Skip to main content
  • Submit
  • About
    • Editorial Board
    • PNAS Staff
    • FAQ
    • Accessibility Statement
    • Rights and Permissions
    • Site Map
  • Contact
  • Journal Club
  • Subscribe
    • Subscription Rates
    • Subscriptions FAQ
    • Open Access
    • Recommend PNAS to Your Librarian
  • Log in
  • My Cart

Main menu

  • Home
  • Articles
    • Current
    • Special Feature Articles - Most Recent
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • List of Issues
  • Front Matter
  • News
    • For the Press
    • This Week In PNAS
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Editorial and Journal Policies
    • Submission Procedures
    • Fees and Licenses
  • Submit
  • About
    • Editorial Board
    • PNAS Staff
    • FAQ
    • Accessibility Statement
    • Rights and Permissions
    • Site Map
  • Contact
  • Journal Club
  • Subscribe
    • Subscription Rates
    • Subscriptions FAQ
    • Open Access
    • Recommend PNAS to Your Librarian

User menu

  • Log in
  • My Cart

Search

  • Advanced search
Home
Home

Advanced Search

  • Home
  • Articles
    • Current
    • Special Feature Articles - Most Recent
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • List of Issues
  • Front Matter
  • News
    • For the Press
    • This Week In PNAS
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Editorial and Journal Policies
    • Submission Procedures
    • Fees and Licenses

New Research In

Physical Sciences

Featured Portals

  • Physics
  • Chemistry
  • Sustainability Science

Articles by Topic

  • Applied Mathematics
  • Applied Physical Sciences
  • Astronomy
  • Computer Sciences
  • Earth, Atmospheric, and Planetary Sciences
  • Engineering
  • Environmental Sciences
  • Mathematics
  • Statistics

Social Sciences

Featured Portals

  • Anthropology
  • Sustainability Science

Articles by Topic

  • Economic Sciences
  • Environmental Sciences
  • Political Sciences
  • Psychological and Cognitive Sciences
  • Social Sciences

Biological Sciences

Featured Portals

  • Sustainability Science

Articles by Topic

  • Agricultural Sciences
  • Anthropology
  • Applied Biological Sciences
  • Biochemistry
  • Biophysics and Computational Biology
  • Cell Biology
  • Developmental Biology
  • Ecology
  • Environmental Sciences
  • Evolution
  • Genetics
  • Immunology and Inflammation
  • Medical Sciences
  • Microbiology
  • Neuroscience
  • Pharmacology
  • Physiology
  • Plant Biology
  • Population Biology
  • Psychological and Cognitive Sciences
  • Sustainability Science
  • Systems Biology
Research Article

Signatures of slip in dewetting polymer films

View ORCID ProfileDirk Peschka, Sabrina Haefner, Ludovic Marquant, View ORCID ProfileKarin Jacobs, Andreas Münch, and Barbara Wagner
PNAS May 7, 2019 116 (19) 9275-9284; first published April 19, 2019; https://doi.org/10.1073/pnas.1820487116
Dirk Peschka
aWeierstrass Institute for Applied Analysis and Stochastics, 10117 Berlin, Germany;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Dirk Peschka
  • For correspondence: dirk.peschka@wias-berlin.de
Sabrina Haefner
bExperimental Physics and Center for Biophysics, Saarland University, 66041 Saarbrücken, Germany;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Ludovic Marquant
bExperimental Physics and Center for Biophysics, Saarland University, 66041 Saarbrücken, Germany;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Karin Jacobs
bExperimental Physics and Center for Biophysics, Saarland University, 66041 Saarbrücken, Germany;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Karin Jacobs
Andreas Münch
cMathematical Institute, University of Oxford, Oxford OX2 6GG, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Barbara Wagner
aWeierstrass Institute for Applied Analysis and Stochastics, 10117 Berlin, Germany;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  1. Edited by David A. Weitz, Harvard University, Cambridge, MA, and approved March 18, 2019 (received for review December 3, 2018)

  • Article
  • Figures & SI
  • Info & Metrics
  • PDF
Loading

Significance

Dewetting is the hydrodynamic process where a uniform layer of liquid destabilizes and decays into distinct patterns of stationary droplets by virtue of interfacial and intermolecular energies. These patterns can now be predicted theoretically, and their evolution can be followed numerically in striking similarity to experimental results. The droplet arrays observed during the quasistatic evolution are generated by two distinct types of instabilities: droplet shedding of moving liquid rims and satellite droplet pinch-off during the decay of stationary ridges. The mechanism causing the emergence of different droplet patterns is the dissipation at the liquid–solid interface modeled by an effective Navier slip.

Abstract

Thin polymer films on hydrophobic substrates are susceptible to rupture and hole formation. This, in turn, initiates a complex dewetting process, which ultimately leads to characteristic droplet patterns. Experimental and theoretical studies suggest that the type of droplet pattern depends on the specific interfacial condition between the polymer and the substrate. Predicting the morphological evolution over long timescales and on the different length scales involved is a major computational challenge. In this study, a highly adaptive numerical scheme is presented, which allows for following the dewetting process deep into the nonlinear regime of the model equations and captures the complex dynamics, including the shedding of droplets. In addition, our numerical results predict the previously unknown shedding of satellite droplets during the destabilization of liquid ridges that form during the late stages of the dewetting process. While the formation of satellite droplets is well known in the context of elongating fluid filaments and jets, we show here that, for dewetting liquid ridges, this property can be dramatically altered by the interfacial condition between polymer and substrate, namely slip. This work shows how dissipative processes can be used to systematically tune the formation of patterns.

  • pattern formation
  • dewetting
  • Navier slip
  • thin films
  • finite elements

The no-slip condition is widely accepted as an appropriate boundary condition for flows of Newtonian liquids sheared along a solid surface. A notable exception arises in the presence of a moving contact line between a viscous liquid and a rigid solid substrate, where the use of the no-slip condition leads to a nonintegrable singularity in the stress field (1, 2). In the past decades, however, it has been shown that thin films of polymer melts can exhibit significant slip when sheared along a substrate, where slip lengths much larger than the film thickness have been observed (3⇓⇓⇓⇓⇓–9).

For retracting rims as they emerge after a hole or trench has opened, the magnitude of slip has a direct impact on the dewetting dynamics. When the slip length is very small or zero, the retraction rate is independent of the size of the growing rim and hence, approximately constant, except for logarithmic corrections (10). For slip that is large compared with the film thickness, viscous dissipation increases with the rim size and gives rise to a t−1/3 power law in time t for the retraction (dewetting) rate (11), which has been confirmed experimentally (12⇓⇓–15). In both cases, the moving rim is susceptible to spanwise instabilities (16⇓–18), but for the case where slip is large compared with the film thickness, the dependence of the retraction velocity on the local rim size provides a crucial amplifying mechanism for the instability, which is absent in the no-slip situation (19). As a consequence, the repeated shedding of droplets leaves a characteristic pattern, which is not present for systems with no slip. In either case, the dewetting rims eventually meet to form residual ridges, which in either case, are susceptible to a Rayleigh–Plateau-type instability with a similar dominant wavelength (20). Eventually, this leads to the break up into droplets. As a result of this long-time process, strikingly different droplet patterns are obtained as shown by the experiments depicted in Fig. 1.

