You are on page 1of 12

Downloaded from http://rsif.royalsocietypublishing.

org/ on May 5, 2016

Enhanced adhesion of bioinspired


rsif.royalsocietypublishing.org
nanopatterned elastomers via colloidal
surface assembly
Sabine Akerboom1, Jeroen Appel1, David Labonte2, Walter Federle2,
Research Joris Sprakel1 and Marleen Kamperman1
1
Cite this article: Akerboom S, Appel J, Laboratory of Physical Chemistry and Colloid Science, Wageningen University, Dreijenplein 6,
Labonte D, Federle W, Sprakel J, Kamperman 6703 HB Wageningen, The Netherlands
2
Insect Biomechanics Workgroup, Department of Zoology, University of Cambridge, Downing Street,
M. 2015 Enhanced adhesion of bioinspired Cambridge CB2 3EJ, UK
nanopatterned elastomers via colloidal surface
assembly. J. R. Soc. Interface 12: 20141061. We describe a scalable method to fabricate nanopatterned bioinspired dry
http://dx.doi.org/10.1098/rsif.2014.1061 adhesives using colloidal lithography. Close-packed monolayers of polystyrene
particles were formed at the air/water interface, on which polydimethyl-
siloxane (PDMS) was applied. The order of the colloidal monolayer and the
immersion depth of the particles were tuned by altering the pH and ionic
Received: 23 September 2014 strength of the water. Initially, PDMS completely wetted the air/water interface
outside the monolayer, thereby compressing the monolayer as in a Langmuir
Accepted: 17 October 2014
trough; further application of PDMS subsequently covered the colloidal
monolayers. PDMS curing and particle extraction resulted in elastomers
patterned with nanodimples. Adhesion and friction of these nanopatterned
surfaces with varying dimple depth were studied using a spherical probe as a
Subject Areas: counter-surface. Compared with smooth surfaces, adhesion of nanopatterned
biomimetics, nanotechnology surfaces was enhanced, which is attributed to an energy-dissipating mechanism
during pull-off. All nanopatterned surfaces showed a significant decrease in
friction compared with smooth surfaces.
Keywords:
bioinspiration, adhesion, colloidal lithography,
self-assembly, nanopatterned surfaces

1. Introduction
Many animals have evolved adhesive organs on their feet enabling them to
Author for correspondence: strongly attach to and quickly detach from nearly any kind of surface. These
Marleen Kamperman organs come in two basic designs: (i) pads densely covered with micro- or nano-
e-mail: marleen.kamperman@wur.nl sized fibrils (hairs) with a wide variety of tip shapes, observed in, for example,
spiders, geckos, beetles and flies, and (ii) pads with macroscopically smooth
surface profiles, observed in, for example, tree frogs, ants and grasshoppers [1].
Regardless of the design, both types use non-covalent surface forces to achieve
adhesion, and research suggests that they rely primarily on van der Waals
forces [2]. The surface structure, not the chemistry, is therefore dominating the
function of natural adhesive systems.
Over the last decade, synthetic patterned surfaces have been developed to
mimic the geometry of natural attachment systems [3,4]. Initial research focused
on the development of vertical pillars with diameters in the micrometre range
[5 7]. To fabricate more complex structures, closer in design and performance
to natural systems, relatively expensive semiconductor technologies and multi-
step patterning techniques have been used [8 13]. Alternatively, wrinkling
techniques have been explored as a scalable, potentially cheaper production
method. Wrinkled surfaces have been shown to increase or decrease adhesion
depending on the testing geometry and wrinkle dimensions [1418]. Ideally,
however, scalable methods should not be limited to two-dimensional patterns,
as more complex, three-dimensional sub-micrometre structures would enable
Electronic supplementary material is available the fabrication of surfaces with different property profiles [19].
In this study, we describe a novel approach based on colloidal lithography
at http://dx.doi.org/10.1098/rsif.2014.1061 or
to make bioinspired, patterned, dry adhesives. Colloidal lithography is a techni-
via http://rsif.royalsocietypublishing.org. que that uses self-assembled colloidal monolayers for fabricating patterned

& 2014 The Author(s) Published by the Royal Society. All rights reserved.
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

surfaces with nanometre precision. This technique has proved The product was filtered over glass wool to discard 2
to be a fast and cost-effective alternative to other nanofabrica- some coagulum that was formed during the reaction. The

rsif.royalsocietypublishing.org
tion processes such as electron beam lithography or focused filtered particles were treated using two different protocols.
ion beam lithography [20]. Recently, several groups have Colloidal dispersion CP (carboxylated particles) was directly
reported methods to produce large-area, highly ordered, col- exchanged to ethanol by three cycles of centrifugation at
loidal crystals. For example, Pan et al. [21] introduced a 4500 r.p.m. for 1 h followed by resuspension in ethanol.
dynamic interface, by stirring the aqueous subphase, to facili- Colloidal dispersion CP-W (carboxylated particles
tate colloidal crystal formation at the air/water interface. washed) was washed in water by three cycles of centrifugation
Retsch et al. [22] reported a two-step procedure in which at 4500 r.p.m. for 1 h followed by resuspension in Milli-Q
colloids are applied to a glass slide by spin-coating. water. After this washing step, CP-W was exchanged to
Subsequently, a colloidal monolayer was formed upon immer- ethanol, using the same procedure. The smaller particles
sion of the glass slide in an aqueous surfactant solution [22]. were treated similarly to CP-W.

