TOPICAL REVIEW
arXiv:0909.1946v1 [cond-mat.soft] 10 Sep 2009
Slip effects in polymer thin films
O Bäumchen and K Jacobs
Experimental Physics, Saarland University, Campus, D-66123 Saarbrücken, Germany
E-mail: k.jacobs@physik.uni-saarland.de
Abstract. Probing the fluid dynamics of thin films is an excellent tool to study the
solid/liquid boundary condition. There is no need for external stimulation or pumping
of the liquid due to the fact that the dewetting process, an internal mechanism, acts
as a driving force for liquid flow. Viscous dissipation within the liquid and slippage
balance interfacial forces. Thereby, friction at the solid/liquid interface plays a key role
towards the flow dynamics of the liquid. Probing the temporal and spatial evolution
of growing holes or retracting straight fronts gives, in combination with theoretical
models, information of the liquid flow field and especially the boundary condition at the
interface. We review the basic models and experimental results obtained during the last
years with exclusive regard to polymers as ideal model liquids for fluid flow. Moreover,
concepts that aim on explaining slippage on the molecular scale are summarized and
discussed.
PACS numbers: 68.15.+e, 83.50.Lh, 83.80.Sg, 47.15.gm
Submitted to: J. Phys.: Condens. Matter
Slip effects in polymer thin films
2
1. Introduction
Understanding liquid flow in confined geometries plays a huge role in the field of microand nanofluidics [1]. Nowadays, microfluidic or so-called lab-on-chip devices are utilized
in a wide range of applications. Pure chemical reactions as well as biological analysis
performed on such a microfluidic chip allow a high performance while solely small
amounts of chemicals are needed. Thereby, analogies to electronic large-scale integrated
circuits are evident. Thorsen et al fabricated a microfluidic chip with a high density of
micromechanical valves and hundreds of individually addressable chambers [2]. Recent
developments tend to avoid huge external features such as pumps to control the flow by
designing analogues to capacitors, resistors or diodes that are capable to control currents
in electronic circuits [3].
By reducing the spacial dimensions of liquid volume in confined geometries, slippage
can have a huge impact on flow dynamics. Especially the problem of driving small
amounts of liquid volume through narrow channels has drawn the attention of many
researchers on slip effects at the solid/liquid interface. The aim is to reduce the pressure
that is needed to induce and to maintain the flow. Hence, the liquid throughput is
increased and, what is important in case of polydisperse liquids or mixtures, a low
dispersity according to lower velocity gradients perpendicular to the flow direction is
generated. Confined geometries are realized in various types of experiments: The physics
and chemistry of the imbibition of liquids by porous media is of fundamental interest
and enormous technological relevance [4]. Channel-like three-dimensional structures
can be used to artificially model the situation of fluids in confinement. Moreover, the
small gap between a colloidal probe and a surface filled with a liquid (e.g realized in
surface force apparatus or colloidal probe atomic force microscopy experiments) is a
common tool to study liquid flow properties. Besides the aforementioned experimental
systems, knowledge in preparation of thin polymer films has been extensively gained due
to its enormous relevance in coating and semiconductor processing technology. Such a
homogeneous nanometric polymer film supported by a very smooth substrate, as for
example a piece of a silicon (Si) wafer, exhibits two relevant interfaces, the liquid/air
and the substrate/liquid interface. Si wafers are often used according to their very low
roughness and controllable oxide layer thickness. Yet, also highly viscous and elastically
deformable supports, such as e.g. polydimethylsiloxane (PDMS) layers, are versatile
substrates.
The stability of a thin film is governed by the effective interface potential φ as
function of film thickness h. In case of dielectric systems, φ(h) is composed of an
attractive van-der-Waals part and a repulsive part [5, 6, 7, 8]. For the description of
van-der-Waals forces of a composite substrate, the layer thicknesses and their respective
polarisation properties have to be taken into account [9]. Thereby, three major situations
of a thin liquid film have to be distinguished: stable, unstable and metastable films. As
illustrated in Fig. 1 by the typical curves of φ(h), a stable liquid film is obtained if
the effective interface potential is positive and monotonically decaying (cf. curve (1) in
Slip effects in polymer thin films
3
Figure 1. Different shapes of the effective interface potential φ(h) associated with
different wetting conditions. Curve (1) characterizes a stable liquid film. Curve (2)
represents a metastable, curve (3) and (4) an unstable situation.
Fig. 1). The equilibrium film thickness heq is infinite and the liquid perfectly wets the
substrate. In case of a global minimum of φ(h), curves (2) and (3) in Fig. 1, the system
can minimize its energy and a finite value for heq results. The metastable situation is
furthermore characterized by a potential barrier that the system has to overcome to
reduce its potential energy (cf. curve (2) in Fig. 1). Curve (3) and (4) characterize
unstable conditions, since every slight fluctuation in film height will drive the system
towards the global minimum. The wettability of the substrate by the liquid is correlated
to the depth of the minimum at φ(heq ). The deeper the global minimum of φ, the larger
is the equilibrium contact angle of the liquid on the surface. A nearly 180 ◦ situation is
depicted in curve (4) of Fig. 1.
For a 100 nm polystyrene (PS) film on a hydrophobized Si wafer with a native oxide
layer, dewetting starts after heating the sample above the glass transition temperature
of the polymer. Holes nucleate according to thermal activation or nucleation spots (dust
particles, inhomogeneities of the substrate or of the polymer film) and grow with time,
cf. Fig. 2. The subsequent stages of dewetting are characterized by the formation
of liquid ridges by coalescence of growing holes and traveling fronts. Very thin films
in the range of several nanometers may become unstable and can dewet according to
thermally induced capillary modes, that are amplified by forces contributing to the
effective interface potential. This phenomenon is characterized by the occurrence of
a preferred wavelength and is called spinodal dewetting. To study thin film flow with
regard to the influence of slippage, especially nucleated holes enable an easy experimental
access for temporal and spatial observation.
While fronts retract from the substrate and holes grow, a liquid rim is formed at the
three-phase contact line due to conservation of liquid volume. A common phenomenon
Slip effects in polymer thin films
4
Figure 2. Dewetting of a 80 nm PS(65 kg/mol) film at T = 135 ◦C from a
hydrophobized Si substrate captured by optical microscopy (adapted from [10]).
is the formation of liquid bulges and so-called ”fingers” due to the fact that the liquid
rim becomes unstable, similar to the instability of a cylindrical liquid jet that beads up
into droplets (e.g. in case of a water tap). This so-called Rayleigh-Plateau instability
is based on the fact that certain modes of fluctuations become amplified and surface
corrugations of a characteristic wavelength become visible. If two holes coalesce, a
common ridge builds up that in the end decays into single droplets due to the same
mechanism. The final stage is given by an equilibrium configuration of liquid droplets
arranged on the substrate exhibiting a static contact angle. Actually, the final state
would be one single droplet, since the Laplace pressure in droplets of different size is
varying. Yet then, a substantial material transport must take place, either over the gas
phase or via an equilibrium film between the droplets. This phenomenon is also called
Ostwald ripening. In polymeric liquids, these transport pathways are usually extremely
slow so that a network-like pattern of liquid droplets is already termed ”final stage”.
Besides the dynamics and stability of thin liquid films driven by intermolecular forces,
a recent article by Craster and Matar reviews further aspects such as e.g thermally or
surfactant driven flows [11].
In the first part of this topical review, basic concepts from hydrodynamic theories in
the bulk situation to corresponding models with regard to confinement are introduced.
Polymers are regarded as ideal model liquids due to their low vapor pressure, the
available chemical pureness and furthermore the fact that their viscosity can be
controlled in a very reliable manner. By setting the viscosity via temperature, the
experimental conditions can be tuned so that dewetting dynamics can be easily captured.
Mass conservation can safely be assumed, which consequently simplifies the theoretical
description. The dewetting dynamics governed by the driving forces and the mechanisms
of energy dissipation will be discussed also with regard to the shape of liquid ridges. The
second part summarizes experimental studies concerning the dynamics of two different
dewetting geometries: the straight front geometry and the growth of holes. Especially
the influence of parameters such as dewetting temperature, viscosity and molecular
weight of the polymer will be discussed in detail. Moreover, we focus on various scenarios
at the solid/liquid interface on the molecular level. Simulations such as for example
molecular-dynamic (MD) studies help to obtain more information and are supportive
to gain insight into the molecular mechanism of slippage.
Slip effects in polymer thin films
5
2. Basic theoretical concepts
In this section we aim to describe the main concepts of fluid dynamics especially in
a confined geometry. Depending on the type of liquid, viscous or even viscoelastic
effects have to be considered and deviations from Newtonian behavior might become
non-negligible. Since we concentrate in this article on polymer melts, viscosity and
viscoelasticity can be varied by chain length and branching of the polymer. An
important aspect of a moving liquid is the velocity profile.
2.1. Polymer properties
For a comprehensive understanding of slippage, some important polymer properties
like glass transition temperature, viscosity and viscoelasticity have to be taken into
account. Especially in geometries like in a liquid film, confinement effects are a concern.
A detailed description can be found in textbooks [12, 13].
2.1.1. Polymer physics Polymers are synthesized by polymerization of monomers of
molar mass Mmono . To characterize the polydispersity of polymer chains in a solution
or in a melt, the polydispersity index Mw /Mn is calculated as the ratio of mean values
given by the weight Mw and number averaged Mn molecular weight. Polymers are
able to change their conformations. The radius of gyration characterizes the spatial
dimension of the polymer and is given as the mean square displacement between
monomers and the polymer’s center of mass. In an isotropic configuration, the shape of
the polymer chain can be approximated as a spherical entity. Polystyrene, abbreviated
PS, which is commonly used as a melt in dewetting experiments, is a linear homopolymer
with Mmono = 104 g/mol. Besides other properties concerning the micro-structure of
polymer chains such as the tacticity and the architecture (linear, branched, ring-shaped),
physical properties are of special interest. If a melt is heated above its glass transition
temperature Tg , a phase transition from the glassy phase to the liquid phase occurs
and the polymer becomes liquid. Randomly structured macromolecules such as atactic
polymers avoid the formation of semi-crystalline domains below Tg and exhibit a pure
amorphous phase. The glass transition of a bulk polymer or of a polymeric thin film
can be observed e.g. via probing its linear expansion coefficient. Although, the glass
transition is based on a kinematic effect and does not occur due to a rearrangement
process of polymer chains. Therefore, the glass transition usually takes place on a
specific temperature range and not at a exactly allocable temperature. According to
the increased mobility of shorter chains, their glass transition temperature is decreased
significantly. For PS with a sufficiently large chain-length, Tg =100 ◦ C.
The viscosity η of a polymer melt measures the inner friction of polymer chains
and governs the time scale of flow processes. Due to the fact that the mobility of chains
increases while the temperature increases, the viscosity decreases. Internal stresses relax
and dynamical processes proceed faster. The most common description of the functional
Slip effects in polymer thin films
6
dependency of the viscosity η of a polymer on temperature T was developed by Williams,
Landel and Ferry:
η = ηg exp
B(Tg − T )
.
fg (T − T∞ )
(2.1)
In this so-called WLF-equation, ηg denotes the viscosity at Tg , B an empirically obtained
constant, T∞ the so-called Vogel temperature and fg the free liquid volume fraction.
Besides the above discussed impact of temperature on the viscosity of a polymer
melt, also the molecular weight Mw strongly influences η. While Mw increases, the
chain mobility decreases and therefore the relaxation times and thus η increase. For
sufficiently small Mw , the Rouse model predicts a linear increase of η for increasing Mw .
At this point, another characteristic number of polymer physics has to be introduced:
the critical chain length for entanglements Mc . Above Mc , the polymer chains form chain
entanglements exhibiting a specific mean strand length called entanglement length Me .
For PS, Mc =35 kg/mol and Me =17 kg/mol is found [12]. According to the reptation
model of de Gennes [14], the viscosity increases stronger in the presence of chain
entanglements and an algebraic behavior of η ∝ Mw3 is expected. Empirically, the linear
regime below Mc is well reproduced, above Mc an exponent of 3.4 is found experimentally
for different polymers. For a detailed description of chemical and physical properties of
polymers as well as for the Rouse or the reptation theory we refer to the book Polymer
physics by Rubinstein and Colby [12].
2.1.2. Properties in confined geometries In contrast to the bulk situation, where
volume properties of the polymers are measured, liquids in confined geometries such as
a thin film often show deviations from the behavior in the volume due to additional
interface effects. One of these properties is the aforementioned glass transition
temperature. It has been shown in numerous studies, that Tg changes with film thickness
h. On the one hand, in case of free-standing or supported films exhibiting no or
repulsive interactions with the substrate, Tg (h) decreases with decreasing film thickness
[15, 16, 17, 18]. On the other hand, Keddie and Jones have shown that an increase of
the glass transition temperature with decreasing film thickness is possible for attractive
interactions between substrate and polymer film [19]. The influence of the interfacial
energy on the deviation of Tg from its bulk value has been studied and quantified by
Fryer and co-workers for different polymers [20]. In case of PS on a solid support, a
significant change of Tg is found for films thinner than 100 nm (see Fig. 3) [21]. For
PS(2k) below 10 nm for instance, the glass transition temperature and the viscosity of
the polymer film are affected such that these films are liquid at room temperature and
may dewet spinodally.
