You are on page 1of 17

Invited Feature Article

pubs.acs.org/Langmuir

Bioinspired Structured Surfaces


Bharat Bhushan*
Nanoprobe Laboratory for Bio- & Nanotechnology and Biomimetics, The Ohio State University, Columbus, Ohio 43210-1142,
United States

ABSTRACT: Nature has evolved objects with desired functionality using commonly found
materials. Nature capitalizes on hierarchical structures to achieve functionality. The
understanding of the functions provided by objects and processes found in nature can guide
us to produce nanomaterials, nanodevices, and processes with desirable func-
tionality. Various natural objects which provide functionality of commercial interest have
been characterized to understand how a natural object provides functionality. We have
modeled and fabricated structures in the lab using nature’s route and developed optimum
structures. Once it is understood how nature does it, optimum structures have been fabricated
using smart materials and fabrication techniques. This feature article provides an overview of
four topics: Lotus effect, rose petal effect, gecko feet, and shark skin.

1. INTRODUCTION nature using nanofabrication techniques for commercial applica-


Nature has gone through evolution over the 3.8 billion years tions. Biomimetics has spurred interest across many disciplines.
since life is estimated to have appeared on earth. Nature has Various biomimetics-inspired materials and objects are being
evolved objects with desired functionality using commonly fabricated in laboratories around the world, and some have found
found materials. These function on the macroscale to the molec- industrial applications.
ular scale. The understanding of the functions provided by In this feature article, an overview of four topics as shown in
objects and processes found in nature can guide us to produce Figure 1 is provided: Lotus effect, rose petal effect, gecko feet,
nanomaterials, nanodevices, and processes with desirable func- and shark skin. Each section starts with an introduction fol-
tionality. Biologically inspired design or adaptation or deriva- lowed by some recent contributions to the field and a summary
tion from nature is referred to as “biomimetics”. It means with future directions. The objective of our research is to select
mimicking biology or nature. “Biomimetics” is derived from the objects from nature which provide functionality of commercial
Greek word biomimesis. Other words used include bionics, interest. We characterized natural objects to understand how
biomimicry, and biognosis. The field of biomimetics is highly they provide functionality, modeled them, and fabricated struc-
interdisciplinary. It involves the understanding of biological tures in the lab using nature’s route and developed optimum
functions, structures, and principles of various objects found in structures using our models. Nature has a limited toolbox
nature by biologists, physicists, chemists, and material scientists and uses rather basic materials and routine fabrication methods.
and the design and fabrication of various materials and devices Once we understand how nature does it, we then fabricate
of commercial interest from bioinspiration. optimum structures using smart materials and fabrication
Biological materials are highly organized from the molecular techniques.
scale to the nanoscale, microscale, and macroscale, often in a hierar-
2. LOTUS EFFECT
chical manner with intricate nanoarchitecture that ultimately makes
up a myriad of different functional elements. Nature uses commonly Superhydrophobic surfaces with a high static contact angle above
found materials. The properties of the materials and surfaces result 150° and contact angle hysteresis (the difference between the
from a complex interplay between the surface structure and the advancing and receding contact angles) of less than 10° exhibit
morphology and physical and chemical properties. Many materials, extreme water repellence and self-cleaning properties.3 At a low
surfaces, and devices provide multifunctionality. Molecular scale value of contact angle hysteresis, water droplets may roll in
devices, superhydrophobicity, self-cleaning, drag reduction in fluid addition to sliding, which facilitates the removal of contaminant
flow, antifouling, energy conversion and conservation, reversible particles. Surfaces with low contact angle hysteresis have a low
adhesion, aerodynamic lift, materials and fibers with high mechanical water roll-off (tilt) angle, which denotes the angle to which a
strength, biological self-assembly, antireflection, structural coloration, surface must be tilted for water droplets to roll off. Super-
thermal insulation, self-healing, and sensory aid mechanisms are hydrophobic and self-cleaning surfaces are of interest for vari-
some of the examples found in nature which are of commercial ous applications including self-cleaning windows, windshields,
interest.1,2
Various features found in natural objects are on the nanoscale. Received: August 30, 2011
The major emphasis on nanoscience and nanotechnology since Revised: December 20, 2011
the early 1990s has provided a significant impetus in mimicking Published: January 10, 2012

