app}} =& \frac {\partial g_0(0,{\cdot })}{\partial y} = -\frac {1}{\gamma _1} \frac {3V}{a^2\big(1 + \gamma _1^{-1}\big)}, \\ \theta _{\textit{app}} =& \alpha _{\textit{app}} - \beta _{\textit{app}} = \frac {3V}{a^2}. \end{aligned} \end{equation}
Next, we analyse the problem in the bending region.
3.2. Bending region
We rescale the variables in the bending region as
$y = \epsilon \tilde {y}$
,
$h = \epsilon \tilde {h}$
and
$g = \epsilon \tilde {g}$
. The governing equations, written in terms of these variables, are
To determine the appropriate boundary conditions as
$\tilde {y} \rightarrow \pm \infty$
, we rewrite the outer solutions (3.16) and (3.17) in terms of
$\tilde {y}$
, and arrange terms according to powers of
$\epsilon$
. Accordingly, we impose the far-field behaviour
as
$\tilde {y} \rightarrow + \infty$
. In addition, from the outer solutions we have
as
$\tilde {y} \rightarrow -\infty$
. At the contact line, we impose
We omit the contact line conditions (3.3b ) and (3.3c ) at this stage, as they are only relevant in the immediate vicinity of the contact line, i.e. the inner region.
We expand
$\tilde {h}$
and
$\tilde {g}$
in powers of
$\epsilon$
:
The leading-order equations for
$\tilde {h}_0$
and
$\tilde {g}_0$
are
To solve these equations, we first consider (3.24a
) for
$\tilde {h}_0$
. We integrate the equation once, set the constant of integration to zero, then integrate further and match the far-field condition in (3.20a
), to obtain
To solve (3.24b ) and (3.24c ), it is convenient to introduce
\begin{align} \begin{gathered} l_i = \sqrt {C_b / \gamma _i}, \quad i=1, 2, \\ \tilde {y}_1 = \tilde {y} / l_1, \quad \tilde {y} \geqslant 0, \\ \tilde {y}_2 = \tilde {y} / l_2, \quad \tilde {y} \leqslant 0. \end{gathered} \end{align}
The general solutions to the fourth-order (3.24b ) and (3.24c ) are given by
where
$i=1, 2$
corresponds to the solution when
$\tilde {y}\gt 0$
and
$\tilde {y}\lt 0$
, respectively. The eight constants are to be determined using the far-field conditions (3.20b
) and (3.21), and the contact line conditions in (3.22). From the far-field conditions, we find that
$C_1^+ = 0$
,
$C_2^- = 0$
,
$A_2 = B_2 = 0$
and
$B_1 = \beta _{\textit{app}}$
. We then apply the continuity conditions in (3.22) to obtain
\begin{equation} \tilde {g}_0 = \begin{cases} C_1^-\, e^{-\tilde {y}_1} + \beta _{\textit{app}} \tilde {y} + A_1, & \tilde {y} \geqslant 0, \\ C_2^+\, e^{\tilde {y}_2}, & \tilde {y} \lt 0, \end{cases} \end{equation}
where
For future reference, we compute the leading-order slope of the sheet at the contact line:
This slope is independent of
$C_b$
.
We then consider the next-order problem:
We integrate (3.31a
) for
$\tilde {h}_1$
once, and set the constant of integration to zero, to obtain
\begin{equation} \partial _{\tilde {y}}^3 \tilde {h}_1 = -\frac {\dot {a}_0}{\big(\tilde {h}_0 - \tilde {g}_0\big)^2} = -\frac {\dot {a}_0}{\left (\theta _{\textit{app}} \tilde {y} + C_1^- - C_1^-\, e^{-\tilde {y}_1}\right)^2}. \end{equation}
This equation does not admit a closed-form solution; however, it is possible to find approximate solutions in the far field as
$\tilde {y} \rightarrow +\infty$
, and the near field as
$\tilde {y}\rightarrow 0$
. To find the far-field behaviour, we neglect the exponentially small term on the right-hand side, and consider the far-field approximation
$\tilde {h}_1^\infty$
satisfying
where the constant is
$\tilde {c} = C_1^-/\theta _{\textit{app}} \lt 0$
. We integrate this equation and use the far-field condition (3.20a
) to determine the constant of integration. We obtain
where
$E_1 = 3 + \log (\epsilon /2a)$
.
Similarly, we consider the far-field approximation
$\tilde {g}_1^{\infty }$
for
$\tilde {g}_1$
, satisfying
To solve this equation, we integrate it twice and use the far-field condition (3.20b ) to determine the constant of integration. We obtain
where
$D_1$
is a constant to be determined, and
$\varphi$
is a specific solution to a related second-order ordinary differential equation; see (B3) in Appendix B. We integrate once more to obtain
where the constant of integration is determined by matching to the far-field condition (3.20b
). We leave the constant
$D_1$
undetermined, as it does not contribute to the far-field behaviour.
So far, we have obtained the far-field behaviour of
$\tilde {h}_1$
and
$\tilde {g}_1$
. We next turn to their near-field behaviour as
$\tilde {y} \rightarrow 0^+$
. To solve (3.32) for
$\tilde {h}_1$
in the near field, we first notice from the leading-order solutions (3.25)–(3.28) that
$\tilde {h}_0 - \tilde {g}_0$
is analytic in a neighbourhood of
$\tilde {y} = 0$
. Using a power series to represent the right-hand side of (3.32), we have
\begin{align} \partial _{\tilde {y}}^3 \tilde {h}_1 = -\frac {\dot {a}_0}{\big(\alpha _{\textit{app}} - \tilde {\beta }\big)^2 \tilde {y}^2} + \operatorname {O}\left (\frac {1}{\tilde {y}}\right)\!, \quad \tilde {y} \rightarrow 0^+, \end{align}
where the remainder is a power series consisting of terms
$\tilde {y}^k\ (k\geqslant -1)$
. We perform a term-by-term integration to obtain the near-field approximation
where
$\tilde {C}_1$
is a constant of integration, and the remainder series consists of the terms
$\tilde {y}^k$
and
$\tilde {y}^k \log \tilde {y} \ (k\geqslant 2)$
. We follow a similar procedure and get the equation for
$\tilde {g}_1$
near
$\tilde {y} = 0$
:
We perform a term-by-term integration to obtain the near-field approximation
where the remainder series consists of terms
$\tilde {y}^k\ (k\geqslant 3)$
and
$\tilde {y}^k\log \tilde {y}\ (k\geqslant 4)$
. In deriving this solution, we have imposed the continuity conditions at the contact line (see the next paragraph) and required matching with the intermediate solution in (3.59) (see § 3.4). Comparing (3.39) and (3.41), we see that
$\tilde {h}_1^0$
exhibits a stronger singularity than
$\tilde {g}_1^0$
.