Several attempts have been made to explain the change of Tg according to the
film thickness. Besides interface-related effects such as reorientation of polymer chains
or accumulation of chain-ends at the interface, finite-size effects have been proposed
to be responsible. Herminghaus et al discussed the strain relaxation behavior of thin
Slip effects in polymer thin films
glass transition temperature T
g
[K]
0
7
film thickness [nm]
50
100
150
200
320
10
300
280
100
320
300
280
260
260
Figure 3. Glass transition temperature Tg of polystyrene films of 2 kg/mol against
film thickness (adapted from [21]).
viscoelastic polymer films with regard to surface melting and the shift of the glass
transition temperature [21]. Kawana and Jones studied the thermal expansivity of
thin supported polymer films using ellipsometry and attributed their results concerning
Tg to a liquid-like surface layer [22], a result that was also found by other authors
[23, 24]. Besides confinement effects on Tg , further interface-related phenomena have
been studied: Si et al have shown that polymers in thin films are less entangled than
bulk polymers and that the effective entanglement molecular weight Me is significantly
larger than the bulk value [25].
2.1.3. Viscosity and viscoelasticity One of the major characteristics of a liquid in
general or a polymer in particular is its viscosity η. Applying shear stress σ to a liquid,
it usually reacts with a strain γ. If stress and strain rate γ̇ are proportional, the fluid is
called Newtonian. The constant of proportionality is identified as the viscosity η of the
liquid.
σ = η γ̇.
(2.2)
Liquids such as long-chained polymers show a shear rate dependent viscosity η(γ̇)
due to the fact that the liquid molecules are entangled. If the viscosity increases while
shearing the liquid, we call this behavior shear thickening, whereas in case of lowered
viscosity so-called shear thinning is responsible. In contrast to the elastic deformation
of a solid, a deformation of a viscoelastic liquid might induce an additional flow and can
relax on a specific time scale τ . On short time scales (t < τ ), the liquid behaves in an
elastic, on long time scales (t > τ ) in a viscous manner. Thereby, strain γ is connected
to stress σ via the elastic modulus G of the liquid:
σ = Gγ.
(2.3)
Slip effects in polymer thin films
8
Figure 4. Maxwell model represented by a dashpot and a spring in a serial connection.
To cover the stress relaxation dynamics of a polymer film, several modelling
attempts have been proposed. Mostly, so-called Maxwell or Jeffreys models are applied.
The simplest model is the Maxwell model (see Fig. 4), which assumes a serial connection
of a perfectly elastic element (represented by a spring) and a perfectly viscous one
(represented by a dashpot). Consequently, the total shear strain γ is given by the sum
of the corresponding shear strains γe and γv of both mechanisms:
γ = γe + γv .
(2.4)
With (2.3) and (2.2) we get
σ = GM γe = ηM γ̇v
(2.5)
since both react on the same shear stress σ. Thereby, the ratio of viscosity of the viscous
element to the elastic modulus of the elastic one can be identified with a specific time
scale, the relaxation time τM :
ηM
.
(2.6)
τM =
GM
The relaxation of stress after a step strain γ leads to a time-dependent stress function
σ(t) for a viscoelastic liquid. Due to the fact that the total strain γ is constant, a first
order differential equation for the time-dependent strain γv (t) is obtained:
τM γ̇v = γ − γv (t).
(2.7)
Solving this differential equation using the initial condition γv (t = 0) = 0 gives a simple
exponential decay of σ(t) on the time scale of the stress relaxation time τM in the
Maxwell model:
γe (t) = γ exp (−t/τM ),
σ(t) = GM γe (t) = GM γ exp (−t/τM ).
(2.8)
A situation of special interest is the linear response region: For sufficiently small values
of γ, the stress/strain-relation (2.3) is valid and the stress relaxation modulus G(t) is
independent of the strain γ. In this regime, a linear superposition of stresses resulting
from an infinite number of strain steps can be used to model a steady simple shear flow
of a viscoelastic liquid. For larger applied shear rates, linear response and the linear
superposition fails. The viscosity is still defined as the ratio of stress and strain rate,
but it has to be regarded as an apparent viscosity which differs from the above described
Slip effects in polymer thin films
9
”zero shear rate” viscosity. Polymers with shear-thinning or shear-thickening properties
can be described by the function
σ ∝ γ̇ n ,
(2.9)
where the exponent n can be extracted from experimental data. These type of fluids are
also called ”power law fluids”. Moreover, also other nontrivial stress-strain relations can
be considered or alternatively non-linear extensions can be applied to the linear Maxwell
models. In case of the linear Jeffreys model, the stress tensor σij relaxes according to
the following constitutive relaxation equation:
(1 + λ1 ∂t )σ = η(1 + λ2 ∂t )γ̇,
(2.10)
where the strain rate is given by the gradient of the velocity field γ̇ij = ∂j ui + ∂i uj .
Hereby, λ1 governs the relaxation of stress, whereas λ2 (λ2 < λ1 ) describes the
relaxation of the strain rate, respectively. This model accounts for the viscous and
the elastic properties of a fluid and was used by Blossey, Rauscher, Wagner and Münch
as basis for the development of a thin-film equation that incorporates viscoelastic effects
[26, 27, 28]. For a more elaborate description of non-Newtonian flows we refer e.g. to
the correspondent work of Te Nijenhuis et al [29].
2.1.4. Reynolds and Weissenberg number The flow of a liquid can be characterized by
specific numbers. One of these numbers is the so-called Reynolds number Re, which
describes the ratio of inertia effects to viscous flow contributions. In case of thin liquid
films, Re can be written as
Re =
ρuh
,
η
(2.11)
where ρ denotes the density of the liquid, u describes the flow velocity and h stands
for the film thickness [30]. For thin dewetting polymer films, the Reynolds number is
very small, i.e. Re ≪ 1, and a low-Re lubrication theory can be applied. To quantify
and to judge the occurrence of viscoelastic effects versus pure viscous flow, the so-called
Weissenberg number W i has been introduced as
W i = τ γ̇.
(2.12)
Thereby, τ denotes the relaxation time and γ̇ the strain rate as introduced in the previous
section. If W i ≪ 1, an impact of viscoelasticity on flow dynamics can be neglected and
viscous flow dominates.
2.2. Navier-Stokes equations
The Navier-Stokes equations for a Newtonian liquid mark the starting point for the
discussion of fluid dynamics in confined geometries. According to conservation of mass,
the equation of continuity can be formulated as
∂t ρ + ∇ · (ρu) = 0,
(2.13)
Slip effects in polymer thin films
10
where u = (ux , uy , uz ) is the velocity field of the fluid. For an incompressible liquid,
which implies a temporally and spatially constant liquid density ρ, (2.13) can be
simplified to
∇u = 0.
(2.14)
With the conservation of momentum, the Navier-Stokes equations for an incompressible
liquid can be written as
ρ(∂t + u · ∇)u = −∇p + η△u + f,
(2.15)
with the pressure gradient ∇p and the volume force f of external fields acting as driving
forces for the liquid flow. We already stated that for small Reynolds numbers, i.e.
Re ≪ 1, the terms on the left hand side of (2.15) can be neglected as compared to terms
describing the pressure gradient, external volume forces and viscous flow. By that, we
can simplify (2.15) to the so-called Stokes equation
0 = −∇p + η△u + f.
(2.16)
In section 2.5, we will demonstrate how these basic laws of bulk fluid dynamics can be
applied to the flow geometry of a thin film supported by a solid substrate.
2.3. Free interface boundary condition
At the free interface of a supported liquid film, i.e. at the liquid/gas or usually the
liquid/air interface, no shear forces can be transferred to the gas phase due to the
negligible viscosity of the gas. In general, the stress tensor σij∗ is given by the stress
tension σij , see (2.2), and the pressure p:
σij∗ = σij + pδij = η(∂j ui + ∂i uj ) + pδij .
(2.17)
The tangential t and normal n (perpendicular to the interface) components of the stress
tensor are:
(σ ∗ · n) · t = 0
(σ ∗ · n) · n = γlv κ,
(2.18)
where κ denotes the mean curvature and γlv the interfacial tension (i.e. the surface
tension of the liquid) of the liquid/vapor interface. If the liquid is at rest, i.e. the
stationary case u = 0, the latter boundary condition gives the equation for the Laplace
pressure pL :
pL = γlv κ = γlv (
1
1
+
).
R1 R2
(2.19)
R1 and R2 are the principal radii of curvature of the free liquid/gas boundary; the
appropriate signs of the radii are chosen according to the condition that convex
boundaries give positive signs. Such convex liquid/gas boundaries lead to an additional
pressure within the liquid due to its surface tension. In the next section, the solid/liquid
boundary condition will be discussed, which yields a treatment of slip effects.
Slip effects in polymer thin films
partial slip
no slip
u
u
11
full slip
”apparent” slip
u
u
z0
b=0
b
b=¥
b
Figure 5. Different velocity profiles in the vicinity of the solid/liquid interface and
illustration of the slip (extrapolation) length b. The situation of so-called ”apparent”
slip is illustrated on the right: According to a thin liquid layer of thickness z0
that obtains a significantly reduced viscosity, the slip velocity ux |z=0 is zero, but a
substantial slip length is measured.
2.4. Slip/no-slip boundary condition
2.4.1. Navier slip boundary condition In contrast to fluid dynamics in a bulk volume,
where the assumption that the tangential velocity u|| at the solid/liquid interface
vanishes (no-slip boundary condition), confined geometries require a more detailed
investigation as slippage becomes important. In 1823, Navier [31] introduced a linear
boundary condition: The tangential velocity u|| is proportional to the normal component
of the strain rate tensor; the constant of proportionality is described as the so-called
slip length b:
u|| = b n · γ̇
(2.20)
In case of simple shear flow in x -direction, the definition of the slip length can be
alternatively written as
ux η
η
ux
|z=0 =
= ,
(2.21)
b=
∂z ux
σ
ξ
where ξ = σ/ux denotes the friction coefficient at the solid/liquid interface. The xyplane thereby represents the substrate surface. According to these definitions, the slip
length can be illustrated as the extrapolation length of the velocity profile ”inside” the
substrate, cf. Fig. 5. Moreover, both limiting cases are included within this description:
For b = 0, we obtain the no-slip situation, whereas b = ∞ characterizes a full-slip
situation. The latter case corresponds to ”plug-flow”, where the liquid behaves like a
solid that slips over the support.
2.4.2. How to measure the slip length? In recent years, numerous experimental studies
were published using diverse methods to probe the slip length at the boundary of
different simple or complex liquids and solid supports. For details concerning these
experimental methods we refer to the review articles from Lauga et al [32], Neto et
Slip effects in polymer thin films
12
al [33] and Bocquet and Barrat [34] (and references therein). To probe the boundary
condition, scientists performed either drainage experiments or direct measurements of
the local velocity profile using e.g. tracer particles.
In case of drainage experiments, the liquid is squeezed between two objects, e.g. a
flat surface and a colloidal probe at the tip of an AFM cantilever, and the corresponding
force for dragging the probe is measured (colloidal probe AFM ). Alternatively, in an
surface force apparatus (SFA), two cylinders arranged perpendicular to each other are
brought in closer contact and force/distance measurements are performed to infer the
slip length.
The use of tracer particles as a probe of the local flow profile might bring some
disadvantages. The chemistry of these particles is usually different from the liquid
molecules and their influence on the results might not be negligible. A similar method
is called fluorescence recovery after photo bleaching. Thereby, a distinct part of a
fluorescent liquid is bleached by a laser pulse and the flow of non-bleached liquid into
that part is measured. The disadvantage of this method is that diffusion might be a
further parameter that is hard to control. Recently, Joly et al showed that also thermal
motion of confined colloidal tracers in the vicinity of the solid/liquid interface can be
used as a probe of slippage without relying on external driving forces [35].
2.4.3. Which parameters influence slippage? Of course, many interesting aspects in the
field of micro- and nanofluidics are related to intrinsic parameters that govern slippage
of liquid molecules at the solid/liquid interface. For simple liquids on smooth surfaces,
the contact angle is one of the main parameters influencing slippage [36, 37, 38, 39].