J. R. Soc. Interface 12: 20141061


Assembly at the air/water interface was also successfully
achieved by using interfacial deposition via a glass slide [23]
and a needle tip [24]. In these examples, the liquid interface 2.3. Particle characterization
acts as the assembly scaffold after which the monolayers are Static light-scattering measurements were performed on an
transferred to a solid substrate for further processing. Alterna- ALV instrument equipped with an ALV7002 external correlator
tively, material can be deposited via chemical reactions directly and a 300 mW Cobolt Samba-300 DPSS laser operating at a
onto the colloidal monolayer at the air/liquid interface. This wavelength of 532 nm. The scattering intensity was recorded
method was used to fabricate free-standing porous films [25] at 158  u  1208 in intervals of 28, and the curve was fitted
and nanobowls [26], by reacting monomers or inorganic precur- to a form factor of a solid sphere. From this measurement, an
sors at the gas/liquid and liquid/solid interface, respectively. average radius of 574 nm and a polydispersity index (Rw/Rn)
Here, we fabricated nanopatterned polydimethylsiloxane of 1.01 were obtained (see the electronic supplementary
(PDMS) surfaces, using colloidal lithography directly at the material, figure S1a).
air/water interface. We show that close-packed monolayers The surface charge of the particles was determined by titra-
at the air/water interface can be used as a template without tion of 50 ml colloidal dispersion in Milli-Q water (26 g l21 for
the need to transfer the monolayer to a solid substrate. By omit- CP and 11 g l21 for CP-W) with 0.1 M NaOH and using the
ting transfer and drying steps in the process, pattern disruption titration of 50 ml Milli-Q water (brought to low pH with 2 ml
was minimized. Colloidal monolayers were obtained for a 1 M HCl) as a reference (see the electronic supplementary
range of particle sizes. Moreover, the subphase was used to material, figure S1b). The difference in NaOH required to
systematically tune the order of the colloidal monolayer and reach a pH of 9.4 (well above the pKa values of itaconic acid,
to change the immersion depth of the particles. This led to which are pKa1 3.85 and pKa2 5.44 [28], but below the
direct control over the topography of PDMS surfaces. In plateau in the pH curve) between the dispersion and the refer-
addition, the adhesion and friction of the nanostructured ence sample is taken as the number of itaconic acid groups in the
surfaces were measured, demonstrating that structuring sample. The weight per particle is calculated using the density
resulted in enhanced adhesion and decreased friction. of polystyrene. As the weight of all particles during titration is
known, the number of charges per particle can be calculated.
The parking area is then (particle surface area)/(number of
charges per particle) and is 0.11 nm2 per charge for both CP
2. Experimental section and CP-W.
2.1. Materials
All materials were purchased from Sigma Aldrich and used as
received. PDMS (Sylgard 184) was mixed in a ratio of 1 : 10 2.4. Surface structure fabrication
of cross-linker: prepolymer by weight and degassed by cen- Samples were prepared in polystyrene Petri dishes. This pro-
trifugation. Deionized water, from Barnstead Easypure UV cedure is shown in figure 1. The pH of the subphase was
system (Thermo Scientific), was used in all experiments. varied between 1 and 11, keeping the ionic strength constant
at either 1023 or 0.1 M. A piranha-cleaned coverslip was
placed diagonally against the side, and a suspension of
2.2. Particle synthesis particles in ethanol (20 wt%) was applied via this coverslip
Carboxylated polystyrene particles were synthesized in a onto the water subphase using a pipette (figure 1a(i)). The
one-step synthesis according to [27]. Briefly, for the biggest area covered by the particles at the air/water interface was
particles, water (120 g), itaconic acid (0.5 g) and styrene easily visible, and thus the amount of particles could be
(24.5 g) were heated to 808C in a round bottom flask and adjusted during preparation. The monolayer was left to equili-
flushed with nitrogen for 45 min. Meanwhile, 4,40 -azobis brate for at least 1 h before further processing, to ensure
(4-cyanovaleric acid) (ACVA; 252.4 mg) was dissolved in evaporation of all ethanol from the air/water interface.
water (6.3 g) and 1 M NaOH (2 g). After adding this mixture A syringe with a blunt tip needle was filled with PDMS
to the round bottom flask, the solution reacted overnight and a small droplet was applied to the side of the Petri dish
under continuous stirring at 500 r.p.m. containing the monolayer of particles. Upon reaching the
The smaller particles were synthesized using 125 ml water surface, the droplet spread quickly (figure 1a(ii)). Next,
water, 5 or 10 g styrene and 0.1 g (for 5 g styrene) or 0.2 g PDMS was applied to the complete circumference of the Petri
(for 10 g styrene) itaconic acid, and a temperature of 858C. dish. This resulted in a ring of PDMS on water. By slowly
The reaction was initiated with 0.05 g ACVA (for 5 g styrene) adding more PDMS to this ring in a spiralling manner,
or 0.1 g ACVA (for 10 g styrene). PDMS eventually covered the whole interface. PDMS was
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

(a) (b) 3
(i)

rsif.royalsocietypublishing.org
water polystyrene

(ii)
itaconic
acid

J. R. Soc. Interface 12: 20141061


water

(c)
(iii)

PDMS

water

(iv)

PDMS

Figure 1. Fabrication of nanopatterned PDMS using colloidal lithography directly at the air/water interface. (a) Schematic outline of the experimental procedure.
(i) Spreading of the particles using a glass slide. (ii) Application of the PDMS against the side of the Petri dish. (iii) Cross-linking of the PDMS. (iv) Nanopatterned
PDMS prepared by extraction of the particles. (b) Sketch of the charge distribution on a particle when submerged in water at sufficient pH. (c) Photograph of a
PDMS sample with a colloidal monolayer. Note the clear PDMS spot in the upper left corner, where the PDMS first contacted the water, thereby pushing the particles
away. (Online version in colour.)

allowed to cross-link at room temperature for 2 days, resulting exactly six nearest neighbours in these hexagonal lattices.
in PDMS layers of 0.20.6 mm thickness (figure 1a(iii)). This method of assessing crystallinity is very strict: for
To remove the template particles from the PDMS example, all six particles surrounding a vacancy are not
mould, we first applied Scotch Magic Tape, which removed regarded as crystalline, while in fact they are part of the crystal
the largest portion of templating particles. The remain- lattice. However, assessing the crystallinity by using the more
ing particles were removed by soaking the sample in elaborate Steinhardt orientational bond order parameter, C6
N-methyl-2-pyrrolidone (NMP) for 1 h, followed by dipp- [30], showed the same trends as simply counting the number
ing the samples 30 times in NMP in an ultrasonic bath of nearest neighbours [31]. Therefore, the nearest neighbour
(figure 1a(iv)). NMP was chosen as the solvent because it dis- method was used here.
solves polystyrene, yet does not swell PDMS. The samples
were stored at room temperature.
2.6. Immersion depth determination
We put Petri dishes with monolayers in a desiccator, added
2.5. Crystallinity determination some droplets of ethyl cyanoacrylate in a separate dish and let
Monolayers were scooped from the air/water interface with the system react under reduced pressure for 2 h. Ethyl cyano-
wet, piranha-cleaned coverslips. The area of the monolayers acrylate starts to polymerize when it comes into contact
on these coverslips ranged between 0.5 and 2 cm2. Twenty stat- with water. The apparent radii of the particles embedded in
istically independent images were acquired with bright field poly(ethyl cyanoacrylate) was determined from scanning elec-
microscopy at random locations across each sample. The cen- tron microscopy (SEM) images (6878 particles per image)
troid positions of all particles in the images were determined using the Analyze Particles function in IMAGEJ. See the electronic
using standard particle-tracking algorithms [29] with a lateral supplementary material, figure S3a, for a definition of apparent
resolution of approximately 20 nm. Approximately 10 000 par- radius. Six two-tailed t-tests were performed to test whether the
ticles were found per image; per experiment we thus analysed differences in apparent radii were significant (with p , 0.001)
20  10 000 200 000 particles to assess the crystallinity of the and the Bonferroni method was used to correct for multiple test-
monolayers. We then computed the number of nearest neigh- ing. This apparent radius was used to calculate the immersion
bours, using a cut-off distance of 1.25 the particle diameter depth, which is the height percentage of the particle that is
(for Matlab code, see the electronic supplementary material, immersed in the subphase (see the electronic supplementary
S7). We considered a particle crystalline only when it had material, figure S3b).
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

4
glass surface
parallel

rsif.royalsocietypublishing.org
perpendicular XYZ
or normal spherical probe strain gauges translational
stage
sample

Figure 2. Schematic of test set-up for adhesion and friction measurements.

