1. Introduction
Partially wetting droplets are ubiquitous in nature and industry, whether they are raindrops on windows or 3D-printed material on substrates. Despite the simplicity of its individual components, liquid drops on solid surfaces are a source of surprisingly rich physics. One such interesting problem arises when a high-Reynolds-number air flow sweeps across a droplet-laden surface. In contrast to partially wetting drops in the low-Reynolds-number flow regime, which has been studied extensively (Dussan Reference Dussan1979; Dussan & Chow Reference Dussan and Chow1983; Dussan Reference Dussan1985, Reference Dussan1987; Basu, Nandakumar & Masliyah Reference Basu, Nandakumar and Masliyah1997; Dimitrakopoulos & Higdon Reference Dimitrakopoulos and Higdon1997, Reference Dimitrakopoulos and Higdon1998; ElSherbini & Jacobi Reference ElSherbini and Jacobi2004a,Reference ElSherbini and Jacobib; Dimitrakopoulos Reference Dimitrakopoulos2007), only a limited number of studies exist in the high-Reynolds-number regime (Durbin Reference Durbin1988; Ding & Spelt Reference Ding and Spelt2007; Hooshanginejad & Lee Reference Hooshanginejad and Lee2017). In this flow regime, the droplet can exhibit complex behaviours that range from pinning, runback and even droplet breakup. This unexplored coupling between the inertial external flow and the droplet dynamics is particularly relevant in aircraft icing and deicing applications, atomised lubrication in micromachining, and discrete coating and 3D-printing.
Recently, Hooshanginejad & Lee (Reference Hooshanginejad and Lee2017) experimentally applied a uniform air flow to a partially wetting droplet that is placed behind a solid hemisphere on a textured substrate. In the same set-up without the hemisphere, the droplet would simply depin and move downstream once the applied air forcing is large enough to overcome the contact-angle hysteresis that pins the droplet in place, as experimentally investigated by Schmucker (Reference Schmucker2013). When the droplet is placed behind a solid hemisphere, the air flow separates on the leeward side of the solid before reattaching onto the substrate, which can alter the droplet's response to incoming air flow. Depending on its location relative to the reattachment length, the droplet is observed to dislodge either (downstream) away from or (upstream) towards the solid. When the partially wetting drop is placed between the separated and reattaching air flows, the droplet splits into two satellite drops. Here, the droplet splitting arises from the stagnation-point pressure exerted on the droplet interface as free streamlines reattach onto the droplet.
To gain a better understanding of the droplet breakup induced by a stagnation line, we consider a model of a droplet under a two-dimensional (2-D) stagnation-point flow. In a stagnation-point flow, the pressure decreases quadratically from the stagnation point and drives the flow inside the droplet away from it. Similarly, a spin-coated droplet experiences a centrifugal force that decreases quadratically from the spinning axis and pushes the fluid towards the edge. Hence, the spreading mechanism of a drop in the stagnation-point flow is inherently similar to a spin-coated droplet in the limit of weak surface tension. Motivated by this analogy, a drop or a thin film subject to a jet of air blowing vertically downwards onto the substrate has been previously investigated. Moriarty, Schwartz & Tuck (Reference Moriarty, Schwartz and Tuck1991) used the method of matched asymptotic expansions to compute free surface shapes in the limit of low surface tension by dividing the flow inside the liquid film into inner and outer solutions near to and far from the edges, respectively. In addition, they compared their asymptotic solution with numerical results for some arbitrary input parameters, but without providing further details on the time evolution of the droplet shape. Later, McKinley, Wilson & Duffy (Reference McKinley, Wilson and Duffy1999) investigated the quasi-static solutions of a spreading drop attached to a substrate in the presence of a jet of air in both 2-D and axisymmetric geometries. In the subsequent series of studies, McKinley & Wilson (Reference McKinley and Wilson2001) performed linear stability analyses for a droplet and a ridge of fluid in similar flow configurations. Despite these pioneering analytical studies on the droplet dynamics subject to a jet of air, no systematic experiments have been performed to validate the model results. In addition, the criterion for the droplet breakup has not been explored in any of the proposed models.
In the present study, we perform experiments to examine partially wetting water droplets subject to a jet of air blowing vertically onto a horizontal substrate. This represents a simplified realisation of a droplet that is caught between the separated and reattaching air flows, as observed in the wake of a solid hemisphere. Our experiments reveal that the droplet undergoes splitting when the speed of the applied jet exceeds a critical value that depends on the droplet size and the relative position of the jet. To rationalise our experimental observations, we use a 2-D mathematical model to describe the droplet behaviour subject to a stagnation-point flow. Distinct from previous studies that focused on the propagating front of the droplet and its stability (Moriarty et al. Reference Moriarty, Schwartz and Tuck1991; McKinley et al. Reference McKinley, Wilson and Duffy1999; McKinley & Wilson Reference McKinley and Wilson2001), the goal of our present study is to develop a criterion for the droplet breakup in the stagnation-point flow, by resolving the draining mechanism from the stagnation point. Our droplet lubrication model incorporates potential flow theory for the external flow to simultaneously compute the internal drop flow and corresponding drop deformations. Our treatment of the external flow and the evolving droplet shape is similar to the model developed by Moore et al. (Reference Moore, Ristroph, Childress, Zhang and Shelley2013), who studied eroding bodies in high-Reynolds-number fluid flows. Our model is also reminiscent of other theoretical works (e.g. Durbin Reference Durbin1988; McKinley et al. Reference McKinley, Wilson and Duffy1999; Smith, Li & Wu Reference Smith, Li and Wu2003) that combine the lubrication theory with inviscid external flows in different geometries. In particular, the work by Smith et al. (Reference Smith, Li and Wu2003) considers the thin air flow between the solid wall and an inviscid body of water, by combining potential theory for water and lubrication approximations for air. Our work will follow the same general formulation, with lubrication theory applied to the internal flow of the droplet, while potential theory is used to describe the inviscid outer air flow.
The rest of the paper is organised as follows. In § 2, we present our experimental set-up followed by our experimental observations. In § 3, we describe our mathematical model, which couples the droplet dynamics with the external flow. The comparison between the model and experiments is presented in § 4, followed by the summary of our findings and future studies in § 5.