This originates from the effect of molecular interactions between liquid molecules
and the solid surface: If the molecular attraction of liquid molecules and surface
decreases (and thereby the contact angle increases), slippage is enlarged. Further
studies aim to quantify the impact of roughness [37, 40, 41, 42] or topographic structure
[43, 44, 45, 46, 47] of the surface on slippage. For different roughness length scales,
a suppression (see e.g. [37, 40]) or an amplification (see e.g. [43]) of slippage can be
observed. Moreover, the shape of molecular liquids itself has been experimentally shown
to impact the boundary condition. Schmatko et al found significantly larger slip lengths
for elongated linear compared to branched molecules [48]. This might be associated
with molecular ordering effects [49] and the formation of layers of the fluid in case of the
capability of these liquids to align in the vicinity of the interface [50]. Cho et al identified
the dipole moment of Newtonian liquids at hydrophobic surfaces as a crucial parameter
for slip [51]. De Gennes proposed a thin gas layer at the interface of solid surface and
liquid as a possible source of large slip lengths [52]. Recently, MD studies for water on
hydrophobic surfaces by Huang et al revealed a dependence of slippage on the amount of
water depletion at the surface and a strong increase of slip with increasing contact angle
[53]. Such depletion layers for water in the vicinity of smooth hydrophobic surfaces
have also been experimentally observed using scattering techniques [54, 55, 56, 57].
Contamination by nanoscale air bubbles (so-called nanobubbles) and its influence on
Slip effects in polymer thin films
13
z
h(x,y,t)
u||(ux,uy)
y
x
Figure 6. Illustration of the nomenclature of the thin film length scales (x and y are
parallel to the substrate) and the velocity contribution u|| = (ux , uy ).
slippage has been controversially discussed in literature (see e.g. [58, 59, 60, 61]). In
case of more complex liquids such as polymer melts further concepts come into play.
They will be illustrated in section 3.4.
2.5. Thin-film equation for Newtonian liquids
2.5.1. Derivation Confining the flow of a liquid to the geometry to the one of a
thin film, we can assume that the velocity contribution perpendicular to the substrate
is much smaller than the parallel one. Furthermore, the lateral length scale of film
thickness variations is much smaller than the film thickness itself. On the basis of these
assumptions, Oron et al [62] developed a thin-film equation from the rather complex
equations of motion, p
(2.13) and (2.15). In case of film thicknesses smaller than the
capillary length lc = γlv /ρg, (which is typically in the order of magnitude of 1 mm)
(2.16) can be written as
0 = −∇(p + φ′ (h)) + η△u
(2.22)
Additional external fields such as gravitation can be neglected, but a secondary
contribution φ′ (h), the disjoining pressure, has been added to the capillary pressure
p. The disjoining pressure originates from molecular interactions of the fluid molecules
with the substrate. The effective interface potential φ(h) summarizes the inter-molecular
interactions and describes the energy that is required to bring two interfaces from infinity
to the finite distance h. As already discussed in the introductory part, the stability of a
thin liquid film is also governed by φ(h). For a further description of thin film stability,
we refer to [8] and the references therein.
The derivation of a thin-film equation for Newtonian liquids starts with the
kinematic condition
Z h
∂t h = −∇||
u|| dz,
(2.23)
0
i.e. the coupling of the time derivative of h(x, y, t) to the flow field, where the index
|| in general denotes the components parallel to the substrate (∇|| = (∂x , ∂y ) and
u|| = (ux , uy )) as illustrated in Fig. 6.
Slip effects in polymer thin films
14
For thin liquid films, film thickness variations on lateral scale L are much larger
than the length scale of the film thickness H. Introducing the parameter ǫ = H/L ≪ 1
yields the so-called lubrication approximation and is used in the following to re-scale
the variables to dimensionless values. In a first approximation, linearized equations are
obtained while neglecting all terms of the order O(ǫ2 ). For reasons of simplicity and due
to translational invariance in the surface plane, a one-dimensional geometry is used:
∂x (p + φ′ ) = ∂z2 ux ,
∂z (p + φ′ ) = 0,
∂x ux + ∂z uz = 0.
(2.24)
While the substrate is supposed to be impenetrable for the liquid, i.e. uz = 0 for z = 0,
friction at the interface implies a velocity gradient ∂z ux = ux /b for z = 0. Moreover,
the tangential and normal boundary condition at the free interface, i.e. z = h(x), can
be simplified in the following manner:
∂z ux = 0,
p + ∂x2 h = 0.
(2.25)
From (2.24) and the boundary conditions, the velocity profile ux (z) can be obtained.
Using the kinematic condition (2.23), the equation of motion for thin films in three
dimensions is derived:
∂t h = −∇[m(h)∇(γlv △h − φ′ (h))],
where m(h) denotes the mobility given by
1
m(h) = (h3 + 3bh2 ).
3η
(2.26)
(2.27)
2.5.2. Lubrication models including slippage As discussed in the previous section, the
derivation of the thin-film equation is based on the so-called lubrication approximation
and the re-scaling of relevant values in ǫ. As a consequence, the slip length b is supposed
to obtain values smaller than the film thickness h, i.e. b ≪ h. To extend this so-called
weak-slip situation with regard to larger slip b ≫ h, Münch et al [30] and Kargupta et
al [63] developed independently so-called strong-slip models. Thereby, the slip length is
defined as b = β/ǫ2 . The corresponding equation of motion together with the kinematic
condition in one dimension for a Newtonian thin liquid film read as:
bh
2b
u = ∂x (2ηh∂x u) + ∂x (γlv ∂x2 h − φ′ (h))
η
η
(2.28)
∂t h = −∂x (hu).
In fact, a family of lubrications models, cf. Tab. 1, accounting for different slip
situations have been derived. In the limit b → 0, i.e. the no-slip situation, the mobility
is given by m(h) = h3 /3η. If the slip length is in the range of the film thickness b ∼ h,
the mobility in the corresponding intermediate-slip model is m(h) = bh2 /η. Recently,
Fetzer et al [64] derived a more generalized model based on the full Stokes equations,
developed up to third order of a Taylor expansion. The authors were able to show that
this model is in good agreement with numerical simulations of the full hydrodynamic
equations and is not restricted to a certain slip regime as the aforementioned lubrication
models.
Slip effects in polymer thin films
15
Table 1. Summary of lubrication models for Newtonian flow and different slip
situations.
model
validity
equation
limiting cases
ref.
weak-slip
b≪h
(2.26), (2.27)
b → 0 (no-slip)
b → ∞ (intermediate-slip)
[62]
strong-slip
b≫h
(2.28)
β → 0 (intermediate-slip)
β → ∞ (”free”-slip)
[63, 30]
2.5.3. Lubrication models including viscoelasticity In the meanwhile, the derivation of
a thin film equation for the weak-slip case including linear viscoelastic effects of Jeffreys
type (such as described by equation (2.10) in section 2.1.3) has been achieved (see [26]).
To cover relaxation dynamics of the stress tensor σ, an additional term ∇ · σ on the
right hand side of (2.22) has to be included to the aforementioned model for Newtonian
liquids. Furthermore, the treatment of linear viscoelastic effects was also achieved for
the strong-slip situation by Blossey et al [27]. To summarize these extensions, the
essential result is the fact that linear viscoelastic effects are absent in the weak-slip case
and the Newtonian thin-film model is still valid. The strong-slip situation, however,
is more complicated. Slippage and viscoelasticity are combined and strongly affect the
corresponding equations. In the meanwhile, the authors were able to fully incorporate
the non-linearities of the co-rotational Jeffreys model for viscoelastic relaxation into
their thin-film model [28].
The extensions of the aforementioned thin-film models for different slip conditions
with or without the presence of viscoelastic relaxation (Newtonian and non-Newtonian
models) affect on the one hand the rupture conditions, but also on the other hand
the shape of a liquid ridge. These two phenomena will be discussed in the next two
subsections. A elaborate description of these theoretical aspects can be found in a
recent review article by Blossey [65].
2.5.4. Application I - Spinodal dewetting One of the main applications of the
theoretical thin-film models is the dewetting of thin polymer films. As introduced in
section 1 and illustrated by Fig. 1, the stability of a thin liquid film is governed by the
effective interface potential. Basically, long-range attractive van der Waals forces add
to short-range repulsive forces. Due to the planar geometry of two interfaces of distance
d, the van der Waals contribution to the potential is φ(d)vdW ∝ −A/d2 , where A is the
Hamaker constant. For the description of the explicit calculation of Hamaker constants
from the dielectric functions of the involved materials we refer to [8] and to the book by
Israelachvili [66]. Experimental systems often exhibit multi-layer situations, cf. Fig. 7.
A hydrophobic film and/or an oxide layer of distinct thicknesses di exhibiting Hamaker
constants Ai require a superposition of contributions to the potential:
φ(d)vdw = −
A2 − A1
A3 − A2
A1
−
−
2
2
12πd
12π(d + d1 )
12π(d + d1 + d2 )2
(2.29)
Slip effects in polymer thin films
16
Figure 7. Two examples for polystyrene films prepared on multi-layer substrates.
Left: Silicon wafer with native oxide layer covered with a hydrophobic layer (e.g. a
self-assembled monolayer (SAM)). Right: Silicon wafer with an increased (compared
to a native silicon oxide) oxide layer thickness.
Figure 8. Effective interface potential φ(h), gained from experimental data, plotted
against the thickness h of a thin PS film prepared on silicon wafers with different oxide
layer thicknesses d1 (adapted from [7]). The cross-hatched rectangle marks the error
for heq and the depth of the global minimum.
Consequently, the shape of the effective interface potential (and thereby also the thin
film stability) is governed by the set of Hamaker constants Ai and film thicknesses di .
E.g. for a thin PS film of thickness h on a Si substrate with a native oxide layer of
2.4 nm, φ(h) shows a global minimum at an equilibrium film thickness heq (c.f. Fig. 8).
Moreover, a local maximum at h > heq is found.
If the PS film is sufficiently thin (φ′′ (h) < 0) , it may become unstable due to
the amplification of thermally induced capillary waves. To track the evolution of small
fluctuations of the film thickness h, i.e. f (x, t) = h(x, t) − h with f (x, t) ≪ h, a
Fourier transform of the linearized thin film equation (2.26) has to be performed. The
amplitudes of the capillary waves grow exponentially in time. Their growth rate α can
be calculated as a function of the wavenumber q and depends on the sign of the local
curvature of the interface potential. If the second derivative of the effective interface
potential φ′′ at the film thickness h is positive, α is negative for all q and the amplitudes
Slip effects in polymer thin films
17
Figure 9. In situ (at elevated temperature T = 53 ◦ C) atomic force microscopy (AFM)
images (corresponding annealing times given in pictures) of a spinodally dewetting
3.9(2) nm PS(2 kg/mol) film on a Si wafer with a thick (191 nm) oxide layer (adapted
from [67]).
of the capillary waves are damped. If φ′′ < 0, the growth rate α is positive for a certain
range of wavenumbers up to a critical wavenumber qc and capillary waves are amplified.
The wavenumber that corresponds to the maximum value of α and therefore exhibits
the fastest amplification, is called preferred wavenumber q0 and is connected to the
preferred wavelength λ0 = 2π/q0 . The latter is also called spinodal wavelength and can
be written as
s
8π 2 γlv
.
(2.30)
λ0 =
−φ′′ (h)
The spinodal dewetting process can be monitored e.g. by atomic force microscopy
(AFM) as shown in Fig. 9. By measuring λ0 as a function of film thickness, φ′′ (h)
can be inferred and conclusions can be drawn with regard to the effective interface
potential φ(h) [7]. For further details concerning stability of thin films we refer to [8]
and references therein.
Using the strong-slip model while taking slip into account (2.28), Rauscher and
coworkers could theoretically show that slippage is supposed to influence the capillary
wave spectrum due to a different mobility at the solid/liquid interface [68]. The position
of the maximum q0 shifts to smaller wavenumbers and larger wavelengths for increasing
slip length. As for today, to the best of our knowledge experimental studies concerning
the impact of slippage on the spinodal wavelength are not available.
The description of the influence of thermal noise on the temporal and spatial dynamics
of spinodally dewetting thin polymer films has been recently achieved by Fetzer and
coworkers [69]. A stochastic Navier-Stokes equation with an additional random stress
fluctuation tensor that accounts for thermal molecular motion is utilized to model the
flow while assuming a no-slip boundary condition at the solid/liquid interface. The
stochastic model matches the experimentally observed spectrum of capillary waves and
thermal fluctuations cause the coarsening of typical length scales.
Slip effects in polymer thin films
18
Figure 10. Left: AFM image (scan size 10 µm) of a liquid rim formed during hole
growth in a PS film dewetting from a hydrophobic substrate (Si wafer covered with
a 21 nm hydrophobic Teflonr coating (AF 1600)). Right: Different rim morphologies
for PS on AF 1600 (AFM cross-sections): Profile exhibiting a trough (depicted in the
inset) for PS(35.6 kg/mol) at 120 ◦C and a monotonically decaying rim shape for
PS(125 kg/mol) at 130 ◦ C.
2.5.5. Application II - Shape of a dewetting rim Besides the implications on spinodal
dewetting, the one-dimensional thin-film model has been successfully applied to the
shape of the rim along the perimeter of e.g. nucleated holes. Experimentally, researchers
have studied and observed different types of rim profiles [70, 71, 72]. As shown in Fig.