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

Stationary droplet patterns emerging from a uniform PS(10.3k) film of thickness h∞=110 nm after dewetting at T=120○C from a hydrophobically coated Si wafer: (Left) AF1600 coating and (Right) DTS coating (21).

The evolution of the polymer melts during these dewetting regimes from early stages after hole formation to the late stages of rupture of the ridges, where the hole boundaries meet, is most appropriately modeled by thin-film models. A systematic asymptotic derivation from full Navier–Stokes equations (22, 23) exploiting the separation between the lateral and normal length scales revealed that the resulting dimension-reduced thin-film model depends on the order of magnitude of slip, leading to two asymptotic distinguished limits: a weak-slip regime and a strong-slip regime.

While thin-film models have been shown to be of great advantage for the analysis of free boundary problems, predicting the evolution over long time and large spatial scales deep into the nonlinear regimes is still a major computational challenge. The primary problem is to resolve the length scales associated with nanoscopic residual layers that remain after the film has dewetted, typically about ∼0.1−1 nm up to thickness of about ∼10−1,000 nm of the growing rim, and to account for the slip length in the range of 1−1,000 nm and the length scale of the resulting instability of 103−104 nm. By far, the greatest challenge is to make predictions regarding phenomena on the length scales of the instabilities while using numerical solutions with a fine spatial resolution on the length scale of the residual layer.

In this article, we present a numerical algorithm that is able to answer this need featuring a strategy for local adaptivity and an optimized treatment of the intermolecular potential. We will show that the difference in slip lengths indeed leads to the instability patterns seen in experiments. Our numerical solutions also confirm the Rayleigh–Plateau-type instability of the residual ridges during the late phases of the dewetting for both cases, the no-slip and the intermediate-slip cases, with similar wavelengths at the onset, which had been predicted previously based on a linear stability analysis (17, 20). Similar studies were concerned with cases of infinite ridges with (24) and without (25) gravity followed by a broad range of investigations in the literature for this situation using different contact line models and approximations. Numerically, the work by Diez et al. (26) focuses on finite-length ridges but also, includes a review and elucidating comparison of results for the infinite case. Here, most unstable modes are due to varicose perturbations, and the preferred wavelength of the instability is set by the balance of the destabilizing capillary forces, contact line conditions, and for sufficiently viscous liquids, the viscous dissipation. Stability analysis of the effect of slip vs. no-slip on the instability of a stationary ridge shows that both are linearly unstable and have similar wavelengths (17, 20).

Our numerical method allows us to follow the evolution until rupture and predicts significant morphological differences. In particular, it reveals that, for the no-slip case, the breakup is accompanied by the formation of a cascade of satellite droplets that has never been studied before in this context. Moreover, for the intermediate-slip regime, no satellite droplets appear. Interestingly, a closer look at experimental results confirms these predictions. In other contexts, such as liquid jets or fluid filaments, formation of satellite droplets during rupture is well known, and their destabilization was observed to have a rich structure of intermediate asymptotic regimes [e.g., the works by Tjahjadi et al. (27), Eggers and Villermaux (28), and Castrejón-Pita et al. (29)].

We start by introducing the dissipative free surface flow and the relation to pattern formation. Then, we introduce the nonlinear model for dewetting flow and revisit results for dewetting rates and their impact on instabilities in the early stages of the process. We introduce a highly adaptive numerical method that bridges the multiple length scales and timescales and is able to resolve the dewetting with droplet pinch-off. The solutions are discussed, and we compare with experiments.

Emergence of Dewetting Patterns

Before going into the details of thin-film dewetting, we make some general remarks and simple observations. Assume that the shape of the fluid volume Ω(t)⊂R3 over a substrate at z=0 and time t is described by a function ht,x such thatΩ(t)={(x,z)∈R2×R+:0<z<h(t,x)}.[1]The wetted substrate area is ω(t)={x∈R2:h(t,x)>0}. Then, the surface energy driving the evolution is defined asE(h)=∫ω(γℓs−γgs)+γℓg1+|∇h|2 dx,with γ{ℓs,gs,ℓg} being the surface tensions of the liquid–solid, gas–solid, and liquid–gas interfaces, respectively. For incompressible fluids, the evolution of the film height conserves the total volume V(h)=∫ωh dx. For Newtonian fluids with velocity field u(t):Ω→R3, this energy decreases along solutions: that is,ddtE=−D=−∫Ω2μ|Du|2 dx dz−∫ωμb−1|u|z=02 dx≤0,decomposing the dissipation D into a bulk and substrate part with viscosity μ, Navier slip length b, and symmetric gradient Du=12(∇u+∇u⊤). Thereby, stationary states are energy minimizers, and a straight-forward minimization of E with constant V shows that minimizers are droplets (i.e., up to translational invariance ω is a disc, and h has constant curvature). Apparently, this assumes that contact lines can slide freely and that no pinning is present (2). However, it is also clear that arrays of droplets as shown in Fig. 1 are not global but local minimizers of E and thereby, also represent viable stationary states. While droplet patterns emerging from the spinodal dewetting feature seemingly random distributions of droplets with details depending on intermolecular interactions (30, 31), droplet distributions in other flow scenarios are reminiscent of the process that leads to their creation (e.g., heterogeneous nucleation, surface instabilities on patterned substrates, droplet production in confined environments, or even dynamically for sliding droplets) (32⇓⇓⇓⇓⇓–38). This raises the natural question: Which mechanism decides what pattern is generated?

This question is even more justified when realizing that the surface energies E leading to both patterns in Fig. 1 are qualitatively identical. Therefore, in the following, we show that physical systems with the same wetting energy but different magnitudes of dissipation (i.e., viscosity and Navier slip) will produce qualitatively different droplet patterns. While the exploration of the nonlinear pattern formation will mainly compare experiments and simulations, the general mechanism responsible for different patterns is the switching of different instabilities due to the dissipation (Movie S1).

Thin-Film Models and Instability

Problem Formulation.

After a film has ruptured by nucleation or by external forcing, forming a hole or in a planar-symmetric setting, a trench, the viscous fluid retracts to reduce the overall energy of the liquid–gas, solid–liquid, and solid–gas interfaces. The dewetting process is driven by the intermolecular potential ϕ between the film and the substrate. In the simplest case, it consists of a sum of attractive long-range van der Waals forces and short-range Born repulsion forces, the minimum of which yields the height h⋆ that remains in the dry region from which the film of uniform height h∞≫h⋆ has dewetted. Motivated by Lennard–Jones potentials, one often finds intermolecular potentials (31, 39) of the standard form ϕ(h)=ϕ¯h/h⋆, whereϕ¯(h)=Sn−mnh−m−mh−n,[2]with m=2,n=8, ϕ′(h⋆)=0, and ϕ(h⋆)=S. For partial wetting, the spreading coefficient S=γgs−γℓs−γℓg is negative.

Due to the slow dewetting rates of the polymer films with chain lengths below the entanglement length, the Navier–Stokes equations serve as the underlying model for the viscous fluid with a Navier slip boundary condition,t⋅(2μDu)n+μb−1t⋅u=0,[3]for tangential velocity implied by the previous energy dissipation balance. The scale separation of the characteristic height scale H=[z] and typical length scales L=[x] allows for a consistent thin-film approximation using the small parameterε=HL,[4]which leads to the reduction of the Stokes free boundary problem to a problem for the free boundary h(t,x) in closed form. In the following, all quantities are nondimensionalized with this scaling as explained along with details of derivation in ref. 22, which we here only summarize. In particular, we introduce the nondimensional slip length B=b/H.

