You are on page 1of 37

4.

Parameters influencing contact angle measurements ............................................... 1


4.1. The Measurement ................................................................................................ 2
4.1.1. Introduction .................................................................................................. 2
4.1.2. Experimental ................................................................................................. 9
4.1.3. Results and Discussion ............................................................................... 11
4.1.4. Conclusions and Summary ......................................................................... 19
4.2. Surface functionalisation with Trichlorosilanes................................................ 20
4.2.1. Introduction ................................................................................................ 20
4.2.2. Experimental ............................................................................................... 20
4.2.3. Results and Discussion ............................................................................... 21
4.2.4. Summary ..................................................................................................... 23
4.3. PDMS (Polydimethylsiloxane).......................................................................... 23
4.3.1. Introduction ................................................................................................ 23
4.3.2. Experimental ............................................................................................... 25
4.3.3. Results and Discussion ............................................................................... 26
4.3.4. Summary and Conclusions ......................................................................... 30
4.4. Fit routines ......................................................................................................... 31
4.4.1. Introduction ................................................................................................ 31
4.4.2. Experimental ............................................................................................... 32
4.4.3. Results and Discussion ............................................................................... 33
4.4.4. Conclusions and Summary ......................................................................... 34
4.5. Acknowledgements ........................................................................................... 34
4.6. Literature ........................................................................................................... 34

4. Parameters influencing contact angle measurements


The aim of this chapter is to summarise measurements and data that offer a basic understanding of the
materials and methods used in the two larger (published) chapters (6 and 7).
Using standard, model surfaces (thiols on gold/silicon wafer), the influence of several parameters
(except profound surface roughness, covered in chapters 6 and 7) on contact angle measurements is
investigated, different measurement techniques are compared and a few theoretical assumptions tested.
Perfluorinated silanes were used throughout the thesis either as a demoulding or a hydrophobising
agent. A quick contact-angle survey shows which materials relevant to this thesis can be
functionalised by silanes, adsorbed from the vapour phase.
PDMS (polydimethylsiloxane) is a very versatile elastomeric polymer and is used in many fields of
research (such as micro-contact printing[Love_2005], micro-fluidics[Whitesides_2006], cell
studies[Charnley_2009] and anti-fouling surfaces[Schumacher_2007, Genzer_2000]). It is known that
a native PDMS surface[Morra_1990] experiences a change in conformation of the surface layer upon

contact with water, influencing contact angle measurements. A few tests were performed to better
understand the effect of different treatments on the contact angle.
In the last section some issues that arise when measuring very high contact angles are presented and
discussed.

4.1.

The Measurement

4.1.1. Introduction
In the literature, several different ways of measuring contact angles have been described. In order to
find out how these measuring methods can be compared, different methods were tested on standard
substrates (thiols on gold/silicon wafer) with a standard liquid (water).
A standard substrate for contact angle measurements needs to be clean, stiff and rigid, flat and
smooth, chemically homogeneous and inert towards the test liquid. For lab purposes, mica sheets
(layer silicate), polished silicon wafers and glass microscope slides (float glass) are often used for
contact angle measurements. They all have the advantage that they are reasonably smooth, can be
modified by surface chemistry or can be coated with any metal or polymer of choice. They are
resistant to ultra-high vacuum and to many acids, bases and solvents. They are readily available and
affordable. The choice of the thiol/gold/silicon wafer system was made because self-assembled
monolayers of thiols on gold (chapter 2.2) are well-known and well-characterised[Love_2005] and the
surface energy can easily be tailored by using either an OH-terminated (hydrophilic) or CH3terminated (hydrophobic) thiol species or mixtures of them (see chapter 2.4).
An ideal liquid for contact angle measurements has a low vapour pressure, is pure, does not take up
any substances from the atmosphere (no self-contamination), has a low viscosity and is preferentially
non-toxic[deGennes_2004]. Silicone oils best meet these conditions; additionally they can be tailored
in viscosity by varying the chain length of the silicone species and are therefore often used to study
wetting phenomena. But they have one big disadvantage: silicones are known to contaminate any
surface with low-molecular weight species[Graham_2002]. Due to their low surface tension (fully
spreading on nearly every surface) and non-volatility, they form a precursor film ahead of the drop via
a surface-diffusion process, which is one monolayer to 10 nm in thickness, and will propagate over
any surface until the drop is entirely depleted, covering the surface with a layer of silicone
[Bonn_2009, deGennes_1985]. This silicone contamination is to be avoided in every surface science
lab that is involved in any activity besides working with silicones. Therefore, water is often chosen as
test fluid. It is of great technical interest and readily available. Due to its high surface tension it shows
a finite contact angle on many surfaces. When measuring with water one must take care to always use
clean and fresh water (minimise self contamination) and not to prolong the measurement longer than

necessary (to minimise the effect of evaporation). If these conditions are met, contact angles will show
values close to that predicted by the Youngs equation[deGennes_2004].
Other parameters that may influence the contact-angle measurements are temperature, atmospheric
pressure and humidity (relative vapour pressure).
As already shown in chapter 3, contact angles can be measured either statically or dynamically. Figure
4 shows a compilation of different ways in which contact angles can be measured. This compilation of
methods is far from complete; it only shows the methods that were tested in the present work and are
reported in this chapter. More approaches can be found in the books by de Gennes et. al (chapter
2)[deGennes_2004] and Adamson and Gast (mainly chapter 10)[Adamson_1997].
Generally, the static measurement is performed by producing a specific drop volume that is still
hanging from a syringe needle, which is then gently brought into contact with the solid. Upon contact
with the solid, the drop will detach from the syringe and spread on the surface, decreasing the contact
angle from initially 180 (prior to contact) to the static (equilibrium) contact angle s. For small
drops, the driving forces are only capillary forces (surface tension). During the measurement the
syringe tip does not remain within the drop. If one uses drop volumes larger than the maximum
volume that stays attached to the syringe, then the drop needs to be formed between surface and
syringe. This method can lead to a slightly higher contact angle, but it reduces the influence of impact
of falling droplets.
Static measurements are usually quicker than dynamic measurements, but the single value yields less
information about the surface than the dynamic measurement. It is mainly used as a rapid
characterisation tool. If, by a chemical reaction or other type of surface coating, a significant change in
surface energy can be expected, then s provides rapid information on whether, and within certain
limits, to what extent the experiment was successful. The spatial resolution of a sessile-drop
measurement is defined by the drop-base diameter. The resulting contact angle is an average over the
whole length of the contact line[Gao_2007a, Spori_2010], (whole area of contact[Cassie_1948]), if the
surface features (chemical or topographical) are small compared to the drop. Usually micro-litre-sized
droplets are used for measurements. The spatial resolution can be increased by the use of pico-litresized droplets. Pico-litre droplets have to be measured immediately after deposition to minimise the
influence of evaporation (reduction of contact angle), but are indeed comparable to data measured
with micro-litre drops[Taylor_2007].
If surfaces are smooth and homogeneous, static measurements provide data for calculations in
thermodynamic equilibrium, and the surface tensions SL and SA can be calculated by applying the
appropriate model (see comment further below).

On surfaces where chemical heterogeneities or surface roughness are homogeneously distributed along
the surface and are small compared to the drops footprint, repeated static contact angle measurements
can lead to values with small standard deviations, if the drop is always set on the surface by the same
method and with care. Such contact angles most probably are far from the thermodynamic equilibrium
but nevertheless allow the comparison of different surfaces. s measurements suffer from the
disadvantage that heterogeneities and surface roughness cannot be distinguished from intrinsic surface
energy. In the course of the measurement, swelling or dissolution may also occur and may not be
detected due to the limited time scale of the measurement. Therefore, static contact angle
measurements must always be treated with caution[Kwok_1998].
Influence of drop size V on contact angle s, drop base diameter r and roll-off angle :
Surprisingly, the literature does not provide a conclusive answer to the question as to the influence of
the droplet volume V on contact angle s, drop base diameter r and roll-off angle . The shape of the
drop is determined by the surface tensions involved and by gravity. For every liquid there exists a
particular length, the capillary length -1, beyond which gravity becomes important[deGennes_2004].
When capillary forces and hydrostatic forces are in equilibrium, the capillary length can be calculated
by comparing the two forces; with g being the acceleration due to gravity and the density of the
liquid:

(1)

LA
g

While on the one hand the capillary length defines the characteristic length within which a
perturbation of a surface decays, on the other hand it splits the observation of sessile drops into two
regimes. If the radius r of the drop is smaller than the capillary length, then gravitational effects on the
shape of the drop can be neglected. In the opposite case when r > -1, gravity flattens the drop to a
puddle with an equilibrium height e. e itself is a function of -1 and the equilibrium contact angle Y.
In contact-angle measurements it defines the maximum radius that a drop can adopt before gravity
starts to distort its profile from the ideal truncated sphere (see Figure 1).

Figure 1: For small droplets it can be assumed that gravity has no effect on their profile. The limit for the maximum
radius is given by the capillary length -1. a.) small droplet showing no sign of gravitational effects, b.) large puddle
with a diameter larger than twice the capillary length, which is flattened by gravity.

In fluid mechanics, the Bond number is a numerical expression of the ratio of gravitational (volume)
to surface forces. For a sessile drop measured in air the Bond number Bo is defined as
follows[Adamson_1997]:

Bo

(2)

g r2
LA

If the Bond number is equal to one, the expression is equal to the capillary length if solved for the
drop-base radius r. Numbers well below unity describe static systems that are only governed by
surface forces, and in these cases, gravitational forces can be neglected; for very high numbers it is the
opposite. Numbers around unity describe a system where the two forces are in a non-trivial balance.
As mentioned above, if a drop is unperturbed by gravity, it assumes the form of a truncated sphere. By
knowing the dispensed volume V0 and the measured contact angle, the drop-base radius r can be
calculated (see equations 4 and 5). For a given (or expected) wettability () of a surface the drop-base
diameter for a certain volume can be predicted, which provides useful information for the selection of
specimen size or what can be the maximum spatial resolution for contact-angle measurements. By
actually measuring the radius r on a standard substrate, conclusions can be drawn as to how accurate
the predictions are and whether an influence of gravity can be detected and whether it is significant. In
Figure 2, the different geometrical parameters are outlined and the diameter (2r) of the drops footprint
is calculated for the hydrophilic and hydrophobic case as follows:

rphilic

6 V0 sin 3

2
2
1 cos 3 sin 1 cos

rphobic

3 V0 sin 3

2
1 cos 2 cos 4

(3)

(4)

3
, where = 180 -

Figure 2: Schematic illustration of which parameters were used to calculate the radius of the drops footprint from
the dispensed volume.