10, profiles either decay monotonically into the undisturbed film or they show a more
symmetrical profile exhibiting a trough (termed ”oscillatory shape”). If the depth of
this through is in the range of the film thickness, a ring of so-called satellite holes can
be generated [73, 74].
The shape of a dewetting rim can be understood by the aforementioned thin film
theory for Newtonian liquids: Introducing a small perturbation δh(x, t) ≪ h of the film
thickness h(x, t) and small velocities u(x, t) leads to linearized thin-film equations that
describe the temporal and spatial evolution of δh. Thereby, the disjoining pressure φ′ (h)
can be neglected due to the fact that films thicker than 10 nm are considered. To obtain
stationary solutions of the linearized equations, a frame that is co-moving with the rim
ζ(x, t) = x − s(t) is introduced. Thereby, s(t) denotes the position of the three-phase
contact line; ṡ stands for the dewetting velocity V as described in the next sections.
Fetzer et al used a normal modes ansatz δh(ζ) = δh0 exp kζ and u(ζ) = u0 exp kζ in
the linear stability analysis, which leads to a characteristic polynomial of third order.
Depending on the ratio of slip length to film thickness b/h and on the capillary number
Ca = η ṡ/γlv , the parameter k obtains complex or real solutions. Fetzer and coworkers
successfully identified the morphological transition from oscillatory to monotonically
decaying rims and were able to extract slip lengths and capillary numbers from diverse
experiments on dewetting surfaces [75, 76, 64, 77].
Slip effects in polymer thin films
19
Figure 11. Temporal series of optical micrographs showing the growth of a hole in
a 120 nm PS(13.7 kg/mol) film at T = 120 ◦ C prepared on a Si wafer covered with
a 21 nm hydrophobic Teflonr coating (AF 1600). In the last stage of hole growth
(right image) perturbations of the three-phase contact line (according to the liquid
rim instability) become visible.
3. Flow dynamics of thin polymer films - Experimental studies and
theoretical models
One of the main aspects of experimental studies concerning the flow dynamics of thin
polymer films is to obtain a comprehensive view on the molecular mechanisms of slippage
and on the responsible parameters. Although these insights are rather indirect, several
models have been proposed to explain diverse experimental results. In this section, we
will focus on these studies, with special regard to the proposed mechanisms of slippage
at the solid/polymer interface. In general, we have to distinguish two different dewetting
geometries: growth of holes (cf. Fig. 11) and receding straight fronts. On the one hand
driving forces and on the other hand dissipation mechanisms have to be considered.
3.1. Dewetting dynamics - Driving forces
According to the description of the effective interface potential in section 2.5.4, a global
minimum of φ(h) occurs at heq in case of unstable or metastable films. This means that
the film will thin until a thickness of heq is reached. In other words, a thin wetting
layer remains on top of the substrate, if heq > 0 and if heq has a size that is not below
the size of the molecules ‡. After dewetting has taken place (the dynamics of which
is not covered by φ(h) but depends on viscosity, viscoelasticity, contact angle and the
solid/liquid boundary condition), single droplets remain on top of the wetting layer.
The droplets - in equilibrium - exhibit the Young’s contact angle θY . Parallel to the
Young’s equation that characterizes the contact angle via the involved surface tensions,
another characteristic parameter can be defined to specify the wettability of a surface
by a liquid. This is the so-called spreading parameter S given by
S = γsv − (γsl + γlv + φ(heq )),
(3.1)
‡ Note that the concept of the effective interface potential is a continuum approach. It may fail if
molecular sized phenomena are to be captured, such as extremely thin films close to the size of the
molecules.
Slip effects in polymer thin films
20
where γij denotes the interface tension of the solid (s), liquid (l ) and the vapor (v )
phase. S describes the energetic difference per unit area between a dry surface and a
surface that is covered by a liquid layer. Consequently, if S < 0 the system can gain
energy by reducing the contact area with the substrate. Thereby, the liquid forms a
√
dynamic contact angle θd = θY / 2 at the three-phase contact line [81].
The link between the contact angle in equilibrium and the depth of the minimum of
the effective interface potential becomes obvious, if the equation for the Young’s contact
angle is inserted in (3.1):
φ(heq ) = γlv (cos θY − 1) = S
(3.2)
Driving forces for wetting and dewetting are capillary forces (see [78]). Thereby,
the capillary force per unit length of the contact line is in general given by
Fc
= γlv (cos θY − cos θd ).
(3.3)
l
As long as θd < θY remains valid, the capillary force is negative and the three-phase
contact line will recede; the liquid film dewets. Dewetting ends as soon as droplets
exhibit their Young’s contact angle θY on the surface. During dewetting, a rim which
is formed from accumulated material and retracts from the substrate while further
growing. De Gennes developed a theoretical description of the resulting driving force
for dewetting [78]:
On the ”dry” side of the rim, which denotes the side where the three-phase contact
line is located, a negative capillary force pulls at the contact line:
Fc
= S + γlv (1 − cosθd ).
(3.4)
l
On the other side, where the rim decays into the liquid film (the ”wet” side of the rim),
a positive capillary force occurs:
Fc
= γlv (1 − cosθd ).
(3.5)
l
We have to remark that in experimental systems, the Laplace pressure of course avoids
edges and therefore forces the dry side of the rim to decay smoothly into the film.
Summing up both forces per unit length gives the effective driving force for retracting
contact lines in dewetting systems:
Fc
1
= |S| = γlv (1 − cos θY ) ≃ γlv θY .
(3.6)
l
2
Thereby, it is assumed that the system is in a quasi-stationary state, which means that
changes in the shape of the rim occur much slower that the dewetting velocity. All in
all, this result implies that the driving force for dewetting is given by the absolute value
of the spreading parameter, which solely depends on the surface tension γlv of the liquid
(a property which is purely inherent to the liquid) and the Young’s contact angle of
the liquid on the surface. The same conclusion is drawn if considering energies instead
of driving forces. The depth of the minimum of the effective interface potential φ(h)
is equal to S and gives the energy per unit area that is set free during dewetting [79].
Thereby, S also stands for the force per unit length of the contact line (see (3.6)).
Slip effects in polymer thin films
21
3.2. Dewetting dynamics - Dissipation mechanisms
In the previous section, we focused on the driving forces for dewetting. The
experimentally observed dynamics represents a force balance of driving forces and
friction forces. The resulting dewetting velocity V is connected to the spreading
parameter S and the occurring energy dissipation mechanisms (Fi gives the friction
force per unit length of the contact line) and their corresponding velocity contributions
vi via the power balance
X
|S|V =
Fi vi .
(3.7)
i
In case of dissipation due to viscous friction within the liquid (Fv ) and dissipation due
to friction of liquid molecules at the solid/liquid interface (Fs ), i.e. slippage with a
finite velocity vs = u|z=0, we end up with the following balance (the index v stands for
’viscous’, s denotes ’slip’):
|S|V = Fv vv + Fs vs .
(3.8)
3.2.1. Viscous friction According to the work of Brochard-Wyart et al [80], the no-slip
boundary condition for fluid flow implies a friction force that is proportional to the liquid
viscosity and the dewetting velocity v and means that the dissipation is solely due to
viscous friction within the liquid. The largest strain rates occur in the direct vicinity of
the three-phase contact line. The dissipation is therefore mainly independent from the
size of the dewetting rim. However, the flow geometry (given by the dynamic contact
angle θd ) at the contact line influences viscous dissipation. In case of pure viscous flow
V = vv , the following expression for the velocity contribution is obtained for viscous
flow [80]:
vv = Cv (θd )
|S|
.
η
(3.9)
Cv (θd ) denotes the constant of proportionality and displays a measure for the flow field
in the vicinity of the contact line. According to the description of Redon et al [81], the
velocity in case of viscous dissipation strongly depends on θY and can also be written
as
γlv 3
vv ∝
θ ,
(3.10)
η Y
where the constant of proportionality accounts for the flow singularity near the contact
line. In detail, this constant represents a logarithmic factor that includes i.a. a shortdistance cutoff and accounts for the fact that viscous dissipation diverges near the
contact line. Moreover, a small impact of the slip length b on this logarithmic factor
has been experimentally found [81] by variation of molecular weight Mw (cf. (3.12)) by
Redon [82]. Consequently, only in the case of a no-slip boundary condition, the constant
of proportionality in (3.10) and also Cv (θd ) in (3.9) is purely independent of slip.
Slip effects in polymer thin films
22
H0
w/2
R or D
w/2
x0
h
Figure 12. Schematic representation of a rim cross-section illustrating the rim width
w (usually obtained as the distance between the three-phase contact line and the
position where the rim height is dropped to 1.1 h), the distance x0 for a given film
thickness h and hole radius R (or dewetted distance D of a straight front). (Height
scale in sketch is exaggerated as compared to the lateral length scale.)
For growing holes of radius R and dewetting velocity V = dR/dt, integration of
(3.9) gives a linear proportionality of the radius versus dewetting time t:
R∝t
(3.11)
3.2.2. Slippage - Friction at the solid/liquid interface Besides viscous dissipation within
the rim, de Gennes expected a long-chained polymer film to exhibit an exceptional
amount of slippage. For entangled polymer melts on a non-adsorbing surface, de Gennes
stated that the slip length should strongly increase with the chain length of the polymer
[83]:
N3
b = a 2,
Ne
(3.12)
where a is the size of the monomer, N the polymerization index and Ne the entanglement
length. In case of dominating slippage, a minor contribution of viscous dissipation is
expected due to very small stresses within the rim. The corresponding analytical model
(see [80, 82]) is based on the linear friction force per unit length of the contact line given
by Fs ∝ ξvs . Dissipation occurs along the distance of the solid/liquid interface, where
liquid molecules are moved over the substrate. In general, this distance is identified as
the width w of the rim (cf. Fig. 12).
Consequently, Fs ∝ w can be assumed. If the slip length is introduced according
to Navier’s model by b = η/ξ (see (2.21) in section 2.4), the slip velocity contribution
vs is given by:
vs =
1 |S| b
.
3 η w
(3.13)
Let’s consider the conservation of mass for a growing hole of radius R,
π(R + x0 )2 h = 2πQ(R + w/2),
(3.14)
Slip effects in polymer thin films
23
and for a straight dewetting front of dewetted distance D,
(D + x0 )lh = Ql,
(3.15)
where x0 stands for the distance as depicted in Fig. 12, Q denotes the cross sectional
area of a rim (approximated as a half-circle §, i.e. Q = π(x0 /2)2 ≈ π(w/2)2), and l
is the length of a straight front, which cancels out. Based on the assumption of selfsimilar growing
√ √ rims and the fact that x0 /R ≪ 1 and x0 /D ≪ 1, the width of the
rim w ∝ h R for both geometries. The constant of proportionality, however, which
we name Cs in the following, depends on the shape of the rim and furthermore on the
dewetting geometry itself:
√ √
w = Cs h R.
(3.16)
Values for Cs can be obtained collecting a temporal series of AFM snapshots of rim
profiles and fitting (3.16) to the measured rim width w plotted versus the hole radius
R. The initial film thickness h can also be measured by AFM. Of course, the validity
of this simplified model and especially the approximation w ≈ x0 becomes questionable
for asymmetric rims and should lead to systematic differences in Cs . We point out that,
as described in section 2.5.5, the degree of asymmetry of the cross-section of the rim is
strongly correlated to the capillary number and the ratio of slip length to film thickness.
Replacing the above derived relation of rim width w and hole radius R in (3.13) gives
the slip velocity contribution in terms of the hole radius R for purely slipping liquids:
vs =
1
1 |S| b
√ √ .
3 η Cs h R
(3.17)
Via integration, a characteristic growth law for the radius R of a dewetting hole with
time t is found (separation of variables),
R ∝ t2/3 .
(3.18)
Using the lubrication models described in section 2.5, Münch compared numerical
simulations of polymer melts dewetting from hydrophobized substrates to the above
described dewetting dynamics obtained from scaling arguments (based on energy
balances) [84]: In the no-slip situation (mobility m(h) ∝ h3 ), an exponent α = 0.913
is reported instead of 1. The deviation is traced back to the fact that the logarithmic
factor in the constant of proportionality of equation (3.10) also depends on the width of
the rim (which evolves with time t). Numerical simulations of the slip-dominated case
(m(h) ∝ h2 ) give α = 0.661, which captures α = 2/3 very well.
§ Note that usually a dynamic contact angle θd < 90 ◦ at the three-phase contact line is obtained in
experiments. The difference between a segment of a circle and the half-circle approximation is marginal
and contributes to the constant of proportionality Cs .
Slip effects in polymer thin films
24
Figure 13. Left: Optical measurement of the hole radius versus time in 130 nm
PS(13.7 kg/mol)
films on different substrates. Right: Plot of the dewetting velocity V
√
versus 1/ R. The y-axis intercept is identical for all substrates. The difference in the
slope K indicates substantial differences in slip for dewetting PS (adapted from [86]
and [87]).