We seek the scalar function h:[0,T]×R2→[0,∞) as in Eq. 1, where the film height h(t,x) depends on time t∈[0,T] and on space x=(x,y)∈R2 with given initial data h0(x)=h(t = 0,x) (Fig. 2). It has been shown that, depending on the magnitude of slip length B, there exist two asymptotic distinguished limits (22) for ε≪1. The first distinguished limit assumes B=O(1) and leads to the weak-slip thin-film model given by the fourth-order parabolic partial differential equation (PDE)∂th−∇⋅(m(h)∇π)=0,[5a]where the mobility ism(h)=13h3+Bh2,[5b]and the generalized pressure π is defined as the functional derivative of the energy functionalE(h)=∫R212|∇h|2+ϕ(h) dx[5c]with respect to h: that is,π=δEδh=−∇2h+Π(h),[5d]and Π(h)=∂hϕ(h) is the derivative of the nondimensional intermolecular potential. Effectively, by introducing Π, one enforces positivity of h:[0,T]×ω¯→[0,∞) and replaces the complement of the set ω(t)⊂ω¯⊂R2 by the region where h≈h⋆ as it is sketched in Fig. 2.

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Sketch of solution h(t,x) in the x−z plane.

The second distinguished limit assumes that B=ε−2β and leads to the strong-slip thin-film model given by the system of PDEs for the film height h:[0,T]×ω¯→[0,∞) and the lateral velocity u¯:[0,T]×ω¯→R2:∂th+∇⋅hu¯ = 0,[6a]Re∂tu¯+(u¯⋅∇)u¯ = 1h∇⋅σ−∇π−β−1u¯h,[6b]where Re is the rescaled Reynolds number and the effective shear stress isσ=h∇u¯+(∇u¯)⊤+2(∇⋅u¯)I.[6c]Two important limiting cases are the no-slip and the intermediate-slip thin-film models. The no-slip model is obtained for B→0 so that the degenerate mobility becomes m(h)=13h3. The latter intermediate-slip model is obtained in the limit B→∞ but ε−2B→0, which on rescaling of time, produces the mobility m(h)=h2. In this sense, no-slip m=13h3 mobility and intermediate-slip mobility m=h2 are two extremal limits in the slip length B.

It has been shown experimentally and near the linear regime it was also demonstrated theoretically that details of the dynamic and morphological evolution strongly depend on the magnitude of slip at the solid–liquid interface (40). For no slip or weak slip, the liquid accumulates in a growing rim in front of the contact line and destabilizes. The case of strong slip is characterized by very asymmetric retracting rims with a monotone spatial decay toward the unperturbed film for particularly large slip, which undergoes a transition to oscillatory decay as the slip length decreases below a critical value (22, 41). In fact, this observation has been used to identify the strong-slip regime in experiments and also to determine the slip length quantitatively (42⇓–44). The molecular origin of slip in polymer melts was investigated in ref. 45. More details about this topic can be found in the work by Bäumchen and Jacobs (46).

Dewetting Rates.

We first discuss the dewetting dynamics of a straight rim, where we assume translational invariance ∂yh≡0. In the case of no slip, m(h)=13h3, the evolution is determined by the region near the contact line where the rim meets the residual film of thickness h=h⋆ that remains behind the advancing dewetting front. A careful asymptotic analysis and comparison with long-time numerical solutions (22, 47) reveal that the dewetting rate ṡ(t) is nearly constant. In fact, the position of the contact line s(t) is to leading order given by the dewetting laws(t)∼t tan3(𝜗)ln3(h∞/h⋆)t[7]as t→∞, where h∞=limx→∞h. Physically, this reflects the fact that the size of the rim only has a weak effect on the total friction and hence, in turn, on the dewetting velocity.

For intermediate slip, where m(h)=h2, the evolution of the unperturbed rim is different. At any given size, the rim behaves like a traveling wave but with a wave speed that depends on the width of the rim, which by mass conservation, is proportional to the square root of the distance traveled. A detailed asymptotic analysis (22, 47) that matches the rim to the unperturbed and residual film gives the leading order results(t)∼9Mb⁡tan5⁡𝜗4h∞1/3t2/3[8]in the limit t→∞ with M≈0.0272. A similar prediction had been made by Reiter and Khanna (12).

Rims with capillary humps, like the ones that appear here, are known to be subject to Rayleigh–Plateau-like instabilities (20, 24, 26), where the higher capillary pressure in thinner parts squeezes even more liquid into the thicker parts, hence promoting the growth of undulations along the rim. The linear stability of dewetting rims is complicated by the fact that the base state itself grows in time, giving rise to a linearized PDE that cannot be solved exactly using separation of variables. Instead, the linearized PDE can be solved numerically, and the amplification of an initial perturbation can be tracked over time (48). Interestingly, the perturbation evolves into a universal long-time shape that is not sensitive to the initial perturbation. Comparing these shapes reveals an important difference between the no-slip and intermediate-slip cases. The former is much more symmetric and closer to the classical varicose mode observed in the Rayleigh–Plateau instability than the latter (48). Moreover, the maximum amplification is significantly higher for the intermediate-slip case.

These results were analyzed further using an asymptotic sharp interface approach for large rims and a Wentzel–Kramers–Brillouin analysis (also known as WKB analysis, see ref. 49), which established that the long-time dominant wavenumber is given by an equal area rule and is shorter than the prediction by a frozen-mode analysis (50, 51). Furthermore, theory indicates that perturbations of the rim remain small in the no-slip case and for moderate values of B, while in the intermediate-slip case 1≪B≪ε−2 perturbations grow and develop fingers that eventually pinch off; then, the process repeats itself. The physical explanation for these different manifestations relates to the size dependence of the friction. In the intermediate-slip case, thicker parts of a perturbed rim have a smaller velocity than thinner ones and tend to lag farther behind. This supplies an additional nonlinear enhancement that reinforces the linear Rayleigh–Plateau instability but is essentially absent in the no-slip case, where the dewetting rate is largely independent of the rim size (18, 48, 51). The size dependence also adds another instability mechanism by causing thicker parts of the rim to lag behind thinner parts, thus reinforcing the Rayleigh–Plateau instability and making it more asymmetric (18). A linear stability analysis of the thin-film model predicts the instability to be much more pronounced in the intermediate-slip case than in the no-slip case (48⇓⇓–51).

For slip lengths that are much larger than the film height, the dynamics of the evolution change yet again. In the strong-slip regime, the fluid flow is a plug flow, and the evolution is described by a system of PDEs for the film height and the lateral velocity (6), where the contribution from elongational stresses enters to the same order as the effects from the friction due to slip. Dewetting rim solutions of this model were explored numerically and asymptotically (22, 41). In this regime, the shape of the profile becomes highly asymmetric, with a steep side facing the dewetted area and a much flatter decay to the unperturbed film h∞. This is well reflected in the experimental data for melts with larger polymer chains, where slip is expected to be larger. Interestingly, the change in the balance of stresses leads to an approximately linear dewetting laws(t)∼b1/2⁡tan2(𝜗)t42 h∞1/2⁡ln1/2⁡t,[9]which is another case with an approximately constant velocity of the retracting rim just as in the no-slip case. This suggests that, for very large slip lengths, the rim should become stable again.