J. R. Soc. Interface 12: 20141061


2.7. Atomic force microscopy between 0 and 5 s. Pull-off forces were measured at nine
The PDMS samples were cut and embedded in Norland Optical different preloads, ranging between 0.2 and 5 mN, and
Adhesives 60 on a silicon wafer. Samples were cleaned with using a motor velocity during pull-off of 1 mm s21, at the
Scotch tape and were imaged using a Nanoscope V in Scan same spot. At least five different areas per sample were
Asyst imaging mode, using non-conductive silicon nitride measured, and all measurements can be found in the elec-
probes (Veeco, NY, USA) with a spring constant of tronic supplementary material, figure S6.
0.32 N m21. Images (2  2 mm) were recorded at 0.990 Hz Friction was measured at least 10 times per sample by
with 512 lines and further processed with NANOSCOPE ANALYSIS applying a normal force of 0.5, 1 and 2 mN for 5 s, followed
v. 1.40 software. This software calculates the area per projected by shearing the probe with a motor velocity of 100 mm s21 for
area by creating a three-dimensional tessellation of the surface. 20 s, while keeping the normal force constant via a 20 Hz
More specifically, the actual area is defined as the sum of the force feedback mechanism. For the friction measurements of
areas of every triangle formed by three adjacent data points flat PDMS, a stiffer bending beam (spring constant in the in-
using the (x, y, z) position of each datum point. Three indepen- plane direction: 27.6 N m21) and a normal force of 0.5 mN
dent atomic force microscopy (AFM) images were used to were used. However, the friction force required to initiate slid-
determine the actual area per sample. After this, the height ing was not met, and thus we can only say that the static
with the highest frequency was subtracted, such that the peak friction force of smooth PDMS is more than 70 mN. S2 could
was positioned at zero, and all values were multiplied by 21, not be measured at a normal load of 2 mN for the same
so that negative values denote negative asperities and positive reason: the friction force required to initiate sliding was not met.
values denote positive asperities. The dimple depth was
determined from the histogram at 0.1 arb. units.
3. Results
2.8. Light microscope 3.1. Surface structure fabrication and characterization
Images were recorded with a Zeiss Axiovert 200 equipped with
Monolayers were prepared as shown in figure 1a. Upon
a 100 oil-immersion objective and a Phantom V91 camera.
applying the particle suspension in ethanol on water, a surface
tension gradient was generated that caused the dispersion to
2.9. Scanning electron microscopy rapidly spread outwards and cover the interface with particles
Samples were sputtered with gold (30 mA, 40 s) using a JEOL [24]. Regions with distinct Bragg reflections formed spon-
JFC-1300 autofine coater and imaged using a JEOL JAMP- taneously, which indicated the formation of close-packed
9500F Field Emission Auger Microprobe. High magnification and organized domains. The self-assembly of charged micro-
and tilt images were measured using a FEI Magellan FESEM. metre-sized colloidal particles at an air/water interface is
governed by a competition between attractive capillary forces
and repulsive electrostatic dipolar forces [34]. We observed
2.10. Adhesion measurements the spontaneous formation of islands of close-packed particles,
Adhesion measurements were performed using the force- which suggests that attractive forces dominated. The charges on
controlled set-up shown in figure 2 as described in [32]. the particles originated from co-polymerization with itaconic
Forces were measured using a bending beam with a spring acid (figure 1b). The charge density increased with increasing
constant in the normal direction of 27.3 N m21, and in the pH, as more itaconic acid groups were deprotonated. The park-
in-plane direction of 9.8 N m21. A spherical Al2O3 probe ing area at pH 9.4 was calculated to be 0.11 nm2 per COOH
with a diameter of 4.76 mm (Edmund Optics) was glued group. This value is relatively small compared with values
with Norland Optical Adhesives 61 to a glass coverslip and in the literature [35], but the substantial ionic strength of the
this coverslip was attached to the distal end of the bending subphase screened the electrostatic repulsion between the
beam. The probe was brought into contact with the PDMS particles, and thus attractive forces dominated nonetheless.
50 times before the measurements started, and not cleaned To use a monolayer of particles at the air/water interface
in between measurements. This avoids variation in pull-off as a template, the deposition material should (i) be immisci-
forces due to transfer of PDMS oligomers between the sur- ble and non-reactive with water, (ii) be able to cross-link in
faces [33]. For all measurements, the probe was moved the presence of water, (iii) be able to replicate micrometre
towards the sample until a predefined preload was reached, to nanoscale features, and (iv) have a lower density than
which took between 0 and 5 s, and the time lag thus varies water. PDMS was selected because it fulfils these criteria
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

pH 3 pH 7 pH 10 5
(a) (b) (c)

rsif.royalsocietypublishing.org
10 m 10 m 10 m

(d) (e) (f)

J. R. Soc. Interface 12: 20141061


(g) (h) (i)

10 m 10 m 10 m

Figure 3. Light microscopy images of CP-W (for a definition, see text) colloidal monolayers self-assembled onto water at (a) pH 3, (b) pH 7 and (c) pH 10.
Crystallinity diagrams obtained from light microscopy images at (d ) pH 3, (e) pH 7 and (f ) pH 10. Particles in light grey have six neighbouring particles and
are considered crystalline, while particles in red (dark grey in printed version) do not have six neighbouring particles. SEM images of colloidal monolayers embedded
in PDMS at (g) pH 3, (h) pH 7 and (i) pH 10. (Online version in colour.)

and is a well-characterized material. PDMS completely wet analysis [41], Fourier transformation [24,42] and microscopy
the air/water interface [36], and thus formed a homogeneous to see grain sizes [43]. In these methods, subtle differences in
film at the interface. During spreading of the initial PDMS dro- order are not easily picked up, while our approach allowed
plet, the particles were pushed together, similar to compression for straightforward comparison of the crystallinity of different
in a Langmuir trough. This resulted in close-packed mono- monolayers.
layers composed of the islands formed during deposition. Figure 3gi shows SEM images of the monolayer
A representative example of nanopatterned PDMS is shown embedded in PDMS. We found that PDMS did not disrupt
in figure 1c, in which the Bragg reflections can be clearly seen. the monolayer, as a similar order was observed in light
The crystallinity of the monolayers was assessed by microscopy and SEM. Figure 4 presents SEM images of the
scooping monolayers, which were not cast in PDMS, and resulting nanopatterned PDMS surfaces after particle removal.
imaging them with brightfield microscopy (figure 3ac). Colloidal monolayers were obtained for a range of particle
We confirmed the presence of a monolayer, rather than multi- sizes, which resulted in nanostructured PDMS with different
layers, by changing the focus and thus scanning through feature sizes (electronic supplementary material, figure S2).
the sample. The positions of all particles in the images were These results indicate that our method allows direct control
found using a particle-tracking algorithm, and this spatial over the topography of PDMS by tuning the properties of the
information was used to quantify the number of crystalline colloidal monolayer.
particles, i.e. particles with exactly six nearest neighbours
(see Experimental section and the electronic supplementary
material, S7). An exemplary result of this analysis is shown
3.2. Influence of subphase
Figure 3 clearly shows that the degree of crystallinity of the
in figure 3d f, where crystalline particles are shown in red
colloidal monolayer was strongly dependent on the pH of
(dark grey in printed version), and non-crystalline particles
the subphase. The sample at pH 3 (figure 3a) was largely
are shown in light grey. The normalized crystallinity per
disordered, while large domain sizes were obtained for
image was then calculated using
samples prepared at pH 10 (figure 3c). Samples prepared
crystalline particles at intermediate pH value ( pH 7) exhibited intermediate
normalized crystallinity : (3:1) order (figure 3b). We then systematically investigated the
total number of particles
degree of order of the samples as a function of the compo-
sition of the subphase for two types of particle dispersions,
This quantitative assessment of the crystallinity of the CP and CP-W (figure 5). The CP dispersion contained
monolayer proved insightful to study the effects of subphase residual styreneitaconic acid oligomers, whereas the CP-W
characteristics on the ordering of particles. Common, more dispersion was thoroughly washed with ultrapure water to
qualitative methods include visual inspection of micrographs remove these oligomers to a great extent. To study the pH
or iridescence of the monolayer [22,26,3739], observation of effect on monolayer formation, the pH was adjusted with
moire patterns upon thin film folding [40], diffraction pattern HCl and NaOH, while keeping the ionic strength constant.
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