3.2.3. Models based on the superposition of dissipation mechanisms Considering the
models derived before for pure viscous flow and for pure slippage, a combination of
these models seems to be reasonable for situations, where viscous dissipation as well as
dissipation at the solid/liquid interface act together. Jacobs et al proposed an additive
superposition of the corresponding velocity contributions, i.e. V = vs +vv [85], according
to the fact that both mechanisms counteract the same driving force |S|. Separation of
variables leads to the implicit function:
√
Ks R
Kv 2
Kv √
Kv
))
(3.19)
R + 2( ) ln (1 +
(R − 2
t − t0 =
|S|
Ks
Ks
Kv
with
Kv =
η
,
Cv (θd )
Ks =
3η w
√ ,
b R
√
Kv
∝ b R.
Ks
(3.20)
In that
√ notation, the velocity contributions are given by vv = |S|/Kv and vs =
|S|/( RKv ). One of the preconditions of this superposition was the successful inclusion
of both border cases: For b → 0, R ∝ t is obtained, whereas for b → ∞, R ∝ t2/3 follows.
As a more practicable alternative for fitting this relation between radius R and time
t to experimental data (typical data is shown in Fig. 13), Fetzer and Jacobs recently
proposed a more simple way of visualizing slip effects [86]: According to the additive
superposition, the dewetting velocity can be written in terms of
K
V = vv + √ ,
R
K=
1 |S| b
√ .
3 η Cs h
(3.21)
√
which leads to a linear relationship when plotting the dewetting velocity v versus 1/ R
(c.f. Fig. 13). Then, the y-axis intercept of the straight line can be identified as
the viscous velocity contribution vv , whereas the the slope K is connected to the slip
Slip effects in polymer thin films
25
length as illustrated in (3.21). The validity of this model, which assumes the linear
superposition of velocity contributions, was checked by plotting the viscous velocity
contribution vv versus the reciprocal viscosity η. An excellent agreement has been
demonstrated in view of melt viscosity data obtained from independent measurements.
Moreover, the slope has been used to calculate the slip length b. In the experiments,
substantial differences in slip lengths for different substrates (cf. Fig. 13) at identical
liquid properties could be detected [86].
3.2.4. Molecular-kinetic theory and further approaches Besides theoretical models
based on continuum hydrodynamics (see 3.2.1), contact line dynamics such as the
spreading of a liquid droplet on a planar surface has been analyzed according to the
so-called molecular-kinetic theory (MKT) [88, 89, 90]. The movement of a contact
line is described as an activated rate process, where the liquid molecules close to the
substrate jump from one potential well to another. The resulting contact line friction
coefficient is proportional i.a. to the viscosity η of the liquid and to the exponential
of the reversible work of adhesion γlv (1 + cos θY )/kB T . As a consequence, increasing
the temperature reduces the friction coefficient and increases the velocity. In case of
a squalane droplet spreading on gold surfaces covered with different self-assembled
monolayers (SAM) of thiols, the corresponding contact line friction coefficient turns
out to increase linearly with the chain length of the SAM [91]. SAM surfaces have also
been target of friction experiments using AFM tips. Barrena et al explained discrete
changes in the frictional behavior with discrete molecular tilts of the chains of the SAM
[92]. Viscoelastic deformation of the substrate at the contact line due to the normal
component of the liquid surface tension has also attracted attention with regard to fluid
dynamics on the nanoscale and energy dissipation. The phenomenon of a reduced
velocity of contact-line displacements compared to a corresponding non-deformable
substrate (termed ”viscoelastic braking”) was extensively studied and described by
Shanahan and Carré [93, 94, 95, 96, 97]. In this context, Long et al theoretically
examined the static and dynamic wetting properties of liquids on thin rubber films
[98] and grafted polymer layers [99]. To the best of our knowledge, there is to-date
no experimental evidence of deformations and tilting on the molecular scale in case of
SAM surfaces due to retracting contact-lines of a dewetting polymer film. However,
dewetting on soft deformable rubber surfaces has recently been shown to be influenced
by viscoelastic deformations of the substrate [100]. In case of non-volatile liquids such
as polymers, evaporation into and condensation from the vapor phase can be excluded
as relevant mechanism of energy dissipation.
3.3. Dynamics of growing holes
Besides very early theoretical and experimental studies of thin film rupture [5, 101,
102, 103, 104, 105, 106], one of the first studies concerning dewetting experiments had
been published by Redon et al [81]. Films of alkanes and polydimethylsiloxane (PDMS)
Slip effects in polymer thin films
26
have been prepared on different hydrophobized silicon wafers. The authors observed a
constant dewetting velocity which was inversely proportional to the viscosity and very
sensitive to the equilibrium contact angle of the liquid on a corresponding surface. In
1994, Redon et al investigated the dewetting of PDMS films of different thicknesses
on silanized hydrophobic surfaces [82]. They found that for films larger than 10 µm a
constant dewetting velocity is found, whereas R(t) ∝ t2/3 in films thinner than 1 µm
is observed. These results corroborate the models represented by (3.11) and (3.18) and
indicate that for sufficiently thick films (h ≫ b) viscous dissipation dominates whereas
slippage becomes relevant in case of thin films (h < b) [82].
3.3.1. Stages of Dewetting While studying the growth of holes, Brochard-Wyart et
al proposed a set of subsequent stages that can be attributed to distinct growth
laws [107]: Starting with the birth of the hole, the radius of the hole is supposed
to grow exponentially
√ with time, i.e. R(t) ∝ exp (t/τ ), as long as the radius R is
smaller than Rc ≈ bh. Experimentally, Masson et al [108] found relaxation times
orders of magnitude larger than the largest translational reptation times expected. For
Rc < R < Rc′ ≈ b, the rim is formed and viscous dissipation
√ dominates the hole growth
dynamics, i.e. R ∝ t. The special case of R ≫ Rc = bh moreover gives an analytic
expression for the hole growth:
|S| b 1/2
( ) t
(3.22)
R(t) ≈
η h
√
During this linear regime, a dynamic (receding) contact angle θd = θY / 2 is formed at
the three-phase contact line. In the subsequent stage (for hole radii approximately larger
than the slip length b), the rim is fully established and grows in a self-similar manner.
As discussed in section 3.2.1, this results in a characteristic R ∝ t2/3 growth law if large
slip is present. That stage of self-similar growing rims is often called ”mature” regime.
These distinct regimes have been experimentally observed by Masson et al [108] and
Damman et al [72]. Moreover, a dissipation dominated by slippage has been shown to
be restricted to sufficiently small hole radii: If the rim has accumulated a very large
amount of liquid (so that the height of the rim H0 is much larger than the slip length
b, i.e. H0 ≫ b), viscous dissipation dominates again. This consequently results in a
transition to a linear growth law for the radius with respect to the dewetting time, i.e.
R ∝ t.
In this article, we mainly focus on dewetting phenomena in terms of growing holes
or retracting straight fronts. Additionally, also the subsequent regime influenced by the
fingering instability has attracted the interest of researchers [109]. Thereby, facets such
as the onset of the instability and the morphology of liquid profiles have also shown
to be sensitive to rheological properties of thin films and the hydrodynamic boundary
condition at the solid/liquid interface [110, 111, 112, 113].
3.3.2. ”Mature” holes Recently, Fetzer et al experimentally studied the dewetting
dynamics of ”mature” holes in thin PS films on two different types of hydrophobized
Slip effects in polymer thin films
27
surfaces. The aforementioned assumption of linear superposition of viscous dissipation
and slippage (see 3.2.3) was used to gain information about the slip contribution [86].
Hydrophobization was achieved by the preparation of a dense, self-assembled monolayer
of silane molecules (silanization). In that case two different silanes have been utilized,
octadecyltrichlorosilane (OTS) and dodecyltrichlorosilane (DTS). The slip length turns
out to depend strongly on temperature (and thereby on the melt viscosity); in fact, the
slip length decreased with increasing temperature. Furthermore, slippage was about one
order of magnitude larger on DTS compared to OTS. The results are in good agreement
with the results for the slip length gained by the analysis of the shape of rim profiles of
the corresponding holes (see 2.5.5). The molecular mechanisms of slippage are widely
discussed and several models have been proposed with regard to different experimental
conditions (shear rates, polymers, surfaces, ...). This will be discussed in the next
section.
3.4. Molecular mechanisms of slippage
For small flow velocities, slip lengths lower than expected have been found (see e.g.
[114]). This has been explained by the adsorption of polymer chains at the solid
surface. These adsorbed chains are able to inhibit slippage due to entanglements with
chains in the melt. At larger shear strains, disentanglement of adsorbed and melt
chains occurs, the friction coefficient decreases rapidly and slippage is ”switched on”
[115, 116, 117, 118]. The authors interpret the results as follows: Strongly adsorbed
chains stay at the solid/liquid interface followed by a coil-stretch transition of these
tethered chains. Thereby, the interaction of anchored chains (which are stretched under
high flow velocities) with the chains in the melt are reduced and slippage is enhanced.
This has been experimentally observed by Migler et al [114] and Hervet et al [119].
Thereby, the critical shear strain depends on the molecular weight of chains attached to
the surface and on their density [39]. This situation is also called ”stick-slip transition”
and has moreover been reported for pressure-controlled capillary flow of polyethylene
resins [120]. The authors find a characteristic molecular weight dependence of the slip
length b ∝ Mw3.4 , whereas the critical stress σc for the transition described before scales
as σc ∝ Mw−0.5 [121]. Also de-bonding of adsorbed chains might occur.
Molecular dynamics (MD) studies aiming to investigate the slip length in thin,
short-chained polymer films subject to planar shear revealed a dynamic behavior of
the slip length b(γ̇) upon the shear rate according to b = b∗ (1 − γ̇/γ̇c )−1/2 , where γc
represents a critical shear rate [122]. Slippage strongly increases as γ̇/γ̇c → 1. This
relation was also found for simple liquids. Additionally, the authors studied the border
case γ̇ → 0 and especially the molecular parameters affecting the slip length.
Moreover, a special situation has been discussed in the literature: A thin polymer
film prepared on a surface decorated with end-grafted polymer chains of the same
species. Due to entropic reasons, an interfacial tension between identical molecules
occurs and dewetting can take place. This phenomenon is called autophobicity or
Slip effects in polymer thin films
28
autophobic dewetting [123, 124]. Reiter and Khanna observed that PDMS molecules
slipped over grafted PDMS brushes. They determined a slip length on the order of
10 µm for film thicknesses between 20 and 850 nm of the dewetting PDMS film (high
Mw of 308 kg/mol in most cases). For lower grafting densities, slippage is reduced,
indicating a deeper interpenetration of free melt chains [123].
3.5. Impact of viscoelasticity and stress relaxation
As already introduced in 2.1.3 and furthermore theoretically described in section 2.5,
viscoelasticity can be included in the description of thin film dynamics. In case of thin
film flow with strong slippage, we already mentioned that the impact of viscoelasticity
might be not distinguishable from slippage effects. Numerous experimental studies have
been published by Reiter, Damman and coworkers describing the influence of viscoelastic
effects on the flow dynamics of thin polymer films. These studies have been supported
with theories by Raphaël and coworkers. In the following, a snap-shot of recent results
in chronological order is presented. It reflects the progress from simple to more and
more sophisticated models and experiments. Some of the results, however, may have
tentative character, as brought up by a very recent manuscript by Coppée et al (cf.
section 3.7).
3.5.1. Thin film rupture Reiter and de Gennes [125] pointed out that the usual
spin-casting preparation (based on fast solvent evaporation) of thin, long-chained
polymer films may induce a cascade of distinct states. Annealing the samples to high
temperatures for times larger than the reptation time of the polymer (depending on
Mw and the temperature T ) might induce a complete healing of the preparation effects.
Non-equilibrated conformational states and residual stresses have shown to be capable
to cause rupture of thin films [126]. Thereby, the areal density of holes appearing at a
certain temperature above Tg has been measured: After storage at elevated temperatures
below Tg , an exponential reduction of the hole density can be observed. Vilmin and
Raphaël were able to show that lateral stress reduces the critical value for surface
fluctuations initiated by an anisotropic diffusion of polymer molecules to induce the
formation of holes [127].
3.5.2. Stages of distinct dewetting dynamics - Experimental results Reiter and coworkers furthermore studied the temporal and spatial evolution of dewetting PS fronts
on Silicon wafers covered with a PDMS monolayer at times shorter than the relaxation
(reptation) time τrept in equilibrated bulk samples [126]. At the beginning of their
experiments, the dewetted distance and the rim width w increased in a logarithmic way
with respect to the dewetting time t, consequently V ∝ t−1 is found. They correlated
a maximum of the rim width w occurring at a distinct time t1 (nearly independent on
Mw and significantly shorter than τrept ) of the experiment to a change in dewetting
√
dynamics, namely V ∝ t−1/2 (i.e. R ∝ t).