Experimental Setup and Methods

For the experiments, atactic polystyrene (PS; purchased from PSS; molecular weights are as listed in the experiments) is used as a model viscous liquid. The films were produced by spin casting a toluene solution (Selectipur or LiChrosolv; Merck) of PS on freshly cleaved mica sheets. The glassy thin films were then floated onto an ultrapure water (organic impurities of <6 ppb, resistance at 25○C: <18.2 M Ω cm) surface and were then picked up with hydrophobic Si wafers.

Hydrophobic Si wafers were achieved by two different preparation methods: (i) on the cleaned Si surface, a self-assembled monolayer of silane molecules [dodecyltrichlorosilane (DTS); Sigma Aldrich/Merck] was prepared (52), or (ii) the cleaned Si wafer was dipped into a solution of a fluoropolymer layer (AF1600; Sigma Aldrich/Merck).

Dewetting is initiated by heating the glassy polymer film above its glass transition temperature. The dewetting of the retracting straight fronts was monitored in situ by optical microscopy on a heating plate (Linkam) or by atomic force microscopy (AFM; Dimension ICON; Bruker).

The dewetted distance was typically obtained from optical micrographs. In AFM experiments, the dewetted distance can also be calculated from 3D scans of the rim on the basis of volume preservation. The values resulting from both approaches were checked for consistency.

Slip lengths have been calculated using the rim profile analysis method (53, 54); structural details, surface roughness values, and wetting properties of the coatings are given in the supplementary material of ref. 40. Polymer film thicknesses have been determined by ellipsometry on the glassy film or by AFM on the edge of a film. The equilibrium contact angles are 𝜗=(88±2)○ and 𝜗=(66±2)○ for AF1600- and DTS-coated substrates, respectively. Comparing typical values for rim height and width, we find typical values in the range ε=0.15−0.3 for the dewetting systems.

In Fig. 3, we show retracting rims with four different chain lengths of the PS film, each thereby corresponding to drastically different slip length b. For the smallest chain length PS(65k), the slip length b is smaller than the rim height H and hence, should correspond to a no-slip situation. As chain length and correspondingly, slip length b increase, we observe a more pronounced instability, as we would expect for a transition from a no-slip scenario to an intermediate-slip scenario. However, if b is increased even further by two orders of magnitudes for PS(390k), the instability is suppressed again. Interestingly, this was predicted for the transition to a strong-slip regime.

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Comparisons of the dewetting behavior of h∞=110-nm-thick PS films for different chain lengths dewetting from AF1600-coated substrate after the rim traveled the distance given above the images. Films are annealed to different temperatures to speed up dewetting in case of films with larger viscosities: PS(65k) at T=140○C, PS(101k) at T=140○C, PS(186k) at T=150○C, and PS(390k) at T=150○C. In the chosen systems, the slip length increases with increasing molecular weight, leading to the values indicated next to the corresponding images obtained by fitting the rim profile as reported by Bäumchen et al. (45) or McGraw et al. (55).

For the remaining part of the paper, we will only consider experiments that can be captured with a no-slip (AF1600 coating) or intermediate-slip (DTS coating) regime with PS(13.7k) or PS(10.3k) and use corresponding theoretical models with cubic m=13 h3 or quadratic m=h2 mobility.

Numerical Methods

We now explain the intricacies of numerically resolving the multiple length scales of the problem. We focus on the two cases of no slip and intermediate slip, as most of the experimental results regarding the various instabilities fall into either of these two regimes.

Our solution of the thin-film model in Eq. 5 is based on a P2 finite element method (FEM), where the fourth-order equation is split into a system of two second-order equations. The FEM uses piecewise quadratic elements and local mesh refinement. We use a semiimplicit time discretization, where only the highest-order derivative is treated implicitly. Therefore, the thin-film equation (Eq. 5) is multiplied with a test function v and integrated by parts to obtain∫ω¯∂thv+m(h)∇π⋅∇v dx=0,[10a]where boundary terms vanish due to the no-flux boundary condition n⋅∇π=0 imposed on ∂ω¯. We also rewrite the pressure π in the weak form as∫ω¯πv dx=∫ω¯∇h⋅∇v+Π(h)v dx,[10b]where again, we used integration by parts and n⋅∇h=0 on the boundary ∂ω¯. In particular, this statement of the PDE implies conservation of volumeddt∫ω¯h(t,x) dx=0,which can be seen when selecting v=1 in the continuous or discrete weak formulation (Eq. 10a). Evaluating the solution at discrete times hn(x)=h(nτ,x), we use the time discretization ∂th=τ−1(hn−hn−1). This allows us to rewrite the weak form of the thin-film model in Eq. 10 so that we seek (hn,π)∈W such that∫ω¯hnv+τ m* ∇π⋅∇v dx=∫ω¯hn−1v dx,[11a]∫ω¯πw−∇hn⋅∇w dx=∫ω¯Π*w dx,[11b]which needs to hold for all (v,w) from a suitable function space W and with initial data h0. When we define the mobility m*=m(hn−1) and Π*=Π(hn−1), this becomes a semiimplicit time discretization.

The FEM constitutes a method for the construction of a finite-dimensional subspace Wh⊂W, where we have an admissible decomposition ω¯=⋃k=1NTk of the domain into triangles Tk, on which we define Wh as the space of continuous functions that are piecewise quadratic on each triangle (i.e., P2 finite elements). Then, we seek a discrete solution (hn,π)∈Wh of Eq. 11 valid for all (v,w)∈Wh. The integrals appearing in Eq. 11 are solved exactly or using a seven-point Gauss quadrature. A typical droplet pattern emerging from the simulation with m=h2 is shown in Figs. 4 and 5.

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Simulated pattern formation with intermediate-slip m=h2: 3D view with light areas showing dry regions, whereas elevated dark areas show the liquid–air interface.

Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

(Left) Dewetting rim during pinch-off of a single droplet highlighted with dashed lines and (Right) magnification of this droplet and corresponding locally refined triangulation. The mesh consists of 92, 272 vertices, which require us to solve for 2×368, 423=736, 846 unknowns (hn,π) for the P2 FEM discretization at each time step.

The precursor thickness h⋆ in the potential of Eq. 2 needs to be chosen much smaller than any droplet size we intend to resolve. Then, the solution h will feature large regions, where the solution is either almost constant h≈h⋆ or the solution is smooth and h≫h⋆ as shown in Fig. 2. However, where those regions meet, the solution features a kink, which should be resolved in the triangulation (compare with Fig. 5). We perform a heuristic local mesh refinement in these connecting regions where the contact line is situated. When constructing a new mesh, we start from a coarse base mesh and determine which of its elements are crossed by the contact line of the previous solution. Those elements are refined by inserting additional vertices to an extent that the kink is resolved again. Furthermore, we make sure that neighboring elements are also refined, so that the contact line remains in the refined region of the mesh for a number of time steps. Based on the set of newly created vertices, we perform a new triangulation and interpolate the old solution onto the new triangulation.