© 2012 American Chemical Society 1698 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

droplet, and both the contact area and the adhesion to the
surface are dramatically reduced.3
Based on the so-called Lotus effect, one of the ways to increase
the hydrophobic property of the surface is to increase surface
roughness, so roughness-induced hydrophobicity has become a
subject of extensive investigations. Wenzel10 suggested a simple
model predicting that the contact angle of a liquid with a rough
surface is different from that with a smooth surface. Cassie and
Baxter11 showed that a gaseous phase including water vapor,
commonly referred to as “air” in the literature, may be trapped in
the cavities of a rough surface, resulting in a composite solid−
liquid−air interface, as opposed to the homogeneous solid−
liquid interface. These two models describe two possible wetting
regimes or states: the homogeneous (Wenzel) and the composite
(Cassie−Baxter) regimes.
2.1. Modeling of Contact Angle for a Liquid in
Contact with a Rough Surface. As stated earlier, super-
hydrophobic and self-cleaning surfaces should have both a high
contact angle and a low contact angle hysteresis. Liquid may
form either a homogeneous interface with a solid or a composite
interface with air pockets trapped between the solid and liquid.
In this section, mathematical models which provide the
relationships between roughness and contact angle are discussed.
2.1.1. Homogeneous (Wenzel) Interface. Consider a rough
solid surface with a typical size of roughness details smaller than
the size of the droplet as shown in Figure 2 (left). For a droplet
in contact with a rough surface without air pockets, referred to
as a homogeneous interface with complete wetting, the contact
angle is given as10
cos θ = R f cos θ0 (1)
Figure 1. Montage of four examples from nature. where θ is the contact angle for the rough surface, θ0 is the
contact angle for the smooth surface, and Rf is a roughness
exterior paints for buildings and navigation of ships, utensils, roof factor defined as the ratio of the solid−liquid area ASL to its
tiles, textiles, solar panels, and applications requiring a reduction projection on a flat plane, AF
of drag in fluid flow, e.g., in micro/nanochannels. They also ASL
exhibit antifouling which is of interest such as in membranes Rf =
used for desalination and water purification. These surfaces AF (2)
can also be used for energy conversion and energy conserva- The model predicts that a hydrophobic surface (θ0 > 90°)
tion. Condensation of water vapor from the environment and/or becomes more hydrophobic with an increase in Rf, and a
process liquid film can form menisci, leading to high adhesion hydrophilic surface (θ0 < 90°) becomes more hydrophilic with
in devices requiring relative motion.4 Superhydrophobic sur- an increase in Rf.
faces are needed to minimize adhesion between a surface and a 2.1.2. Composite (Cassie−Baxter) Interface. For a rough
liquid. surface, a wetting liquid will be completely absorbed by the
A model surface for superhydrophobicity and self-cleaning is rough surface cavities while a nonwetting liquid may not pene-
provided by the leaves of the Lotus plant (Nelumbo nucifera) trate into surface cavities, resulting in the formation of air
(Figure 1, top left).5−9 The leaf surface is very rough due to so- pockets, leading to a composite solid−liquid−air interface as
called papillose epidermal cells, which form asperities or shown in Figure 2 (middle). Cassie and Baxter11 extended the
papillae. In addition to the microscale roughness, the surface of Wenzel equation for the composite interface, which was
the papillae is also rough with submicrometer-sized asperities originally developed for the homogeneous solid−liquid inter-
composed of 3-D epicuticular waxes. The waxes of the Lotus face. For this case, there are two sets of interfaces: a solid−
are tubules, but on other leaves waxes also exist in the form of liquid interface with the ambient environment surrounding the
platelets or other morphologies. Lotus leaves have hierarchical droplet and a composite interface involving liquid−air and
structures, which have been studied by Bhushan and Jung.8 The solid−air interfaces. In order to calculate the contact angle for
water droplets on these surfaces readily sit on the apex of the the composite interface, Wenzel’s equation can be modified by
nanostructures, because air bubbles fill the valleys of the struc- combining the contribution of the fractional area of wet
ture under the droplet. Therefore, these leaves exhibit consider- surfaces and the fractional area with air pockets (θ = 180°)
able superhydrophobicity. The water droplets on the leaf surfaces cos θ = R f cos θ0 − fLA (R f cos θ0 + 1) (3)
remove any contaminant particles present when they roll off,
leading to self-cleaning. It has been reported that nearly all where f LA is the fractional flat geometrical area of the liquid−air
superhydrophobic and self-cleaning leaves consist of an intrinsic interface under the droplet. In this regime, a hydrophobic
hierarchical structure.6 Water on such a surface forms a spherical smooth surface can be changed to superhydrophobic with an
1699 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