(b) 6
(a)

rsif.royalsocietypublishing.org
10 m 10 m

J. R. Soc. Interface 12: 20141061


(c) (d)

10 m 100 m

Figure 4. SEM images of PDMS replicas of colloidal monolayers self-assembled onto water at (a) pH 3, (b) pH 7 and (c,d ) pH 10.

(a) (b) (c)


1.0

0.8
normalized crystallinity

0.6

0.4

0.2

0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
pH pH pH
Figure 5. Degree of order as a function of subphase composition. (a) Normalized crystallinity as a function of pH for CP (triangle) and CP-W (diamond) at [salt]
0.1 M. Every point represents a different sample and standard deviations were obtained from crystallinity determination at 20 different positions within a sample.
(b) Normalized crystallinity as a function of pH for CP at [salt] 0.1 M. Solid symbols denote samples for which the pH of the subphase was constant during
processing. Open symbols denote samples for which the pH of the subphase was kept at pH 10 during self-assembly and adjusted afterwards. (c) Normalized
crystallinity as a function of pH for CP-W at [salt] 1023 M. Solid symbols denote samples for which the pH of the subphase was constant during processing.
Open symbols denote samples for which the pH of the subphase was kept at pH 10 during self-assembly and adjusted afterwards.

The pH was calculated from the known composition of the particles. The increased electrostatic repulsion counteracts
subphase and not determined experimentally, but we the attractive van der Waals and capillary forces. Consequen-
expect no strong deviations between calculated and actual tly, the particles remained more mobile at the interface,
values due to the low particle concentration in the overall resulting in more ordered monolayers [20,22,23]. Thus, at
system. The crystallinity was again determined by scooping pH 3, particles that approach each other closely (i.e. by capil-
monolayers and subsequent light microscopy imaging. lary forces) immediately feel the attractive van der Waals
Figure 5a shows that the crystallinity increased with forces that bring them in close contact. This process immobi-
increasing pH for both CP and CP-W. At higher pH, the lizes the particles and results in monolayers with relatively
charge density at the particle surface was increased. Itaconic low crystallinities. At pH 10, the electrostatic repulsion is
acid groups contain two carboxyl groups, with pKa1 3.85 greatly increased. This increases the energy barrier for a
and pKa2 5.44 [28]. The onset of increasing crystallinity close contact between particles and the particles have suffi-
indeed coincided with dissociation of the first carboxyl cient time and mobility to crystallize in a low free energy
group at pH higher than 4. The increased charge density position. Hence, monolayers with higher crystallinities are
resulted in larger electrostatic repulsion between the obtained [23].
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

It was observed that monolayers made from CP were (a) Milli-Q water (b) pH = 3, [salt] = 103 M 7
lessmobile and mechanically more stable than CP-W mono-

rsif.royalsocietypublishing.org
layers. More specifically, when part of the CP monolayer
was scooped with a coverslip, the hole left in the monolayer
remained square, whereas for CP-W the particles within the
monolayer rearranged after scooping, leaving an irregular-
1 m 1 m
shaped hole in the monolayer. We speculate that the surface
active styrene itaconic acid oligomers left in the CP disper- (c) pH = 10, [salt] = 103 M (d ) [NaCl] = 0.1 M
sions glue the particles together, in a similar fashion as
was observed by Ho et al. [44] for polyethylene oxide poly-
mers. To study the difference in mobility, CP and CP-W
samples were prepared at high pH, which resulted in well-

J. R. Soc. Interface 12: 20141061


ordered monolayers. HCl was added to the subphase to 1 m 1 m
lower the pH while keeping the salt concentration constant.
For CP monolayers, the order was maintained, even when (e) 98
the pH was as low as 2 (figure 5b). By contrast, lowering
the pH of CP-W monolayers decreased the crystallinity sig- 96
nificantly (figure 5c). The cohesive layer between the CP

immersion depth (%)


particles thus played an important role in preserving the 94
order during further processing of the monolayer.
We also altered the properties of the monolayer through 92
changes in ionic strength of the subphase at constant pH.
We determined the immersion depth of CP-W particles in 90
subphases at different pH and ionic strength by fixing their
position at the interface in a polymeric matrix (figure 6). 88
The matrix was formed by interfacial polymerization of A B C D
ethyl cyanoacrylate, which was introduced via the gas 86
phase and started to polymerize when it came into contact
( f ) Milli-Q water (g) pH = 3, [salt] = 103 M
with water. After drying, immersion depths were determined
by imaging the poly(ethyl cyanoacrylate) layer with SEM.
Figure 6a d shows top-view images of monolayers
embedded in the cyanoacrylate matrix, where deeper immer-
sion of the particle resulted in a smaller spherical cap and
1 m 1 m
thus a smaller apparent radius. The boxplots of the immer-
sion depth of the particles in figure 6e show that both
(h) pH = 10, [salt] = 103 M (i) [NaCl] = 0.1 M
salt concentration and pH influenced the immersion depth.
Statistical analysis showed that the differences between
the apparent radius of all four samples were significant
(with p , 0.001). Figure 6b,c shows that an increase in pH
led to a deeper particle immersion. Figure 6a,d also shows
1 m
that an increase in ionic strength (at pH 7, thus above 1 m

the pKa values of itaconic acid) resulted in a deeper immer-


sion, which illustrates that the influence of ionic strength Figure 6. Effect of pH and ionic strength of the subphase on the immersion depth
on immersion depth can be more pronounced than that of particles. SEM images of monolayers embedded in poly(ethyl cyanoacrylate) using
of pH. As discussed earlier, the charge density increases (a) Milli-Q water, (b) pH 3, [salt] 1023 M, (c) pH 10, [salt] 1023 M
with pH, making the particle more hydrophilic, thus lead- and (d) [NaCl] 0.1 M as subphase. (e) Boxplot of immersion depth of particles.
ing to deeper immersion into the water phase. Similarly, SEM images of nanopatterned PDMS surfaces fabricated using (f) Milli-Q
increasing the ionic strength leads to screening of the densely water, (g) pH 3, [salt] 1023 M, (h) pH 10, [salt] 1023 M and
packed charges at the particle surface, and thus enhances (i) [NaCl] 0.1 M as subphase.
deprotonation. This in turn also increases the surface charge
density and thus the immersion depth in the aqueous
subphase [45].
CP-W monolayers using the same four subphases were 3.3. Adhesion
prepared and used as templates for PDMS. Figure 6f i To study the effect of topographical differences on adhesion,
shows SEM images of the resulting nanopatterned PDMS three nanopatterned PDMS surfaces were prepared by
surfaces. Subtle differences in dimple depth were observed, varying the pH of the subphase, as well as a flat control,
yet were not as pronounced as those observed with ethyl cya- and force displacement curves were recorded. We used pH
noacrylate. Upon application of PDMS, the immersion depth values 3, 5 and 10 to obtain samples with different degrees
of the particles may change because of PDMS PS inter- of crystallinity, according to the results shown in figure 5a.
actions, thereby reducing the effect of pH and ionic Sample characteristics are the same as described in the
strength. Nevertheless, the results showed that ionic strength previous sections (see also AFM characterization in the elec-
and pH of the subphase, as well as particle size, can be used tronic supplementary material, figure S4) and summarized
to tune the final topography of the PDMS surface. in table 1. The cross section through the samples was
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