To allow the control of the minimum and the derivatives of ϕ separately via h⋆ and ε⋆, it is advantageous to work with an alternative potential representation of the form ϕ(h)=ϕ^(h−h⋆)/ε⋆ withϕ^(s)=12γ1+s−(1+γ)exp(−s2),[12]which has similar properties as the rescaled version of Eq. 2. In particular, with this potential, we still have the same Γ convergence property as h⋆,ϵ⋆→0. This abstract statement ensures that equilibrium contact angles are maintained and that the energetic contributions of ϕ from Eq. 12 and Eq. 2 coincide in this limit. We note that, for 0<γ≪1, the minimum slightly shifts away from h⋆; however, one gains a slightly stabilizing potential with ϕ″ > 0 for h≫h⋆. In addition, we point out that we monitor that the minimum of the solution minxh(t,x) never violates the nonnegativity requirement for the given choice of ϕ(h) and parameters h⋆,ε⋆ (Fig. 6), in particular, by keeping the time step size τ sufficiently small. We also refer to previous mathematical studies (56, 57) that incorporate the nonnegativity property into their numerical scheme.

Fig. 6.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 6.

Lennard–Jones potential Eq. 2 compared with the much shorter-ranged exponential potential Eq. 12, both with h⋆=1/20 and ε⋆=1/16. Note that, in the numerics we rescaled so that at the contact line the slope is |∇h|=1, leading to a nondimensional potential with ϕ(h⋆)=−1/2.

Initial Data for Rims

We still need to specify the initial data h0(x) for the numerical simulation to describe the various stages of the dewetting process. In experiments, the dewetting is initiated from a uniform flat layer bounded by a nearly straight edge. When the sample is heated, the layer liquefies, and the edge becomes a moving contact line of the dewetting process. In simulations, we choose the supporting domain sufficiently large ω¯=[0,500]×[0,600]; using h⋆=1/20, we represent the uniform layer with the nearly straight contact line at x0 with initial heighth0(x)=h⋆+12(h∞−h⋆)1+tanhx−x0(y)Δ,[13]and the smooth initial contact line position is represented by x0(y)=20+∑n=150an⁡cos(nπy/600). With Δ=1/2, one can interpret h0 to be an approximation of a step-like profile, where slight corrugations of x0 along the y direction are introduced using the normally distributed coefficients an with zero mean and SD 1/10. This choice is then maintained throughout all dewetting simulations for both mobilities. Due to the introduced scaling, we have h∞=1.

To study the systematic effect of the slip boundary condition on dewetting patterns, we study the contact line instability of moving rims and stationary ridges for the mobility functions m(h)=13h3 (no slip) and m(h)=h2 (intermediate slip). In Fig. 7, the evolution of the rims for no slip and intermediate slip is shown at times where they have accumulated a similar volume, starting with the same initial film thickness h∞ as given by the experiments. This corresponds to the experimental results in Fig. 8, where the evolution of the rims is shown at the same distance measured from the initial coordinates. The distinctive dewetting patterns for the intermediate-slip case are observed, while for the no-slip case, only shallow oscillations will occur in excellent comparison with the experimental results in Fig. 8.

Fig. 7.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 7.

Numerical solutions on ω¯=[0,500]×[0,600] (h⋆=1/20, γ=1/10) for (Upper) m=h2 (intermediate-slip B→∞) and (Lower) m=13h3 (no-slip B=0) for initially slightly perturbed straight front and time progressing from left to right shown for similar rim progression (Movie S1).

Fig. 8.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 8.

Series of micrographs from experiments with thin films (Upper) dewetting from a hydrophobized Si wafer covered with a silane monolayer DTS with b=1 μm (intermediate slip) and (Lower) dewetting from an AF1600-covered Si wafer with b=40 nm (no slip). In both series, a h∞=100 nm thin PS(13.7k) film dewets at T=120○C. In both series, undulations are formed along the rim; however, only in the intermediate-slip case above, budding is observed that later leads to fingers and a pinch-off of droplets. Comparing rims that have traveled a similar distance ensures that only rims of similar volumes are evaluated. Reprinted with permission from ref. 19. Copyright 2015 by the American Physical Society.

Breakup of Liquid Ridges

Toward the final stages of dewetting process, for long times and domain size far beyond ω¯, the rims approach each other and merge, yielding a polygonal network of ridges. As is known from previous studies, stationary ridges are susceptible to a Rayleigh–Plateau instability for both no-slip and intermediate-slip boundary conditions at the substrate. Both the no-slip case and the intermediate-slip case have been extensively investigated in the literature (20, 24, 26).

To study the instability of a capillary ridge, we first compute a 1D stationary solution hstat(x) (i.e., a time-independent 2D solution of Eq. 5 with ∂th≡0, ∂yh≡0, and constant thickness h→h⋆ as x→±∞ as shown in Fig. 9, with a symmetric and approximately parabolic shape around the maximum). Then, we study the linear stability with respect to perturbationsh(t,x)=hstat(x)+δh1(x;k)exp(iky+σt)[14]with fixed wavenumber k, which returns an eigenproblem for the perturbation h1 of the base state (20, 51). We select the most unstable mode kmax and define as initial datah0(x)=hstat(x)+δh1(x;k)cos(ky)[15]with a sufficiently small δ so that the nonnegativity of h0 is also not violated. In SI Appendix , SI Text and Figs. S1–S3, we present the details of the linear stability analysis and show that solutions of the nonlinear problem follow the predicted exponential amplification while the linear regime is valid. Furthermore, we checked that the alternative choice of potential Eq. 12 compared with the standard choice Eq. 2 has no visible impact on the linear stability as long as the minimal value ϕ(h⋆) and the minimal height h⋆ are the same.

Fig. 9.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 9.

The 1D steady-state hstat with ε⋆=1/20 and h⋆=1/80.

However, we are interested in using the numerical simulation as a tool to observe features in the nonlinear regime that are distinctive features of the mobility law. Therefore, to resolve finer features of the ridge shape (e.g., secondary or tertiary droplets), the general form of the intermolecular potential is the same, but we use a smaller h⋆=1/80.

We now consider nonlinear simulations using the stationary ridge with very small monochromatic perturbations as initial data. Specifically, we consider an unstable perturbation with kmax and carry out simulations with the mobilities h2, 13h3.

As long as the perturbations are small enough, the evolution of the corrugations in the nonlinear model closely follows the predictions from the linear stability analysis. After the perturbations become comparable with the size of the rim, nonlinear terms become relevant, and the growth rate changes.

Regarding the morphological evolution of the ridges for the no-slip and intermediate-slip cases, our numerical simulations show qualitatively and quantitatively very similar behavior during the linear regime. However, deep into the nonlinear regime, the breakup into droplets follows different scenarios. In the no-slip case, a cascade of satellite droplets emerges, while for the intermediate-slip case, they disappear. Starting with the same initial condition, the evolution for both cases is shown in Fig. 10. Interestingly, this different behavior is also observed in our experimental results as seen in Fig. 11. The instability mechanisms causing these different patterns are encoded in the behavior near the self-similar pinch-off (e.g., refs. 58 and 59 are related theoretical studies of thread breakup).

Fig. 10.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 10.