Figure 2. Schematic of a liquid droplet in contact with smooth and rough solid surfaces (left), schematic of the formation of a composite solid−
liquid−air interface for a rough surface (middle), and schematic of a tilted surface (with tile angle α) with a moving liquid droplet (θadv and θrec are
advancing and receding contact angles, respectively), showing the definition of contact angle hysteresis.

increase in Rf but a lower value as compared to that for the Nosonovsky and Bhushan3 derived a relationship for contact
Wenzel regime. Even for a hydrophilic surface, the contact angle hysteresis as a function of roughness, given as
angle increases with an increase of f LA, and at a high value of
cos θadv0 − cos θrec0
f LA, a surface can become hydrophobic. However, the value of θadv − θrec = (1 − fLA )R f
f LA required may be unachievable, or the formation of air pockets − sin θ
may become unstable. Using the Cassie−Baxter equation, the cos θrec0 − cos θadv0
value of f LA at which a hydrophilic surface could turn into a hydro-
= ( )
1 − fLA R f
2(R f cos θ0 + 1) (5)
phobic one is given as12
For the homogeneous interface, f LA = 0, whereas for a com-
R f cos θ0 posite interface f LA is a nonzero number. It is observed from eq
fLA ≥ for θ0 < 90°
R f cos θ0 + 1 (4) 5 that, for a homogeneous interface, increasing roughness (high Rf)
leads to increasing contact angle hysteresis (θadv − θrec), while
High Rf can be achieved by both micropatterns and nano- for a composite interface, an approach to unity of f LA provides
patterns. For a high f LA, a nanopattern is desirable because both a high contact angle and a small contact angle hysteresis.
generation of the liquid−air interface depends upon the ratio of Therefore, the composite interface is most desirable for super-
the distance between two adjacent asperities and the droplet hydrophobicity and self-cleaning.
radius. Furthermore, nanoscale asperities can pin liquid droplets 2.1.4. Stability of a Composite Interface and Role of
and thus prevent liquid from filling the valleys between Hierarchical Structure. Stability of the composite interface is
asperities. Transition between Wenzel and Cassie-Baxter an important issue.14−17,3 Even though it may be geometrically
regimes is dependent upon the size of the droplet and the possible for the system to become composite, it may be energet-
pitch values of the micro/nanopattern.13 ically possible for the liquid to penetrate into the valleys between
2.1.3. Contact Angle Hysteresis. The contact angle hysteresis asperities and form a homogeneous interface.
is another important characteristic of a solid−liquid interface. If a The formation of a composite interface is a multiscale phenom-
droplet sits over a tilted surface (as shown in Figure 2, right), the enon which depends upon the relative sizes of the liquid
contact angle at the front and back of the droplet corresponds to droplet and roughness details. A composite interface is metas-
the advancing contact angle and the receding contact angle, table, and its stability is an important issue. Even though it may
respectively. The advancing angle is greater than the receding be geometrically possible for the system to become composite,
angle, which results in contact angle hysteresis. Contact angle it may be energetically profitable for the liquid to penetrate into
hysteresis occurs due to surface roughness and heterogeneity. the valleys between asperities and form the homogeneous inter-
Although for surfaces with the roughness carefully controlled on face. The composite interface is fragile and can be irreversibly
the molecular scale it is possible to achieve a contact angle transformed into the homogeneous interface, thus damaging
hysteresis as low as <1°, hysteresis cannot be eliminated com- superhydrophobicity. Many authors have investigated the stabil-
pletely, since even atomically smooth surfaces have a certain ity of artificial superhydrophobic surfaces and showed that whether
roughness and heterogeneity. The contact angle hysteresis is a the interface is homogeneous or composite may depend on the
measure of energy dissipation during the flow of a droplet along history of the system, in particular whether the liquid was applied
a solid surface. Surfaces with low contact angle hysteresis have a from the top or condensed at the bottom. Nosonovsky and Bhushan3
very low water roll-off angle, which denotes the angle to which a have identified mechanisms which lead to the destabilization of the
surface must be tilted for water droplets to roll off. A low water composite interface, namely, the capillary waves, condensation and
roll-off angle is important in liquid flow applications such as in accumulation of nanodroplets, and surface inhomogeneity. They also
micro/nanochannels and surfaces with self-cleaning ability. reported that a convex surface leads to a stable interface and high
1700 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