2. Experiments
We first discuss laboratory experiments in which we apply a uniform 2-D jet of air to a partially wetting water droplet placed on an aluminium surface. Our choice of the 2-D jet leads to a stagnation streamline impinging on the droplet surface, with the primary droplet dynamics occurring in the $x$–
$z$ plane normal to the solid surface and the jet, as shown in figure 1(a). The quasi-2-D nature of our experiments has two important implications. First, under certain conditions, the jet induces a 2-D splitting of the partially wetting droplet as observed in the wake of a solid hemisphere (Hooshanginejad & Lee Reference Hooshanginejad and Lee2017). Second, it allows us to qualitatively model the observed phenomena using a 2-D droplet model, to be discussed in § 3.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig1.png?pub-status=live)
Figure 1. The experimental set-up to study water drops under a high-speed stagnation-point flow. (a) A 2-D jet provides a unidirectional flow, and the side-view image ($x$–
$z$ plane) of the droplet is photographed every 0.0005 s. The jet speed
$U_0$ is controlled by a back-pressure regulator and measured by a hot-wire anemometer. (b) The side-view schematic of a water drop under a 2-D jet (
$H_0=2\ {\rm cm}$,
$D_0=0.2\ {\rm mm}$,
$\theta _0=30^{\circ }$).
2.1. Experimental method
The schematic in figure 1(a) illustrates the key features of our experimental apparatus. The air flow is supplied by an air compressor through a rectangular nozzle (width $25.4\ \textrm {mm} \times \textrm {length}\ 0.2\ \textrm {mm}$). The flow pressure is controlled by a back-pressure regulator (Wilkerson), while a diffuser and contraction region ensure a low-turbulence flow (i.e. turbulence intensity below 2 %) of nearly uniform velocity at the outlet. In each experimental run, we first deposit a distilled water droplet of volume
$V$ on a polished aluminium surface (240 grit SiC); we allow the droplet to reach an equilibrium shape before applying the jet of air. We measure the droplet's static contact angle,
$\theta _0$, with the solid substrate by first identifying the water–air interface via image processing and computing its slope close to the contact line;
$\theta _0$ is approximately equal to
$30^{\circ }$ for all experiments reported herein. The initial droplet half-width is given by
$L$, and the distance between the nozzle outlet and the substrate,
$H_0$, is fixed at 2 cm, as illustrated in figure 1(b). Then, prior to opening the air compressor valve, we set a desired pressure for the regulator to ensure that the jet speed at the outlet is kept constant throughout the entire experiment. We measure the flow speed,
$U_0$, along the centreline at the outlet using a hot-wire anemometer (Testo). For the range of
$U_0$ in this study, typical temporal fluctuations in air speed are measured to be 3 %. Similarly, variations in jet speed along the centreline at different lateral positions are less than 5 % for locations at least 5 mm from the sides of the outlet.
The experiments are backlit by a light-emitting diode panel (ASD Lighting) and recorded with a high-speed camera (Photron, 2000 frames per second, $1280\times 1024$ resolution) and a macro lens (Nikon). The camera is placed in front of the droplet (see figure 1a) to capture the side-view images of the evolving droplet. We systematically vary the droplet volume (i.e.
$V=20\text {--}200\ \mathrm {\mu }\textrm {l}$) as well as the speed of air flow
$U_0$, which can go up to 40 m s
$^{-1}$. The characteristic Reynolds number of the air flow corresponds to
$Re=U_0L/\nu _{a}=10^{3}$ to
$8\times 10^{3}$, where
$\nu _{a}=15\times 10^{-6}\ \textrm {m}^{2}\ \textrm {s}^{-1}$ is the kinematic viscosity of air at room temperature.
2.2. Experimental observations
In the first set of experiments, we examine the dynamic response of a partially wetting droplet when a 2-D jet of air is applied along the drop's centreline for varying $V$ and
$U_0$. For given
$V$, the droplet exhibits two different behaviours – ‘hanging’ versus ‘splitting’ – depending on the value of
$U_0$; the transition between the two regimes is marked by the threshold air speed
$U_{cr}$. When the air flow first reaches the drop, the droplet deforms into a concave shape at the centre (see figure 2a,ii), which corresponds to the stagnation point of the jet with the maximum external pressure. The pressure difference between the high-pressure centreline and the low-pressure edges drives an internal flow towards the two opposite sides of the droplet, which results in the observed necking in the centre. For
$U_0<U_{cr}$ (‘hanging’ regime), the neck continues to thin in the
$z$-direction until the droplet reaches an equilibrium state. The droplet continues to oscillate about this equilibrium configuration, which is related to the propagation of capillary waves along the droplet surface (Milne et al. Reference Milne, Defez, Cabrerizo-Vílchez and Amirfazli2014). However, the unsteady oscillations that are observed after the droplet reaches the hanging state are outside the scope of the current study. When
$U_0>U_{cr}$ (‘splitting’ regime), the necking along the centreline of the droplet thins further until the droplet splits into two (figure 2a,iv).
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig2.png?pub-status=live)
Figure 2. (a) Visualising the time evolution of droplet shapes held against a 2-D stagnation-point flow. (i) A partially wetting droplet of 200 $\mathrm {\mu }$l is initially in the equilibrium state (
$t=0$). (ii) As the droplet is subject to a 2-D jet of air with velocity
$U_0=10.8\ \textrm {m}\ \textrm {s}^{-1}$, it forms a neck along the centreline (
$t=0.3\ \textrm {s}$). (iii) The thinning of the meniscus continues until the drop reaches the onset of breakup (
$t=0.55\ \textrm {s}$). (iv) The two resultant drops depin and move in opposite directions (
$t=0.75\ \textrm {s}$). (b) The plot of the critical jet speed for splitting,
$U_{cr}$, as a function of the droplet volume
$V$; the value of
$U_{cr}$ has been extracted from the experiments.
Figure 2(b) shows the plot of $U_{cr}$ as a function of the droplet volume
$V$. As
$V$ increases, the droplet splits at increasingly lower jet speeds. This can be qualitatively explained in terms of the capillary pressure. As the droplet size increases, the magnitude of the mean curvature of the drop must decrease, corresponding to an overall decrease in the drop's internal pressure. Therefore, a lower stagnation pressure is required to drive the internal flow from the drop's centre towards the edges and to result in a breakup, compared with smaller droplets.
In the second series of experiments, we examine the effect of systematically varying the stagnation-point position relative to the droplet's centre for given $V$ and
$U_0$. Here,
$\alpha$ characterises the dimensionless distance between the centre of the jet and that of the drop normalised by
$L$ (i.e.