For
$\tilde {y} \lt 0$
, the next-order sheet solution is simply
$\tilde {g}_1 \equiv 0$
, by noting the far-field condition (3.21) and the continuity conditions (3.22) up to the second derivative at the contact line. To summarise, we have obtained the interface and the sheet solutions up to
$\operatorname {O} (\epsilon)$
in the bending region. The leading-order solutions are given in (3.25) and (3.28). The far-field behaviour of the next-order solutions is given in (3.34) and (3.37), while the near-field behaviour is given in (3.39) and (3.41).
3.3. Inner region
We now consider the inner region by introducing the rescaled variables
$\bar {y} = y/\lambda$
,
$\bar {h} =h/\lambda$
and
$\bar {g} = g/\lambda$
. We rewrite the governing equation (3.2) in terms of these inner variables:
At the contact line
$\bar {y} = 0$
, both the compatibility condition (3.3d
) and the continuity conditions in (3.3a
) apply. In addition, we impose the contact line conditions (3.3b
) and (3.3c
). They now become
We note that since
$\lambda$
is exponentially small compared to
$\epsilon$
, the equations for
$\bar {g}$
reduce to
$\partial _{\bar {y}}^4 \bar {g} = 0$
, and the contact line condition (3.43a
) reduces to
$[\![ \partial _{\bar {y}}^3 \bar {g} ]\!]=0$
. For the purpose of matching, we require the sheet curvature to vanish at infinity, and we obtain the solution
which is accurate up to all orders of
$\epsilon$
.
We only need to expand the interface profile
From (3.42a
), we see that the leading-order equation for
$\bar {h}$
is
and from (3.43b ), the leading-order contact line condition is
We further require the interface curvature to vanish at infinity, and hence obtain
For the next-order problem, the governing equation is
and the contact line condition is
We integrate (3.49) and set the constant of integration to zero to obtain
We integrate further, require the curvature to vanish at infinity, and impose the contact line condition to obtain
From here, we integrate one more time to obtain
\begin{equation} \begin{aligned} \bar {h}_1 ={} & \frac {\dot {a}_0}{\theta _Y} \biggl \{ \left (\mu _{\varLambda } - \frac {1}{2\theta _Y}\right) \bar {y} + \frac {1}{2\theta _Y^2} \log \left (\theta _Y \bar {y} + 1\right) + \frac {1}{\theta _Y} \bar {y} \log \left (\theta _Y \bar {y} + 1\right) \\ & {}+ \frac {1}{2} \bar {y}^2 \left [\log \left (\theta _Y \bar {y} + 1\right) - \log \left (\theta _Y \bar {y}\right)\right ] \biggr \} \\ \sim{} & \frac {\dot {a}_0}{\theta _Y} \left \{ \frac {1}{\theta _Y}\bar {y}\log \bar {y} + \left (\mu _{\varLambda } + \frac {\log \theta _Y}{\theta _Y}\right) \bar {y} + {\cdots } \right \} \quad \text{as } \bar {y} \rightarrow +\infty . \end{aligned} \end{equation}
We note that the inner solutions derived here are valid in an
$\operatorname {O} (\lambda)$
neighbourhood of the contact line, or when
$\bar {y} = \operatorname {O} (1)$
.
3.4. Intermediate region
We proceed to the intermediate region that bridges the inner and bending regions. We apply the transformation
such that
$y(0) = \lambda$
corresponds to the inner region, and
$y(1 + \epsilon \log \epsilon) = \epsilon$
corresponds to the bending region. Following the usual practice, we also set
These intermediate variables are related to the inner variables via
$\bar y = e^{s/\epsilon }$
,
$\bar h = H(s, t)\,\bar y$
and
$\bar g = G(s, t)\, \bar y$
.
To derive the equation for
$H$
, we drop the exponentially small term in (3.42a
), integrate once, and set the constant of integration to zero. We then rewrite the equation in terms of the intermediate variables to obtain
To derive the equation for
$G$
, we rewrite (3.42b
) in terms of the intermediate variables:
Here we note that when
$0 \lt s \lt 1 + \epsilon \log \epsilon$
, the term
$\epsilon ^2\, e^{2(1-s)/\epsilon }$
is positive and exponentially large. Therefore, the above equation reduces to
with some exponentially small corrections. This equation is linear, independent of
$H$
, and can be solved exactly. We obtain
$G \equiv \text{const}$
by discarding solutions that lead to an exponentially large curvature. By matching to the inner solution (3.44) and the bending solutions (3.28)–(3.30), we find
which is accurate to all orders of
$\epsilon$
.