In 1962 Furmidge[Furmidge_1962] was looking into the conditions necessary for drop retention on
inclined surfaces and, by equating all the forces involved, defined a correlation between the inclination

of the surface with the drop weight m, drop diameter perpendicular to the sliding direction w,

advancing and receding contact angles (A, R) and the surface tension LA of the liquid.

m g sin w LA cos R cos A

(2)

Furmidge (see Figure 3) assumed the footprint of the drop to be rectangular during sliding, which may
be a source of error, but he found the equation to be sufficiently accurate to be able to account for
sliding drops on a flat surface.

Figure 3: Sketch of a sliding drop in the side and top view with indication of all the parameters involved in the
Furmidge equation: drop weight m, acceleration due to gravity g, inclination angle , drop width w, liquid-air
surface tension LA and dynamic contact angles a and r

The Youngs equation correlates the involved surface tensions by using one specific contact angle

Y. Therefore, on an ideal surface, no hysteresis would occur[Kwok_1999]. However, measuring on


real surfaces reveals the fact that on every surface the contact angle can adopt any contact angle
between a minimum (receding, r) and a maximum (advancing a) contact angle. The difference (a

r) between those two quantities is known as the hysteresis. But even if a surface were perfectly
smooth, chemically homogeneous and perfectly elastic, hysteresis could only be prevented if the
experiment was carried out infinitely slowly, to ensure thermodynamic reversibility. Since this is not
possible, energy-dissipating processes occur. Upon contact, chemical effects, such as interdiffusion,
interdigitation, molecular reorientations and exchange processes occur. Due to the finite time of
measurement, the work of adhesion W (Young-Dupr: W = LA(1+cos)) of the separation (receding)
is larger than that of the approach (advancing)[Ruths_2008].
More significant causes of hysteresis are attributed to chemical heterogeneities and surface
roughness[deGennes_1985]. Hysteresis occurs due to contact-line pinning, e.g. on strong and sparse
defects[deGennes_1985, deGennes_2004, Priest_2007, Reyssat_2009a]. More information on the
hysteresis induced by roughness can be found in chapters 1, 6 and 7. In general, hysteresis is still a
subject that is not fully understood. Many issues remain to be solved on what the nature and location
of energy dissipation is, and how solid-liquid interactions, surface heterogeneity and roughness,
substrate viscoelasticity and stiffness influence wetting dynamics[Fetzer_2009].

Even if no comprehensive model for hysteresis has been found, its experimental value can be
determined by dynamic contact-angle measurements.

Figure 4: Measuring contact angles: Contact angle measurements can basically be divided into static and dynamic
measurements. a.) depicts a static measurement. The syringe is out of the drop and the drop is at rest. b.-e.) show
dynamic measurements where the contact line of the drop is forced to move: b.) by changing the drop volume
incrementally; c.) by changing the drop volume at a constant speed; d.) by dragging the drop along the surface and
e.) by tilting the substrate until the drop starts to move.

Dynamic measurement: Influence of choice of method


During a dynamic measurement the contact line is forced into movement. The contact angle facing
fresh surface (larger contact angle) is known as the advancing contact angle a while the contact angle
moving over surface which had already been wet by the drop (smaller contact angle) is denoted as the
receding contact angle r. In principle, during wetting the contact line of the drop measured statically
encounters the same underlying mechanisms as the advancing contact angle, and therefore usually s
and a generally show similar effects[Blake_2006]. Some have even replaced s with a in their
calculations[Kwok_1999]a procedure that others do not endorse[Fetzer_2009].

Figure 5: Schematic drawing of contact-angle-hysteresis evolution with increasing contact line speed
U.[Blake_2006]

The main parameters that quantify the dynamics of wetting are the contact-line speed U and the
dynamic contact angles a and r. a and r are macroscopic quantities, since they are usually
measured by relatively simple optical setups. a and r can get pinned, and sometimes displays a
stick-slip behaviour that renders its interpretation difficult and the achievement of steady velocities
experimentally challenging. Advancing (receding) contact angles increase (decrease) with increasing
contact line speed U (see Figure 5). Two main approaches (models) were developed to explain
dynamic wetting. They differ from each other both in terms of conceptual framework (mechanical

versus chemical model[deGennes_2004]) and the channel of energy dissipation[Blake_2006]. The first
is known as hydrodynamic theory and the energy dissipation occurs due to viscous flow within the
wedge of the liquid near the moving contact line. The second model is known as molecular-kinetic
theory. Liquid transport is described as a statistical process of thermally activated local displacements
at the contact line. Energy is mainly dissipated by contact-line friction[Blake_2006, Bonn_2009,
Fetzer_2009].
There are indications e.g. in the process of dewetting (receding), that the hydrodynamic theory fits
better for higher contact line speeds (speeds occurring right after placement of a drop),whereas the
molecular-kinetic theory describes the slower range shortly before complete relaxation. Additionally,
models that involve pinning have been proposed[Fetzer_2009].
Researchers have developed several different approaches to moving the contact line. Figure 4b shows
the case where the drop volume is enlarged incrementally until the contact line moves and then the
contact angle is measured. It is employed in cases when no automated machine is
available[Drelich_1994, Morra_1990, Extrand_2004a]. The automated version of this type of
measurement is shown in Figure 4c where the drop volume is changed by enlarging and reducing the
drop at a constant dosing speed. This is the type that has been used most frequently in our
group[Tosatti_2002, Spori_2008, Spori_2007], and will be referred to as automated in the
subsequent discussion. Figure 4d forces the contact line to move by moving the substrate underneath
the drop[Callies_2005a, Priest_2009] and will be referred to as drawing. If the substrate is tilted to
the point where the drop starts to move, the front angle can be considered as the advancing contact
angle and the back angle as the receding contact angle[Furmidge_1962, Long_2009]. This type of
measurement will be abbreviated as tilting.

A comment on the possibilities of measuring SA and SL:


The Youngs equation consists of four parameters, but only and LA can be measured independently,
which leaves SA and SL undefined. From the static contact angle data of different, carefully chosen
test liquids measured on the surface of interest, the surface tensions SA and SL can be
calculated[KrssGmbH_2004, Walther_2007, Gerhardt_2009] by employing one of the models
developed by Zisman[Fox_1952], Fowkes[Fowkes_1964], Owens and Wendt[Owens_1969] and
vanOss et al.[VanOss_1988]. These models, starting with Fowkes, made the assumption that the kinds
of interactions occurring at the liquid-solid interface play an important role. They distinguish the
interactions as either polar or dispersive [Fowkes_1964, Owens_1969], or, within dispersive, acid and
base interactions[VanOss_1988]. Thus, only interactions of the same nature will combine, the
resultant being calculated from the geometric mean of the fraction of the total surface tensions LA and

SA. For example, the surface tension LA of water (72.8 mN/m) consists only of 30% dispersive

interactions[Fowkes_1964], meaning that only these 30% of the total surface tension LA will
attractively interact with a dispersive surface such as polyethylene (purely dispersive, SA 35
mN/m[Fowkes_1964]), leading to a low adhesion and large contact angles.
The group of Neumann has cast serious doubt on the accuracy of these models[Kwok_1999] and
proposes an alternative, equation-of-state approach. They claim that other authors developed their
approaches because they did not have reliable contact angle data, and further state that the models
contradict physical reality. Their main criticism was that these models were developed after
measuring mainly static contact angles. By measuring static contact angles, effects like stick-jump
behaviour, dissolution of the surface, adsorption of molecules onto the surface cannot be recognised as
they can by measuring slow advancing contact angles. Thus, to explain the scatter in their data,
Fowkes and others came up with the idea that it is not only the surface tension between the phases
which defines the contact angle, but also the type of intermolecular forces at work. This assumption is
what the Neumann group strongly doubts, since they can show that by choosing the liquids that
provide (in their eyes) reliable data, polar and apolar liquids fall smoothly on the same curve in the

LVcos vs. LV plot, which would not be expected, if the specific intermolecular interactions indeed
played a role
4.1.2. Experimental

Substrates:
Single-side-polished silicon wafers (Si-Mat Silicon Materials, Germany) of dimensions 10 x 40 mm2
were cleaned by ultrasonication, first for 10 min in toluene (for HPLC, Acros Organics, Belgium),
then twice 10 min in ethanol (analytical grade, Scharlau, Spain) to remove glue residues from the
cutting and then blown dry in a stream of nitrogen. Additionally, in order to be able to measure large
drops on a substrate with a hydrophilic coating, one microscope slide (40 x 40 mm2, custome size,
Menzel-Glser, Germany) was cleaned with Piranha solution (30 v% H2O2 in 70 v% fuming sulphuric
acid 95 97 %, both pro analysi, Merck, Germany). Subsequently the silicon wafers were cleaned for
2 min in air plasma (Harrick Plasma, USA) and then, together with the glass slide, coated by resistance
evaporation (MED 020 coating system, Baltec, Liechtenstein) with 10 nm of Cr and 80-100 nm of Au
(purity >99.99 %, Umicore, Liechtenstein).

Surface modification:
Directly after coating, the samples were immersed in a 1.2 mM ethanolic solution of 11mercaptoundecanol or dodecanethiol (Aldrich Chemicals, USA). Self-assembly took place overnight.
Directly before the measurement the samples were removed from the solution, rinsed with ethanol and
dried under a stream of nitrogen.