$\alpha =1$ at the contact line). Analogous to the
$\alpha =0$ case, at low jet speeds, the droplet reaches the ‘hanging’ equilibrium state and does not split, as shown in figure 3(a). However, at intermediate
$U_0$ between the hanging and splitting regimes, a new droplet behaviour emerges. As shown in figure 3(c), droplets depin towards the side of the stagnation point with a larger droplet volume, for a range of
$U_0$ that depends on
$\alpha$ and
$V$. In this ‘depinning’ regime, the droplet initially forms the neck at the stagnation position. Once the neck reaches a minimum thickness, the fluid on the smaller side of the drop drains towards the larger side due to relative pressure differences inside the droplet. Then, at higher values of
$U_0$, drops split into two, as shown in figure 3(b). Note that
$U_{cr}$ marks the onset of the splitting regime and is now a function of
$V$ and
$\alpha$.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig3.png?pub-status=live)
Figure 3. (a–c) Different behaviours of a $50\ \mathrm {\mu }\textrm {l}$ droplet under an off-centred high-speed jet flow. (a) A jet with
$U_0=6.5\ \textrm {m}\ \textrm {s}^{-1}$ located at
$\alpha =0.2$ deforms the droplet until it reaches the equilibrium. (b) The droplet is subject to a jet with
$U_0=17.8\ \textrm {m}\ \textrm {s}^{-1}$ at
$\alpha =0.2$. The neck that is formed under the jet thins until it splits into two smaller drops of different sizes. (c) The initial extensional flow inside the droplet caused by
$U_0=8.7\ \textrm {m}\ \textrm {s}^{-1}$ at
$\alpha =0.2$ transitions to a unidirectional flow from the smaller side to the larger side of the drop, resulting in a drifting behaviour. (d,e) Summary of different droplet responses to a high-speed jet for varying jet velocity
$U_0$ and non-dimensional jet position
$\alpha$, for (d)
$V=50\ \mathrm {\mu } \textrm {l}$ and (e)
$V=200\ \mathrm {\mu } \textrm {l}$.
Figure 3(d,e) summarises the three droplet behaviours in an $\alpha$–
$U_0$ phase diagram for
$V=50$ and
$200\ \mathrm {\mu }\textrm {l}$, respectively. For
$V=200\ \mathrm {\mu }\textrm {l}$, the ‘depinning’ regime becomes suppressed for small
$\alpha$, and the droplet transitions from hanging to splitting regimes directly with increasing
$U_0$. When the stagnation point moves further from the initial centre of the drop (i.e.
$\alpha >0.2$), the depinning behaviour emerges. By contrast, for
$V=50\ \mathrm {\mu }\textrm {l}$, all three droplet behaviours are observed with an increase in
$U_0$, for all non-zero
$\alpha$ values considered (see figure 3e). Furthermore, for both small and large droplets, an increase in
$\alpha$ consistently shifts the onset of the splitting regime to a higher jet speed (i.e.
$U_{cr}$ increasing with
$\alpha$), which will be explored in a later section.
2.3. Capillary pressure inside the evolving droplet
The internal flows that lead to ‘hanging’, ‘splitting’ or ‘depinning’ regimes are driven by the non-uniform pressure distribution inside the droplet. Hence, we now estimate the capillary pressure inside the evolving drop, by measuring the local in-plane curvature $\kappa (t)$ at the stagnation point, and at the left and right corners of the droplet, respectively (denoted with the subscripts ‘
$s$’, ‘
$l$’ and ‘
$r$’) and applying the Young–Laplace equation. The dimensionless external pressure over the droplet in the
$x$–
$z$ plane is approximated by a quadratic function,
$P^{*}=P^{*}_{s}-(L/H_0)^{2}(x^{*}-\alpha )^{2}$, where
$x^{*}=x/L$ and
$P^{*}_{s}$ is the dimensionless stagnation pressure at
$x^{*}=\alpha$, normalised by
$(1/2)\rho _{a}U^{2}_0$. Here,
$\rho_a$ corresponds to the density of air. By neglecting the drop's out-of-plane curvature, the non-dimensional internal pressure
$p^{*}$ can be approximated as
$p^{*}(t)=P^{*}-2\sigma \kappa (t)/(\rho _{a} U^{2}_0)$, where
$\sigma$ is the surface tension coefficient. Hence, the values of
$p^{*}-P^{*}_{s}$ at a given time are determined by extracting
$\kappa _{s}(t)$,
$\kappa _{l}(t)$ and
$\kappa _{r}(t)$ from the experimental images.
We focus on the case of $\alpha = 0.2$ and
$V=50\ \mathrm {\mu }\textrm {l}$, as presented in figure 3. At the initial time
$t=0$,
$\kappa _{s}(0)\approx \kappa _{r}(0)\approx \kappa _{l}(0)$, so that we obtain
$p^{*}_{s}(0)>p^{*}_{r}(0)>p^{*}_{l}(0)$, mirroring
$P^{*}_{s}(0)>P^{*}_{r}(0)>P^{*}_{l}(0)$ (see figure 4a). This internal pressure gradient must drive an asymmetric diverging flow from the stagnation point towards both sides of the droplet. As fluid drains from the stagnation point,
$\kappa _s$ increases, while
$\kappa _{r}$ and
$\kappa _{l}$ decrease, as illustrated in figure 4(b). The resultant shape deformations cause time-dependent changes in the internal pressure. Figure 4(c–e) shows the plots of
$p^{*}-P^{*}_{s}$ over time at the stagnation point (
$x^{*}=\alpha$), and at the left (
$x^{*}=-1$) and right (
$x^{*}=1$) corners of the drop.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig4.png?pub-status=live)
Figure 4. (a) The schematics of a droplet's initial shape at the moment we apply the jet to an arbitrary off-centred point on the droplet interface. (b) The droplet shape at time $t$ can be divided into three zones: the stagnation zone with a concave shape, the smaller convex side with a higher mean curvature, and the larger convex side with a lower mean curvature. (c–e) The relative internal pressure in the different droplet zones is computed from the extracted mean curvature, for (c) the hanging regime in figure 3(a), (d) the splitting regime in figure 3(b), and (e) the depinning regime in figure 3(c).
When $U_0$ is low (i.e.