We then turn to solve (3.56). We drop the exponentially small term
$e^{-s/\epsilon }$
. Since
$\partial _sG \equiv 0$
, we can rewrite the equation in terms of
$H-G$
:
From here we solve for
$H-G$
:
where
$c_0$
and
$c_1$
are the integration constants to be determined. Further manipulation gives the two-term expansion
3.5. Matching
The outer and bending solutions have been matched to each other. We now consider the matching between the intermediate and inner solutions. To this end, we rewrite the intermediate solutions (3.62) in terms of the inner variables and expand them in powers of
$\epsilon$
:
\begin{align} \begin{aligned} \bar {h} - \bar {g} &= \left (H - G\right) \bar {y} \\ &\sim \left (3\epsilon \dot {a}_0\log \bar {y} + c_0\right)^{1/3} \bar {y} + \epsilon \frac {3\epsilon \dot {a}_1 \log \bar {y} + c_1}{3 \left (3\epsilon \dot {a}_0 \log \bar {y} + c_0\right)^{2/3}} \bar {y} + {\cdots } \\ &\sim c_0^{1/3} \bar {y} + \frac {\epsilon }{c_0^{2/3}} \left (\dot {a}_0 \bar {y} \log \bar {y} + \frac {c_1}{3} \bar {y} \right) + {\cdots } . \end{aligned} \end{align}
This expansion is then matched to the inner solution
Matching the two expansions term by term, we obtain
To match the intermediate solutions to the bending ones, we rewrite the intermediate solutions (3.62) in terms of the bending variables and re-expand the expression to get
\begin{align} \begin{aligned} \tilde {h} - \tilde {g} &= \left ( H - G \right) \tilde {y} \\ &\sim \left (3\dot {a}_0\left (1 + \epsilon \log \epsilon \tilde {y}\right) + c_0\right)^{1/3} \tilde {y} + \epsilon \frac {3\dot {a}_1\left (1 + \epsilon \log \epsilon \tilde {y}\right) + c_1}{3 \left (3\dot {a}_0\left (1 + \epsilon \log \epsilon \tilde {y}\right) + c_0\right)^{2/3}} \tilde {y} + {\cdots } \\ &\sim \left (c_0 + 3\dot {a}_0\right)^{1/3} \tilde {y} + \frac {\epsilon }{\left (c_0 + 3\dot {a}_0\right)^{2/3}} \left \{ \dot {a}_0 \tilde {y} \log \tilde {y} + \left (\dot {a}_0 \log \epsilon + \dot {a}_1 + \frac {c_1}{3}\right) \tilde {y} \right \} + {\cdots } . \end{aligned} \end{align}
This is matched to the near-field bending solution (cf. (3.25)–(3.28), (3.39)–(3.41)):
By matching the corresponding terms, we obtain
Combining (3.65) and (3.68), we deduce
where we have used (3.30) for the angle
$\tilde {\beta }$
. Equation (3.69a
) provides an estimate for the leading-order contact line speed in terms of the apparent contact angles
$\alpha _{\textit{app}}$
and
$\beta _{\textit{app}}$
. We note that this equation was derived under the assumption
${\textit{Ca}} = \epsilon$
; the general case, which can be obtained by a straightforward rescaling of the governing equations, is given in (1.3). Equation (3.69b
) gives an estimate for the next-order correction to the contact line speed; however, this correction is of limited practical use, as the constant
$\tilde {C}_1$
(cf. (3.39)) remains undetermined. Nevertheless,
$\tilde {C}_1$
can, in principle, be determined by solving the full next-order problem in the bending region either analytically or numerically, which we will not consider in this study.
Before concluding this section, we make three remarks concerning (3.69a
). First, the relation can be viewed as a generalisation of the classical Cox–Voinov relation. Taking the limit
$\gamma _1 \rightarrow +\infty$
, we find that
$\beta _{\textit{app}} \rightarrow 0$
(cf. (3.18)), and we formally recover the classical result
This limiting case corresponds to an infinitely pre-stretched sheet, which excludes any vertical displacement and behaves effectively as a rigid substrate. Second, as long as the friction coefficient
$\mu _{\varLambda }$
is an
$\operatorname {O} (1)$
parameter, it does not appear in the leading-order relation (3.69a
), but only affects the next-order relation (3.69b
) through the moving contact line condition applied in the inner region. Finally, (3.69a
) is independent of the bending modulus
$C_b$
. This is because the slope of the sheet in the bending region near the contact line,
$\tilde {\beta }$
(cf. (3.30)), does not depend on
$C_b$
. This observation is further supported by the numerical results presented in the next section.
4. Numerical results and discussion
We compare the asymptotic results obtained in the previous section with numerical solutions of the thin-film model. In § 4.1, we show that the interface and sheet profiles, along with their slopes, in the numerical solutions agree well with the asymptotic predictions. In § 4.2, we verify the asymptotic relation (3.69a ), and examine the long-time behaviour of the moving contact line in spreading and receding films.
We solve the governing equations (2.18)–(2.20) using a finite difference method with a moving mesh. The numerical method and its validation are detailed in Appendix A. In the simulations, we choose the friction coefficient
$\mu _{\varLambda }=1$
, the equilibrium angle
$\theta _Y=1$
, the slip length
$\lambda = 10^{-4}$
, which corresponds to
$\epsilon \approx 0.109$
, and the capillary number
${\textit{Ca}} = \epsilon$
. We vary the parameters
$\gamma _1$
,
$\gamma _2$
and
$C_b$
.
4.1. Interface and sheet profiles
We perform the simulations for the spreading and receding dynamics of a liquid film. The parameters in (2.18)–(2.20) are chosen as
$\gamma _1 = 2 / \epsilon$
,
$\gamma _2 = 2.95 / \epsilon$
and
$C_b = \epsilon$
. The initial interface and sheet profiles are given by
\begin{align} h(x,0) = c_V \left (1 + 0.2 \cos \frac {2\pi x}{a(0)}\right) \left \{1 - \left (\frac {x}{a(0)}\right)^2\right \}\!, \quad g(x,0) \equiv 0, \end{align}
where the constant
$c_V$
is chosen so that the liquid volume is
$V=1$
. We use the initial condition
$a(0) = 1.0$
for a spreading film, and
$a(0) = 2.0$
for a receding film. In figure 3, we present the simulation results of the interface and sheet profiles at several time instants.