Quality control:
Before starting the measurement, the quality of the water was checked by a pendent-drop experiment
(DSA100, Krss GmbH, Germany) and found to be very pure with a surface tension of above 72.4
mN/m[DellaVolpe_1997]. Room temperature was 23 C and the relative humidity 51 % while
measuring the dodecanethiol SAMs and 26 % for the 11-mercaptoundecanol SAMs. Additionally, the
crystallinity and the organization of the alkyl chains of the SAM were controlled by polarizationmodulation infrared reflection absorption spectroscopy (PMIRRAS). The position of the peak maxima
and the broadness of the absorption bands indicate that the SAMs were of high
quality[Venkatamaran_2006].

Contact-angle measurements:
All contact-angle and roll-off angle measurements were performed on the DSA100 (Krss GmbH,
Germany) and evaluated by the Tangent method 2 of the DSA3 software (Krss GmbH, Germany).
Contact angles were then exported into Excel (Microsoft, USA) for further evaluation (average,
standard deviation, deletion of artefacts occurring from faulty profile detection).
The static measurements were carried out with drops of between 1 and 30 l volume(Table 1), which
were gently placed onto the substrates. An image was taken of each drop and subsequently the contact
angles were evaluated. Drop-base diameters were extracted with the straight-line-selection tool in the
ImageJ program (rsbweb.nih.gov/ij). Pixels were calibrated by the width of the syringe (0.516 mm),
which was also visible in the image.
Table 1: Volumes used to measure static and roll-off measurements on the two different substrates.

SAM/volume [l]

CH3

OH

12

15

30

During the automated measurement a movie is recorded while enlarging or reducing the drop volume
at a constant flow speed. The contact angles are then subsequently evaluated from the movie. The
measurements were performed at different flow speeds (5, 10, 15, 20, 25, 30 l/min). First, a drop of 5

l was set onto the substrate, with the syringe still within the drop. Then, the drop was enlarged in two
steps of 4 l to 13 l at the corresponding speed. The reason to divide the advancing measurement into
two movies was to provide the possibility of re-centreing the drop to the syringe, in case it had initially
grown asymmetrically. The experiment with the hydrophilic substrates was repeated with additional
measurements of the advancing contact angle by increasing the drop in one step by 8 l at the
corresponding speed.

10

The recording settings were chosen such that at least 60 images of the drop could be evaluated (up to
170). The receding measurement consisted of only one movie. Only those images were evaluated in
which the side of the drop was actually seen to be moving.
For the drawing measurement a drop of 8 l was placed on a substrate coated with the hydrophobic
SAM, with the syringe close to the surface. Then, the stage was moved underneath the drop at three
different speeds offered by the machine: fast (11.1 mm/sec), normal (6.3 mm/sec) and slow (2.9
mm/sec). During the movement, a movie was recorded.
The tilting measurement was performed directly after the static measurement. Therefore the same drop
volumes as mentioned above were investigated. All samples were tilted with a speed of 1.2 /sec
(machine parameter slow) and during tilting a movie was recorded. As soon as the drop started to
move, tilting was stopped at the corresponding angle and advancing and receding contact angles were
evaluated from the moment where the drop started to move on both sides. Images for evaluation varied
between 3 to 10 images per drop, depending on the distance available for sliding and the sliding speed
of the drop.

4.1.3. Results and Discussion

Static contact angles: The first row in Figure 6 shows the results from the static-contact-angle
measurements versus droplet volume. The same contact-angle range is shown for the hydrophilic and
hydrophobic SAMs (Figure 6a and b), to allow comparison of the standard deviations. For both
substrate types, no distinct dependence of the contact angles on the investigated drop volumes can be
observed. All data points scatter around one specific value (20 2 for the hydrophilic, 108.7 0.6
for the hydrophobic substrate). The same behaviour was observed in another study[Taylor_2007]
where contact angles were also investigated by the tangent method. Others[Amirfazli_1998,
Tadmor_2008] have reported a slight decrease with drop size, a tendency which might be hidden
within the (small) standard deviation of the data presented here .
On the hydrophilic substrate, the contact angle of the 30 l drop seems to be slightly higher than
expected. Since the 30 l drop had a diameter of nearly 10 mm, it could only be measured on the large
samplethe OH-thiol/gold-coated glass slide. The different substrate is indicated by an open instead
of a full symbol in Figure 6a. Other drop sizes (data not shown) were also measured on the very same
sample and they all showed the same trend: they scattered around 23 instead of 20. Since the
standard deviations from the data measured on the silicon wafer and the glass slide slightly overlap, no
definite conclusion can be drawn that this influence is due the choice of the substrate. But all other
parameters (gold coating, solvent preparation, glass cleaning) were held constant.

11

The standard deviation on the hydrophilic substrate is significantly larger than that observed on the
hydrophobic substrate. This may have several origins: hydrophilic surfaces with their low contact
angles are more difficult to measure[KrssGmbH_2004]. Also, the apex of the drop reflects the light,
which can confuse the software. Additionally, hydrophilic surfaces are more prone to airborne
contamination than hydrophobic surfaces, even though all the surfaces used here were exposed to air
as briefly as possible. They were measured immediately after removal from the thiol solution.
On each substrate coated with the hydrophobic SAM, one additional drop of 6 l was measured. It was
intended as a control to be sure that all substrates had indeed the same quality. But because space was
limited this 6 l drop was always placed at the end where the sample was held with the tweezers.
When evaluating the data it was clear that the contact angles measured on this spot was again higher
than on the untouched areas, namely 110.1 0.4 rather than 108.7 0.6. In order to achieve very
reproducible results it is necessary to always measure on untouched surfaces.

Figure 6: Influence of drop volume on static contact angle measured on a) OH- and b) CH3-terminated thiols on
gold/silicon. For the observed drop volume range no significant influence of drop size on contact angle was found.

The data points in Figure 7a were obtained by extracting the drop-base diameters from the images of
the static measurements . The dashed line represents the limit given by the capillary length (equation
1) and the solid lines show the evolution of the drop-base diameter when calculated from the contactangle average from the static measurement (philic = 20, phobic = 109) and the dispensed drop volume
as shown in equations 3 and 4. For both substrates and for drop sizes up to 15 l, the calculated curves
lie within the standard deviation of the measured drop-base diameters. Distinct underestimation occurs
for 30-l-drops, and this was slightly more pronounced on the hydrophilic than on the hydrophobic
substrate.

12

For the hydrophilic case, the diameter already intercepts the capillary length criterion at a volume of
~6l and indeed, the curve starts to deviate, even though, as mentioned, the deviation only starts to
become obvious for drops larger than 15 l. The intercept for the calculated hydrophobic case with the
capillary length would only be at a drop volume of 73 l.
For the hydrophobic SAMs, the Bond numbers (equation 2) vary from 0.06 to 0.63, and for the
hydrophilic SAMs from 0.26 to 3.27. None of these values is far away from unity (as is the case for
pico-litre droplets, for example[Taylor_2007]), so all measurements are probably influenced by
surface and volume forces. The two SAM species (OH- and CH3-terminated) were chosen such that
contact angles on the respective surface are close to the lower and higher limit of finite contact-angle
measurements observable on smooth surfaces. Contact-angle measurements are usually performed
with micro-litre droplets. Therefore the Bond number may not be the correct criterion to choose a
suitable size for sessile-drop measurements.
In Figure 7b, drop images of 1 and 30 l drops on the hydrophilic (top row) and the hydrophobic SAM
(bottom row) are depicted. The outside diameter of the syringe is 0.516 mm, and serves as a scale bar.
By eye, no gravitational influences can be detected for the hydrophilic case, but in the hydrophobic
case the large 30 l drop no longer resembles a spherical drop. The deformation is probably even more
pronounced for a drop of 73 l, as suggested by the capillary length criterion to be the maximum drop
size without gravitational influence.
To conclude, the influence of the drop volume on the static contact angle is very small. For drop
volumes below 15 l, the drop diameter does not strongly deviate from the calculation with the
truncated sphere assumption. Nevertheless, it should be mentioned that Taylor et al.[Taylor_2007]
investigated the accuracy of contact-angle determination of micro-litre-sized drops by evaluating the
drop profile with the truncated sphere assumption (DSA 3: circle method) on a surface with medium
wettability ( 69). They found that with the circle method, contact-angle underestimation increases
with increasing drop volume. Thus, even though the radius does not strongly deviate when calculated
with the sphere assumption, the contact angle cannot be fitted adequately by the circle method, if used
for medium or low wettability.

13

Figure 7: Drop-base diameter as a function of drop volume. a.) The data points represent diameters actually
measured on the sessile drop, the full lines show the diameter calculated with the truncated sphere assumption and
the dashed line corresponds to the capillary length. b.) Examples of drop images, 1 and 30 l drops on hydrophilic
(top row) and hydrophobic (bottom row) substrates. The syringe outside diameter is 0.516 mm and can be used as a
scale bar.

Roll-off angle : The substrates for all drops used in the static measurement were tilted to determine
the roll-off angle as a function of drop volume (Figure 8). The smallest drops did not move at all, the
weight being too small to overcome the pinning of the drop. All other drops moved and the roll-off
angle was recorded. All parameters used in the Furmidge[Furmidge_1962] equation (equation 5) were
measured independently; drop width w as in Figure 7, dynamic contact angles (r, a) defined by the

automated method as in Figure 11, and the drop weight calculated from the dispensed droplet
volumes. The Furmidge equation fits the data reasonably well. By solving the Furmidge equation for
the weight of the drop and defining the drop width as a function of the volume of a spherical drop, one
can calculate which would be the smallest volume to still be able to move on the surface: for the
hydrophobic surface it would be 2.0 l and on the hydrophilic 2.7 l.

14

Figure 8: Roll-off angles versus droplet size. The equation of Furmidge[Furmidge_1962] (lines) correlates well with
the measured roll-off angles.