(a) 600 (b) 100 8

rsif.royalsocietypublishing.org
500 86 nm
10 S1
400
1

arb. units
300 135 nm
nm

S2
200 0.1

100 S3
0.01 173 nm
0

100 0.001
0 0.5 1.0 1.5 2.0 250 200 150 100 50 0 50 100 150

J. R. Soc. Interface 12: 20141061


m nm

(c) 0.6 (d ) 1.05 (e) 1.25

0.4 1.00
1.20

normalized pull-off force


0.2 0.95
pull-off force (mN)

1.15
0 0.90
force (mN)

0.2 0.85 1.10

0.4 0.80
1.05
0.6 0.75

0.70 1.00
0.8

1.0 0.65 0.95


0.25 0.20 0.15 0.10 0.05 0 0 1 2 3 4 5 6 flat S1 S2 S3
displacement (mm) preload (mN)

Figure 7. Different (nanopatterned) surface structures and adhesion thereof. (a d) (circle) Smooth PDMS; (diamond) sample 1; (triangle) sample 2; and (square)
sample 3. (a) Cross section from AFM image for smooth PDMS and S1 S3. (b) Histogram of height differences from an AFM picture for smooth PDMS and S1 S3.
(c) Representative force displacement curve of nanopatterned PDMS. Preload is the maximum on the y-axis (5 mN), and pull-off force is the minimum on the y-axis
(0.92 mN). Note the loading curve consists of only 14 data points due to the quick approach. (d ) Pull-off force of smooth PDMS and S1 S3 as a function of preload
with 95% confidence intervals. (e) Normalized pull-off force against normalized actual area for smooth PDMS and S1 S3 with 95% confidence intervals.

Table 1. Sample characteristics of samples S1 S3.

l
a
sample subphase crystallinity a a (nm) l (nm)

S1 pH 10, [salt] 0.1 M 0.54 + 0.04 86 1150


S2 pH 5, [salt] 0.1 M 0.35 + 0.04 135 1150

S3 pH 3, [salt] 0.1 M 0.13 + 0.02 173 1150

a
Determined by AFM.

approximated by hemispherical dimples separated by flat relatively large inter-dimple regions to obtain flat inter-
inter-dimple regions, with a as the dimple depth and l as dimple regions. As expected, upon decreasing the pH, the
the dimple wavelength. The wavelength l is defined as the dimple depth increased from sample 1 (S1; dark grey), to
diameter of the particles for a cross section through the sample 2 (S2; grey) and to sample 3 (S3; light grey) (see
centre of the particles. Actual average wavelengths of the also the electronic supplementary material, figure S4).
samples are a function of the degree of crystallinity and Second, histograms of the height distributions of the three
will deviate slightly from this value. Dimple depths were nanopatterned samples and smooth PDMS were determined
determined from the histogram in figure 7b. by AFM and show that the same trend in depth was also
As expected, the normalized crystallinity differed found over larger areas (5  5 mm). The histogram for
between the three samples and decreased with decreasing smooth PDMS (black) peaks around the normalized zero
pH value of the subphase: 54% for S1, 35% for S2 and 14% (see Experimental section for details) and extends slightly
for S3. Apart from the crystallinity, the dimple depth also to the right, indicating the presence of some small bumps
differed between the three samples, which was determined or dirt particles (figure 7b). The histograms of the nanopat-
by AFM. First, a cross section through the centre of a terned structures show a broad distribution and extend to
dimple of each of these samples was drawn, as shown in the left, indicating dimples. These data were used to
figure 7a. Profiles were created by selecting dimples with determine dimple depths.
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

Force displacement curves were obtained by moving a Substituting parameter values as obtained by Jin et al. [49] 9
smooth spherical probe (of 4.76 mm diameter) against the for PDMS with similar material properties, i.e. E* 1.5 MPa
and Wad 0.163 J m22, we obtain a maximum dimple depth

rsif.royalsocietypublishing.org
sample to a defined preload. Subsequently, the probe was
retracted with a constant velocity of 1 mm s21 until the final of 0.6 mm, which is a larger value than the depths of our
detachment event occurred giving the pull-off force, which samples. Thus, the dimples in this study are most likely to
is a measure of the adhesion performance. A representative be shallow enough to spontaneously jump to contact.
force displacement curve is given in figure 7c. At least Although theoretical considerations suggest that complete
five series of nine force displacement curves with nine contact formation may occur for our dimple dimensions, in
different preloads were measured on at least five different practice air can be trapped in the dimples and prevent complete
areas per sample. The example in the electronic supplemen- contact from being achieved. A difference between our
tary material, figure S5, shows that the pull-off force did dimpled structures and wrinkles or an array of pillars is that
not change within the series. This indicates that we did dimples are isolated and air can get locked in. However,