In the spreading case, the contact line advances to the right, the apex heights of both the interface and the sheet decrease, and the liquid film becomes flatter. By
$t=32$
, the system has almost reached equilibrium, with the contact line speed dropping below
$3 \times 10^{-3}$
. In the receding case, the contact line retreats to the left, and the liquid becomes more concentrated. By
$t=24$
, the system has almost reached equilibrium, with the contact line speed falling below
$5 \times 10^{-3}$
.
Next, we compare the computed interface and sheet profiles at
$t=1$
and
$t=4$
with the asymptotic solutions derived in § 3. The asymptotic solutions invlove the parameters
$\gamma _i\ (i=1,2)$
,
${\textit{Ca}}$
,
$C_b$
,
$\theta _Y$
,
$\mu _{\varLambda }$
,
$\lambda$
,
$\epsilon$
, and the liquid volume
$V$
, together with the apparent contact angles
$\alpha _{\textit{app}}$
and
$\beta _{\textit{app}}$
, the contact line position
$a(t)$
, and its velocity components
$\dot {a}_0$
and
$\dot {a}_1$
. All physical parameters (
$\gamma _i$
,
${\textit{Ca}}$
,
$C_b$
,
$\theta _Y$
,
$\mu _{\varLambda }$
,
$\lambda$
,
$\epsilon$
,
$V$
) are prescribed in the simulations. The contact line position
$a(t)$
at
$ t=1, 4$
is obtained directly from the numerical solutions. Given
$a(t)$
, the apparent contact angles
$\alpha _{\textit{app}}$
and
$\beta _{\textit{app}}$
are computed using (3.18). The leading-order prediction of the contact line speed
$\dot {a}_0$
follows from (3.69a
). The next-order contact line speed
$\dot {a}_1$
cannot be determined directly from (3.69b
) due to the undetermined constant
$\tilde {C}_1$
, so we estimate it via
$\dot {a}_1 = (\dot {a}_{{num}} - \dot {a}_0) / \epsilon$
, where
$\dot {a}_{{num}}$
is the numerically obtained contact line speed. We note that no parameters or variables are adjusted by fitting in these comparisons.
We first focus on the spreading dynamics discussed above. In figure 4(a), the two-term outer solutions
$h_0 + \epsilon h_1$
and
$g_0 + \epsilon g_1$
from (3.9) and (3.14) are compared with the numerical solutions from the thin-film model at
$t=1.0$
and
$t=4.0$
. We observe good overall agreement between the asymptotic and numerical solutions, except that the outer solutions do not capture the smooth bending of the sheet near the contact line. We then examine the bending region. In figure 4(b), the leading-order bending solutions
$\tilde {h}_0$
and
$\tilde {g}_0$
, given by (3.25) and (3.28), respectively, are compared with the numerical results. These asymptotic solutions agree well with the numerical solutions near the contact line, even though only the leading-order terms are used.

Figure 4. Interface and sheet profiles for a spreading film at
$t=1.0$
and
$t=4.0$
, shown in (a) the outer region and (b) the bending region. Circles denote numerical solutions of the thin-film model (2.18)–(2.20), while solid lines show the asymptotic approximations: the two-term outer solutions from (3.9) and (3.14) in (a), and the leading-order bending solutions from (3.25) and (3.28) in (b).
We also compare the numerical and asymptotic solutions in all four regions; in particular, we check the agreement of their slopes. For the same spreading dynamics discussed above, we plot the interface and sheet slopes,
$\partial _{y}h$
and
$\partial _{y}g$
, against the intermediate variable
$s = \epsilon \log y + 1$
in figures 5(a) and 5(b), respectively. In figure 5(a), we show the interface slopes in their respective regions of validity: (i) the near-field outer solution
$\partial _y h_0 + \epsilon\, \partial _y h_1$
from (3.16); (ii) the far-field bending solution
$\partial _{\tilde {y}}\tilde {h}_0 + \epsilon\, \partial _{\tilde {y}}\tilde {h}_1^{\infty }$
from (3.25) and (3.34); (iii) the near-field bending solution
$\partial _{\tilde {y}}\tilde {h}_0 + \epsilon\, \partial _{\tilde {y}}\tilde {h}_1^0$
from (3.25) and (3.39); (iv) the intermediate solution
$H + \epsilon\, H^{\prime}(s)$
from (3.59) and (3.62); and (v) the inner solution
$\partial _{\bar {y}} \bar {h}_0 + \epsilon\, \partial _{\bar {y}} \bar {h}_1$
from (3.48) and (3.53). For the intermediate solution, the next-order contact line speed
$\dot {a}_1$
cannot be directly obtained from (3.69b
), so we estimate it using the numerical contact line speed
$\dot {a}_{{num}}$
via
$\dot {a}_1 = (\dot {a}_{{num}} - \dot {a}_0) / \epsilon$
. The two-term bending solution
$\tilde {h}_0 + \epsilon \tilde {h}_1^{\infty }$
is a rearrangement of the outer solution in the bending region (with some higher-order corrections). This explains the smooth connection observed between the outer and bending solutions. Overall, we observe good agreement between the numerical and asymptotic solutions at both
$t=1.0$
and
$t=4.0$
. In figure 5(b), we overlay the far-field solution
$\partial _{\tilde {y}} \tilde {g}_0 + \epsilon\, \partial _{\tilde {y}} \tilde {g}_1^{\infty }$
(with
$D_1=0$
; cf. (3.37)) and the near-field solution
$\partial _{\tilde {y}} \tilde {g}_0 + \epsilon\, \partial _{\tilde {y}} \tilde {g}_1^0$
from (3.28) and (3.41). These asymptotic solutions agree well with the numerical ones in their respective regions of validity. In particular, this validates our prediction of the sheet slope
$\tilde {\beta }$
in the bending region, as given in (3.30). The discrepancy towards the inner region is possibly due to the fact that the bending length is not sufficiently small, which limits the accuracy of the asymptotic solution. We also note that the far-field and near-field solutions generally do not match each other near
$\tilde {y} = \operatorname {O} (1)$
, since this lies outside their region of validity.