Dynamic measurements: Since the automated approach is the most commonly used method for
measuring dynamic contact angles in our group, the technique was scrutinised a bit more closely.
Initially the advancing contact angle was measured in two steps, since sometimes the drop did not
grow axio-symmetrically to the syringe but tended to grow to one side. A pause between two movies
(successive measurements) allows the syringe to be readjusted to the middle of the drop and thus leads
to a better measurement while recording the second movie. Examining the contact angles derived from
two such consecutive movies on the hydrophobic substrate did not reveal any influence of the break on
the contact angle behaviour, but there was clearly an influence seen on the hydrophilic substrate.
Therefore it was decided to repeat the measurement on the hydrophilic SAM, measuring drop growth
in one continuous movie. In Figure 9a, the contact-angle data of the 1-movie and the 2-movie
measurements at a dosing speed of 15 l/min are plotted versus time. By looking at the contact angles
of the 2-movie dataset, the break can easily be seen at 15 seconds. There is a possibility that there was
an air bubble trapped unnoticed in the dosing system, which would change the dosing speed due to its
compressibility, but also in this case one longer movie would be more favourable, because there would
be a longer time to build up a steady flow, despite the trapped air bubble. In the end, the number of
movies really did not have an influence on the averaged contact angle, but the standard deviation with
one movie appears slightly lower.
When measuring receding angles, the values of the hydrophilic substrate did not show any dependence
on the decreasing volume. On the hydrophobic SAM, another contact angle behaviour was observed,
see Figure 9b. With decreasing volume the data seems to show a downward tendency. Since it is right
from the beginning and the drop was still pretty large, the syringe argument does probably not hold

15

here. The scatter and the tendency do not seem to be velocity dependent therefore it is assumed that
indeed the volume has an influence on the receding contact angle on this hydrophobic substrate.

Figure 9: Datasets from the automated measurement. a.) comparison of the data measured at a speed of 15 l/min
on the hydrophilic surface. The 2-movie-data set was measured in two steps, which is clearly visible in the contact
angles. This artefact does not occur with the long single measurement (1 movie). b.) With decreasing volume, the
receding contact angle is not constant, but also increases on the hydrophobic surface.

Both hydrophilic and hydrophobic surfaces were measured with the same series of dosing speeds. By
extracting the drop-base diameter from the movies and plotting the radius versus time, the contact-line
speed could be extracted from the slope of the linear regression (Figure 10). Contact line speeds from
0.002 to 0.025 mm/sec are considered as low-rate dynamic contact angle measurements[Kwok_1999].
Such velocities lead to data extremely close to the y-axis of Figure 5. In other words, with these
velocities the inherent, speed-independent hysteresis can be characterised. The upper limit (0.025
mm/sec) is indicated with a dashed line in Figure 10. When measuring advancing angles with speeds
in this range on relatively smooth surfaces, they are thought to be essentially identical to the static
contact

angles[Kwok_1998]

and

could

therefore

be

used

for

surface-energy

calculations[Kwok_1999]. More generally it means that contact angles measured with speeds within
the range given above will not show influences of the dosing speed. Interestingly, the contact-line
speed on a hydrophilic substrate is significantly higher for the same dosing speed than on a
hydrophobic substrate, and crosses in the advancing case the limit of 0.025 mm/sec at a dosing speed
of slightly larger than 15 l/min. On both surfaces the advancing contact-line speed increases linearly
with the dosing speed. The same is true for the receding contact line on the hydrophobic surface. The
receding contact-line speed for the hydrophilic sample was harder to define, since it was only the very
last portion of the drop volume that moved over the surface, and the influence of the syringe was
readily detectable by a change in the decreasing rate of the radius r. Due to the very small residual
volume that is actually moving, starting from a large contact diameter, the contact line speed is

16

naturally a lot higher than on the hydrophobic surface. Additionally, contact angles as low as 5 (see
Figure 11a) are very hard to extract reliably with the sessile-drop measurement. The placement of the
horizontal line can also have a strong influence on the extraction of r. As a conclusion, 15 l/min is
clearly the upper limit for dynamic contact-angle measurements with the automated method. On
strongly hydrophilic samples that have a a of about 30 35, even lower dosing speeds are advisable.

Figure 10: Contact-line speed versus dosing speed (automated measurement). The contact-line speed increases
linearly with the dosing speed, and much more steeply on the hydrophilic (OH) than on the hydrophobic (CH3)
substrate. The receding contact line speeds on the hydrophilic sample are significantly higher than on the
hydrophobic sample since movement only starts for the very last portion of residual volume.

Comparison of three different techniques to measure dynamic contact angles: The three different
techniques, automated, tilting and drawing (see Figure 4) are compared with each other in Figure 11.
Full symbols correspond to advancing, open symbols to receding contact angles. The first row was
measured on the hydrophobic SAM, the second on the hydrophilic SAM.
The drawing method was only tested on the hydrophobic substrate. The data points (Figure 11) clearly
show a strong dependence on stage speed, which is essentially equal to the contact-line speed. The
contact-line speed is therefore about 100 to 1000 times higher than with the automated method (Figure
10). Since the receding contact angle deviates strongly from the values measured with the other
methods, the stage speeds offered by the machine are clearly too fast to obtain representative results.
The automated method leads to contact angles (a and r) that scatter around one specific angle on
both substrates, quite independently of the dosing speeds investigated. The tilting method also leads to
consistency of the contact angle as a function of the drop volume. However, on the hydrophilic sample
in particular, the scatter in the a values is larger than that obtained with the automated method. The
automated method does not lead to advancing contact angles that are equal to static contact angles,
even though many of the data points fulfil the requirement for slow dynamic measurements. The same
is true for the tilting measurement on the hydrophobic surface. On the hydrophilic substrate however,

17

there are indeed two measurements that are close to 20 and thus similar to the results obtained from
the static measurement. However, since only 2 out of 6 measurements fit, no general conclusion can
be drawn.

Figure 11: Dynamic contact angles measured with three different methods: automated, tilting and drawing. Full
symbols denote a, open symbols represent r. top row: hydrophobic SAM; bottom row: hydrophilic SAM.

Subtracting the value of r from a, the hysteresis can be calculated (Table 2). The contact angle
hysteresis on the hydrophobic substrate is very similar for the two types of measurement. On the
hydrophilic substrate however, a considerable difference is visible. One reason for the smaller
hysteresis achieved by the tilting method might be found in the fact that only two spots of the drops
footprint are fully in the advancing and receding mode. All other spots are a mixture of both modes.
And since all spots are connected especially via the liquid-air interface, which acts like a membrane,
the hysteresis might be slightly reduced. On the hydrophilic substrate, the larger adhesion and
substrate contact area may cause this effect to become visible.
Table 2: Hysteresis calculated by subtracting the average of the a minus the average of the r measured by the
automated and tilting methods.

Hydrophobic SAM

Hydrophilic SAM

automated

10 2

27 2

tilting

10 2

17 4

18

A few closing remarks: With the tilting method it is harder to control contact-line speed. As
Furmidge[Furmidge_1962] has described, the drop usually increases in speed once it has left its
original position. This effect is more pronounced the longer the drop was at rest. Therefore, Furmidge
reduced the tilt angle after the drop started moving to achieve a speed of about 2 mm/sec, which is
quite fast. Precise tilt-angle reduction while the drop is moving would be a challenge on the DSA 100,
and was therefore not done. If the drop was initially slightly pinned, e.g. as seen on superhydrophobic
surfaces (data not shown), the drop moves very fast once it has overcome the pinning energy barrier.

a and r determination then become difficult and require a high-speed camera.


A few effects were not discussed in this chapter, but which are also believed to have an influence on
the contact angle:
- Some authors claim to see an effect of the contact-line tension on the contact angle[Amirfazli_1998],
while others do not[Taylor_2007], and the literature is not even conclusive on the sign and the
magnitude of the contact-line tension. According to de Gennes[deGennes_2004] it is in the range of
10-11 J/m and too small to influence the macroscopic contact angle as determined here.
-Usually, contact angles are not measured in vacuum, and therefore the substrate is most probably not
completely dry but has a certain amount of the adsorbed molecules from the test liquid. If such a film
is present, an additional film pressure [Adamson_1997] (also called surface pressure[Bonn_2009]),
has to be considered, which has the same units as the surface tension. The influence of is more
pronounced if liquid and solid attract each other. In the case of water it is more likely to play a role on
the hydrophilic than on the hydrophobic SAM. But again, this value is important for solid surface
tension (SA) determination, and not so much for comparative contact-angle measurements. Control of

(e.g. humidity) increases the reproducibility of contact-angle measurements.


4.1.4. Conclusions and Summary
Water contact angles were measured on standard substrates coated with two different SAMs:
Mercaptoundecanol on gold/silicon renders the surface very hydrophilic and dodecanethiol very
hydrophobic. Thus, investigations were performed on two substrates with a large difference in surface
energy.

Static measurements: The influence of drop volume on contact angles is minimal for the investigated
volumes (1 to 30 l). Up to drop volumes of 15 l, drop-base radii can be approximated by assuming
the dispensed volume takes up the shape of a truncated sphere. The capillary-length criterion is of the
right magnitude to distinguish between drops that are or are not influenced by gravity. However, it
seems, especially on the hydrophobic substrate, that it greatly underestimates the influence of gravity
on the drop profile. The Furmidge equation leads to a reasonable fit of roll-off angles.

19

Dynamic measurements: For the automated method, the contact-line speed was evaluated as a function
of the dosing speed. The contact-line speed at the same dosing speed is considerably higher on
hydrophilic substrates than on hydrophobic surfaces. In order to stay in the range of slow dynamic
measurements, a dosing speed of 15 l/min is the upper limit for hydrophilic surfaces. It might even
be advisable to reduce the speed to 10 l/min when measuring on hydrophilic surfaces.