$6.5\ \textrm {m}\ \textrm {s}^{-1}$), the shape deformations continue until the internal pressure becomes uniform, reaching the steady state at time
$t_{f}$. This is evidenced by
$p^{*}_{s}$,
$p^{*}_{l}$ and
$p^{*}_{r}$ converging to a single value in figure 4(c) and corresponds to the ‘hanging’ regime (see figure 3a). If
$U_0$ is increased to 17.8 m s
$^{-1}$, the deformations are not adequate to equalise the internal pressure before the necking yields breakup at
$t_{f}$, as evident in figure 4(d). This exemplifies the ‘splitting’ regime, previously shown in figure 3(b).
At $U_0=8.7 \ \textrm {m}\ \textrm {s}^{-1} < U_{cr}$,
$p^{*}_{r}-P^{*}_{s}$ gradually increases past
$p^{*}_{s}-P^{*}_{s}$. This reverses the direction of the part of the internal flow from the positive
$x$-direction (from the stagnation point to the right-hand side) to the negative
$x$-direction (from right to left), which leads to droplet depinning. This flow reversal is evident in the time-sequential images in figure 3(c): fluid first drains away from the stagnation point (i), then temporarily stalls as the internal pressure on the right-hand side marginally surpasses that of the stagnation point at time
$t_{f}$ (ii), and finally drains from right to left (iii, iv) before depinning (v). Therefore, the ‘depinning’ regime is marked by the gradual pressure buildup above the stagnation pressure on one side of the droplet, which drives the net internal flow in one direction before the internal pressure can equalise. Figure 4(e) demonstrates the evolution of the internal pressure up to the moment when the droplet stalls, as illustrated in figure 3(c,ii). Hence,
$t_{f}$ in figure 4(e) marks the transition from the diverging flow to a unidirectional flow inside the droplet.
The direct correlation between $p^{*}-P^{*}_{s}$ and the experimental observation signifies that the external pressure field (modelled as a stagnation-point flow) and the droplet curvature in the
$x$–
$z$ plane constitute the two dominant physical effects in our current system. Hence, we will construct a theoretical model of the evolving droplet that incorporates these two effects to leading order.
3. Theory
In order to rationalise the experimental observations, we develop a mathematical model of a partially wetting droplet under a 2-D jet flow. Owing to the high-$Re$ flow exterior to the drop, we divide the external fluid domain into the outer and inner regions (Moore et al. Reference Moore, Ristroph, Childress, Zhang and Shelley2013). Here, the external air flow comprises a far-field irrotational stagnation-point flow with characteristic velocity
$U_0$ and a viscous boundary layer surrounding the droplet, as illustrated in figure 5.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig5.png?pub-status=live)
Figure 5. The schematic of our 2-D droplet model that couples the droplet internal flow with the external flow. A thin droplet is subject to a high-speed stagnation-point flow that comprises an outer (irrotational) flow, which is described by a potential flow theory, and an inner (boundary layer) flow, where we use the Prandtl boundary layer equations. The table provides the range of the dimensionless parameters that are based on the experiments and applied in our model.
We make three major assumptions in our model. First, we model the system as completely 2-D, neglecting variations in the $y$-direction. While the 2-D model cannot quantitatively resolve the flow inside the three-dimensional (3-D) droplet, it qualitatively captures the key physical features of the droplet dynamics due to the quasi-2-D nature of the external flow. In § 2.3, we have demonstrated that the in-plane curvature of the droplet alone can sufficiently capture the droplet behaviour, while the out-of-plane curvature has been neglected. This helps validate the use of the 2-D model. In addition, the working limits of the 2-D approach are further explained in the appendix.
Second, we assume that the 2-D droplet, with height $h(x,t)$ and half-width
$L$, is ‘thin’ (i.e.
$h/L \ll 1$) and employ lubrication approximations to reduce the governing equations. In the lubrication limit, the time-dependent droplet deformations do not affect the far-field external flow. We acknowledge that the lubrication approximation is strictly valid for
$\theta _0 \ll~1$, which may break down as we employ
$\theta _0$ as large as
$30^{\circ }$. However, previous studies (e.g. Krechetnikov Reference Krechetnikov2010; Espín & Kumar Reference Espín and Kumar2017) have demonstrated that the lubrication approximation can work beyond the strictly small-slope limits. Furthermore, we emphasise only qualitative comparisons with the experiments, instead of quantitative ones. The model limitations stemming from this assumption are later addressed in the results section.
Third, we assume that the droplet is pinned at the contact line, which deviates from experiments in the splitting and depinning regimes. Despite its clear limitations, the 2-D pinned model retains the leading-order mechanism of the droplet splitting – namely, the fluid drainage from the centre towards the edges, without the complexity of the moving contact line. Existing moving contact line models introduce an additional empirical parameter incorporating van der Waals forces between the liquid drop and the solid surface (e.g. Bertozzi Reference Bertozzi1998; Schwartz & Eley Reference Schwartz and Eley1998; Schwartz et al. Reference Schwartz, Roy, Eley and Princen2004; Park & Kumar Reference Park and Kumar2017), which will be considered in a future study.
3.1. Lubrication equation
Under lubrication approximations, the linear momentum equations inside the droplet reduce to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn1.png?pub-status=live)
where $\mu _{d}$ and
$\rho _{d}$ are the dynamic viscosity and density of the droplet, respectively, while the pressure inside the drop is given by
$p_{d}(x,z,t)$, and the gravitational acceleration by
$g$. We solve for the horizontal velocity inside the drop,
$u_{d}(x,z,t)$, by integrating (3.1a,b) subject to boundary conditions on the droplet surface,
$z=h(x,t)$:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn2.png?pub-status=live)
in addition to $u_{d}(z=0)=0$. Here, the first corresponds to the continuity of shear stress at the fluid–air interface, where
$\tau _{s}$ denotes the external shear stress acting on the droplet surface. The second condition relates the jump between
$p_{d}$ and external pressure
$P$ due to surface tension (with coefficient
$\sigma$); here
$\kappa (x,t)$ denotes the planar curvature of the droplet surface, which reduces to
$h''(x)$ in the lubrication limit. We non-dimensionalise the governing equations and boundary conditions using the following rescalings:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn3.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn4.png?pub-status=live)
Here $\varepsilon =\tan (\theta _0/2)\approx \theta _0/2\ll 1$ is the expansion parameter and a measure of the surface wettability; and
$\mu _{a}$ refers to the dynamic viscosity of air. Combining (3.1a,b) and (3.2a,b) with the continuity equation yields the dimensionless evolution equation for
$h^{*}(x,t)$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn5.png?pub-status=live)
with capillary number $Ca=\mu _{d}U_0/\sigma$, Weber number
$We=\rho _{a}U_0^{2}L/\sigma$ and Bond number
$Bo=\rho _{d}gL^{2}/\sigma$. The values of all the relevant dimensionless numbers, including the Reynolds number
$Re$, are tabulated in figure 5, based on the experimental parameters. Equation (3.5) requires four boundary conditions, which are given as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn6.png?pub-status=live)
where $\kappa ^{*}_0$ is the initial dimensionless curvature of the droplet at the corners. The boundary conditions specify a fixed contact line with a constant curvature at the edge of the droplet, as previously implemented by Constantin et al. (Reference Constantin, Dupont, Goldstein, Kadanoff, Shelley and Zhou1993) to simulate droplet breakup inside a Hele-Shaw cell. The constant-curvature condition leads to an internal flow towards the sides of the drop that could cause splitting (Constantin et al. Reference Constantin, Dupont, Goldstein, Kadanoff, Shelley and Zhou1993; Bertozzi Reference Bertozzi1998).