Figure 5. (a) Interface slope
$\partial _y h$
and (b) sheet slope
$\partial _y g$
in a spreading film at
$t=1.0$
and
$t=4.0$
. Circles denote numerical solutions of the thin-film model (2.18)–(2.20), while lines indicate asymptotic solutions (see text for details). In (a), both far-field and near-field bending solutions are shown as black dashed lines, while in (b), only the bending region asymptotic solutions are displayed.
A similar comparison for the receding film is shown in figures 6(a) and 6(b). The slope profiles are visually similar at all time instants, so we only present the case
$t=4.0$
. Again, we observe good agreement between the numerical results and the asymptotic predictions. We conclude that the asymptotic solutions in all four regions accurately capture the interface and sheet behaviours.

Figure 6. (a) Interface slope
$\partial _y h$
and (b) sheet slope
$\partial _y g$
for a receding film at
$t=4.0$
. Circles denote numerical solutions of the thin-film model (2.18)–(2.20), while lines indicate asymptotic solutions (see text for details). In (a), both far-field and near-field bending solutions are shown as black dashed lines, while in (b), only the bending region asymptotic solutions are displayed.
It is interesting to note that the interface does not exhibit a clear bending region. In the numerical solutions, although the bending effect dominates the change in the sheet slope over the interval
$s \in (0.5, 0.8)$
, the interface slope remains relatively unchanged in the same interval. This indicates that the interface is largely insensitive to small-scale variations in the substrate. We expect similar behaviour in liquid films spreading over substrates with variable height.
4.2. Contact line motion
We now compare the contact line speed predicted by the asymptotic relation (3.69a
) with that obtained from numerical solutions of the thin-film model (2.18)–(2.20). The initial conditions and parameters are the same as in the previous subsection, except that we now vary the tension:
$\gamma _1 = 1/\epsilon , 2/\epsilon , 4/\epsilon , 8/\epsilon$
and
$\gamma _2 = \gamma _1 + 0.95/\epsilon$
. Results for the spreading film and the receding film are shown in figures 7(a) and 7(b), respectively. Overall, the theoretical predictions agree well with the numerical results, especially for more rigid sheets, i.e. when
$\gamma _1$
is large. The classical Cox–Voinov relation (3.70) is also shown in the same figures. As
$\gamma _1$
increases, the results approach the classical one.

Figure 7. Contact line speed
$\dot {a}$
versus film radius
$a(t)$
for (a) a spreading film and (b) a receding film. Markers denote numerical solutions of the thin-film model (2.18)–(2.20). Solid lines show predictions from the asymptotic relation (3.69a
), while dashed lines are predictions using the classical Cox–Voinov relation (3.70) for rigid substrates.
Another important observation from these figures is that the spreading process becomes slower, and the receding process becomes faster, as the sheet becomes softer. The reason behind this behaviour can be understood from the dependence of the contact angle in the bending region,
$\alpha _{\textit{app}} - \tilde {\beta }$
, on the sheet tension. Using (3.18) and (3.30), we can rewrite this angle as
\begin{gather} \rho (\gamma _1, \gamma _2) = \frac {1}{1 + \gamma _1^{-1}} \left (1 + \frac {1}{\sqrt {\gamma _1} \left (\sqrt {\gamma _1} + \sqrt {\gamma _2}\right)}\right)\!, \end{gather}
where
$\theta _{\textit{app}}$
depends only on the volume
$V$
and the film radius
$a(t)$
.
Recall that
$\sigma _{\textit{pre}}$
is the stretching tension applied to the sheet (cf. (2.1)). When the stretching tension is large, we have
\begin{align} \rho = 1-\frac {1}{2\sigma _{\textit{pre}}}+\operatorname {O}\left (\frac {1}{\sigma _{\textit{pre}}^2}\right)\!. \end{align}
In the limit of a rigid substrate, i.e. as
$\sigma _{\textit{pre}}\rightarrow +\infty$
, we have
$\rho \rightarrow 1$
, and recover the classical Cox–Voinov relation. For large but finite
$\sigma _{\textit{pre}}$
, however,
$\rho \lt 1$
and it increases monotonically with
$\sigma _{\textit{pre}}$
. This reduction in the effective contact angle on softer substrates explains the observed phenomena of retarded spreading and enhanced receding in figures figures 7(a) and 7(b). It is worth noting that the retarded spreading of drops has also been observed on viscoelastic solids (Carré et al. Reference Carré, Gastel and Shanahan1996; Tamim & Bostwick Reference Tamim and Bostwick2023), where it is mainly attributed to viscous dissipation within the substrate. This mechanism, however, is not relevant here, since purely elastic solids do not exhibit such dissipation.
Next, we examine the effect of the bending modulus
$C_b$
on contact line motion. In figure 8, we show numerical results from the thin-film model (2.18)–(2.20) for the contact line speed of a spreading film with
$\gamma _1 = 1/\epsilon , 2/\epsilon , 4/\epsilon , 8/\epsilon$
and
$C_b = 0.01\epsilon , 0.1\epsilon , \epsilon$
. We observe that the speed is barely affected by variations in
$C_b$
. Similar behaviour is observed for the receding case (not shown). These results are consistent with the predictions from (3.69a
), where the contact line speed is independent of
$C_b$
to leading order. We expect the relation to remain valid over a wide range of bending moduli, provided that the bending length satisfies
$\lambda \ll l_1 \lesssim \epsilon$
.