Comparison of three different methods to characterise dynamic contact angels: For each method,
different parameters were varied to investigate the parameter influence and compare the methods
among each other. The drawing method soon proved not to be applicable with the DSA 100, due to the
excessive speed of the stage. The automated and tilting methods lead to consistent contact-angle
values, which were not strongly influenced by the investigated parameters. The tilting method seems
to give slightly lower hysteresis on hydrophilic substrates. Generally, with both methods, advancing
contact angles were not found to be equal to static contact angles.

4.2.

Surface functionalisation with Trichlorosilanes

4.2.1. Introduction
Alkyl-trichlorosilanes are known to react with a number of oxide surfaces, such as silicon, glass and
metallic oxides[Maoz_1984]. Also polymers, such as Polydimethylsiloxane (PDMS), which form a
superficial silicate layer upon plasma-oxidation, are known to be good substrates for silane-based
SAMs[Chaudhury_1995, Morra_1990]. Other polymers such as polyethylene have been successfully
functionalised by creating a thin silicate layer (adsorption and hydrolysis of SiCl4) before exposure to
the silanes[Chaudhury_1995].
Silanisation is not a trivial reaction. Under ambient conditions, a clean oxide surface, such as that of
silica, is hydrated (silanol groups) and a thin layer of adsorbed water is present. It is believed that a
trichlorosilane head group arriving on the surface from the gas phase becomes hydrolysed once the
molecule gets close enough to the oxide surface, where it forms hydrogen bonds to the surface and
neighbouring silane groups. By means of water elimination, this unstable situation then leads to the
formation of a network in which each chain head is covalently linked to the surface and neighbouring
silane groups[Silberzan_1991].
In this chapter, the static water contact angles of a few materials of relevance to this thesis (silicon
wafer, glass, polydivinylsiloxane (PROVILnovo) and epoxy) were measured before and after exposure
to a perfluorinated trichlorosilane.
4.2.2. Experimental

Static water contact angle measurements:

20

All measurements were performed on a simple Ram-Hart contact-angle goniometer (Ram Hart
model 100, Ram Hart Inc., USA). Prior to air-plasma treatment and after exposure to the
perfluorinated silanes, static contact angles were measured with 6 l drops of high-purity water (18.3
M, EASYpure, Barnstead, USA) on two different spots per sample.

Substrates:
A single-side-polished silicon wafer piece (Si-Mat Silicon Materials, Germany) was cleaned with
toluene and ethanol (see 4.1.2) to remove glue residues arising from the cutting process. A microscope
slide (Menzel GmbH & Co KG, Germany) was measured directly from the box. Two negatives
(imprints) made from a commercial polydivinyl siloxane species (PROVILnovo, Light C.D. 2, fast set;
Heraeus Kulzer GmbH, Germany) were prepared from a freshly cleaved mica surface. One of the
negatives was used for contact-angle measurements; the other was used as a mould to cast a replica
from an epoxy blend (EPO-TEK 302-3, Epoxy Technology, USA). The replication technique is more
closely described in chapter 5 and 6.2.

Surface functionalisation:
All substrates were exposed to air plasma (RF Level high, 0.1 torr, PDC32G, Harrick Plasma, USA)
for 2 min and then put in a desiccator. The humidity of the air formed the necessary silanol groups for
the reaction. A small glass vial containing 1H, 1H, 2H, 2H-Perfluorooctyltrichlorosilane (ABCR
GmbH, Germany) was placed in the desiccator. A gentle vacuum was applied and the SAM was
allowed to form overnight. Contact angles were measured immediately after removal from the
desiccator.
4.2.3. Results and Discussion
The results of the static water-contact-angle measurements are summarised in Table 3. Since all
surfaces are fully wettable after plasma treatment, Before denotes measurements on the native
surfaces as received, while After refers to the samples after functionalisation. Only the two surfaces
with a silicon oxide layer (glass and silicon wafer) show a distinct change in surface energy. Their
contact angle increases to ~110. This value is commonly found on CF2-rich surfaces[Shafrin_1964];
therefore it can be assumed that the monolayers formed by this experimental procedure did not lead to
crystalline structures, exhibiting mostly their CF3-end group, which would correspond to a contact
angle of 120[Shafrin_1964]. This conclusion is supported by the fact that alkylsilanes of short alkyl
chainlengths (i.e. C10 or lower) are known to be in a liquid-like state on the surface[Chaudhury_1995].
The main component of the commercial material PROVILnovo is a polydivinylsiloxane, a silicone
species closely related to PDMS (whose contact angle data on the native and the perfluorinated surface
is displayed in chapter 4.3). Therefore, it was interesting to see whether it could be modified in a

21

similar way to PDMS. But even by its colour, opaque green instead of transparent, it is clear that it is a
complex mixture of many components. It is commonly used by dentists to get an imprint of the
dentition. Due to its low viscosity, short curing times (minutes) and ability to be removed from any
surface, it is a useful tool for replication of any surface of interest. When depositing a drop of water on
the PROVILnovo substrate, the contact angle changed from hydrophobic to ~60 within seconds, most
probably revealing the reason why it does not stick to any surface: it contains a high concentration of
surfactants. The surfactants can also be made evident by immersing a PROVILnovo substrate in water
and shaking it: A foam is generated.
Plasma treatment and exposure to the perfluorinated silanes did not change this behaviour upon
contact with water.
The case is slightly different with the epoxy sample. The epoxy surface does show a stable static
contact angle at ~73. After functionalisation the contact angle starts, very similar to the
PROVILnovo, at a hydrophobic value and then it decreases to slightly below its initial contact angle.
This could be interpreted that, indeed, a layer of perfluorinated silanes was physisorbed on the epoxy,
but reacted immediately upon contact with water. Therefore, it can be concluded, with the chosen
experimental setup, epoxy surfaces cannot be functionalised with trichlorosilanes via the vapour
phase. This finding increases the understanding of the situation when photolithographic structures
made of SU-8 (epoxy polymer) are used as a template for PDMS replication. These templates consist
of SU-8 patterns on a silicon wafer. If such a template were to be used for PDMS replication without
further treatment, the PDMS would react with the substrate and could not be removed anymore.
Therefore it is common to functionalise the template with perfluorinated silanes. However, as is seen
here, these only stick covalently to the bottom of the template, where the silicon surface is exposed.
The fact that PDMS removal still works would lead to the conclusion that the silicone either does not
stick to the epoxy or the silane layer also works as a mould release agent when it is only weakly
adsorbed.
Table 3: Static water contact angle measurements on different substrates before and after fluorosilanisation.

Substrate / water s

Before

After

Si-wafer

64 2

111 1

Microscope Slide

35 3

107 3

PROVILnovo

61 6 *

57 5 *

Epoxy

73 3

68 2 *

22

* all these contact angles started initially at a contact angle > 90, but decreased within approximately
60 seconds to the values stated here .
4.2.4. Summary
In this short study, four different smooth surfaces were activated by air plasma and then exposed to a
vapour of perfluorinated silanes. It was found that with this treatment silicon wafers and glass slides
can be hydrophobised. The contact angle is equal to a contact angle expected for a disordered, CF2rich surface. The commercial PROVILnovo material reacts with water and cannot be modified by
trichlorosilanes. Contact-angle measurements on epoxy revealed that the silanes only adsorb weakly
on the substrate.

4.3.

PDMS (Polydimethylsiloxane)

4.3.1. Introduction
Cured PDMS is a polymeric elastomer consisting of an inorganic backbone with alternating silicon
and oxygen atoms. To each silicon atom two methyl groups (CH3-groups) are bound (see Figure 12).

Figure 12: Simplified chemical structure of cured PDMS.

Ambiguity of PDMS:
This structure imparts a certain ambiguity to PDMS; the silicone backbone is rather hydrophilic
whereas the methyl-side chains are hydrophobic. When exposed to air, the methyl groups of the native
PDMS form a close-packed layer to shield the high-energy silicone backbone. Upon contact with
water, the chains reorient such that the hydrophilic silicone chain is facing the polar
liquid[Morra_1990]. Thus, the high flexibility of the loosely cross-linked backbone allows the
polymer to adopt the conformation leading to the lowest energy state. If the PDMS faces air, which
can be considered as a very hydrophobic environment, the CH3-groups will face the atmosphere. If the
PDMS is in contact with the highly polar water then the silicone backbone will favourably be at the
interface. The methyl groups bend away from the surface normal[Chen_2004]. This process is fast and
occurs already during the course of a dynamic contact angle measurement[Morra_1990], leading to a
large hysteresis.
However, this is not the only ambiguity occurring with PDMS. One of the preconditions for common
contact-angle measurements is that the solid has to be rigid. The rigidity of the surface has to balance
the force in y-direction of the liquid-air surface tension, LA-Y. Therefore if a drop is placed on another

23

immiscible liquid, it assumes the shape of a floating lens[deGennes_2004], instead of a drop with a
flat base. If placed on a soft solid such as a hydrogel or an elastomer with a small E-modulus E, LA-Y
can pull up a ridge around the edge of a drop (see Figure 13). The height h of this ridge is given by the
formula developed by Shanahan and de Gennes[Extrand_1996, Shanahan_1995, Shanahan_1987]:

3 LA sin
E

(6)

denotes the apparent contact angle. By measuring dynamic contact angles (tilting method),
Extrand[Extrand_1996] investigated the maximum E-modulus of elastomers for which such a ridge is
formed. To this end he prepared a series of natural and polybutadiene rubbers with different crosslinking densities to modify the E-modulus. He found that the limit for contact-angle-influencing ridge
formation lies at an E-modulus of about 5 MPa. The ridge is a local increase of roughness and
therefore it increases the advancing and decreases the receding contact angle, thus increasing
hysteresis. The PDMS used here has an E-modulus of 1.448 MPa[Dusseiller_2005], i.e. well below 5
MPa. The ridge height h would be 146 nm. It is generally expected that feature sizes above 100 nm
have an influence on the contact angle[Bonn_2009].

Figure 13: Sketch of a drop on a soft elastomer. The y-direction of the liquid-air surface tension LA-Y pulls up a
ridge of height h around the edge of the drop.