contact angle. They have suggested the effects of a droplet’s weight was deposited by thermal evaporation to produce superhydro-
and curvature among the factors which affect the transition. It has phobic and oleophobic flat and micropatterned surfaces. For
been demonstrated that a combination of microroughness and nano- water and oil droplets in three-phase interfaces, the experimental
roughness (multiscale roughness) with convex surfaces can help to observations showed that there is a good agreement between the
resist destabilization by pinning the interface. High asperities resist the measured contact angle and the predicted values of models. It is
capillary waves, while nanobumps prevent nanodroplets from filing also found that the transition can occur for hydrophobic and
the valleys between asperities and pin them. The effect of roughness oleophobic micropatterned surfaces with a larger distance bet-
on wetting is scale dependent, and mechanisms that lead to the ween pillars.
destabilization of a composite interface are also scale-dependent. To 2.2. Fabrication and Characterization of Micro-, Nano-,
effectively resist these scale-dependent mechanisms, it is expected that and Hierarchical Structured Surfaces. In this section, we
a multiscale roughness (such as a hierarchical structure) is optimum describe measurements of the wetting properties of micro-,
for superhydrophobicity. nano-, and hierarchical structured surfaces fabricated using
2.1.5. Ideal Surfaces with Hierarchical Structure. For different techniques.
stability of a composite interface, the structure of an ideal hierar- 2.2.1. Fabrication using Nature’s Route of Hierarchical
chical surface is shown in Figure 3. The asperities should be high Surfaces. We first fabricated hierarchical structured surfaces
using nature’s route. A two-step molding process was used to
replicate microstructures, and the self-assembly of wax platelets
and tubules was used to create different nanostructures. Then,
measurements of the wetting properties of micro-, nano-, and
hierarchical structures fabricated using these techniques were
made.18
A two-step molding process was used to fabricate several
structurally identical copies of a micropatterned Si surface and
Figure 3. Schematic of the structure of an ideal hierarchical surface. Lotus leaves. The technique was used to mold a microstruc-
Microasperities consist of circular pillars with diameter D, height H, tured Si surface with pillars of 14 μm diameter and 30 μm
and pitch P. Nanoasperities consist of pyramidal nanoasperities of height with 23 μm pitch fabricated by photolithography. This
height h and diameter d with rounded tops.18 Copied with permission micropatterned geometry was selected as it provides a super-
from Phil. Trans. R. Soc. hydrophobic surface.18 Before replication of the Lotus leaf, the
epicuticular wax tubules were removed in areas of approximately
enough so that the droplet does not touch the valleys. As an 6 cm2. To remove the wax, a two-component fast-hardening glue
example, for a droplet with a radius on the order of 1 mm or was applied on the upper side of the leaves and was carefully
larger, based on the transition model between the Wenzel and pressed onto the leaf. After hardening, the glue with the embedded
Cassie−Baxter regimes, a value of H on the order of 30 μm, D waxes was removed from the leaf, and the same procedure was
on the order of 15 μm, a P on the order of 130 μm is repeated. A polyvinylsiloxane dental wax was used to create a
optimum.13,18 Nanoasperities can pin the liquid−air interface and negative replica, and epoxy resin was then used to create a positive
thus prevent liquid from filling the valleys between asperi- replica of microstructured masters. A nanostructure on top of the
ties. They are also required to support nanodroplets, which may microstructure replica was created by self-assembly of synthetic
condense in the valleys between large asperities. Therefore, and plant waxes deposited by thermal evaporation. Tubule
nanoasperities should have a small pitch to handle nanodroplets, forming waxes, which were isolated from Nelumbo nucifera leaves,
less than 1 μm down to a few nanometers in radius. The values in the following referred to as Lotus, were used to create tubule
of h on the order of 10 nm and d on the order of 100 nm can be structures. After coating, the specimens with the wax were placed
easily fabricated. in a desiccator at room temperature for crystallization. For the
2.1.6. Oleophobic/philic Surfaces. Oleophobic surfaces have plant waxes which are a mixture of aliphatic components, different
the potential for self-cleaning and antifouling from biological crystallization conditions were chosen.
and organic contaminants both in air and in underwater ap- Figure 4 shows the scanning electron microscope (SEM)
plications.19,20 A model for predicting the contact angle of water micrographs of hierarchical structures with microstructured
and oil droplets has been presented by Jung and Bhushan.20 The replicas covered with a nanostructure of Lotus wax tubules. The
surface tension of oil and organic liquids is lower than that of recrystallized Lotus wax shows tubular hollow structures, with
water. So, to make the surface oleophobic in a solid-air-oil inter- random orientation on the surfaces. Their shapes and sizes show
face, a material with a surface energy lower than that of oil should only a few variations. The tubular diameter varied between 100
be used. Based on the model, it is found that, for a hydrophilic and 150 nm, and their length varied between 1500 and 2000 nm.
surface, an oleophobic surface in the solid−water−oil interface They have a morphology comparable to that of the Lotus leaf.
can be created if γOA cos θO is lower than γWA cos θW. γOA and To study the effect of various structures on superhydro-
γWA are surface tensions of oil−air and water−air interfaces, phobicity, static contact angle, contact angle hysteresis, and tilting
respectively, and θO and θW are contact angles when in contact angle were measured on flat, microstructured Lotus replica, and
with oil and water, respectively. For a hydrophobic surface and hierarchical surfaces. Additionally, fresh Lotus leaves were inves-
an oleophobic surface in a solid−air−oil interface, an oleophobic tigated to compare the properties of the fabricated structures with
surface in solid−water−oil interface can be created if γOA cos θO the original biological model. Figure 5 shows that the hierarchical
is higher than γWA cos θW. structured Lotus leaf replica showed a static contact angle of 171°,
To validate the model for predicting the oleophobic/philic a contact angle hysteresis of 2°, and tilt angles of 1−2°, with
nature of the surfaces, flat and micropatterned surfaces with properties most favorable as compared to the micro- and nano-
varying pitch values were produced by soft lithography. structures. The fresh Lotus leaf surface investigated here showed a
n-Pefluoroeicosane (C20F42) with low surface energy (6.7 mN/m) static contact angle of 164°, a contact angle hysteresis of 3°, and a
1701 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