Numerical solutions showing ridges with (Upper) no slip m(h)=13h3 and (Lower) intermediate slip m(h)=h2. The initial data are h0(x)=hstat(x)+δh1(kmax;x)cos(ky), with time increasing from left to right. The wavenumber kmax is the one with the largest amplification from Eq. 14, and the domain ω¯=[−12,0]×[0,π/kmax] is then extended to [−12,12]×[−π/kmax,3π/kmax] using the implied symmetry of the solution (Movie S1).

Fig. 11.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 11.

Experimental results. Close-up view of the late-stage ridge of PS(10.3k) on AF1600 with time progressing from left to right showing decay into satellite and subsatellite droplets. The initial film height is h∞=115 nm, and b=40 nm. With B=b/H and H corresponding to rim height, this gives rise to the no-slip regime (21).

Conclusion and Discussion

We have introduced a highly adaptive finite element-based numerical approach that correctly captures the complex dewetting process described by a class of thin-film models with degenerate mobilities. We showed that, for the no-slip condition, the droplet pinch-off is absent during the retraction of the rim, while for the intermediate-slip case, self-repeating droplet pinch-off occurs in excellent agreement with experimental observations. The ability to resolve the different length scales for long timescales also enables the prediction of phenomena, such as the formation of satellite droplets, as a function of the mobility. The emergence of satellite droplets is well known during the break up of liquid jets and the related problem of liquid filaments. For the latter problem, destabilization is due to the difference in the axial contribution to the capillary pressure between thicker and thinner parts. In this system, the pressure is higher in the thinner parts and squeezes the liquid into the bulges, thus increasing the perturbation until the filament breaks up. Apart from the huge literature on experimental studies, the problem has sparked numerous numerical and analytical investigations (28, 60, 61). In particular, the work by Tjahjadi et al. (27), where the emergence of satellite droplets was captured numerically, and the highly the accurate numerical schemes developed by Kim et al. (62) improved the understanding the underlying physical processes considerably.

For the situation of a ridge that destabilizes on a solid substrate, the additional influence of the substrate enters the linear stability analysis, in particular through the contact angle dynamics. It will be interesting to investigate analytically the rupture behavior for this problem to help understand the influence of the boundary condition at the substrate. Similarly, it remains an interesting open question how the dissipation due to a dynamic contact angle would affect the whole pattern formation process. Such a study would certainly require even more sophisticated finite element techniques (63, 64).

The decision about the pathway leading to specific droplet patterns is then often decided by the specific influence that the dissipation has on flow instabilities. In this study, these instabilities are the shedding of droplets from moving rims and the symmetry of the Rayleigh–Plateau instability leading to satellite droplets, and they are influenced by the magnitude of the interface dissipation encoded in the Navier slip condition.

We conclude that, by controlling dissipative effects in this dewetting films system, we can steer pattern formation without changing the driving forces.

Acknowledgments

We thank Kirstin Kochems for technical assistance. D.P. acknowledges the financial support of the Einstein Foundation Berlin in the framework of the research center Matheon.

Footnotes

  • ↵1To whom correspondence should be addressed. Email: dirk.peschka{at}wias-berlin.de.
  • Author contributions: D.P., K.J., A.M., and B.W. designed research; D.P., S.H., and L.M. performed research; and D.P., K.J., A.M., and B.W. wrote the paper.

  • The authors declare no conflict of interest.

  • This article is a PNAS Direct Submission.

  • This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1820487116/-/DCSupplemental.

  • Copyright © 2019 the Author(s). Published by PNAS.

This open access article is distributed under Creative Commons Attribution License 4.0 (CC BY).