To solve (3.5), we must model the external flow to find expressions for $P$ and
$\tau _{s}$. Consistent with the lubrication approximation, the external flow is modelled as a steady stagnation-point flow over a flat surface to leading order; the presence of the thin droplet is incorporated into the external flow model at
$O(\varepsilon )$. Based on the high-Reynolds-number nature of the air flow (i.e.
$Re \sim 10^{3}\text {--}10^{4}$), we further divide the external flow into outer and inner regions (see figure 5). To characterise the inner region, we estimate the thickness of the boundary layer,
$\delta /L$, as
$Re^{-1/2}\sim 10^{-2}$, which is an order of magnitude smaller than
$\varepsilon \sim 10^{-1}$, or the dimensionless droplet height. In the current limit of
$Re^{-1/2} \ll \varepsilon$, the effects of viscous shear stress
$\tau _{s}$ must be an order of magnitude smaller than the external pressure
$P$ (Smith et al. Reference Smith, Brighton, Jackson and Hunt1981; Durbin Reference Durbin1988). Therefore, we reasonably assume that the droplet dynamics is governed by
$P$ to leading order, which matches our experimental observations in § 2.3. Notably, the current problem is in a completely different physical regime from the well-established work of Fry (Reference Fry2012), in which a droplet or a protrusion is fully submerged inside the inner deck of the viscous boundary layer, with no imposed external pressure gradient (see also Smith et al. Reference Smith, Brighton, Jackson and Hunt1981).
The relative sizes of $P$ and
$\tau _{s}$ can also be deduced by directly comparing the first and last terms on the right-hand side of (3.5). The first term corresponds to the flux inside the droplet that is induced by the external pressure gradient, while the last term corresponds to that due to external shear stress. The ratio between the two terms scales as
$Re\,\varepsilon ^{2}\sim \varepsilon ^{-1}$, confirming that the shear stress term is indeed at least
$O(\varepsilon )$ smaller than the external pressure gradient term. Based on the scaling argument, we calculate the external pressure up to
$O(\varepsilon )$, which accounts for the linearised effect of the droplet, and shear stress to
$O(1)$. The latter is equivalent to computing shear stress due to a stagnation-point flow over a flat plate only. Our method consists of three steps:
(1) Determine the external pressure
$P$ in the outer flow and shear stress
$\tau _{s}$ in the inner flow along the solid surface to
$O(1)$.
(2) Given the droplet shape at time
$t$, compute
$P$ to
$O(\varepsilon )$.
(3) Evolve the droplet shape by solving (3.5) and return to step 2.
Owing to the importance of the $O(1)$ outer velocity field for resolving
$\tau _s$, we first describe our approach for computing
$P$ in the irrotational outer flow (steps 1 and 2). Next, we discuss the solution to the Prandtl boundary layer equations (step 1). In § 4, we present our numerical simulations (step 3). Finally, we include the steady-state analysis of the governing equations and the corresponding analytical results to justify experimental observations.
3.2. External flow: outer region
In the outer region sufficiently far from the boundaries, we take the flow as inviscid. As illustrated in figure 6(a), we rescale the outer flow with $[\tilde {x}]= [\tilde {z}]=H_0$, where
$H_0$ is the distance between the flow source and the solid surface (i.e.
$\tilde {x}=(L/H_0)x^{*}$). Accordingly, the droplet height in the rescaled coordinates is given by
$\tilde {z}=\xi h^{*}$, where
$\xi =\varepsilon L/H_0$ is a modified small parameter, and
$L/H_0=O(1)$. The external velocity
$U$ is scaled by
$U_0$, such that
$\tilde {U}=U/U_0$. In general, the tilde denotes dimensionless variables in the external flow, while the asterisk indicates those from the original lubrication equation.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig6.png?pub-status=live)
Figure 6. (a) The schematic of the potential flow past the drop and the flat plate with zero-flux condition at the boundaries. (b) The computed values of the dimensionless pressure gradient over the droplet interface for different droplet shapes, when a jet of $U_0=10\ {\rm m}\ {\rm s}^{-1}$ is applied to a droplet with
$L=7.7\ \textrm {mm}$ and
$\theta _0=10^{\circ }$.