J. R. Soc. Interface 12: 20141061


not damage our structures during attachment and detach- PDMS is known to have good air permeability, and air will
ment, and thus that our nanopatterned structures can be be removed through the material [52].
used repeatedly. Our results thus support complete contact formation. As
Figure 7d shows the average pull-off forces plotted stated above, for complete contact the true contact area is
against the applied preload for smooth PDMS and for pat- defined by the probe and, therefore, our results cannot be
terned surfaces. The pull-off force values for patterned explained by a difference in contact area. Instead, the increase
surfaces were higher than those for smooth surfaces at in adhesion of our materials may be attributed to a mechanical
almost any preload. Only S2 showed a preload dependency instability resulting from crack trapping as proposed by Guduru
and exhibited a lower pull-off force than the smooth surfaces [53] and observed experimentally for surfaces consisting of con-
at the highest preload. Owing to the relatively low preload centric wrinkles [54]. The waviness of this axisymmetric surface
forces used in this study, and the small dimple size, this structure resulted in a detachment process with stable and
decreased pull-off force is unlikely to be due to mechanical unstable regions. The crack remains stable as it climbs towards
damage. For all other samples, the pull-off force did not the top of a wrinkle. The crack is trapped and an increased
vary significantly with the preload. Normalized pull-off force is needed to move it forward. Just after passing the top,
forces in the preload-independent regime (i.e. less than the crack becomes unstable and quickly propagates past the
5 mN) are plotted in figure 7e. trough. Energy is dissipated during detachment from the
Previous studies have shown that wrinkles and surface unstable regions, resulting in increased separation energy [53].
roughness typically decrease adhesive properties [17,18,46,47], The origin of this dissipation, namely the specific surface struc-
but in some studies adhesion enhancement was found [14,48]. ture, is also present in our materials, and for both concentric
Adhesion enhancement between compliant smooth surfaces wrinkles and dimples the crack can be trapped. By contrast,
and rigid rough surfaces is often attributed to an increase in for two-dimensional wrinkled surfaces, it has been observed
true contact area [47]. As long as the roughness of this rigid sur- that the crack first moves parallel to the wrinkles after which
face is small enough, the positive effect from the true contact it jumps in the direction perpendicular to the wrinkles [17].
area will outweigh the negative effect from the stored elastic Therefore, for two-dimensional wrinkles, this crack-trapping
energy, and adhesion will be enhanced. In this study, the maxi- mechanism only applies to the last part of the detachment
mum true contact area is defined by the smooth rigid probe, and and may explain why no adhesion enhancement was found
hence does not change. for two-dimensional wrinkled surfaces with similar wrinkle
Moreover, full contact was likely to always be achieved dimensions to our dimple dimensions [17].
for our structures. Although we did not observe the contact An alternative energy-dissipating mechanism may be pro-
area directly due to the relatively small size of our surface vided by discrete crack initiation, which has been found for
structures, we conclude that full contact was achieved by small wrinkle amplitudes of 0.55 mm and an aspect ratio of
comparing our surface feature dimensions and material prop- 0.1, but not for slightly bigger wrinkles (amplitude of 3 mm
erties with similar surface structures in previous reports with the same aspect ratio) [17]. For these small wrinkles, the
[16,49,50]. For example, Jin et al. [49] experimentally observed individual wrinkles detach first, during pull-off. This discrete
complete contact independent of preload for PDMS wrinkles crack initiation throughout the surface dissipates energy.
with a wavelength of 27 mm and amplitudes (analogous to Additional energy could thus be dissipated ahead of the
dimple depth) smaller than 0.6 mm, while wrinkles with crack front. It should be pointed out that this crack initiation
amplitudes higher than 3.3 mm showed partial contact. The mechanism was found for a planar wrinkled surface in contact
sample with an amplitude of 1.8 mm showed only complete with a flat probe, whereas in our case the probe is spherical.
contact upon applying preload [49]. In addition, in a theore- Therefore, in our case, the middle region will stay under com-
tical study by McMeeking et al. [51], an approximate pressive stress much longer during detachment than when
condition for complete contact of dimpled surfaces without measured with a flat probe. This means that energy dissipation
the application of preload, i.e. spontaneous jump-to-contact, due to discrete crack initiation can only occur in the vicinity of
was derived. They showed that the depth of a dimple in a the crack front.
compliant material below which full contact occurs depended
on dimple diameter (approximated here as dimple wave-
length, l ), work of adhesion, Wad, and effective modulus, 3.4. Influence of dimple depth and density
E*, as [51] The difference in adhesion enhancement for the three struc-
r tured samples can be explained considering two parameters:
plWad [1] dimple depth and [2] dimple density. The dimple depth
a, : (3:2)
E determines both the magnitude of the crack-trapping effect
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

(a) (b) 10
15

rsif.royalsocietypublishing.org
10

friction (mN)

J. R. Soc. Interface 12: 20141061


0 0.5 1.0 1.5 2.0 0 0.5 1.0 1.5 2.0 2.5
motor movement (mm) normal load (mN)
Figure 8. Effect of nanopatterns on the frictional force at different normal loads. (a) Three typical friction curves for S1 (dark grey), S2 (grey) and S3 (light grey).
Normal load of 0.5 mN (solid line), 1 mN (dashed line) and 2 mN (dotted line). (b) Average static friction with 95% confidence interval for (diamond) sample 1,
(triangle) sample 2 and (square) sample 3.

and the elastic penalty upon compression. These two effects PDMS resulted in pronounced stickslip motion, patterned
counteract each other. A deeper dimple results in more surfaces with vastly different patterns did not show stickslip
energy dissipation [53]. However, the larger the dimple instabilities and showed decreased friction [55,56]. Lorenz &
depth, the more elastic energy is stored due to surface defor- Persson [57] explained the disappearance of stickslip motion
mation to achieve complete contact. This stored elastic energy for patterned surfaces by the occurrence of sequential interfacial
is (partly) recovered and may help to break interfacial bonds. slip. They showed that surface patterns with millimetre-sized
As all three structured samples show enhancement of adhesion surface features will reduce the elastic coupling between the
compared with the flat control, the energy dissipation through different surface features, which can cause these features to
unstable detachment outweighs the elastic penalty, and the net move more independently of each other than surface patches
effect of one dimple is an increase in adhesion. on the flat homogeneous surface [57]. This reduces the tendency
In addition to the dimple depth, the dimple density will also for macroscopic stickslip motion. Whereas our dimples are
affect the adhesion. More specifically, a higher density may relatively small and shallow they still result in a heterogeneous
have a larger influence on adhesion. For our samples, the stress distribution and a decrease in lateral stiffness, as
dimple density can be assessed by the crystallinity. A higher explained by Rand & Crosby [58]. This decreases the friction.
(normalized) crystallinity stems from a more efficient packing
and thus a higher dimple density. Indeed, if we count all par-
ticles in figure 3df and compare S1 ( pH 10, normalized
crystallinity 0.54) and S2 (pH 5, normalized crystallinity 0.35),
4. Summary
In summary, we developed a scalable method to fabricate
a decrease in the number of particles of 6% is found. S3 (pH 3,
dimpled dry adhesives. Particles formed close-packed islands
normalized crystallinity 0.13) shows a decrease of 13%
at the air/water interface that transitioned to well-ordered
compared with S2.
monolayers by adding PDMS. Colloidal monolayers at the
In our experiments, both the dimple depth and the
air/water interface were used directly as a template for
dimple density are varied for each sample. We find
PDMS. The composition of the subphase was altered to tune
the largest increase in adhesion for the sample with interme-
the order and immersion depth of the monolayer, which was
diate dimple density and depth. The increase in adhesion per
translated into the PDMS nanopatterned surfaces. These sur-
dimple should therefore be higher with increasing depth in
faces showed an increase in adhesion, which is attributed to
this range of depths, but the decrease in the number of
an energy-dissipating mechanism during pull-off, and a
dimples with increasing depth counters this effect.
decrease in friction. These results show that interfacial templat-
ing of colloidal particles provides a promising route for creating
3.5. Friction functional surfaces of which the three-dimensional topology
The friction of the samples was measured by lowering the can be tuned without elaborate and energy-consuming litho-
probe onto the sample, applying a defined constant normal graphic methods. With the emergence of a wide variety of
load, and subsequently moving the probe in plane with new methods to tailor the size and shape of colloids, and
100 mm s21 for 20 s. The static friction of flat PDMS was too their organization at liquid interfaces, this opens a wealth of
high (more than 70 mN) to be measured with our set-up. For opportunities for lithography-free design of new adhesive and
the nanopatterned samples, the shear force first rose up to a low-frictional materials.
peak (static friction), and subsequently dropped to a nearly
constant value (dynamic friction), which indicates a sliding Acknowledgements. S.A. would like to thank Elma Vroege for her work on
monolayers as templates during her Bachelors thesis and Marcel Gies-
frictional mechanism (figure 8a) [55]. Figure 8b shows that, com- bers for his help with Auger experiments and SEM at such short notice.
pared with flat surfaces, friction was decreased by the patterns. Funding statement. M.K., J.A., J.S. and S.A. acknowledge the Nether-
This is in agreement with several previous publications on lands Organization for Scientific Research (NWO) for financial
patterned elastomers: whereas friction measurements on flat support.
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