View Abstract

References

  1. ↵
    1. Dussan EB
    (1979) On the spreading of liquids on solid surfaces: Static and dynamic contact lines. Annu Rev Fluid Mech 11:371–400.
    OpenUrlCrossRef
  2. ↵
    1. de Gennes PG
    (1985) Wetting: Statics and dynamics. Rev Mod Phys 57:827–863.
    OpenUrlCrossRef
  3. ↵
    1. Brochard-Wyart F,
    2. de Gennes PG
    (1992) Shear-dependent slippage at a polymer/solid interface. Langmuir 8:3033–3037.
    OpenUrlCrossRef
  4. ↵
    1. Brochard-Wyart F,
    2. de Gennes PG
    (1993) Sliding molecules at a polymer/polymer interface. C R Acad Sci Ser II 327:13–17.
    OpenUrl
  5. ↵
    1. Brochard-Wyart F,
    2. de Gennes PG,
    3. Hervert H,
    4. Redon C
    (1994) Wetting and slippage of polymer melts on semi-ideal surfaces. Langmuir 10:1566–1572.
    OpenUrlCrossRef
  6. ↵
    1. Migler KB,
    2. Hervet H,
    3. Léger L
    (1993) Slip transition of a polymer melt under shear stress. Phys Rev Lett 70:287–290.
    OpenUrlCrossRefPubMed
  7. ↵
    1. Léger L
    (2003) Friction mechanisms and interfacial slip at fluid-solid interfaces. J Phys Condens Matter 15:S19–S29.
    OpenUrlCrossRef
  8. ↵
    1. Tropea C,
    2. Yarin AL,
    3. Foss JF
    1. Lauga E,
    2. Brenner MP,
    3. Stone HA
    (2007) Microfluidics: The no-slip boundary condition. Handbook of Experimental Fluid Dynamics, eds Tropea C, Yarin AL, Foss JF (Springer, Berlin), pp 1219–1240.
  9. ↵
    1. Bocquet L,
    2. Barrat JL
    (2007) Flow boundary conditions from nano- to micro-scales. Soft Matter 3:685–693.
    OpenUrlCrossRef
  10. ↵
    1. Redon C,
    2. Brochard-Wyart F,
    3. Rondelez F
    (1991) Dynamics of dewetting. Phys Rev Lett 66:715–718.
    OpenUrlCrossRefPubMed
  11. ↵
    1. Redon C,
    2. Brzoska JB,
    3. Brochard-Wyart F
    (1994) Dewetting and slippage of microscopic polymer films. Macromolecules 27:468–471.
    OpenUrlCrossRef
  12. ↵
    1. Reiter G,
    2. Khanna R
    (2000) Kinetics of autophobic dewetting of polymer films. Langmuir 16:6351–6357.
    OpenUrlCrossRef
  13. ↵
    1. Jacobs K,
    2. Seemann R,
    3. Schatz G,
    4. Herminghaus S
    (1998) Growth of holes in liquid films with partial slippage. Langmuir 14:4961–4963.
    OpenUrlCrossRef
  14. ↵
    1. Damman P,
    2. Baudelet N,
    3. Reiter G
    (2003) Dewetting near the glass transition: Transition from a capillary force dominated to a dissipation dominated regime. Phys Rev Lett 91:216101.
    OpenUrlPubMed
  15. ↵
    1. Fetzer R,
    2. Jacobs K
    (2007) Slippage of Newtonian liquids: Influence on the dynamics of dewetting thin films. Langmuir 23:11617–11622.
    OpenUrlPubMed
  16. ↵
    1. Masson JL,
    2. Olufokunbi O,
    3. Green PF
    (2002) Flow instabilities in entangled polymer films. Macromolecules 35:6992–6996.
    OpenUrl
  17. ↵
    1. Brochard-Wyart F,
    2. Redon C
    (1992) Dynamics of liquid rim instabilities. Langmuir 8:2324–2329.
    OpenUrl
  18. ↵
    1. Reiter G,
    2. Sharma A
    (2001) Auto-optimization of dewetting rates by rim instabilities in slipping polymer films. Phys Rev Lett 80:166103.
    OpenUrl
  19. ↵
    1. Haefner S, et al.
    (2015) Influence of slip on the Plateau–Rayleigh instability on a fibre. Nat Commun 6:7409.
    OpenUrlCrossRefPubMed
  20. ↵
    1. King JR,
    2. Münch A,
    3. Wagner B
    (2006) Linear stability of a ridge. Nonlinearity 19:2813–2831.
    OpenUrlCrossRef
  21. ↵
    1. Marquant L
    (2013) Rim instability in dewetting of thin polymer films. PhD thesis (Saarland University, Saarbrücken, Germany).
  22. ↵
    1. Münch A,
    2. Wagner B,
    3. Witelski TP
    (2006) Lubrication models with small to large slip lengths. J Eng Math 53:359–383.
    OpenUrl
  23. ↵
    1. Kargupta K,
    2. Sharma A,
    3. Khanna R
    (2004) Instability, dynamics and morphology of thin slipping films. Langmuir 20:244–253.
    OpenUrlCrossRefPubMed
  24. ↵
    1. Davis SH
    (1980) Moving contact lines and rivulet instabilities. Part 1. The static rivulet. J Fluid Mech 98:225–242.
    OpenUrlCrossRef
  25. ↵
    1. Sekimoto K,
    2. Oguma R,
    3. Kawasaki K
    (1987) Morphological stability analysis of partial wetting. Ann Phys 176:359–392.
    OpenUrl
  26. ↵
    1. Diez JA,
    2. González AG,
    3. Kondic L
    (2009) On the breakup of fluid rivulets. Phys Fluids 21:082105.
    OpenUrl
  27. ↵
    1. Tjahjadi M,
    2. Stone HA,
    3. Ottino JM
    (1992) Satellite and subsatellite formation in capillary breakup. J Fluid Mech 243:297–317.
    OpenUrl
  28. ↵
    1. Eggers J,
    2. Villermaux E
    (2008) Physics of liquid jets. Rep Prog Phys 71:036601.
    OpenUrlCrossRef
  29. ↵
    1. Castrejón-Pita JR, et al.
    (2015) Plethora of transitions during breakup of liquid filaments. Proc Natl Acad Sci USA 112:4582–4587.
    OpenUrlAbstract/FREE Full Text
  30. ↵
    1. Xie R,
    2. Karim A,
    3. Douglas JF,
    4. Han CC,
    5. Weiss RA
    (1998) Spinodal dewetting of thin polymer films. Phys Rev Lett 81:1251–1254.
    OpenUrlCrossRef
  31. ↵
    1. Seemann R,
    2. Herminghaus S,
    3. Jacobs K
    (2001) Gaining control of pattern formation of dewetting liquid films. J Phys Condens Matter 13:4925–38.
    OpenUrl
  32. ↵
    1. Reiter G
    (1992) Dewetting of thin polymer films. Phys Rev Lett 68:75–78.
    OpenUrlCrossRefPubMed
  33. ↵
    1. Konnur R,
    2. Kargupta K,
    3. Sharma A
    (2000) Instability and morphology of thin liquid films on chemically heterogeneous substrates. Phys Rev Lett 84:931–934.
    OpenUrlPubMed
  34. ↵
    1. Seemann R,
    2. Brinkmann M,
    3. Kramer EJ,
    4. Lange FF,
    5. Lipowsky R
    (2005) Wetting morphologies at microstructured surfaces. Proc Natl Acad Sci USA 102:1848–1852.
    OpenUrlAbstract/FREE Full Text
  35. ↵
    1. Dangla R,
    2. Kayi SC,
    3. Baroud CN
    (2013) Droplet microfluidics driven by gradients of confinement. Proc Natl Acad Sci USA 110:853–858.
    OpenUrlAbstract/FREE Full Text
  36. ↵
    1. Eggersdorfer ML,
    2. Seybold H,
    3. Ofner A,
    4. Weitz DA,
    5. Studart AR
    (2018) Wetting controls of droplet formation in step emulsification. Proc Natl Acad Sci USA 115:9479–9484.
    OpenUrlAbstract/FREE Full Text
  37. ↵
    1. Cuellar I, et al.
    (2018) Self-assembly of a drop pattern from a two-dimensional grid of nanometric metallic filaments. Phys Rev E 98:043101.
    OpenUrl
  38. ↵
    1. Wilczek M,
    2. Tewes W,
    3. Engelnkemper S,
    4. Svetlana V
    (2017) Gurevich, and Uwe Thiele. Sliding drops: Ensemble statistics from single drop bifurcations. Phys Rev Lett 119:204501.
    OpenUrl
  39. ↵
    1. Seemann R,
    2. Herminghaus S,
    3. Jacobs K
    (2001) Dewetting patterns and molecular forces: A reconciliation. Phys Rev Lett 86:5534.
    OpenUrlCrossRefPubMed
  40. ↵
    1. Bäumchen O, et al.
    (2014) Influence of slip on the Rayleigh-Plateau rim instability in dewetting viscous films. Phys Rev Lett 113:014501.
    OpenUrlCrossRefPubMed
  41. ↵
    1. Evans PL,
    2. King JR,
    3. Münch A
    (2006) Intermediate-asymptotic structure of a dewetting rim with strong slip. Appl Math Res Express 2006:1–25.
    OpenUrl
  42. ↵
    1. Fetzer R,
    2. Jacobs K,
    3. Münch A,
    4. Wagner B,
    5. Witelski TP
    (2005) New slip regimes and the shape of dewetting thin liquid films. Phys Rev Lett 95:127801.
    OpenUrlCrossRefPubMed
  43. ↵
    1. Fetzer R,
    2. Münch A,
    3. Wagner B,
    4. Rauscher M,
    5. Jacobs K
    (2007) Quantifying hydrodynamic slip: A comprehensive analysis of dewetting profiles. Langmuir 23:10559–10566.
    OpenUrlCrossRefPubMed
  44. ↵
    1. Fetzer R,
    2. Rauscher M,
    3. Münch A,
    4. Wagner BA,
    5. Jacobs K
    (2006) Slip-controlled thin-film dynamics. EPL 75:638–644.
    OpenUrl
  45. ↵
    1. Bäumchen O,
    2. Fetzer R,
    3. Jacobs K
    (2009) Reduced interfacial entanglement density affects the boundary conditions of polymer flow. Phys Rev Lett 103:247801.
    OpenUrlCrossRefPubMed
  46. ↵
    1. Bäumchen O,
    2. Jacobs K
    (2010) Slip effects in polymer thin films. J Phys Condens Matter 22:033102.
    OpenUrlPubMed
  47. ↵
    1. Flitton JC,
    2. King JR
    (2005) Surface-tension-driven dewetting of Newtonian and power-law fluids. J Eng Math 50:241–266.
    OpenUrl
  48. ↵
    1. Münch A,
    2. Wagner B
    (2005) Contact-line instability of dewetting thin films. Physica D 209:178–190.
    OpenUrlCrossRef
  49. ↵
    1. Holmes MH
    (2012) Introduction to Perturbation Methods (Springer Science & Business Media, New York), Vol 20.
  50. ↵
    1. Dziwnik M,
    2. Korzec M,
    3. Münch A,
    4. Wagner B
    (2014) Stability analysis of unsteady, nonuniform base states in thin film equations. Multiscale Model Simul 12:755–780.
    OpenUrl
  51. ↵
    1. King J,
    2. Münch A,
    3. Wagner B
    (2009) Linear stability analysis of a sharp-interface model for dewetting thin films. J Eng Math 63:177–195.
    OpenUrl
  52. ↵
    1. Lessel O, et al.
    (2015) Self-assembled silane monolayers: An efficient step-by-step recipe for high-quality, low energy surfaces. Surf Interf Anal 47:557–564.
    OpenUrl
  53. ↵
    1. Bäumchen O, et al.
    (2012) Slippage and nanorheology of thin liquid polymer films. J Phys Condens Matter 24:325102.
    OpenUrlPubMed
  54. ↵
    1. Ellero M, et al.
    1. Bäumchen O,
    2. Fetzer R,
    3. Münch A,
    4. Wagner B,
    5. Jacobs K
    (2009) Comprehensive analysis of dewetting profiles to quantify hydrodynamic slip. IUTAM Symposium on Advances in Micro-and Nanofluidics, ed Ellero M, et al. (Springer, New York), pp 51–65.
  55. ↵
    1. McGraw JD, et al.
    (2014) Nanofluidics of thin polymer films: Linking the slip boundary condition at solid–liquid interfaces to macroscopic pattern formation and microscopic interfacial properties. Adv Colloid Interf Sci 210:13–20.
    OpenUrl
  56. ↵
    1. Zhornitskaya L,
    2. Bertozzi AL
    (1999) Positivity-preserving numerical schemes for lubrication-type equations. SIAM J Numer Anal 37:523–555.
    OpenUrl
  57. ↵
    1. Grün G,
    2. Rumpf M
    (2000) Nonnegativity preserving convergent schemes for the thin film equation. Numer Math 87:113–152.
    OpenUrl
  58. ↵
    1. Demetrios TP
    (1995) On the breakup of viscous liquid threads. Phys Fluids 7:1529–1544.
    OpenUrlCrossRef
  59. ↵
    1. Dallaston MC,
    2. Fontelos MA,
    3. Tseluiko D,
    4. Kalliadasis S
    (2018) Discrete self-similarity in interfacial hydrodynamics and the formation of iterated structures. Phys Rev Lett 120:034505.
    OpenUrl
  60. ↵
    1. Eggers J
    (1997) Nonlinear dynamics and breakup of free-surface flows. Rev Mod Phys 69:865–930.
    OpenUrlCrossRef
  61. ↵
    1. Bertozzi AL
    (1996) Lubrication approximation for surface tension driven interfaces: Some open problems. Z Angew Math Mech 76:373–419.
    OpenUrl
  62. ↵
    1. Kim J,
    2. Kang K,
    3. Lowengrub J
    (2004) Conservative multigrid methods for Cahn–Hilliard fluids. J Comput Phys 193:511–543.
    OpenUrl
  63. ↵
    1. Peschka D
    (2015) Thin-film free boundary problems for partial wetting. J Comput Phys 295:770–778.
    OpenUrl
  64. ↵
    1. Peschka D
    (2018) Variational approach to contact line dynamics for thin films. Phys Fluids 30:082115.
    OpenUrl
PreviousNext
Back to top
Article Alerts
Email Article