As the outer flow is irrotational, $\tilde {U} = \tilde {\boldsymbol {\nabla }} \phi$, where
$\phi$ is a dimensionless harmonic velocity potential. We then expand
$\phi$ as a power series of
$\xi$:
$\phi =\phi _0+\xi \phi _1+O(\xi ^{2})$. The leading-order velocity potential reduces to the stagnation-point flow over a flat surface and is given by
$\phi _{0}=\frac {1}{2}(\tilde {x}^{2}-\tilde {z}^{2} )$. Then at
$O(\xi )$, the presence of a thin droplet is incorporated into the outer flow through the modified geometry of the boundaries, which must satisfy the no-flux condition. The linearised velocity potential gradient on the droplet surface corresponds to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn7.png?pub-status=live)
Then, at $O(\xi )$, the zero-flux condition on the surface becomes
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn8.png?pub-status=live)
We note that (3.8) neglects the effect of the kinematic boundary condition transmitted through the inner flow, as such corrections appear at the next order (i.e. boundary layer thickness, $\delta /L\sim O(\varepsilon ^{2}$)). By setting
$\phi _{1}=\sum _{k=-\infty }^{\infty} \hat {\phi }_{k}(\tilde {z},t^{*})\,\textrm {e}^{\textrm {i}k{\rm \pi} x^{*}}$, we note that
$\tilde {\nabla }^{2}\phi _1=( (\partial ^{2}/\partial \tilde {z}^{2})-k^{2}{\rm \pi} ^{2} ) \hat {\phi }_{k}=0$, which yields
$\hat {\phi }_{k}(\tilde {z},t^{*})=A_{k}(t^{*})\,\textrm {e}^{| k | {\rm \pi}\tilde {z}}+B_{k}(t^{*})\,\textrm {e}^{-| k | {\rm \pi}\tilde {z}}$. The boundedness of
$\phi _{1}$ requires that
$A_{k}=0$. By substituting
$\phi _1$ into the left-hand side of (3.8) and through the Fourier expansion of the right-hand side, we obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn9.png?pub-status=live)
where $\mathcal {F}(X(\omega ))=\frac {1}{2}\int _{-1}^{+1}X(\omega )\, \textrm {e}^{-\textrm {i}k{\rm \pi} \omega }\,{\textrm {d}}\omega$. In addition, the Bernoulli equation applies throughout the outer region, such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn10.png?pub-status=live)
where $P^{*}_0$ is the dimensionless pressure at the nozzle. After substituting (3.7) into (3.10), the external pressure on the interface can be written as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn11.png?pub-status=live)
The corresponding external pressure gradient ${\textrm {d}}P^{*}/{\textrm {d}}x^{*}$ is plotted as a function of
$x^{*}$ in figure 6(b) at different times. It demonstrates the extent to which external pressure is modified due to the evolving droplet shape.
3.3. External flow: inner region
In this section, we outline our approach for computing the shear stress $\tau _s$ to leading order, which we later incorporate into (3.5). Thus, we only consider the inner region, which consists of a thin layer surrounding the solid surface where the effects of viscosity come to bear, as shown in figure 7. The Navier–Stokes equations in the boundary layer reduce to Prandtl boundary layer equations (Prandtl Reference Prandtl1904; Schlichting Reference Schlichting1960):
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn12.png?pub-status=live)
where $(u,v)$ denote velocity components of the external flow in the inner region, while
$U$ refers to the fluid velocity in the inviscid outer flow. The Prandtl equations hold within a boundary layer of thickness
$\delta (x)$, which has characteristic scale
$\bar \delta =L/\sqrt {Re}$. No-slip boundary conditions and matching conditions to the outer flow include
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn13.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn14.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig7.png?pub-status=live)
Figure 7. (a) The schematic of the leading-order boundary layer flow on the flat substrate. (b) The inner flow velocity profile as a function of the normal distance from the solid boundary at $x=L$ using the similarity solution by Hiemenz (Reference Hiemenz1911) and the VKP method. (c) The plot of the shear stress on the surface as a function of
$x$ based on the similarity solution and the VKP method for
$L=7.7\ \textrm {mm}$,
$U_0=10\ \textrm {m}\ \textrm {s}^{-1}$ and
$\theta _0=10^{\circ }$.
While a similarity solution to (3.12a,b) exists (Hiemenz Reference Hiemenz1911; Howarth Reference Howarth1935), in our simulations, we employ the von Kármán-Pohlhausen (VKP) method to approximate the solution to the Prandtl equations (von Kármán Reference von Kármán1921; Pohlhausen Reference Pohlhausen1921). This method approximates the inner flow with a fourth-order polynomial that satisfies the boundary conditions (3.13a,b). Once we find $\delta (x)$ with the VKP method, we can write the shear stress on the solid surface in this method as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn15.png?pub-status=live)
Figure 7(b,c) compares the velocity profile and the tangential shear stress over the surface for the similarity solution and the VKP method for sample input parameters (i.e. $U_0=10\ \textrm {m}\ \textrm {s}^{-1}$,
$L=7.7\ \textrm {mm}$ and
$\theta _0=10^{\circ }$). More details about the VKP calculations can also be found in the related study by Moore et al. (Reference Moore, Ristroph, Childress, Zhang and Shelley2013).
4. Results
4.1. Time-dependent solutions
In this section, we present the numerical simulations of droplet shapes subject to a stagnation-point flow along its centreline. We use a second-order implicit method of Crank–Nicolson type to integrate (3.5), while satisfying boundary conditions (3.6a,b). Notably, (3.5) becomes singular at $x=0$ if
$h^{*}=0$ (see e.g. Constantin et al. Reference Constantin, Dupont, Goldstein, Kadanoff, Shelley and Zhou1993); hence, we halt the simulations when the minimum droplet thickness reaches
$h^{*}=0.05$, which we set as the numerical criterion for droplet splitting. Given the characteristic length scales in the problem (i.e.
$\varepsilon L\sim O(0.1\ \textrm {mm})$), this criterion corresponds to the limit in which the minimum droplet thickness is of the order of micrometres and must be governed by van der Waals forces. In addition, we note that the second spatial derivative of the numerical solution to (3.5) has a logarithmic singularity at the contact line, as previously discussed by Ren, Trinh & Weinan (Reference Ren, Trinh and Weinan2015) and similarly by Constantin et al. (Reference Constantin, Dupont, Goldstein, Kadanoff, Shelley and Zhou1993). Despite the singularity in the curvature, Ren et al. (Reference Ren, Trinh and Weinan2015) pointed out that their numerical solutions of
$h$ and
$h_x$ remain bounded near the contact line, while Constantin et al. (Reference Constantin, Dupont, Goldstein, Kadanoff, Shelley and Zhou1993) argued that their computations with the Crank–Nicolson scheme converge to the steady weak solution for
$h(x,t)$. Therefore, we presently focus on the solutions for
$h(x,t)$ only in our current analysis and the physical implications of the resulting droplet shapes.
Figure 8 compares the experimental (a) and computational (b) shapes of a droplet with $L=7.7\ \textrm {mm}$ (corresponding to
$200\ \mathrm {\mu }\textrm {l}$ in the experiments) over time for two different jet speeds. The left-hand column corresponds to the jet speed below the splitting threshold (i.e.
$U_0 =9\ \textrm {m}\ \textrm {s}^{-1} < U_{cr}$), while the right-hand side is for
$U_0 =13\ \textrm {m}\ \textrm {s}^{-1}>U_{cr}$. In order to simulate an equivalent 2-D droplet, we use the experimental values for
$L$,
$U_0$ and the fluid properties in the simulations, so that the values of the dimensionless parameters in (3.5), i.e.