Figure 8. Numerical results from the thin-film model (2.18)–(2.20) for the contact line speed
$\dot {a}$
versus the film radius
$a(t)$
in a spreading film. Circle, triangle and square markers correspond to
$C_b = 0.01\epsilon$
,
$0.1\epsilon$
and
$\epsilon$
, respectively. Colours indicate sheet tension: blue for
$\gamma _1 =1/\epsilon$
; orange for
$\gamma _1 = 2/\epsilon$
; green for
$\gamma _1=4/\epsilon$
; and red for
$\gamma _1 = 8/\epsilon$
.

Figure 9. Film radius
$a(t)$
for (a) a spreading film with
$a(0) = 1$
, and (b) a receding film with
$a(0) = 2$
, with bending modulus
$C_b = \epsilon$
. Markers denote numerical solutions of the thin-film model, while solid lines show the leading-order predictions by solving the generalised Cox–Voinov relation (3.69a
).
Finally, we plot the film radius against time in figure 9, with
$C_b = \epsilon$
, varying tensions, and other parameters as prescribed at the beginning of the subsection. In this figure, the markers represent the numerical simulations of the thin-film model, while the solid lines correspond to the predictions from the asymptotic relation (3.69a
). Here, to get the radius
$a(t)$
, we solve (3.69a
) as an ordinary differential equation for
$a$
, with its right-hand side rewritten in terms of
$a$
using (3.18). Overall, the thin-film dynamics is well captured by the asymptotic relation, particularly for large tensions
$\gamma _1$
. The discrepancies are caused by the facts that the relation (3.69a
) is only a leading-order approximation, and that the bending length
$l_1 = \sqrt {C_b/\gamma _1}$
is not sufficiently small for low values of
$\gamma _1$
. Additionally, the assumption
$\dot {a} \sim \operatorname {O} (1)$
is slightly violated during the initial transient dynamics. The errors accumulated in the transient stage carry over to the later stage of the evolution.
For a liquid film spreading on a rigid substrate, the film radius grows algebraically (Tanner’s law) and then approaches its equilibrium value exponentially (Thampi et al. Reference Thampi, Pagonabarraga, Adhikari and Govindarajan2016). A similar quantitative description can be obtained for the present problem. Let
$a_{\infty }$
denote the equilibrium film radius predicted by (3.69a
). By substituting (3.18) and (4.2) into (3.69a
), we obtain
At early times when
$a_0$
is far below
$a_\infty$
, (4.4) can be approximated by
This describes the algebraic regime characterised by the
$t^{1/7}$
scaling. The exponent
$1/7$
is the same as that in the case of a rigid substrate. At later times, when
$\Delta a = a_0 - a_{\infty }$
is small, (4.4) can be approximated by
This is the exponential regime. Both regimes are quantitatively similar to the spreading on a rigid substrate, and differ in some constants only.
5. Conclusion
We studied the motion of a liquid film on an elastic sheet. We first introduced a thin-film model consisting of fourth-order partial differential equations that couple the dynamics of the fluid interface with the deformation of the elastic sheet, and derived the appropriate contact line conditions following the principles of non-equilibrium thermodynamics.
We then carried out a four-region matched asymptotic analysis on this model in the distinguished limit of small slip length
$\lambda \ll 1$
, capillary number
${\textit{Ca}} = \operatorname {O} (\epsilon)$
, and bending length
$l = \operatorname {O} (\epsilon)$
, where
$\epsilon$
is given in (1.1). The four regions are: (a) the outer region located at a distance of
$\operatorname {O} (1)$
from the contact line, where the apparent contact angles are measured; (b) the bending region of size
$\operatorname {O} (l)$
near the contact line where significant sheet bending occurs; (c) the inner region of distance
$\operatorname {O} (\lambda)$
from the contact line where fluid slip occurs; and (d) the intermediate region which connects the bending and inner regions. By matching the two-term asymptotic expansions across these regions, we obtained the central result of this study, namely, the generalised Cox–Voinov relation (1.3), which relates the contact line speed to the apparent contact angles. Predictions from this relation show good agreement with numerical solutions of the full thin-film model.
Notable differences in the film motion on an elastic sheet, compared to a rigid substrate, include retarded spreading and enhanced receding. These behaviours are well explained by the generalised Cox–Voinov relation. The increase of the sheet tension (or rigidity) leads to a decrease in the effective contact angle, which alters the contact line dynamics. However, the long-time spreading and receding behaviours, such as those described by Tanner’s law, remain quantitatively unchanged. Another interesting aspect of the generalised relation is its independence from the bending length. This suggests that the relation applies to a wide range of bending lengths, which is supported by numerical simulations.
The generalised Cox–Voinov relation (1.3) provides an effective model for the contact line motion in terms of measurable physical quantities, such as the apparent contact angles. This circumvents the need to resolve microscopic-scale dynamics, as required in the full model. The relation may serve as a useful tool in applications such as coating (Kajiya et al. Reference Kajiya, Brunet, Royon, Daerr, Receveur and Limat2014) and inkjet printing (Park & Moon Reference Park and Moon2006), where control of contact line motion is essential.
Our present analysis is based on the lubrication approximation and is therefore restricted to thin drops with small interfacial slopes. In addition, we have assumed negligible viscosity of the surrounding air and thus ignored its influence on the liquid dynamics. The generalised lubrication formulation (Snoeijer Reference Snoeijer2006; Chan et al. Reference Chan, Kamal, Snoeijer, Sprittles and Eggers2020) provides an appealing framework to relax these assumptions. By replacing the classical lubrication equation with this generalised formulation, one should be able to extend the analysis to moving contact lines with large contact angles.