Slip effects in polymer thin films
29
Even before, Damman, Baudelet and Reiter compared the growth of a hole and the
dynamics of a dewetting straight front [72]. Thereby, strong influence of the dewetting
geometry on the dynamics of early stages became obvious. Moreover, the authors
correlated their findings to the shape of the corresponding rim and clearly identified
distinct dewetting stages: At the beginning of hole growth, capillary forces dominate
the dynamics and exponential growth is observed. In contrast to that, dewetting of
a straight front at the beginning starts almost instantaneously at a high velocity and
decreases as V ∝ t−1 . As already observed by Redon et al [81], V = const and R ∝ t
due to viscous dissipation is found for growing holes. For both dewetting geometries,
Damman et al find a maximum of the rim width w which they interpret in terms of a
transition of the shape of the rim from very asymmetric profiles towards more symmetric
rims, while the volume still increases during dewetting. In the following dissipationdominated regime, V ∝ t−1/3 and R ∝ t2/3 (slip boundary condition) or V = const and
R ∝ t (no-slip boundary condition) is found, depending on the boundary condition for
the solid/liquid interface. Afterwards, again a constant dewetting velocity (until the
”mature rim” regime is reached and slippage dominates) is obtained.
Recently, Damman and Reiter determined the strain γ from rim shapes of dewetting
fronts, which can be written according to (2.3) by γ = (S/h + σ0 )/G, where σ0 denotes
residual stresses [128]. These values were one order of magnitude larger than for
equilibrated PS films (γ = S/(hGbulk )) and increased with increasing Mw . Consequently,
larger residual stresses σ0 or smaller elastic moduli G can be responsible for this effect.
Indeed, the elastic modulus of a thin film is supposed to be smaller than its bulk value
according to the fact that larger numbers for the entanglement length Me (for larger Mw )
are expected (see section 2.1.2). Additionally, instead of one transition time for both,
two different transitions times tw for the maximum of the rim width and tv for the power
law of the dewetting velocity (to V ∝ t−1/3 ) are found for larger molecular weights. Time
tv is comparable to the reptation time trept (also for larger Mw ). The authors interpret
this as follows: After tv ∼ trept, the equilibrium entanglement conformation has been
reached via interdiffusion and re-entangling of chains. Relaxation of stress, which is
correlated to the rim shape transition and tw , occurs at shorter times.
3.5.3. Stages of distinct dewetting dynamics - Theoretical models These observations
have motivated a theoretical model by Vilmin and Raphaël [129], that takes viscoelastic
properties of the liquid and slippage into account. The logarithmic increase of the
dewetted distance with time can be explained if residual stress is regarded as an extra
driving force for dewetting that has to be added to the capillary forces. Subsequent
relaxation of the stress reduces this additional contribution. For times larger than the
reptation time τrept , Vilmin and Raphaël predict a constant dewetting velocity and
V ∝ t−1/3 as soon as the ”mature regime” is reached (and slippage becomes important).
They consider a simple equation for a viscoelastic film which includes one relaxation
time τ1 (and an elastic modulus G), but two distinct viscosities: η0 is a short-time
viscosity which describes the friction between monomers, and η1 is the melt viscosity
Slip effects in polymer thin films
30
(η1 ≪ η0 ) governed by disentanglements of polymer chains. Theoretically, three regimes
of distinct time-response are expected: For t < τ0 = η0 /G, Newtonian flow accompanied
with a small viscosity η0 . Afterwards (τ0 < t < τ1 = η1 /G), the elastic modulus governs
the dynamics. For longer times, i.e. t > τ1 , Newtonian flow is again obtained, but
with a much larger viscosity η1 . During the first and the last regime, asymmetric rims
and constant velocities are predicted (the latter velocity of course is much smaller than
the initial velocity). The intermediate regime, governed by viscoelastic behavior, is
expected to show V ∝ t−1/2 according to the fact that the width of the rim w will
increase proportional to the dewetted distance D, i.e. w ∝ D (in contrast to the squareroot dependence in case of viscous flow). Including residual stresses into their model,
Vilmin and Raphaël obtained the aforementioned experimentally observed V ∝ t−1 law
instead of V ∝ t−1/2 . Recently, Yang and co-workers experimentally quantified the
molecular recoiling force stemming from non-equilibrium chain conformations. They
obtained values that are at least comparable to (or even larger than) the dispersive
driving forces [130].
3.5.4. Further remarks Concerning studies and models including viscoelastic effects,
one has to bear in mind that stress relaxation dynamics and relaxation times in thin films
have been shown to be different from bulk polymer reptation values. As mentioned in
section 2.1.2 and indicated by the characteristic relaxation times τ1 of elastic constraints
being significantly shorter than bulk values, the entanglement length has shown to be
significantly increased in case of thin films.
Recently, Reiter and coworkers extended their studies with regard to viscoelastic
dewetting on soft, deformable substrates [100]. The essential result was that transient
residual stresses can cause large elastic deformations in the substrate which almost stop
dewetting for times shorter than the relaxation time τrept of the polymer film. For times
longer than τrept , the elastic behavior and the elastic trench in the deformable substrate
vanishes.
Vilmin and Raphaël applied their model for viscoelastic liquids and residual stresses
to the hole growth geometry [131] with regard to the early stage (exponential growth
for Newtonian liquids). They discovered a very fast opening regime followed by a slow
exponential growth of the radius of the holes.
3.6. Non-linear friction
Up to now, concerning energy dissipation at the solid/liquid interface, exclusively linear
friction has been considered. As described in section 3.4, often not only smooth and
passive (non-adsorbing) surfaces have been experimentally considered, but also grafted
or adsorbed polymer layers are used as a support for dewetting experiments. These
systems motivate the theoretical treatment of non-linear friction, as described in the
following subsection.
Slip effects in polymer thin films
31
3.6.1. Theoretical model To cover non-linear friction, Vilmin and Raphaël introduced
a friction force per unit area that is linear below a certain transition velocity vr and
non-linear above [131]:
Fs = ξvr (v/vr )1−r ,
v > vr ,
(3.23)
where r denotes a so-called friction exponent.
Consequently, the effective (velocity-dependent) friction coefficient ξef f can be written
in terms of
ξef f (v) = ξ(vr /v)r
(3.24)
and a velocity-dependent slip length b(v) can be defined as
b(v) = η/ξef f (v).
(3.25)
An increasing velocity leads to a decreasing friction coefficient and to pronounced
slippage. Thereby, the friction at the solid/liquid interface influences the power law
of the velocity decrease. Assuming non-linear friction in the intermediate regime (which
is governed by the viscoelastic behavior) gives (according to w ∝ D) V ∝ t−1/(2−r) .
Including residual stress σ0 to this model leads to a formula for the maximum width
of the rim (depending on r and σ0 ). Experimentally, values for r between 0 (for low
Mw ) and 1 (for large Mw ) have been found [128]. The dependence of the nonlinearity of
friction upon molecular weight Mw could be explained by the influence of chain length
on slippage (see (3.12)).
3.6.2. Variation of substrate properties Hamieh et al focused on the frictional behavior
of dewetting viscoelastic PS films on PDMS-coated (irreversibly adsorbed) silicon wafers
[132]. Thereby, thickness (via the PDMS chain-length) and preparation of the PDMS
support (via the annealing temperature) were varied. In summary, the observations
are consistent with the aforementioned (section 3.5) experiments by Reiter, showing a
characteristic time t1 for the change in dewetting dynamics and rim shape (from highly
asymmetric towards a more symmetric equilibrated shape). Again, this transition is
interpreted by the Laplace pressure that overcomes elastic effects. Probing the maximum
rim width enables the authors to identify the impact of the friction coefficient or the
friction exponent r: If they prepare thicker PDMS layers, the result is a larger maximum
rim width. This result can be explained in terms of a small increase of r. Consequently,
the velocity-dependent slip length b(v) increases (see (3.25)): Thicker PDMS-layers
lead to more slippage. Concerning the characteristic time for stress relaxation t1 , no
significant influence of the preparation procedure (annealing temperature of the PDMS
coating) and thickness of the PDMS layer have been found.
Slip effects in polymer thin films
32
3.7. Temporal evolution of the slip length
Most of the aforementioned dewetting experiments of thin (viscoelastic) PS films are
based on supports consisting of a PDMS layer prepared onto a Si wafer. These PDMS
surfaces, which were assumed to be impenetrable by PS chains, turned out to be
less ideal. Recently, the aforementioned fast decay of the dewetting velocity and the
maximum of the rim width upon dewetting time (ascribed to relaxation of residual
stresses, see section 3.5), has also been observed in case of low molecular weight PS [133].
Furthermore, neutron reflectometry experiments revealed an interdiffusion of polymer
chains at the PS-PDMS interface below the PS bulk glass transition temperature. The
dewetting velocity accelerated as the brush thickness is increased. In view of energy
dissipation due to brush deformation (pronounced by increasing thickness), this is
contradictory to the slower dewetting velocities expected in that case.
Ziebert and Raphaël recently investigated the temporal evolution of the energy
balance (viscous dissipation and sliding friction) of thin film dewetting by numerical
treatment, especially concerning non-linear friction [134]. They point out that both
mechanisms have different time dependencies and propose that simple scaling arguments
such as the mass conservation w ∝ D should be revisited. In case of non-linear friction,
viscous dissipation is even more important than sliding friction for times larger than τ1 ,
whereas for linear friction both mechanisms are approx. equally important. The stage
of ”mature” rims afterwards is again dominated by friction at the solid/liquid interface.
Using scaling arguments and numerical solutions of the thin film model for viscoelastic
liquids [129], they moreover showed that the aforementioned dewetting characteristics
(fast decay of the dewetting velocity, maximum of the rim width in course of dewetting
time) can be explained by the temporal decrease of the slip length during the experiment
instead of stress relaxation [135]. Therefore, roughening of the PS-PDMS interface (as
detected by neutron reflectometry) or potentially also an attachment of very few PS
chains to the silicon substrate might be responsible, if very low PDMS grafting densities
are prepared.
The attachment of melt chains to the substrate has been recently considered by
Reiter et al [136]. They pointed out that the driving force is reduced by a certain pullout force for molecules attached to the surface as they get stretched while resisting to be
pulled out. Consequently, this force (per unit length of the contact line) can be written
in terms of Fp = νLf ∗ , where ν represents the number of surface-connected molecules,
L the length according to the stretching and f ∗ the pull-out force per chain.
4. Conclusions and outlook
To conclude, dewetting experiments can be regarded as a very powerful tool to probe
rheological thin film properties and frictional mechanisms. The validity of the no-slip
boundary condition at the solid/liquid interface, for a long period of time accepted as a
standard approach in fluid dynamics, fails. Even for Newtonian liquids such as polymers
Slip effects in polymer thin films
33
below their critical length for entanglements substantial amount of slippage has been
found. For larger molecular weights, slippage can be even more pronounced if chain
entanglements come into play.
We have reviewed in this article theories and experiments characterizing the statics
and dynamics in thin liquid polymer films, focussing on energy dissipation mechanisms
occurring during dewetting. Experiments are relatively easy to perform, since usually
no very low or very high temperatures are needed, neither are high-speed cameras
necessary. However, careful preparation, preferably in a clean-room environment, of
thin films is inevitable. Concerning the experimental data, diligent interpretation and
consideration of all relevant parameters are of essential importance: Molecular weight,
film thickness (also with regard to the molecular dimensions of a polymer coil in the
respective melt), dewetting temperature, melt viscosity, dewetting velocity and capillary
number, residual stresses, relaxation times and ageing time represent parameters that
are sometimes coupled and not easy to disentangle. Their unique impact on the friction
coefficient or the slip length on a specific, ideally non-adsorbing and non-penetrable
substrate (to reduce the number of parameters of the system) is not always easy to
identify. In particular, the specific stage of dewetting (early or mature stage) and the
dewetting geometry (holes or straight fronts) represent further facets and playgrounds
for experimentalists and theoreticians that, under careful consideration, enable to gain
insight into rheological or frictional mechanisms. The dynamics and the morphology
of the fingering instability can provide additional access to the solid/liquid boundary
condition and the rheology.
To obtain a deeper understanding of the molecular mechanisms at the solid/liquid
interface is one of the main tasks in micro- and nanofluidics. In this context, we like to
highlight two essential questions and possible pathways to answer them:
a) The liquid: What is the impact of the polydispersity of the liquid on dewetting?
Dewetting studies of polymer melts by adding a second chemical component have shown
to influence slippage [137]. However, also the driving force is changed due to the
difference in chemical composition for different species. To overcome this problem,
the influence of chain length distribution on dewetting can be probed by studying
polymer mixtures instead of monodisperse polymer melts [138]. Concerning a theoretical
approach to this question, dissipative particle dynamics (DPD) simulations enable to
locate energy loss in the rim and reveal interesting results in case of two immiscible fluids
of different viscosity: In case of a low-viscosity layer at the solid/liquid interface, faster
dewetting dynamics is found that is attributed to a lubrication effect, i.e. the sliding of
the upper high-viscosity layer [139]. Finally, these aspects lead to the fundamental
discussion, whether ”apparent” slip, possibly induced by the formation of a shortchained layer of low viscosity, is present.
b) The substrate: What is the impact of the molecular structure of the substrate?