References 11

rsif.royalsocietypublishing.org
1. Gorb S. 2001 Attachment devices of insect cuticle. 16. Davis CS, Martina D, Creton C, Lindner A, Crosby AJ. three dimensions. Opt. Expr. 17, 46854704.
New York, NY: Springer. 2012 Enhanced adhesion of elastic materials to (doi:10.1364/OE.17.004685)
2. Loskill P, Puthoff J, Wilkinson M, Mecke K, Jacobs K, small-scale wrinkles. Langmuir 28, 14 899 14 908. 30. Steinhardt PJ, Nelson DR, Ronchetti M. 1983 Bond-
Autumn K. 2013 Macroscale adhesion of gecko (doi:10.1021/la302314z) orientational order in liquids and glasses. Phys. Rev.
setae reflects nanoscale differences in subsurface 17. Lin P-C, Vajpayee S, Jagota A, Hui C-Y, Yang S. 2008 B 28, 784. (doi:10.1103/PhysRevB.28.784)
composition. J. R. Soc. Interface 10, 20120587. Mechanically tunable dry adhesive from wrinkled 31. Higler R, Appel J, Sprakel J. 2013 Substitutional
(doi:10.1098/rsif.2012.0587) elastomers. Soft Matter 4, 1830 1835. (doi:10. impurity-induced vitrification in microgel crystals. Soft
3. Kamperman M, Kroner E, del Campo A, McMeeking 1039/b802848f ) Matter 9, 53725379. (doi:10.1039/c3sm50471a)
RM, Arzt E. 2010 Functional adhesive surfaces with 18. Davis CS, Crosby AJ. 2011 Mechanics of wrinkled 32. Drechsler P, Federle W. 2006 Biomechanics of

J. R. Soc. Interface 12: 20141061


gecko effect: the concept of contact splitting. Adv. surface adhesion. Soft Matter 7, 53735381. smooth adhesive pads in insects: influence of tarsal
Eng. Mater. 12, 335 348. (doi:10.1002/adem. (doi:10.1039/c1sm05146f ) secretion on attachment performance. J. Comp.
201000104) 19. Kraus T, Brodoceanu D, Pazos-Perez N, Fery A. 2013 Physiol. A 192, 1213 1222. (doi:10.1007/s00359-
4. Boesel LF, Greiner C, Arzt E, del Campo A. 2010 Colloidal surface assemblies: nanotechnology meets 006-0150-5)
Gecko-inspired surfaces: a path to strong and bioinspiration. Adv. Funct. Mater. 23, 45294541. 33. Kroner E, Maboudian R, Arzt E. 2010 Adhesion
reversible dry adhesives. Adv. Mater. 22, (doi: 10.1002/adfm.201203885) characteristics of PDMS surfaces during repeated
21252137. (doi:10.1002/adma.200903200) 20. Vogel N, Weiss CK, Landfester K. 2012 From pull-off force measurements. Adv. Eng. Mater. 12,
5. Geim A, Grigorieva SDI, Novoselov K, Zhukov A, soft to hard: the generation of functional and 398404. (doi:10.1002/adem.201000090)
Shapoval SY. 2003 Microfabricated adhesive complex colloidal monolayers for nanolithography. 34. Sirotkin E, Apweiler JD, Ogrin FY. 2010 Macroscopic
mimicking gecko foot-hair. Nat. Mater. 2, 461463. Soft Matter. 8, 40444061. (doi:10.1039/ ordering of polystyrene carboxylate-modified
(doi:10.1038/nmat917) c1sm06650a) nanospheres self-assembled at the water2air
6. Sitti M, Fearing RS. 2003 Synthetic gecko foot-hair 21. Pan F, Zhang J, Cai C, Wang T. 2006 Rapid interface. Langmuir 26, 10 67710 683. (doi:10.
micro/nano-structures as dry adhesives. J. Adhes. fabrication of large-area colloidal crystal monolayers 1021/la1009658)
Sci. Technol. 17, 1055 1073. (doi:10.1163/ by a vortical surface method. Langmuir 22, 35. Weekes SM, Ogrin FY, Murray WA, Keatley PS. 2007
156856103322113788) 7101 7104. (doi:10.1021/la053323n) Macroscopic arrays of magnetic nanostructures from
7. Greiner C, del Campo A, Arzt E. 2007 Adhesion of 22. Retsch M, Zhou Z, Rivera S, Kappi M, Zhao XS, Jonas self-assembled nanosphere templates. Langmuir 23,
bioinspired micropatterned surfaces: effects of pillar U, Li Q. 2009 Fabrication of large-area, transferable 1057 1060. (doi:10.1021/la061396g)
radius, aspect ratio, and preload. Langmuir 23, colloidal monolayers utilizing self-assembly at the 36. Mann E, Langevin D. 1991 Poly (dimethylsiloxane)
34953502. (doi:10.1021/la0633987) air/water interface. Macromol. Chem. Phys. 210, molecular layers at the surface of water and
8. del Campo A, Greiner C, Arzt E. 2007 Contact shape 230 241. (doi:10.1002/macp.200800484) of aqueous surfactant solutions. Langmuir 7,
controls adhesion of bioinspired fibrillar surfaces. 23. Vogel N, Goerres S, Landfester K, Weiss CK. 2011 1112 1117. (doi:10.1021/la00054a016)
Langmuir 23, 10 23510 243. (doi:10.1021/ A convenient method to produce close- and non- 37. Moon GD et al. 2011 Assembled monolayers of
la7010502) close-packed monolayers using direct assembly at hydrophilic particles on water surfaces. ACS Nano 5,
9. Murphy MP, Aksak B, Sitti M. 2009 Gecko-inspired the airwater interface and subsequent plasma- 8600 8612. (doi:10.1021/nn202733f )
directional and controllable adhesion. Small 5, induced size reduction. Macromol. Chem. Phys. 212, 38. Yu J, Yan Q, Shen D. 2010 Co-self-assembly of
170175. (doi:10.1002/smll.200801161) 1719 1734. (doi:10.1002/macp.201100187) binary colloidal crystals at the air2water interface.
10. Gorb S, Varenberg M, Peressadko A, Tuma J. 2007 24. Zhang JT, Wang L, Lamont DN, Velankar SS, Asher ACS Appl. Mater. Interfaces 2, 19221926. (doi:10.
Biomimetic mushroom-shaped fibrillar adhesive SA. 2012 Fabrication of large-area two-dimensional 1021/am100250c)
microstructure. J. R. Soc. Interface 4, 271 275. colloidal crystals. Angew. Chem. Int. Edn 51, 39. Kim MH, Im SH, Park OO. 2005 Rapid fabrication of
(doi:10.1098/rsif.2006.0164) 6117 6120. (doi:10.1002/anie.201105439) two- and three-dimensional colloidal crystal films
11. Jeong HE, Lee J-K, Kim HN, Moon SH, Suh KY. 2009 25. Xu H, Goedel WA. 2003 From particle-assisted via confined convective assembly. Adv. Funct. Mater
A nontransferring dry adhesive with hierarchical wetting to thin free-standing porous membranes. 15, 1329 1335. (doi:10.1002/adfm.200400602)
polymer nanohairs. Proc. Natl Acad. Sci. USA 106, Angew. Chem. Int. Edn 42, 4694 4696. (doi:10. 40. Fujii S, Kappl M, Butt H-J, Sugimoto T, Nakamura Y.
56395644. (doi:10.1073/pnas.0900323106) 1002/anie.200351427) 2012 Soft Janus colloidal crystal film. Angew. Chem. Int.
12. Glassmaker NJ, Jagota A, Hui C-Y, Noderer WL, 26. Chen J, Chao D, Lu X, Zhang W, Manohar S. 2006 Edn 51, 98099813. (doi:10.1002/anie.201204358)
Chaudhury MK. 2007 Biologically inspired crack General synthesis of two-dimensional patterned 41. Vlad A et al. 2012 Direct transcription of two-
trapping for enhanced adhesion. Proc. Natl Acad. conducting polymer-nanobowl sheet via chemical dimensional colloidal crystal arrays into three-
Sci. USA 104, 10 786 10 791. (doi:10.1073/pnas. polymerization. Macromol. Rapid Commun. 27, dimensional photonic crystals. Adv. Funct. Mater.
0703762104) 771 775. (doi:10.1002/marc.200600047) 23, 1164 1171. (doi:10.1002/adfm.201201138)
13. Rong Z, Zhou Y, Chen B, Robertson J, Federle W, 27. Appel J, Akerboom S, Fokkink RG, Sprakel J. 2013 42. Jiang P, Bertone J, Hwang K, Colvin V. 1999 Single-
Hofmann S, Steiner U, Oppenheimer PG. 2013 Bio- Facile one-step synthesis of monodisperse micron- crystal colloidal multilayers of controlled thickness.
inspired hierarchical polymer fiber-carbon nanotube sized latex particles with highly carboxylated Chem. Mater. 11, 2132 2140. (doi:10.1021/
adhesives. Adv. Mater. 26, 14561461. (doi:10. surfaces. Macromol. Rapid Commun. 34, cm990080+)
1002/adma.201304601) 1284 1288. (doi:10.1002/marc.201300422) 43. Li C, Hong G, Yu H, Qi L. 2010 Facile fabrication of
14. Chan EP, Smith EJ, Hayward RC, Crosby AJ. 2008 28. Tasdelen B, Kayaman-Apohan N, Guven O, Baysal BM. honeycomb-patterned thin films of amorphous
Surface wrinkles for smart adhesion. Adv. Mater. 20, 2004 Preparation of poly (N-isopropylacrylamide/ calcium carbonate and mosaic calcite. Chem. Mater.
711716. (doi:10.1002/adma.200701530) itaconic acid) copolymeric hydrogels and their drug 22, 3206 3211. (doi:10.1021/cm100363a)
15. Kundu S, Davis CS, Long T, Sharma R, Crosby AJ. release behavior. Int. J. Pharm. 278, 343351. 44. Ho C-C, Chen P-Y, Lin K-H, Juan W-T, Lee W-L. 2011
2011 Adhesion of nonplanar wrinkled surfaces. (doi:10.1016/j.ijpharm.2004.03.017) Fabrication of monolayer of polymer/nanospheres
J. Polym. Sci. B Polym. Phys. 49, 179185. (doi:10. 29. Gao Y, Kilfoil ML. 2009 Accurate detection and hybrid at a waterair interface. ACS Appl. Mater.
1002/polb.22181) complete tracking of large populations of features in Interfaces 3, 204208. (doi:10.1021/am100814z)
Downloaded from http://rsif.royalsocietypublishing.org/ on May 5, 2016