Figure 4. SEM micrographs taken at 45° tilt angle (shown using three magnifications) of hierarchical structure using Lotus replica and
micropatterned silicon pillars and evaporated Lotus wax (0.8 μg/mm2) after storage for seven days at 50 °C with ethanol vapor.18 Copied with
permission from Phil. Trans. R. Soc.

tilting angle of 3°, which suggests that the artificial hierarchical structures created using nature’s route exhibit properties com-
surfaces showed a higher static contact angle and lower contact parable to that of Lotus leaves. This suggests that we under-
angle hysteresis. The structural differences between the original stand how nature provides functionality.
Lotus leaf and the artificial Lotus leaf produced here are limited to To simulate impact and oscillation of liquid droplets on
a difference in the length of the wax tubules, which are 0.5 to 1 μm hierarchal structured surfaces such as during rain and vibrating
longer in the artificial Lotus leaf.18 situations, respectively, bouncing droplet experiments and
Adhesive force measured using a 15 μm radius borosilicate vibrating droplet experiments have been performed to investi-
tip in an atomic force microscope (AFM) also shows a similar gate the transition from the composite interface to the homo-
trend as the wetting properties for the artificial surfaces (Figure 5).18 geneous interface by Bhushan.9 Bouncing and vibrating droplet
Adhesion force of the hierarchical surface structure was lower than experiments show that, in hierarchical structured and nano-
that of micro- and nanostructured and flat surfaces because the structured surfaces with certain crystal densities, wetting did
contact between the tip and surface was lower as a result of the not occur up to 1.5 m/s. In microstructured surfaces, due to the
contact area being reduced. However, for the fresh Lotus leaf, larger distance between the pillars, composite interface was
there is moisture within the plant material which causes softening destroyed above a certain critical velocity in bouncing droplet
of the leaf, and so when the tip comes into contact with the leaf experiments and above a certain inertia force of the droplet on
sample, the sample deforms, and a larger area of contact between the surfaces in vibrating droplet experiments.
the tip and sample causes an increase in the adhesive force.4 2.2.2. Mechanically Durable CNT-Composite Hierarchical
In order to measure the self-cleaning efficiency of hierarchical Structured Surfaces. Various attempts have been made to
structured surfaces, various structures were contaminated using fabricate mechanically durable structures using multiwalled carbon
silicon carbide (SiC) particles in two different size ranges of 1− nanotube (CNT) arrays with high mechanical strength.21−24 A
10 μm and 10−15 μm.18 For the cleaning test, the specimens simpler approach is to use CNT composites. CNT composites
with the contaminants were subjected to water droplets using a were deposited on flat epoxy resin and a microstructure to create
microsyringe. It was reported that the most particles (70−80%) nano- and hierarchical structures using a spray method. Jung and
remained on smooth surfaces, and 50−70% of particles were Bhushan25 measured the static contact angle and contact angle
found on microstructured surfaces. Most particles were removed hysteresis on nano- and hierarchical structures with CNT. For
from the hierarchical structured surfaces, but approximately 30% static contact angle and contact angle hysteresis, after introducing
of the particles remained. A clear difference in particle removal, the CNT nanostructure on top of the micropatterned Si replica, a
independent of particle sizes, was only found in flat and nano- high static contact angle of 170° and low contact angle hysteresis
structured surfaces, where larger particles were removed with of 2° were found for the hierarchical structures. Both nano- and