Thank you for your interest in spreading the word on PNAS.

NOTE: We only request your email address so that the person you are recommending the page to knows that you wanted them to see it, and that it is not junk mail. We do not capture any email address.

Enter multiple addresses on separate lines or separate them with commas.
Signatures of slip in dewetting polymer films
(Your Name) has sent you a message from PNAS
(Your Name) thought you would like to see the PNAS web site.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
Signatures of slip in dewetting polymer films
Dirk Peschka, Sabrina Haefner, Ludovic Marquant, Karin Jacobs, Andreas Münch, Barbara Wagner
Proceedings of the National Academy of Sciences May 2019, 116 (19) 9275-9284; DOI: 10.1073/pnas.1820487116

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Request Permissions
Share
Signatures of slip in dewetting polymer films
Dirk Peschka, Sabrina Haefner, Ludovic Marquant, Karin Jacobs, Andreas Münch, Barbara Wagner
Proceedings of the National Academy of Sciences May 2019, 116 (19) 9275-9284; DOI: 10.1073/pnas.1820487116
Digg logo Reddit logo Twitter logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Mendeley logo Mendeley
Proceedings of the National Academy of Sciences: 116 (19)
Table of Contents

Submit

Sign up for Article Alerts

Article Classifications

  • Physical Sciences
  • Physics

Jump to section

  • Article
    • Abstract
    • Emergence of Dewetting Patterns
    • Thin-Film Models and Instability
    • Experimental Setup and Methods
    • Numerical Methods
    • Initial Data for Rims
    • Breakup of Liquid Ridges
    • Conclusion and Discussion
    • Acknowledgments
    • Footnotes
    • References
  • Figures & SI
  • Info & Metrics
  • PDF

You May Also be Interested in

Abstract depiction of a guitar and musical note
Science & Culture: At the nexus of music and medicine, some see disease treatments
Although the evidence is still limited, a growing body of research suggests music may have beneficial effects for diseases such as Parkinson’s.
Image credit: Shutterstock/agsandrew.
Scientist looking at an electronic tablet
Opinion: Standardizing gene product nomenclature—a call to action
Biomedical communities and journals need to standardize nomenclature of gene products to enhance accuracy in scientific and public communication.
Image credit: Shutterstock/greenbutterfly.
One red and one yellow modeled protein structures
Journal Club: Study reveals evolutionary origins of fold-switching protein
Shapeshifting designs could have wide-ranging pharmaceutical and biomedical applications in coming years.
Image credit: Acacia Dishman/Medical College of Wisconsin.
White and blue bird
Hazards of ozone pollution to birds
Amanda Rodewald, Ivan Rudik, and Catherine Kling talk about the hazards of ozone pollution to birds.
Listen
Past PodcastsSubscribe
Goats standing in a pin
Transplantation of sperm-producing stem cells
CRISPR-Cas9 gene editing can improve the effectiveness of spermatogonial stem cell transplantation in mice and livestock, a study finds.
Image credit: Jon M. Oatley.

Similar Articles

Site Logo
Powered by HighWire
  • Submit Manuscript
  • Twitter
  • Facebook
  • RSS Feeds
  • Email Alerts

Articles

  • Current Issue
  • Latest Articles
  • Archive

PNAS Portals

  • Anthropology
  • Chemistry
  • Classics
  • Front Matter
  • Physics
  • Sustainability Science
  • Teaching Resources

Information

  • Authors
  • Editorial Board
  • Reviewers
  • Librarians
  • Press
  • Site Map
  • PNAS Updates

Feedback    Privacy/Legal

Copyright © 2021 National Academy of Sciences. Online ISSN 1091-6490