Oxygen Plasma Treatment of PDMS:


By exposing PDMS to an air or oxygen plasma, a superficial silica-like phase is
formed[Hillborg_2001]. This brittle, glassy layer increases in thickness with plasma exposure time
until it reaches its maximum thickness, which is around 100-150 nm. The exposed surface exhibits a
high surface energy, leading to full spreading of water. Thus, upon contact with the humidity of the
air, the high-energy surface of the plasma-treated PDMS will immediately form a wet silanol layer, as
is necessary for the reaction with silane molecules[Silberzan_1991]. PDMS is known to recover its
hydrophobicity over time by conformational changes, condensation of silanol groups[Morra_1990]
and the diffusion of low molar mass PDMS species to the surface[Hillborg_2001]. The low-molar-

24

mass species are residues from the polymerisation process and are formed during plasma treatment
due to chain scission. This recovery is very slow as long as the silica-like layer remains closed. Upon
introduction of cracks, recovery is much faster.
Even though the recovery restores a low-energy surface, it is not equivalent to a native PDMS surface.
The brittle, silica-like layer is still present, only now covered by low-molecular-weight species. The
chains have lost their mobility, and thus contact-angle hysteresis is smaller on a recovered PDMS
surface[Morra_1990].
Extraction:
The silicone elastomer kit that was mostly used in this work (Sylgard 184) consists of a base and a
curing agent. Both components consist of numerous silicone species and solvents. A full description of
the single compounds can be found in the thesis of Raphael Heeb[Heeb_2009].
In this commercial elastomer kit, reinforcing fillers and uncrosslinked silicone species with high
mobility are to be found. These are known to diffuse from the bulk to the surface. By using a strongly
swelling solvent[Lee_2003] these species can be largely removed from the polymer.
4.3.2. Experimental

Substrates:
The two components of the Sylgard 184 elastomer kit (PDMS, SYLGARD 184 silicone elastomer,
base and curing agent, Dow Corning, USA) were mixed in an oil-to-hardener ratio of 10 : 1. The air
bubbles introduced into the mixture due to vigorous stirring with a glass rod were removed by
evacuation in a desiccator. Then the mixture was cast into two different moulds: a single-use
Polystyrene Petri dish (TC Dish 92x17, 150350, NuncTM, Thermo Fisher Scientific, Denmark) and a
perfluorinated (see below) single-side polished silicon wafer (Si-Mat Silicon Materials, Germany) of 2
by 2 cm2. Curing took place at 70 C overnight., The cured PDMS was cut into pieces of 1 x 3 cm2
with a razor blade. Most samples were measured on the surface facing the petri-dish or the
perfluorinated silicon wafer during curing. For one set of measurements the side facing the air was
also used.
A set of samples was functionalised with a monolayer of perfluorinated silanes.
Perfluorination:
The silicon wafer was cleaned with toluene and ethanol before usage to remove glue residues from the
cutting step. The PDMS samples were used directly after cutting, without further treatment. The
conditions for perfluorination were the same for all substrates. They were first treated with an air
plasma (RF Level high, 0.1 torr, PDC-32G, Harrick Plasma, USA) for 2 min and then loaded into a
desiccator. A small glass vial with a small amount of 1H, 1H, 2H, 2H-Perfluorooctyltrichlorosilane

25

(ABCR GmbH, Germany) was also placed into the desiccator. Then, a low vacuum was applied. SAM
formation via the vapour phase was allowed to take place for 1 h.
PDMS extraction:
Low-molecular weight species were extracted from four PDMS samples. The samples were placed in
a piranha-cleaned glass vessel and covered by a large amount of hexane. Hexane strongly swells the
samples and thus extracts non-bound species. The hexane was exchanged after 3.5 and 6 h. After 24 h
the samples were rinsed with hexane and loaded into a desiccator. With a gentle vacuum the absorbed
hexane was removed from the PDMS samples. Contact-angle measurements or fluorination were
performed after another 24 h, to allow residual traces of hexane to evaporate under ambient
conditions.
Cleaning methods:
Intensive rinsing with ultra-pure water (>18.2 M, TKA-GenPure, TKA, Germany) and high-purity
ethanol (analytical grade, Scharlau, Spain) was undertaken, holding the samples with tweezers such
that no liquid could flow back from the tweezers on the sample and possibly contaminate it. The
samples were measured immediately after drying under a stream of nitrogen. A few samples were
ultra-sonicated (Sonorex TK30, Bandelin, Germany) in ethanol for 10 min in piranha-cleaned
glassware. After removal, the samples were rinsed with ethanol and dried with nitrogen. An overview
of the cleaning methods and treatments to the samples is given in Table 5.
Contact-angle measurements:
All measurements were performed on the DSA 100 and evaluated with the DSA3 software (Krss
GmbH, Germany). On each substrate, three 9 l drops were used to measure static and roll-off angles.
The stated static contact angles and roll-off angles are an average over three drops. Dynamic
measurements were conducted with a dosing speed of 15 l/min. For the advancing contact angle, two
movies were recorded, for the receding only one. For the advancing measurement, the initial volume
was 5 l and was increased twice by 4 l to a 13 l drop. After recording the receding movement, 15
l were deposited from the syringe to ensure a high water quality for the next measurement. Only
contact angles where the contact line was actually moving were considered. Roll-off angle
measurements were performed at the slowest tilting speed, which is 1.2 /sec.
4.3.3. Results and Discussion

The aim of this chapter was to learn what influences water-contact-angle measurements on PDMS. In
Figure 14a, dynamic and static contact angles were measured seven times on the very same spot. Even
though it is always better to measure on a fresh spot, the water does not seem to alter the surface such
that it would lead to a significant change in contact angle. For Figure 14b, contact angles were

26

measured on different spots over time after removal from the Petri dish. No change in contact angle
can be observed over the time span of two days.

Figure 14: Repeatability of contact angle measurements on native PDMS. a.) 7 subsequent measurements on the
same spot; b.) influence of time after exposing the sample to air ; measured on the same sample but on different
spots.

The PDMS was also allowed to cure against different mould (negative) materials. Their dynamic
contact angles are summarised in Table 4.
Table 4: Dynamic water contact angles evaluated on the three different mould materials.

Material

Petri dish

76 1

13 2

Si-wafer

114 1

98 1

Air

180

180

Figure 15 shows the dynamic measurements from the native PDMS as a function of the mould
material. The advancing contact angles are very similar for the three moulds, no difference can be seen
here. The case looks very different when comparing the receding contact angles. The PDMS sample
which was cured against the rather hydrophilic Petri dish shows a receding contact angle which is 23
lower than the sample that was cured against air. It is likely that the surface energy of the phase facing
the PDMS during curing has an influence on the superficial microstructure of the PDMS. Low energy
surfaces such as air or the perfluorinated silicon wafer will cause the CH3-groups to orient along the
surface normal, shielding the hydrophilic silicone backbone; on a high-energy surface the opposite
occurs. Even though reorientation of the chains upon contact with water is fast, the initial order of the
polymer chains might play a role for contact-angle measurements. Nevertheless, the results are not

27

unambiguous, since it is expected that the roughness of the moulds decreases from the petri dish to the
silicon wafer and then to air. Reduction in roughness invariably also leads to a reduction in hysteresis.
During the course of several measurements it was also recognised that receding contact angles of
samples cured against a Petri dish are not always identical. Usually they are around 73 75, but also
values as low as 65 were measured. Considering the influence of the E-modulus on ridge formation it
is very well possible that slight variations in the E-modulus strongly influence the receding contact
angle (as it can happen e.g. by using an older, thus less effective hardener, leading to a lower crosslinking density and E-modulus).

Figure 15: Dynamic contact angle measurements as a function of the mould (negative) material during curing of the
PDMS. Only the receding contact angle shows an influence of the mould on the contact angle. The finding is
somewhat ambiguous, because both surface energy and mould roughness decrease from the Petri dish to air.

In the following section it was tried to define how different treatments influence the contact angle. If
this study would have been performed only by measuring static contact angles, no conclusions could
have been drawn except that none of the treatments has any influence on the PDMS. All static contact
angles s scatter around ~105 2, and no influence of the extraction or the perfluorination could be
detected. This experiment is therefore a very good example to emphasise the fact (see 4.1.1) that static
contact angles only carry a limited amount of information.
In Table 5 all treatments carried out and their corresponding advancing and receding contact angles
are presented. For the native PDMS (left column, without perfluorinated layer), the advancing contact
angle scatters reliably around 120 and is therefore independent of the treatment. Solely the extracted
PDMS shows a slightly higher a, which could be interpreted in terms of slightly higher surface
roughness after extraction. This assumption is supported by the receding contact angle, which is about
10 lower than on the unextracted PDMS samples. Surface roughness leads to an increase of the
advancing and a decrease of the receding contact angle due to pinning events (see more in chapter 6).
Rinsing with water or ethanol does not change the wetting properties of PDMS.
Perfluorinated PDMS samples usually show a lower a and a higher r compared to the native PDMS.
At first, the lower a for the perfluorinated species may appear surprising, since it is generally

28

assumed that perfluorinated surfaces show higher contact angles than surfaces consisting only of
hydrocarbons[Shafrin_1964]. However, this observation confirms the fact that the native and
perfluorinated surfaces are very different in nature. On the native surface, the surface tension of water
is able to pull up a rim at the edge of the drop, thus increasing the advancing contact angle. This effect
is a lot more difficult to achieve on the plasma-treated and perfluorinated PDMS. The superficial,
brittle silica-like layer is a lot stiffer (higher E-modulus) than the bulk PDMS, thus the surface tension
of the water is not strong enough to pull up the ridge. Additionally, the chains of the native PDMS are
free to conform to the phase with which they are in contact. This freedom is lost for the perfluorinated
PDMS substrates, resulting in higher receding contact angles.
Rinsing seems to slightly influence the receding contact angle. Water reduces, while ethanol increases
the r.
The increase in contact angle upon ethanol rinsing is somewhat contradictory to the decrease in r for
the sample that was ultrasonicated in ethanol (no. 4). One could speculate that it may indicate that by
rinsing with ethanol, physically adsorbed silane residues are washed away, thus leading to a smoother
surface. However, upon ultrasonication roughness might be introduced by removal of parts of the
brittle silica-like layer. Contact angles on the ultrasonicated perfluorinated substrate (no. 4) are similar
to those on the sample that was extracted and then ultrasonicated in ethanol (no. 6), which could
indicate a similar surface roughness.
The data for the extracted and perfluorinated sample (no. 5) are in brackets because due to an
experimental error this sample could only be measured four days after preparation. The high hysteresis
(compared to the other perfluorinated samples) may indicate that a hydrophobic recovery of low
molecular species through the cracks in the silica-like layer has already occurred. Thus, the contact
angles measured resemble those measured on the native PDMS and can therefore not be reliably used
for interpretation in this study. The numbers were still added into the table to show that pefluorinated
PDMS is not as inert over time as might be expected from other perfluorinated surfaces.
Table 5: dynamic contact angle data evaluated on PDMS and perfluorinated PDMS substrates after different
treatments.