$Ca, We, Bo$ and
$\mu _{a}/\mu _{d}$, match the experiments. In our simulations, we consider three different values for the equilibrium contact angle:
$\theta _0= 10^{\circ }$,
$\theta _0= 20^{\circ }$ and
$\theta _0= 30^{\circ }$. We note that, given the inherent difficulties in comparing simulations of 2-D thin drops to 3-D data, we only focus on qualitative comparisons between theory and experiments.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig8.png?pub-status=live)
Figure 8. Comparison of the droplet shape between the experiments and the simulations. (a) The time-evolution shapes for two drops of $200\ \mathrm {\mu }\textrm {l}$ and
$L=7.7\ \textrm {mm}$, subject to two different flow speeds. The droplet on the left is exposed to
$U_0=9\ \textrm {m}\ \textrm {s}^{-1}$ and reaches an equilibrium shape (
$Re=4600$,
$Ca=0.11$,
$Bo=8$,
$We=10.6$); and
$U_0=13\ \textrm {m}\ \textrm {s}^{-1}$ is applied to the droplet on the right, which yields splitting (
$Re=6620$,
$Ca=0.16$,
$Bo=8$,
$We=22.1$). (b) The same jet speeds are incorporated into the model for the same value of
$L$ and
$\theta _0=10^{\circ }$.
The plots in figure 8 demonstrate that the numerical simulations successfully capture the overall droplet behaviours that are observed in the experiments (i.e. ‘hanging’ versus ‘splitting’) for given $L$ and
$U_0$. The comparison is especially good for
$U_0<U_{cr}$ (left column), as the contact line remains fixed in place in both the experiment and the simulation. In the splitting regime
$U_0>U_{cr}$ (right column), the contact line advances away from its initial position in the experiment, while it is pinned in the simulation. Hence, as a result of the fixed contact line condition, the simulations deviate from the experiments in the splitting regime, for example, in setting the time scale of droplet breakup.
Figure 9 shows the measurement of splitting time $t_{f}$ in comparison to simulations for varying
$L$ and
$U_0$. The computed
$t_{f}$ decreases with increasing
$L$ and
$U_0$, in qualitative agreement with the experiments; however, there are systematic differences that depend on the value of
$\theta _0$. In order to understand these deviations, we emphasise that the pinned contact line and reducing
$\theta _0$ have two counteracting effects on the splitting time scale. Pinning the contact line tends to increase
$t_{f}$, while smaller
$\theta _0$ must decrease
$t_{f}$, as there is less fluid to drain from the centre to reach splitting. We clearly see the manifestation of these competing effects in our numerical results. Simulations that correspond to
$\theta _0=20^{\circ }$ and
$\theta _0=30^{\circ }$ overpredict experimental
$t_{f}$, due to the dominant effects of the fixed contact line. When
$\theta _0$ is further reduced to
$10^{\circ }$ to override the pinned contact line condition, the numerical results are consistently lower than the experimental measurements. We also note that the neglect of the out-of-plane curvature must contribute to the deviation of
$t_{f}$ between theory and experiments.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig9.png?pub-status=live)
Figure 9. The time scale of droplet splitting $t_{f}$ from the experiments and simulations (a) as a function of the initial half-width
$L$ for
$U_0 = 15\ \textrm {m}\ \textrm {s}^{-1}$, and (b) for varying
$U_0$ when
$L=7.7\ \textrm {mm}$.
4.2. Steady-state analytical solutions
While the simulations in § 4.1 can capture the time scale and shapes of evolving droplets, the key distinction between the ‘hanging’ and ‘splitting’ regimes lies in the difference in equilibrium droplet shapes at the steady state. In the splitting regime, the droplet shape may become singular (i.e. $h^{*}\approx 0$) even before the steady state is reached, while the minimum droplet thickness in the hanging regime must be non-zero at the steady state. Based on this physical picture, we focus on the steady-state solutions to (3.5) by setting
$\partial h^{*}/\partial t^{*}=0$. In addition, we neglect shear stress and the coupling between the evolving drop and the outer pressure gradient at
$O(\varepsilon )$, so that (3.5) simplifies to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn16.png?pub-status=live)
where $\widehat {We}=(We/\varepsilon )(L/H_0)^{2}$ is a modified Weber number. The neglect of higher-order terms in the steady state allows us to obtain simple analytical solutions and simultaneously to examine the relative importance of
$O(\varepsilon )$ terms in setting
$U_{cr}$. Physically, the equilibrium condition in (4.1) signifies that the pressure inside the droplet is constant and yields no net internal flow. We solve (4.1) analytically, subject to
$h^{*}(1)=h^{*}(-1)=0$ and
$\int _{-1}^{1} h^{*}\,{\textrm {d}}x^{*} = \mathrm {area}^{*} = \mathrm {area}/\varepsilon L^{2}$, where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn17.png?pub-status=live)
For $\alpha =0$ (centred jet), the equilibrium droplet shape is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn18.png?pub-status=live)
Implementing the criterion required for splitting (i.e. $h^{*}(0)=0$) yields the critical Weber number for the given dimensionless area:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_eqn19.png?pub-status=live)
which relates $U_{cr}$ to the droplet half-width
$L$. Figure 10 shows the plot of
$U_{cr}$ as a function of
$L$ from the experiments, numerical simulations and steady-state analytical predictions for
$\theta _0=10^{\circ }$,
$20^{\circ }$ and
$30^{\circ }$. Both simulations and analytical solutions for
$U_{cr}$ show a decrease with increasing
$L$, which matches the experimental observations. In addition, the model prediction of
$U_{cr}$ grows with increasing
$\theta _0$ and a corresponding increase in the droplet area; the match with experimental data is especially compelling for
$30^{\circ }$. We note that the simulation prediction is lower than the steady-state results, as it incorporates the effects of shear stress and the unsteady droplet shape on pressure gradient inside the drop. In particular, the deviation between theory and simulations is larger for smaller
$L$, indicating that the contributions of
$O(\varepsilon )$ terms must become more significant for smaller droplets.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig10.png?pub-status=live)
Figure 10. Comparison of the critical jet velocity for splitting, $U_{cr}$, as a function of
$L$ between simulation (solid lines) and steady-state prediction (dashed lines) for three different values of
$\theta _0=10^{\circ }$ (blue),
$\theta _0=20^{\circ }$ (dark grey) and
$\theta _0=30^{\circ }$ (maroon). The experimental data are included with a square symbol.