We also expect that the insights and methodology developed in this work can be extended to other soft wetting problems, such as those involving viscoelastic substrates (Kajiya et al. Reference Kajiya, Daerr, Narita, Royon, Lequeux and Limat2013, Reference Kajiya, Brunet, Royon, Daerr, Receveur and Limat2014; Karpitschka et al. Reference Karpitschka, Das, Van Gorcum, Perrin, Andreotti and Snoeijer2015; Kansal et al. Reference Kansal, Bertin, Datt, Eggers and Snoeijer2024). To incorporate viscoelastic effects, one must resolve the wetting ridge deformation on the substrate (Jerison et al. Reference Jerison, Xu, Wilen and Dufresne2011; Karpitschka et al. Reference Karpitschka, Das, Van Gorcum, Perrin, Andreotti and Snoeijer2015). This requires identifying the relevant length scale on which the ridge arises, and determining the appropriate distinguished limit. We leave these extensions to future work.
Funding
The work was supported in part by Singapore MOE Academic Research Fund Tier 1 (A-8001948-00-00) and National University of Singapore (Suzhou) Research Institute.
Declaration of interests
The authors report no conflict of interest.
Code availability
The code used in this study is publicly available at https://github.com/zhixuan-zxli/Thin-Film-Code.
Appendix A. Numerical method
We use a moving mesh to handle the moving contact line, and below formulate (2.18) on a moving mesh. Let
$\xi$
be the coordinate of a time-independent mesh, and let
$x(\xi ,t) = a(t)\, \xi$
be the physical coordinate so that the mesh point
$\xi =1$
is always mapped to the contact line. We denote by
$\bar {h}$
and
$\bar {g}$
the pull-back of
$h$
and
$g$
onto the time-independent mesh (here,
$\bar {h}$
and
$\bar {g}$
should not be confused with the inner variables in § 3.3):
We rewrite the governing equation (2.18) on the time-independent mesh as follows:
where we have introduced the curvature
$\kappa =\partial _{xx}g$
and accordingly the pull-back
$\bar {\kappa }$
as new unknowns. We note that
$\bar {\kappa }$
is continuous across
$\xi =1$
, while
$\partial _\xi \bar {\kappa }$
has a jump at
$\xi =1$
.
For spatial discretisation, we use a bounded domain
$[0, \xi _r]$
, and discretise
$[0,1]$
into
$n_f$
cells, and
$[1,\xi _r]$
into
$n-n_f$
cells. We use integer subscripts, e.g.
$\xi _1, \ldots , \xi _{n_f}, \xi _{n_f+1}, \ldots , \xi _n$
, to denote the cell centres, and the half subscripts, e.g.
$\xi _{({1}/{2})}=0, \ldots , \xi _{n+({1}/{2})}=\xi _r$
, to denote the cell boundaries. They are related by
$\xi _i = (\xi _{i-({1}/{2})} + \xi _{i+({1}/{2})})/2$
for
$1 \leqslant i \leqslant n$
. For temporal discretisation, we use
$t^0 = 0, t^1, \ldots$
to denote the time instants. The function values at the cell centres are
We use
$\mathrm{D}$
and
$\mathrm{D}^2$
to denote the first- and second-order finite difference operators at the cell centres:
We use
$\mathrm{D}^3$
to denote the third-order finite difference operator at the cell boundaries,
In addition, we use
$\Delta$
to denote the forward time difference:
Suppose that we have obtained the numerical solutions
$h^m, g^m,a^m$
at the time step
$m$
. To advance in time, we first update the contact line position according to (2.20c
):
where
\begin{align} \alpha ^m = \frac {h^m_{n_f+1} - h^m_{n_f}}{a^m\left (\xi _{n_f+1} - \xi _{n_f}\right)}, \quad \beta ^m = \frac {g^m_{n_f+1} - g^m_{n_f}}{a^m\left (\xi _{n_f+1} - \xi _{n_f}\right)}, \end{align}
are the explicit estimates of (the tangents of) the dynamic contact angles. We then apply a semi-implicit discretisation on (A2a ) to obtain
\begin{align} \Delta h^m_i - \Delta g^m_i -\xi _i \frac {\Delta a^m}{a^{m+1}} \left (\mathrm{D} h^m_i - \mathrm{D} g^m_i\right) + \frac {1}{a^{m+1}\,{\textit{Ca}}}\frac {F^{m+1}_{i+\frac {1}{2}}-F^{m+1}_{i-\frac {1}{2}}}{\xi _{i+\frac {1}{2}} - \xi _{i-\frac {1}{2}}} = 0, \quad 1 \leqslant i \leqslant n_f, \end{align}
where
\begin{align} F^{m+1}_{i+\frac {1}{2}} = \left (\hat {h}^m_{i+\frac {1}{2}}-\hat {g}^m_{i+\frac {1}{2}}\right)^2 \left (\hat {h}^m_{i+\frac {1}{2}}-\hat {g}^m_{i+\frac {1}{2}}+\lambda \right) \frac {\mathrm{D}^3 h^{m+1}_{i+\frac {1}{2}}}{\left (a^{m+1}\right)^3} \end{align}
is a semi-implicit discretisation of the flux, and
$\hat {h}_{i+\frac {1}{2}}$
is the linear interpolation of
$h_i$
and
$h_{i+1}$
at
$\xi _{i+(1/2)}$
. For (A2b
)–(A2d
), we use a fully implicit discretisation:
\begin{align} \begin{gathered} \kappa ^{m+1}_i = \frac {\mathrm{D}^2 g^{m+1}_i}{\left (a^{m+1}\right)^2}, \quad 1 \leqslant i \leqslant n, \\ \frac {C_b }{\left (a^{m+1}\right)^2} \left (\mathrm{D}^2 \kappa ^{m+1}_i - \eta ^{m+1}\, \mathrm{D}^2 v^{m+1}_i\right) - \gamma _1 \kappa ^{m+1}_i = \frac {\mathrm{D}^2h^{m+1}_i}{\left (a^{m+1}\right)^2{\textit{Ca}}}, \quad 1 \leqslant i \leqslant n_f, \\ \frac {C_b }{\left (a^{m+1}\right)^2} \left (\mathrm{D}^2 \kappa ^{m+1}_i - \eta ^{m+1}\, \mathrm{D}^2 v^{m+1}_i\right) - \gamma _2 \kappa ^{m+1}_i = 0, \quad n_f+1 \leqslant i \leqslant n, \end{gathered} \end{align}
where
$\eta ^{m+1} = -[{\textit{Ca}}^{-1}\,\alpha ^{m+1}+(\gamma _1-\gamma _2)\beta ^{m+1}]/C_b$
discretises the jump
$[\![ \partial _x \kappa ]\!]$
, and
is a correction function for incorporating the jump condition (2.20b ) into the finite difference discretisation.