PDMS
a []

perfluorinated PDMS
r []

a []

r []

(1) No treatment

119.9 0.5

75.7 2.6

115.1 1.0

98.2 2.6

(2) Rinsed with water

119.6 0.6

74.6 1.5

115.3 0.8

95.1 2.0

(3) Rinsed with ethanol

120.0 0.8

74.8 1.8

115.8 0.5

102.8 1

29

(4) 10 min US ethanol

117.2 0.3

93.5 0.8

(5) Extracted

120.9 0.7

66.5 1.5

(120.9 0.7)

(81.5 1.3)

115.0 0.4

92.0 1.4

(6) Extracted, 10 min US ethanol

Also roll-off angles were measured on all PDMS samples. No roll-off angle could be detected on the
native PDMS. Reasons for this may be once more the ridge which is pulled up by the surfaces tension
of water, leading to high pinning and the ability of the PDMS chains to conform to the polar nature of
the water, thus increasing the adhesion of the drop to the surface.
All perfluorinated samples showed a roll-off angle (see Figure 16a). Samples that were not extracted
or ultrasonicated showed significantly lower roll-off angles. Rinsing seems to increase the scatter
(standard deviation of samples 1 to 3) in the measured data. Plotting the roll-off angle versus the
hysteresis shows that generally, a higher hysteresis also causes a higher roll-off angle, as expected
from the Furmidge equation (see equation 5).

Figure 16: a.) Roll-off angles measured on the perfluorinated PDMS samples. The numbers on the X-axis denote the
treatment applied to them and correspond to the numbers given in Table 5. b.) plotting roll-off angles versus
hysteresis shows that generally the roll-off angle is higher with increasing hysteresis. The dashed line represents the
linear regression through the data and is meant as a guide to the eye.

4.3.4. Summary and Conclusions

PDMS is a very interesting, but ambiguous material for contact-angle measurements. The advancing
contact angle is very constant between 119 and 121 and nearly unaffected by any treatment. If no
parameter is changed, the receding contact angle is also very constant. When comparing different
conditions, however, the receding contact angle shows a high variability. There is an indication that
the receding contact angle is influenced by the mould material against which the PDMS was cured.

30

The more hydrophobic the mould material, the higher the receding contact angle of the PDMS.
However, in this study, the effect of roughness was not determined and might have an effect leading in
the same direction as surface energy.
The high mobility of the hydrophilic silicone backbone and the hydrophobic methyl groups allow the
material to adopt the lowest energy state by changing conformation. The low E-modulus of PDMS
was identified as another cause to increase the hysteresis on PDMS. PDMS is not rigid enough to
compensate the force of the water-air-surface tension perpendicular to the surface. Therefore, a ridge
is pulled up around the edge of the water drop, locally increasing the surface roughness and thus
hysteresis. Native PDMS can be rinsed with water or ethanol without changing the contact angle.
Extraction of low molecular species reduces the receding contact angle slightly, possibly due to an
increase in surface roughness.
Plasma-treatment and subsequent perfluorination by adsorption of perfluorinated silanes significantly
changes the properties of the surface. The formation of a silica-like layer during plasma treatment
increases the stiffness of the topmost layer, thus inhibiting the formation of a ridge around the drop.
This reduction in roughness causes the advancing contact angle to be slightly smaller after
functionalisation than that of the native PDMS. Additionally the silica formation destroys the
flexibility of the chains. Thus, reorganisation upon contact with the liquid is no longer possible. The
functionalisation with a perfluorinated monolayer reduces the adhesion of water. These reasons lead to
a greatly reduced hysteresis when compared to native PDMS.
The effect of rinsing with water and ethanol is not very clear. It seems to slightly change the receding
contact angle. Ultrasonication and extraction of low-molecular-weight species also seem to reduce the
receding contact angle, thus increasing the hysteresis.

4.4.

Fit routines

4.4.1. Introduction

While fitting the recorded contact angles in the DSA3 program from Krss it was realised that the fit
was not always optimal. Deviation was mainly observed for contact angles above 135. Therefore an
evaluation was performed, as to which of the analysis tools provided with the software would fit a
superhydrophobic drop best. The circle-fitting method fits the drop contour to a circle segment
function. The tangent-method-1 attempts to fit the drop profile by a general conic-section equation,
whereas the tangent-method-2 fits the slope in the vicinity of the three-phase-contact line with a
function of the type (y=a+bx+cx0.5+d/lnx+e/x2) and was generally used in this work. The
YoungLaplace fitting procedure is the most complicated fitting model and takes interfacial and
gravitational effects into account.

31

Additionally, two methods for the program ImageJ were employed. One was a semi-automated
method, the drop-snake analysis[Stalder_2006] (Snake tool) and the other the simple angle tool
provided in ImageJ.
Criticism of contact-angle data stated in literature including many significant figures was already
pointed out by J. Zimmermann[Zimmermann_2008b] and emphasised in the review of Dorrer and
Rhe[Dorrer_2009]. J. Zimmermann evaluated the reproducibility of data extracted by the YoungLaplace method by small changes in the optical settings during the acquisition of the images. He
found that lighting and contrast, setting of the horizontal line and focus had a great influence on the
extracted contact angle value. The very same drop varied then between 166 and 176. His conclusion
was that the automated evaluation is good and reliable for contact angles between 20 and 140, but has
deficiencies for very high contact angles.
4.4.2. Experimental

Images (see Figure 17) of four 9 l drops were recorded on a perfluorinated PDMS substrate with an
f1 (on a porous substrate, areal fraction in contact with the drop, see chapter 2.1.5) of 5.6 %, thus
clearly showing very high contact angles (around 160). Four scientists were asked to evaluate the
images with the ImageJ tools; drop snake analysis and angle tool.
For the evaluation with the DSA3 software (Krss GmbH, Germany), three different sensitivities in
the optical settings were used for the analysis (gradient Threshold value 20, 40, and 60). The circle fit,
tangent-1 and tangent-2 method and the Young-Laplace method were tested.

32

Figure 17: 4 images on a superhydrophobic perfluorinated PDMS substrate as used in chapter 7 at an f1 value of 5.6
%.

4.4.3. Results and Discussion

The extracted contact angles differ a lot depending on the method that was used. The circle fitting was
developed to measure small contact angles and clearly fails to measure such high contact angles. The
two tangent methods give similar results and the Young-Laplace method extracts very high contact
angles (see Figure 18a).
The values extracted from the four scientists with the snake and the angle tool were clearly higher than
those obtained from the tangent methods, but lower than that from the Young-Laplace fit. The
difference between the use of the snake tool and the simple angle tool are negligible (Figure 18b),
therefore for the time-intensive evaluations of the dynamic data (many drop pictures) it was decided to
use the simpler and faster tool, the angle tool. Thus, if interest lies in the apparent contact angle that
can actually be seen, drops with a contact angle above 135 should be evaluated by the use of ImageJ,
because automated fits either under- (tangent methods) or overestimate (Young-Laplace) the apparent
contact angle.

33

Figure 18: Comparison of different evaluation techniques to extract the contact angle out of a drop picture. a.)
automated methods delivered with the program DSA3; b.) semi-automated (Snake tool) and manual tool provided
with the program ImageJ.

4.4.4. Conclusions and Summary

On superhydrophobic surfaces, reliable contact-angle measurements are difficult. Not only do the
optical settings of a contact-angle measurement strongly influence the experimental values of the
contact angles, but also the choice of method. The tangent methods in DSA3 tend to underestimate,
and the Young-Laplace method overestimates the apparent contact angle. The simple angle tool
provided with ImageJ can be used instead to measure contact angles above 135.

4.5.

Acknowledgements

I would like to thank Eva Beurer for the PMIRRAS measurements of the thiol SAMs on gold, which
were used in chapter 4.1.

4.6.