As indicated before, the set of parameters concerning the topographical and/or chemical
structure of the support is large. Besides parameters such as surface roughness and
surface energy, experiments can be performed on substrates e.g. decorated with
Slip effects in polymer thin films
34
an amorphous coating, a self-assembled monolayer or ever grafted polymer brushes
of the same or different species. Scattering techniques (using neutrons or X-rays)
provide access to the solid/liquid interface, complementing dewetting experiments and
confirming proposed mechanisms of slippage. Simulations based on molecular dynamics
(MD) of near-surface flows can help to compare experimental results from dewetting
studies to molecular parameters, easily tunable in theoretical models, and structural
changes [140, 141]. Moreover, the evolution of coarse-grained polymer brush/melt
interfaces under flow has also been identified as a potential application of MD studies
[142, 143]. In the end, all these facets help to obtain a universal picture of sliding
friction which can potentially lead to surfaces precisely tailored for special microfluidic
applications.
Acknowledgments
The authors acknowledge financial support from the German Science Foundation (DFG)
under grant JA905/3 within the priority program 1165 ”Micro- and nanofluidics”.
References
[1] Squires T M and Quake S R 2005 Microfluidics: Fluid physics at the nanoliter scale Rev. Mod.
Phys. 77 977
[2] Thorsen T, Maerkl S J and Quake S R 2002 Microfluidic large-scale integration Science 298 580
[3] Leslie D C, Easley C J, Seker E, Karlinsey J M, Utz M, Begley M R and Landers J P 2009
Frequency-specific flow control in microfluidic circuits with passive elastomeric features Nature
Physics 5 231
[4] Bear J 1972 Dynamics of fluids in porous media (New York: Elsevier)
[5] Vrij A 1966 Possible mechanism for the spontaneous rupture of thin, free liquid films Discuss.
Faraday Soc. 42 14
[6] Herminghaus S, Jacobs K, Mecke K, Bischof J, Fery A, Ibn-Elhaj M, Schlagowski S 1998 Spinodal
dewetting in liquid crystal and liquid metal films Science 282 916
[7] Seemann R, Herminghaus S and Jacobs K 2001 Dewetting patterns and molecular forces: A
reconciliation Phys. Rev. Lett. 86 5534
[8] Jacobs K, Seemann R and Herminghaus S 2008 Stability and dewetting of thin liquid films
Polymer Thin Films ed O K C Tsui and T P Russell (Singapore: World Scientific)
[9] Seemann R, Herminghaus S and Jacobs K 2001 Gaining control of pattern formation of dewetting
liquid films J. Phys.: Condens. Matter 13 4925
[10] Jacobs K, Seemann R and Mecke K 2000 Dynamics of structure formation in thin films: A special
spatial analysis Lecture Notes in Physics (Statistical Physics and Spatial Statistics) ed D Stoyan
and K Mecke (Heidelberg: Springer)
[11] Craster R V and Matar O K 2009 Dynamics and stability of thin liquid films Rev. Mod. Phys.
81 1131
[12] Rubinstein M and Colby R H 2003 Polymer Physics (New York: Oxford University Press)
[13] Jones R A L and Richards R W 1999 Polymers at surfaces and interfaces (Cambridge University
Press)
[14] De Gennes 1971 Reptation of a polymer chain in the presence of fixed obstacles J. Chem. Phys.
55 572
[15] Keddie J L, Jones R A L and Cory R A 1994 Size-dependent depression of the glass transition
temperature in polymer films Europhys. Lett. 27 59
Slip effects in polymer thin films
35
[16] Mattsson J, Forrest J A and Börgesson L 2000 Quantifying glass transition behavior in ultrathin
free-standing polymer films Phys. Rev. E 62 5187
[17] Dalnoki-Veress K, Forrest J A, de Gennes P-G and Dutcher J R 2000 Glass transition reductions
in thin freely-standing polymer films : A scaling analysis of chain confinement effects J. Phys.
IV France 10 221
[18] Herminghaus S, Jacobs K and Seemann R 2001 The glass transition of thin polymer films: Some
questions, and a possible answer Europ. Phys. J. E 5 531
[19] Keddie J L and Jones R A L 1995 Glass transition behavior in ultrathin polystyrene films J. Isr.
Chem. Soc. 35 21
[20] Fryer D S, Peters R D, Kim E J, Tomaszewski J E, De Pablo J J and Nealey P F 2001
Dependence of the glass transition temperature of polymer films on interfacial energy and
thickness Macromolecules 34 5627
[21] Herminghaus S, Jacobs K and Seemann R 2003 Viscoelastic dynamics of polymer thin films and
surfaces Eur. Phys. J. E 12 101
[22] Kawana S and Jones R A L 2001 Character of the glass transition in thin supported polymer
films Phys. Rev. E 63 021501
[23] Herminghaus S, Seemann R, and Landfester K 2004 Polymer surface melting mediated by
capillary waves Phys. Rev. Lett. 93 017801
[24] Seemann R, Jacobs K, Landfester K, and Herminghaus S 2006 Freezing of polymer thin films and
surfaces: The small molecular weight puzzle J. polym. Sci. B 44 2968
[25] Si L, Massa M V, Dalnoki-Veress K, Brown H R and Jones R A L 2005 Chain entanglement in
thin freestanding polymer films Phys. Rev. Lett. 94 127801
[26] Rauscher M, Münch A, Wagner B and Blossey R 2005 A thin-film equation for viscoelastic liquids
of Jeffreys type Eur. Phys. J. E 17 373
[27] Blossey R, Münch A, Rauscher M and Wagner B 2006 Slip vs. viscoelasticity in dewetting thin
films Eur. Phys. J. E 20 267
[28] Münch A, Wagner B, Rauscher M and Blossey R 2006 A thin-film model for corotational Jeffreys
fluids under strong slip Eur. Phys. J. E 20 365
[29] Te Nijenhuis K, McKinley G H, Spiegelberg S, Barnes H A, Aksel N, Heymann L, Odell J A 2007
Non-Newtonian flows Springer Handbook of experimental fluid mechanics ed C Tropea, A L
Yarin and J F Foss (Berlin, Heidelberg, New York: Springer)
[30] Münch A, Wagner B A and Witelski T P 2005 Lubrication models with small to large slip lengths
J. Eng. Math. 53 359
[31] Navier C L M H 1823 Mémoire sur les lois du mouvement des fluids Mem. Acad. Sci. Inst. Fr. 6,
389 and 432
[32] Lauga E, Brenner M P and Stone H A 2007 Microfluidics: The no-slip boundary condition
Springer Handbook of experimental fluid mechanics ed C Tropea, A L Yarin and J F Foss
(Berlin, Heidelberg, New York: Springer)
[33] Neto C, Evans D R, Bonaccurso E, Butt H-J and Craig V S J 2005 Boundary slip in Newtonian
liquids: A review of experimental studies Rep. Prog. Phys. 68 2859
[34] Bocquet L and Barrat J-L 2007 Flow boundary conditions from nano- to micro-scales Soft Matter
3 685
[35] Joly L, Ybert C and Bocquet L 2006 Probing the nanohydrodynamics at liquid/solid interfaces
using thermal motion Phys. Rev. Lett. 96 046101
[36] Barrat J-L and Bocquet L 1999 Large slip effect at a nonwetting fluid-solid interface Phys. Rev.
Lett. 82 4671
[37] Pit R, Hervet H and Léger L 2000 Direct experimental evidence of slip in hexadecane: Solid
interfaces Phys. Rev. Lett. 85 980
[38] Cottin-Bizonne C, Jurine S, Baudry J, Crassous J, Restagno F and Charlaix E 2002 Nanorheology:
An investigation of the boundary condition at hydrophobic and hydrophilic interfaces Eur.
Phys. J. E 9 47
Slip effects in polymer thin films
36
[39] Leger L 2003 Friction mechanisms and interfacial slip at fluid-solid interfaces J.Phys.: Condens.
Matter 15 S19
[40] Zhu Y and Granick S 2002 Limits of the hydrodynamic no-slip boundary condition Phys. Rev.
Lett. 88 106102
[41] Schmatko T, Hervet H and Léger L 2006 Effect of nano-scale roughness on slip at the wall of
simple fluids Langmuir 22 6843
[42] Kunert C and Harting J 2007 Roughness induced boundary slip in microchannel flows Phys. Rev.
Lett. 99 176001
[43] Cottin-Bizonne C, Barrat J-L, Bocquet L and Charlaix E 2003 Low-friction flows of liquid at
nanopatterned interfaces Nat. Mater. 2 237
[44] Priezjev N V and Troian S M 2006 Influence of periodic wall roughness on the slip behaviour
at liquid/solid interfaces: Molecular-scale simulations versus continuum predictions J. Fluid
Mech. 554 25
[45] Joseph P, Cottin-Bizonne C, Benoı̂t J-M, Ybert C, Journet C, Tabeling P and Bocquet L 2006
Slippage of water past superhydrophobic carbon nanotube forests in microchannels Phys. Rev.
lett. 97 156104
[46] Ybert C, Barentin C, Cottin-Bizonne C, Joseph P and Bocquet L 2007 Achieving large slip with
superhydrophobic surfaces: Scaling laws for generic geometries Phys. Fluids 19 123601
[47] Steinberger A, Cottin-Bizonne C, Kleimann P and Charlaix E 2007 High friction on a bubble
mattress Nature Materials 6 665
[48] Schmatko T, Hervet H and Léger L 2005 Friction and slip at simple fluid/solid interfaces: The
roles of the molecular shape and the solid/liquid interaction Phys. Rev. Lett. 94 244501
[49] Priezjev N V, Darhuber A A and Troian S M 2005 Slip behavior in liquid films on surfaces of
patterned wettability: Comparison between continuum and molecular dynamics simulations
Phys. Rev. E 71 041608
[50] Heidenreich S, Ilg P and Hess S 2007 Boundary conditions for fluids with internal orientational
degrees of freedom: Apparent velocity slip associated with the molecular alignment Phys. Rev.
E 75 066302
[51] Cho J-H J, Law B M and Rieutord F 2004 Dipole-dependent slip of Newtonian liquids at smooth
solid hydrophobic surfaces Phys. Rev. Lett. 92 166102
[52] De Gennes P G 2002 On fluid/wall slippage Langmuir 18 3413
[53] Huang D M, Sendner C, Horinek D, Netz R R and Bocquet L 2008 Water slippage versus contact
angle: A quasiuniversial relationship Phys. Rev. Lett. 101 226101
[54] Steitz R, Gutberlet T, Hauss T, Klösgen B, Krastev R, Schemmel S, Simonsen A C, Findenegg G
H 2003 Nanobubbles and their precursor layer at the interface of water against a hydrophobic
substrate Langmuir 19 2409
[55] Doshi D A, Watkins E B, Israelachvili J N and Majewski J 2005 Reduced water density at
hydrophobic surfaces: Effect of dissolved gases PNAS 102 9458
[56] Mezger M, Reichert H, Schröder S, Okasinski J, Schröder H, Dosch H, Palms D, Ralston J and
Honkimäki V 2006 High-resolution in situ x-ray study of the hydrophobic gap at the wateroctadecyl-trichlorosilane interface PNAS 103 18401
[57] Maccarini M, Steitz R, Himmelhaus M, Fick J, Tatur S, Wolff M, Grunze M, Janeček, Netz R
R 2007 Density depletion at solid-liquid interfaces: A neutron reflectivity study Langmuir 23
598
[58] Tyrrell J W G and Attard P 2001 Images of nanobubbles on hydrophobic surfaces and their
interactions Phys. Rev. Lett. 87 176104
[59] Tretheway D C and Meinhart C D 2004 A generating mechanism for apparent fluid slip in
hydrophobic microchannels Phys. Fluids 16 1509
[60] Poynor A, Hong L, Robinson I K, Granick S, Zhang Z and Fenter P A 2006 How water meets a
hydrophobic surface Phys. Rev. Lett. 97 266101
[61] Hendy S C and Lund N J 2009 Effective slip length for flows over surfaces with nanobubbles:
Slip effects in polymer thin films
[62]
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]
[83]
[84]
[85]
[86]
[87]
[88]
37
The effect of finite slip J. Phys.: Condens. Matter 21 144202
Oron A, Davis S H and Bankoff S G 1997 Long-scale evolution of thin liquid films Rev. Mod.