We implement the boundary conditions (2.19) with the help of ghost cells:
\begin{align} \begin{gathered} h^m_0 = h^m_1, \quad F^m_{\frac {1}{2}} = 0, \\ g^m_0 = g^m_1, \quad \kappa ^m_{0} = \kappa ^m_{1}, \\ g^m_n + g^m_{n+1} = 0, \quad \kappa ^m_n + \kappa ^m_{n+1} = 0. \end{gathered} \end{align}
At the contact line, we impose
To validate the numerical method, we perform a series of numerical simulations with successively refined spatial and temporal step sizes. For spatial discretisation, we find
$\xi _r = 2$
to be a good cut-off distance for the contact line dynamics to be unaffected. We use a uniform spatial mesh, with
$n_f = 128$
,
$n = 2n_f = 256$
and step size
$\Delta \xi = 1 / n_f$
. For temporal discretisation, we use
$t^m = m\, \Delta t$
with
$\Delta t = 1/512$
. The errors for the temporal and spatial refinement are measured by
\begin{align} \begin{aligned} {E}_{\Delta \xi , \Delta t} [h] &= \max _{1\leqslant i\leqslant n_f} \left \vert h^{m, \Delta t}_{i, \Delta \xi } - h^{2m, \Delta t/2}_{i, \Delta \xi } \right \vert\! , \\ {E}^*_{\Delta \xi , \Delta t} [h] &= \max _{1\leqslant i \leqslant n_f} \left \vert h^{m, \Delta t}_{i, \Delta \xi } - \frac {1}{2} \left ( h^{2m, \Delta t/2}_{2i-1, \Delta \xi /2} + h^{2m, \Delta t/2}_{2i, \Delta \xi /2} \right) \right \vert\! , \end{aligned} \end{align}
with
$m = t/\Delta t$
, and the final time
$t$
to be specified later. The numerical errors for the other unknowns are defined similarly. The parameters used in the refinement tests are
$\lambda =10^{-2}$
,
${\textit{Ca}} = \epsilon , \gamma _1 = 2/\epsilon , \gamma _2 = 2.9/\epsilon$
,
$C_b = \epsilon$
and
$\mu _{\varLambda }=1$
. We confirm the first-order convergence in both time and space, as shown in tables 1 and 2.
Table 1. Errors and convergence rates in the time-refinement tests.

Table 2. Errors and convergence rates in the space-refinement tests.

For the numerical results presented in § 4, where
$\lambda$
was taken to be
$10^{-4}$
, we use a mesh locally refined at the contact line. Specifically, we take mesh size
$\Delta \xi = 2^{-i}/32$
in the interval
$[1 - 2^{-(i-1)}, 1-2^{-i}]$
for
$i = 1, 2, \ldots , 9$
, and
$\Delta \xi _{\textit{min}} = 2^{-15}$
in the interval
$[1-2^{-9}, 1]$
. Meanwhile, we use adaptive time stepping by starting from
$\Delta t = 2^{-19}$
, and adjusting it every 128 time steps subject to the CFL-like condition
$\Delta t \leqslant \min \left \{ \Delta \xi _{\textit{min}} / 8\,\Delta a^m, 4\, \Delta \xi _{\textit{min}} \right \}$
.
Appendix B. Solution to a second-order ordinary differential equation
Consider the non-homogeneous second-order ordinary differential equation
The solution is
where
$C_+$
and
$C_-$
are constants of integration depending on initial or boundary conditions. The specific solution is
where
is the exponential integral. The derivatives of the specific solution satisfy
\begin{align} \begin{gathered} \varphi ^{\prime} = \frac {1}{2} \left [ e^x\,\textrm {Ei}(-x) + e^{-x}\,\textrm {Ei}(x) \right]\!, \\ \varphi ^{\prime\prime} = \frac {1}{2} \left [ e^x\,\textrm {Ei}(-x) - e^{-x}\,\textrm {Ei}(x) \right ] + \frac {1}{x} = \varphi + \frac {1}{x}. \end{gathered} \end{align}
There are two asymptotic series for the exponential integral (O’Malley Reference O’Malley2014),
\begin{align} \begin{gathered} \textrm {Ei}(x) \sim \log \left \vert x\right \vert + \gamma _{{E}} + e^{x/2}\left (x - \frac {1}{4}x^2 + {\cdots } \right)\!, \quad x \rightarrow 0, \\ \textrm {Ei}(x) \sim \frac {e^x}{x} \left ( \sum _{k=0} \frac {k!}{x^k} + {\cdots } \right)\!, \quad \left \vert x\right \vert \rightarrow \infty , \end{gathered} \end{align}
for small
$x$
and large
$x$
, respectively. Here,
$\gamma _{{E}} \approx 0.57722$
is Euler’s constant. From these series, we have the following asymptotic expansions for the specific solution:
\begin{align} \begin{gathered} \varphi \sim x \log \vert x\vert + \left (\gamma _{{E}}-1\right)x, \quad \varphi ^{\prime} \sim \gamma _{{E}} + \log \vert x\vert \quad \text{as } x \rightarrow 0, \\ \varphi \sim -\frac {1}{x}, \quad \varphi ^{\prime} \sim \frac {1}{x^2} \quad \text{as } \vert x\vert \rightarrow \infty . \end{gathered} \end{align}












