Literature

Adamson, A. W., and Gast, A. P. Physical Chemistry of Surfaces. sixth edition ed. New
York: John Wiley & Sons, Inc., 1997.
Amirfazli, A., et al. "Line Tension Measurements through Drop Size Dependence of
Contact Angle." Journal of Colloid and Interface Science 205, no. 1 (1998): 111.
Blake, T. D. "The Physics of Moving Wetting Lines." Journal of Colloid and Interface
Science 299, no. 1 (2006): 1-13.
Bonn, D., et al. "Wetting and Spreading." Reviews of Modern Physics 81, no. 2 (2009):
739-805.
34

Callies, M., et al. "Microfabricated Textured Surfaces for Super-Hydrophobicity


Investigations." Microelectronic Engineering 78-79 (2005): 100-05.
Cassie, A. B. D. "Contact Angles." Discussions of the Faraday Society 3 (1948): 11-16.
Charnley, M., et al. "Integration Column: Microwell Arrays for Mammalian Cell
Culture." Integrative Biology 1 (2009): 625-34.
Chaudhury, M. K. "Self-Assembled Monolayers on Polymer Surfaces." Biosensors &
Bioelectronics 10, no. 9-10 (1995): 785-88.
Chen, C., Wang, J., and Chen, Z. "Surface Restructuring Behavior of Various Types of
Poly(Dimethylsiloxane) in Water Detected by Sfg." Langmuir 20, no. 23 (2004):
10186-93.
de Gennes, P. G. "Wetting - Statics and Dynamics." Reviews of Modern Physics 57, no.
3 (1985): 827-63.
de Gennes, P. G., Qur, D., and Brochart-Wyart, F. Capillarity and Wetting
Phenomena: Drops, Bubbles, Pearls, Waves. New York: Springer
Science+Business Media, Inc., 2004.
Della Volpe, C., and Siboni, S. "Some Reflections on Acid-Base Solid Surface Free
Energy Theories." Journal of Colloid and Interface Science 195, no. 1 (1997):
121-36.
Dorrer, C., and Ruhe, J. "Some Thoughts on Superhydrophobic Wetting." Soft Matter 5
(2009): 51-61.
Drelich, J., et al. "Wetting Characteristics of Liquid-Drops at Heterogeneous Surfaces."
Colloids and Surfaces a-Physicochemical and Engineering Aspects 93 (1994): 113.
Dusseiller, M. R. "Micro- and Nanoengineering the 3-Dimensional Environment of
Cells in Culture." ETHZ, 2005.
Extrand, C. W. "Contact Angles and Their Hysteresis as a Measure of Liquid-Solid
Adhesion." Langmuir 20, no. 10 (2004): 4017-21.
Extrand, C. W., and Kumagai, Y. "Contact Angles and Hysteresis on Soft Surfaces."
Journal of Colloid and Interface Science 184, no. 1 (1996): 191-200.
Fetzer, R., and Ralston, J. "Dynamic Dewetting Regimes Explored." Journal of
Physical Chemistry C 113, no. 20 (2009): 8888-94.
Fowkes, F. M. "Attractive Forces at Interfaces." Industrial And Engineering Chemistry
56 (1964): 40-52.
Fox, H. W., and Zisman, W. A. "The Spreading of Liquids on Low-Energy Surfaces. Ii.
Modified Tetrafluoroethylene Polymers." Journal of Colloid Science 7, no. 2
(1952): 109-21.
Furmidge, C. G. "Studies at Phase Interfaces. 1. Sliding of Liquid Drops on Solid
Surfaces and a Theory for Spray Retention." Journal of Colloid Science 17, no. 4
(1962): 309-24.
Gao, L., and McCarthy, T. J. "How Wenzel and Cassie Were Wrong." Langmuir 23, no.
7 (2007): 3762-65.
Genzer, J., and Efimenko, K. "Creating Long-Lived Superhydrophobic Polymer
Surfaces through Mechanically Assembled Monolayers." Science 290, no. 5499
(2000): 2130-33.
35

Gerhardt, L. C., et al. "Fabrication, Characterisation and Tribological Investigation of


Artificial Skin Surface Lipid Films." Tribology Letters 34, no. 2 (2009): 81-93.
Graham, D. J., Price, D. D., and Ratner, B. D. "Solution Assembled and Microcontact
Printed Monolayers of Dodecanethiol on Gold: A Multivariate Exploration of
Chemistry and Contamination." Langmuir 18, no. 5 (2002): 1518-27.
Heeb, R. "Surface Modifications for Improved Aqueous Lubrication under LowContact-Pressure Conditions." ETH Zurich, 2009.
Hillborg, H., Sandelin, M., and Gedde, U. W. "Hydrophobic Recovery of
Polydimethylsiloxane after Exposure to Partial Discharges as a Function of
Crosslink Density." Polymer 42, no. 17 (2001): 7349-62.
Installation and Operation Version V1.6-02. Krss GmbH, Hamburg.
Kwok, D. Y., et al. "Measuring and Interpreting Contact Angles: A Complex Issue."
Colloids and Surfaces a-Physicochemical and Engineering Aspects 142, no. 2-3
(1998): 219-35.
Kwok, D. Y., and Neumann, A. W. "Contact Angle Measurement and Contact Angle
Interpretation." Advances in Colloid and Interface Science 81, no. 3 (1999): 167249.
Lee, J. N., Park, C., and Whitesides, G. M. "Solvent Compatibility of
Poly(Dimethylsiloxane)-Based Microfluidic Devices." Analytical Chemistry 75,
no. 23 (2003): 6544-54.
Long, C. J., Schumacher, J. F., and Brennan, A. B. "Potential for Tunable Static and
Dynamic Contact Angle Anisotropy on Gradient Microscale Patterned
Topographies." Langmuir 25, no. 22 (2009): 12982-89.
Love, J. C., et al. "Self-Assembled Monolayers of Thiolates on Metals as a Form of
Nanotechnology." Chemical Reviews 105 (2005): 1103-69.
Maoz, R., and Sagiv, J. "On the Formation and Structure of Self-Assembling
Monolayers. 1. A Comparative Atr-Wetability Study of Langmuir-Blodgett and
Adsorbed Films on Flat Substrates and Glass Microbeads." Journal of Colloid
and Interface Science 100, no. 2 (1984): 465-96.
Morra, M., et al. "On the Aging of Oxygen Plasma-Treated Polydimethylsiloxane
Surfaces." Journal of Colloid and Interface Science 137, no. 1 (1990): 11-24.
Owens, D. K., and Wendt, R. C. "Estimation of the Surface Free Energy of Polymers."
Journal of Applied Polymer Science 13, no. 8 (1969): 1741-47.
Priest, C., et al. "Asymmetric Wetting Hysteresis on Hydrophobic Microstructured
Surfaces." Langmuir 25, no. 10 (2009): 5655-60.
Priest, C., Sedev, R., and Ralston, J. "Asymmetric Wetting Hysteresis on Chemical
Defects." Physical Review Letters 99, no. 2 (2007): 026103-4.
Reyssat, M., and Qur, D. "Contact Angle Hysteresis Generated by Strong Dilute
Defects." The Journal of Physical Chemistry B 113, no. 12 (2009): 3906-09.
Ruths, M., and Israelachvili, J. N. "Surface Forces and Nanorheology of Molecularly
Thin Films." In Nanotribology and Nanomechanics, 417-515, 2008.
Schumacher, J. F., et al. "Engineered Antifouling Microtopographies - Effect of Feature
Size, Geometry, and Roughness on Settlement of Zoospores of the Green Alga
Ulva." Biofouling 23, no. 1 (2007): 55-62.
36

Shafrin E., G., and Zisman W., A. "Upper Limits to the Contact Angles of Liquids on
Solids." In Contact Angle, Wettability, and Adhesion, 145-57. WASHINGTON,
D.C.: AMERICAN CHEMICAL SOCIETY, 1964.
Shafrin Elaine, G., and Zisman William, A. "Upper Limits to the Contact Angles of
Liquids on Solids." In Contact Angle, Wettability, and Adhesion, 145-57.
WASHINGTON, D.C.: AMERICAN CHEMICAL SOCIETY, 1964.
Shanahan, M. E. R., and Carr, A. "Viscoelastic Dissipation in Wetting and Adhesion
Phenomena." Langmuir 11 (1995): 1396-402.
Shanahan, M. E. R., and De Gennes, P. G. Equilibrium of the Triple Line
Solid/Liquid/Fluid of a Sessile Drop. Edited by K. W. Allen, Adhesion 11. New
York: Elsevier Applied Science, 1987.
Silberzan, P., et al. "Silanation of Silica Surfaces. A New Method of Constructing Pure
or Mixed Monolayers." Langmuir 7, no. 8 (1991): 1647-51.
Spori, D. M., et al. "Beyond the Lotus Effect: Roughness Influences on Wetting over a
Wide Surface-Energy Range." Langmuir 24, no. 10 (2008): 5411-17.
. "Cassie-State Wetting Investigated by Means of a Hole-to-Pillar-Density
Gradient." Langmuir (submitted).
. "Influence of Alkyl Chain Length on Phosphate Self-Assembled Monolayers."
Langmuir 23, no. 15 (2007): 8053-60.
Stalder, A. F., et al. "A Snake-Based Approach to Accurate Determination of Both
Contact Points and Contact Angles." Colloids and Surfaces a-Physicochemical
and Engineering Aspects 286, no. 1-3 (2006): 92-103.
Tadmor, R., and Yadav, P. S. "As-Placed Contact Angles for Sessile Drops." Journal of
Colloid and Interface Science 317, no. 1 (2008): 241-46.
Taylor, M., et al. "Picoliter Water Contact Angle Measurement on Polymers." 23, no.
13 (2007): 6875-78.
Tosatti, S., et al. "Self-Assembled Monolayers of Dodecyl and Hydroxy-Dodecyl
Phosphates on Both Smooth and Rough Titanium and Titanium Oxide Surfaces."
Langmuir 18, no. 9 (2002): 3537-48.
VanOss, C. J., Chaudhury, M. K., and Good, R. J. "Interfacial Lifshitz-Vanderwaals
and Polar Interactions in Macroscopic System." Chemical Reviews 88, no. 6
(1988): 927-41.
Venkatamaran, N. V., Zrcher, S., and Spencer, N. D. "Order and Composition of
Methyl-Carboxyl and Methyl-Hydroxyl Surface-Chemical Gradients." Langmuir
22, no. 9 (2006): 4184-89.
Walther, F., et al. "Stability of the Hydrophilic Behavior of Oxygen Plasma Activated
Su-8." Journal of Micromechanics and Microengineering 17, no. 3 (2007): 52431.
Whitesides, G. M. "The Origins and the Future of Microfluidics." Nature 442, no. 7101
(2006): 368-73.
Zimmermann, J. "Silicone Nanofilaments as Functional Coatings:Properties,
Applications and Modifications." Universitt Zrich, 2008.

37

You might also like