45. Lyklema J, van Leeuwen HP, Vliet M, Cazabat A-M. to full contact. Soft Matter 7, 10 72810 736. Phys. Solids 55, 473 488. (doi:10.1016/j.jmps. 12
2005 Fundamentals of interface and colloid science. (doi:10.1039/c1sm06367g) 2006.09.007)
Degrandi-Contraires E, Beaumont A, Restagno F,

rsif.royalsocietypublishing.org
New York, NY: Academic Press. 50. 55. Rand CJ, Crosby AJ. 2009 Friction of soft elastomeric
46. Fuller K, Tabor D. 1975 The effect of surface Weil R, Poulard C, Leger L. 2013 CassieWenzel- wrinkled surfaces. J. Appl. Phys. 106, 064913
roughness on the adhesion of elastic solids. like transition in patterned soft elastomer adhesive 064914. (doi:10.1063/1.3226074)
Proc. R. Soc. Lond. A 345, 327 342. (doi:10.1098/ contacts. Europhys. Lett. 101, 14001. (doi:10.1209/ 56. Varenberg M, Gorb SN. 2009 Hexagonal
rspa.1975.0138) 0295-5075/101/14001) surface micropattern for dry and wet friction.
47. Persson B, Tosatti E. 2001 The effect of 51. McMeeking RM, Ma L, Arzt E. 2010 Bi-stable adhesion Adv. Mater. 21, 483 486. (doi:10.1002/adma.
surface roughness on the adhesion of elastic of a surface with a dimple. Adv. Eng. Mater. 12, 200802734)
solids. J. Chem. Phys. 115, 5597. (doi:10.1063/ 389397. (doi:10.1002/adem.201000091) 57. Lorenz B, Persson B. 2012 On the origin of why
1.1398300) 52. ChangaKim Y. 2008 Analysis of pressure-driven air static or breakloose friction is larger than kinetic
48. Briggs G, Briscoe B. 1977 The effect of surface bubble elimination in a microfluidic device. Lab Chip friction, and how to reduce it: the role of aging,

J. R. Soc. Interface 12: 20141061


topography on the adhesion of elastic solids. 8, 176 178. (doi:10.1039/b712672g) elasticity and sequential interfacial slip. J. Phys.
J. Phys. D Appl. Phys. 10, 2453. (doi:10.1088/0022- 53. Guduru P. 2007 Detachment of a rigid solid from an Condens. Matter. 24, 225008. (doi:10.1088/0953-
3727/10/18/010) elastic wavy surface: theory. J. Mech. Phys. Solids 8984/24/22/225008)
49. Jin C, Khare K, Vajpayee S, Yang S, Jagota A, Hui C- 55, 445 472. (doi:10.1016/j.jmps.2006.09.004) 58. Rand CJ, Crosby AJ. 2007 Friction of soft elastomeric
Y. 2011 Adhesive contact between a rippled elastic 54. Guduru P, Bull C. 2007 Detachment of a rigid solid surfaces with a defect. Appl. Phys. Lett. 91, 261909.
surface and a rigid spherical indenter: from partial from an elastic wavy surface: experiments. J. Mech. (doi:10.1063/1.2828136)

You might also like