Phys. 69 931
Kargupta K, Sharma A and Khanna R 2004 Instability, dynamics, and morphology of thin slipping
films Langmuir 20 244
Fetzer R, Münch A, Wagner B, Rauscher M and Jacobs K 2007 Quantifying hydrodynamic slip:
A comprehensive analysis of dewetting profiles Langmuir 23 10559
Blossey R 2008 Thin film rupture and polymer flow Phys. Chem. Chem. Phys. 10 5177
Israelachvili J 1992 Intermolecular and surface forces 2nd edition (New York: Academic Press)
Becker J, Grün G, Seemann R, Mantz H, Jacobs K, Mecke K R and Blossey R 2003 Complex
dewetting scenarios captured by thin-film models Nature Materials 2 59
Rauscher M, Blossey R, Münch A and Wagner B 2008 Spinodal dewetting of thin films with large
interfacial slip: Implications from the dispersion relation Langmuir 24 12290
Fetzer R, Rauscher M, Seemann R, Jacobs K and Mecke K 2007 Thermal noise influences fluid
flow in thin films during spinodal dewetting Phys. Rev. Lett. 99 114503
Seemann R, Herminghaus S and Jacobs K 2001 Shape of a liquid front upon dewetting Phys.
Rev. Lett. 87 196101
Reiter G 2001 Dewetting of highly elastic thin polymer films Phys. Rev. Lett. 87 186101
Damman P, Baudelet N and Reiter G 2003 Dewetting near the glass transition: Transition from
a capillary force dominated to a dissipation dominated regime Phys. Rev. Lett. 91 216101
Neto C, Jacobs K, Seemann R, Blossey R, Becker J and Grün G 2003 Correlated dewetting
patterns in thin polystyrene films J. Phys.: Condens. Matter 15 421
Neto C, Jacobs K, Seemann R, Blossey R, Becker J and Grün G 2003 Satellite hole formation
during dewetting: Experiment and simulation J. Phys.: Condens. Matter 15 3355
Fetzer R, Jacobs K, Münch A, Wagner B and Witelski T P 2005 New slip regimes and the shape
of dewetting thin liquid films Phys. Rev. Lett. 95 127801
Fetzer R, Rauscher M, Münch A, Wagner B A and Jacobs K 2006 Slip-controlled thin film
dynamics Europhys. Lett. 75 638
Bäumchen O, Fetzer R, Münch A, Wagner B and Jacobs K 2009 Comprehensive analysis
of dewetting profiles to quantify hydrodynamic slip IUTAM Symp. on Adv. in Micro- and
Nanofluidics ed M Ellero, X Hu, J Fröhlich and N Adams (Springer)
De Gennes P G 1985 Wetting: Statics and dynamics Rev. Mod. Phys. 57 827
Frumkin A N 1938 On the wetting phenomena and attachment of bubbles I J. Phys. Chem. USSR
12 337
Brochard-Wyart F, De Gennes P-G, Hervet H and Redon C 1994 Wetting and slippage of polymer
melts on semi-ideal surfaces Langmuir 10 1566
Redon C, Brochard-Wyart F and Rondelez F 1991 Dynamics of dewetting Phys. Rev. Lett. 66
715
Redon C, Brzoska J B and Brochard-Wyart F Dewetting and slippage of microscopic polymer
films 1994 Macromolecules 27 468
De Gennes P G 1979 Ecoulements viscométriques de polymères enchevêtrés C. R. Acad. Sci. B
288 219
Münch A 2005 Dewetting rates of thin liquid films J. Phys.: Condens. Matter 17 S309
Jacobs K, Seemann R, Schatz G and Herminghaus S 1998 Growth of holes in liquid films with
partial slippage Langmuir 14 4961
Fetzer R and Jacobs K 2007 Slippage of newtonian liquids: Influence on the dynamics of dewetting
thin films Langmuir 23 11617
Bäumchen O, Jacobs K and Fetzer R 2008 Probing slippage and flow dynamics of thin dewetting
polymer films Proc. Eur. Conf. on Microfluidics (Bologna) Dec 10-12, 2008
Blake T D and Haynes J M 1969 Kinetics of liquid/liquid displacement J. Colloid Interface Sci.
30 421
Slip effects in polymer thin films
38
[89] De Ruijter M J, Blake T D and De Coninck J 1999 Dynamic wetting studied by molecular
modeling simulations of droplet spreading Langmuir 15 7836
[90] Blake T D and De Coninck J 2002 The influence of solid/liquid interactions on dynamic wetting
J. Adv. Colloid Interface Sci. 96 21
[91] Voué M, Rioboo R, Adao M H, Conti J, Bondar A I, Ivanov D A, Blake T D and De Coninck
J 2007 Contact-line friction of liquid drops on self-assembled monolayers: Chain-length effects
Langmuir 23 4695
[92] Barrena E, Kopta S, Ogletree D F, Charych D H and Salmeron M 1999 Relationship between
friction and molecular structure: Alkylsilane lubricant films under pressure Phys. Rev. Lett.
82 2880
[93] Shanahan M E R and Carré A 1994 Anomalous spreading of liquid drops on an elastomeric
surface Langmuir 10 1647
[94] Shanahan M E R and Carré A 1995 Viscoelastic dissipation in wetting and adhesion phenomena
Langmuir 11 1396
[95] Carré A and Shanahan M E R 1995 Influence of the ”wetting ridge” in dry patch formation
Langmuir 11 3572
[96] Carré A, Gastel J-C and Shanahan M E R 1996 viscoelastic effects in the spreading of liquids
Nature 379 432
[97] Shanahan M E R and Carré A 2002 Spreading and dynamics of liquid drops involving nanometric
deformations on soft substrates Colloids Surf. A 206 115
[98] Long D, Ajdari A and Leibler L 1996 Static and dynamic wetting properties of thin rubber films
Langmuir 12 5221
[99] Long D, Ajdari A and Leibler L 1996 How do grafted polymer layers alter the dynamics of wetting
Langmuir 12 1675
[100] Al Akhrass S, Reiter G, Hou S Y, Yang M H, Chang Y L, Chang F C, Wang C F and Yang A C-M
2008 Viscoelastic thin polymer films under transient residual stresses: Two-stage dewetting on
soft substrates Phys. Rev. Lett. 100 178301
[101] Ruckenstein E and Jain R K 1974 Spontaneous rupture of thin liquid films J. Chem. Soc. Faraday
Trans. II 132
[102] Brochard-Wyart F, Di Meglio J-M and Quéré D 1987 Dewetting C. R. Acad. Sc. Paris Série II
304 553
[103] Sharma A and Ruckenstein E 1989 Dewetting of solids by the formation of holes in macroscopic
liquid films J. Coll. Int. Sci. 133 358
[104] Brochard-Wyart F and Daillant J 1990 Drying of solids wetted by thin liquid films Can. J. Phys.
68 1084
[105] Brochard-Wyart F, Redon C and Skykes C 1992 Dewetting of ultrathin liquid films C.R. Acad.
Sci. Paris Série II 314 19
[106] Reiter G 1992 Dewetting of thin polymer films Phys. Rev. Lett. 68 75
[107] Brochard-Wyart F, Debregeas G, Fondecave R and Martin P 1997 Dewetting of supported
viscoelastic polymer films: Birth of rims Macromolecules 30 1211
[108] Masson J-L and Green P F 2002 Hole formation in thin polymer films: A two-stage process Phys.
Rev. Lett. 88 205504
[109] Brochard-Wyart F and Redon C 1992 Dynamics of rim instabilities Langmuir 8 2324
[110] Reiter G and Sharma A 2001 Auto-optimization of dewetting rates by rim instabilities in slipping
polymer films Phys. Rev. Lett. 87 166103
[111] Münch A and Wagner B 2005 Contact-line instability of dewetting thin films Physica D 209 178
[112] Gabriele S, Sclavons S, Reiter G and Damman P 2006 Disentanglement time of polymers
determines the onset of rim instabilities in dewetting Phys. Rev. Lett. 96 156105
[113] King J R, Münch A and Wagner B 2009 Linear stability analysis of a sharp-interface model for
dewetting thin films J. Eng. Math. 63 177
[114] Migler K B, Hervet H and Leger L 1993 Slip transition of a polymer melt under shear stress Phys.
Slip effects in polymer thin films
[115]
[116]
[117]
[118]
[119]
[120]
[121]
[122]
[123]
[124]
[125]
[126]
[127]
[128]
[129]
[130]
[131]
[132]
[133]
[134]
[135]
[136]
[137]
[138]
39
Rev. Lett. 70 287
Brochard F and de Gennes P G 1992 Shear-dependent slippage at a polymer/solid interface
Langmuir 8 3033
Ajdari A, Brochard-Wyart F, De Gennes P-G, Leibler L, Viovy J-L and Rubinstein M 1994
Slippage of an entangled polymer melt on a grafted surface Physica A 204 17
Gay C 1999 New concepts for the slippage of an entangled polymer melt at a grafted solid interface
Eur. Phys. J. B 7 251
Tchesnokov M A, Molenaar J, Slot J J M and Stepanyan R 2005 A molecular model for cohesive
slip at polymer melt/solid interfaces J. Chem. Phys. 122 214711
Hervet H and Leger L 2003 Flow with slip at the wall: From simple to complex fluids C. R.
Physique 4 241
Drda PP and Wang S-Q 1995 Stick-slip transition at polymer melt/solid interfaces Phys. Rev.
Lett. 75 2698
Wang S-Q and Drda P A 1996 Stick-slip transition in capillary flow of polyethylene. 2. Molecularweight dependence and low-temperature anomaly Macromolecules 29 4115
Priezjev N V and Troian S M 2004 Molecular origin and dynamic behavior of slip in sheared
polymer films Phys. Rev. Lett. 92 018302
Reiter G and Khanna R 2000 Kinetics of autophobic dewetting of polymer films Langmuir 16
6351
Reiter G and Khanna R 2000 Negative excess interfacial entropy between free and end-grafted
chemically identical polymers Phys. Rev. Lett. 85 5599
Reiter G and De Gennes P-G 2001 Spin-cast, thin, glassy polymer films: Highly metastable forms
of matter Eur. Phys. J. E 6 25
Reiter G, Hamieh M, Damman P, Sclavons S, Gabriele S, Vilmin T and Raphaël E 2005 Residual
stresses in thin polymer films cause rupture and dominate early stages of dewetting Nature
Materials 4 754
Vilmin T and Raphaël E 2006 Dynamic instability of thin viscoelstic films under lateral stress
Phys. Rev. Lett. 97 036105
Damman P, Gabriele S, Coppée S, Desprez S, Villers D, Vilmin T, Raphaël E, Hamieh M, Al
Akhrass S and Reiter G 2007 Relaxation of residual stress and reentanglement of polymers in
spin-coated films Phys. Rev. Lett. 99 036101
Vilmin T and Raphaël E 2005 Dewetting of thin viscoelastic polymer films on slippery substrates
Europhys. Lett. 72 781
Yang M H, Hou S Y, Chang Y L and Yang A C-M 2006 Molecular recoiling in polymer thin film
dewetting Phys. Rev. Lett. 96 066105
Vilmin T and Raphaël 2006 Dewetting of thin polymer films Eur. Phys. J. E 21 161
Hamieh M, Al Akhrass S, Hamieh T, Damman P, Gabriele S, Vilmin T, Raphaël E and Reiter G
2007 Influence of substrate properties on the dewetting dynamics of viscoelastic polymer films
J. of Adhes. 83 367
Coppée S, Gabriele S, Jonas A M, Jestin J, Damman P 2009 Influence of chain interdiffusion
between immiscible polymers on dewetting dynamics arXiv:0904.1675v1
Ziebert F and Raphaël E 2009 Dewetting dynamics of stressed viscoelastic thin polymer films
Phys. Rev. E 79 031605
Ziebert F and Raphaël E 2009 Dewetting of thin polymer films: Influence of interface evolution
Europhys. Lett. 86 46001
Reiter G, Al Akhrass S, Hamieh M, Damman P, Gabriele S, Vilmin T and Raphaël E 2009
Dewetting as an investigative tool for studying properties of thin polymer films Eur. Phys. J.
Special Topics 166 165
Besancon B M and Green P F 2007 Dewetting dynamics in miscible polymer-polymer thin films
mixtures J. Chem. Phys. 126 224903
Fetzer R 2006 Einfluss von Grenzflächen auf die Fluidik dünner Polymerfilme (Berlin: Logos)
Slip effects in polymer thin films
40
[139] Merabia S and Avalos J B 2008 Dewetting of a stratified two-component liquid film on a solid
substrate Phys. Rev. Lett. 101 208304
[140] Servantie J and Müller M 2008 Temperature dependence of the slip length in polymer melts at
attractive surfaces Phys. Rev. Lett. 101 026101
[141] Müller M, Pastorino C and Servantie J 2008 Flow, slippage and a hydrodynamic boundary
condition of polymers at surfaces J. Phys.: Condens. Matter 20 494225
[142] Pastorino C, Binder K, Kreer T and Müller M 2006 Static and dynamic properties of the interface
between a polymer brush and a melt of identical chains J. Chem. Phys. 124 064902
[143] Pastorino C, Binder K and Müller M 2009 Coarse-grained description of a brush/melt interface
in equilibrium and under flow Macromolecules 42 401