In the case of an off-centred jet, we compute the equilibrium droplet shape $h^{*}(x^{*})$ by solving (4.1) for
$\alpha \neq 0$. We subsequently find the position of the minimum droplet thickness
$x^{*}_{m}$ by setting
$\partial h^{*}/\partial x^{*}=0$. Then, based on the splitting criterion
$h^{*}(x^{*}_{m})=0$, we numerically compute
$U_{cr}$ from the equilibrium droplet shape for given
$L$ and
$\theta _0$. Figure 11 shows the theoretical prediction for
$U_{cr}$ for changing position of the stagnation point
$\alpha$, in comparison with the experimental data. The plots correspond to
$L=7.7$ and
$4.9\ \textrm {mm}$, or equivalently to
$V=200$ and
$50\ \mathrm {\mu } \textrm {l}$ in the experiments, respectively. For all parameters considered, an increase in
$\alpha$ increases the jet speed that is required to split the given droplet, both experimentally and theoretically. Despite the qualitative agreement in
$U_{cr}$ between theory and experiments, note that we are unable to describe the depinning regime with the current theoretical framework, due to the fixed contact line boundary condition.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig11.png?pub-status=live)
Figure 11. Comparison of $U_{cr}$ as a function of the jet position
$\alpha$ between the experiments (squares) and the steady-state analytical predictions with
$\theta _0=10^{\circ }$ (blue),
$\theta _0=20^{\circ }$ (dark grey) and
$\theta _0=30^{\circ }$ (maroon), for (a)
$L=7.7\ \textrm {mm}$ and (b)
$L=4.9\ \textrm {mm}$.
5. Summary and conclusions
In summary, we conducted a series of experiments to examine the behaviour of a partially wetting droplet with radius $L$ subject to a stagnation-point flow. We extracted the temporal evolution of the side-view droplet shapes when the jet of air is applied either at the droplet's centre or at a distance
$\alpha L$ from the centre. Our experimental results indicate that the droplet may exhibit different behaviours depending on its size, the strength of the stagnation-point pressure, and the position of the stagnation point on the droplet's interface. Three distinct droplet behaviours include hanging, depinning and splitting; in particular, the regime in which a droplet depins to one side emerges only when
$\alpha \ne 0$. For all cases, there exists a critical air flow velocity
$U_{cr}$ at which the droplet splits into two satellite drops. We experimentally extracted the value of
$U_{cr}$ for given
$L$ and
$\alpha$ and demonstrated that
$U_{cr}$ tends to decrease with
$L$, while increasing with
$\alpha$.
The experimental observations can be explained by considering the depth-averaged flow inside the droplet and the corresponding spatial changes in the droplet internal pressure. Hence, we use the lubrication approximation to model the thin, 2-D droplet, whose motion is primarily governed by the inviscid pressure field in the stagnation-point flow. Despite the major simplifications of the model including a fixed contact line and the neglect of the out-of-plane curvature, our numerical results qualitatively match the interfacial shapes observed experimentally. Consistent with the experimental measurements, our model shows a decline in the stagnation pressure threshold for splitting as the droplet size increases. Furthermore, increasing our choice of the static contact angle, or a measure of the surface wettability, yields a larger critical velocity for splitting. In addition to the time-dependent simulations, our simplified steady-state solution successfully captures a qualitative increase in the critical splitting velocity as the jet position is placed closer to the contact line.
Extensions of this work include incorporating a dynamic contact line and free streamline theory to capture a possible boundary layer separation over the droplet. In addition, droplet inertia needs to be accounted for to understand the unsteady oscillations of the windswept droplet. Ideally, such improved models would be compared with a fully 3-D computational analysis to gain quantitative understanding of the droplet dynamics under high-speed air flows. Finally, additional experiments and improved visualisation techniques are required to verify the dependence of droplet regimes on varying degrees of wettability and surface roughness.
Acknowledgements
We acknowledge Dr N. Wilkinson, Ms S. Narayan and Dr Y. Chen for their help in our experiments. We also thank Professor E. Nazockdast, Professor H. Stone and Professor E. Kanso for fruitful discussions. This work is partially supported by the National Science Foundation (Grant No. CBET-1605947) and the Simons Foundation.
Declaration of interests
The authors report no conflict of interest.
Appendix. 2-D approximation limitations
The 2-D approach is one of the major simplifying assumptions of our present analysis. In order to test the limits of its validity, we first consider the 3-D volume conservation inside the droplet, which is written as $\partial h/\partial t=\partial Q_x/\partial x +\partial Q_y/\partial y$. The flow rate in the lateral direction,
$Q_y$, can be derived in a similar fashion to what we have explained in § 3.1; the resultant expression is equivalent to (3.5), after the
$x$-derivative has been replaced with
$y$. The important exception is that the external pressure gradient in the
$y$-direction is an order of magnitude smaller than its counterpart in the
$x$-direction, since the applied jet is uniform in the lateral direction.
In addition, the pressure inside the droplet can be approximated as $p_{d}=P-\sigma (\kappa _x+\kappa _y)$ based on the Young–Laplace equation, where
$\kappa _x$ and
$\kappa _y$ denote the droplet's interfacial curvatures in the axial and lateral directions, respectively. Our 2-D approach is deemed reasonable as long as
$\partial \kappa _y/\partial x \approx 0$. Figure 12(a,b) illustrates a droplet shape in the
$z$–
$y$ plane at different times in the splitting and hanging regimes; it can be seen that the droplet deformations are minimal in the
$y$-direction. Hence, the pressure inside the droplet must be constant in the
$y$-direction, which corresponds to a negligible internal flow. However, we acknowledge that the assumption of the constant pressure and the 2-D analysis must break down near the droplet corners.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200825221839362-0944:S0022112020005601:S0022112020005601_fig12.png?pub-status=live)
Figure 12. The $z$–
$y$ plane view of a
$100\ \mathrm {\mu }{\textrm {l}}$ droplet under a jet of air flow applied to the droplet's centreline (
$\alpha =0$). (a,i) For
$U_0=12.5\ \textrm {m}\ \textrm {s}^{-1}$, (ii) the droplet deforms until (iii) it undergoes a breakup (hidden from the
$z$–
$y$ view). (b,i) For
$U_0=7.5\ \textrm {m}\ \textrm {s}^{-1}$, (ii) the droplet deforms until (iii) it reaches the steady state. In both cases, the water–air interface remains relatively flat in the
$y$-direction.