higher efficiency. Observations of the droplet behavior during the hierarchical structures created with CNT showed superhydropho-
movement on the surfaces showed that droplets were rolling bicity and self-cleaning ability.
only on the hierarchical structured surfaces. On flat, micro-, and To investigate the durability of the nanostructure fabricated
nanostructured surfaces, the droplets first applied were not mov- using CNT, Jung and Bhushan25 conducted wear tests by
ing, but the continuous application of water droplets increased creating wear scars with a 15 μm radius borosilicate ball using
the droplet volume and led to a sliding of these large droplets. an AFM for 1 cycle at two normal loads of 100 nN and 10 μN
During this, some of the particles had been removed from the using AFM. Figure 6 shows surface height maps before and
surfaces. However, the rolling droplets on hierarchical structures after wear tests for nanostructures with CNT. As the normal
did not collect the dirt particles trapped in the cavities of the load of 100 nN was applied on the nanostructure with CNT,
microstructures. The data clearly show that hierarchical the wear scar induced on the surface was not found or could
structures have superior cleaning efficiency. not be measured. It was also hard to quantify a wear depth on
All the results presented in this section concerning the nanostructure with CNT scanned with a borosilicate ball.
superhydrophobicity, adhesion, and self-cleaning show that With increasing the normal load to 10 μN, it was found that the
1702 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

Figure 6. Surface height maps before and after wear tests with a 15 μm
radius borosilicate ball at 100 nN and 10 μN for nanostructures with
CNT using an AFM.25 Copied with permission from ACS Nano.

2.3. Summary. Surfaces with micro-, nano-, and hierarchical


structures have been produced by replication of the micro-
structure of a Lotus leaf and by self-assembly of wax platelets
and tubules (Lotus) to create nanostructures. The influence of
micro-, nano-, and hierarchical structures on static contact
angle, contact angle hysteresis, tilt angle and air pocket forma-
tion, adhesive force, as well as efficiency of self-cleaning, has
been studied. It has been shown that for micro-, nano-, and
hierarchical structures, introduction of roughness increased the
hydrophobicity of the surfaces. Micro-, nano-, and hierarchical
structures led to a high static contact angle, e.g., for Lotus wax
on the order of 160°, 167°, and 173°, respectively. Contact
angle hysteresis for the hierarchical structure was the lowest,
about 3°. Structured surfaces provide air pocket formation. For
a hierarchical structure, air pockets inside the grooves under-
neath the liquid reduce the contact area between the liquid and
the surface, resulting in the reduction of contact angle
hysteresis, tilt angle, and adhesive force. Hierarchical structure
provides self-cleaning property even at low tilt angle.
To develop Lotus-inspired mechanical durable surfaces, CNT-
composite and micro- and nanoparticle composite hierarchical
structures have been fabricated. They exhibit high durability for
Figure 5. Bar chart showing the measured static contact angle, contact industrial applications.
angle hysteresis, and tilt angle on various structures. The bar chart also
shows adhesive forces for various structures measured using a 15 μm 3. ROSE PETAL EFFECT
radius borosilicate tip. Hierarchical structures were fabricated using
Lotus and micropatterned Si replicas superimposed with nanostruc-
Unlike the Lotus leaf, some rose petals, scallions, and garlic exhibit
tures of Lotus wax tubules. The error bar represents ±1 standard superhydrophobicity with high contact angle hysteresis.27−29 While
deviation.18 Copied with permission from Phil. Trans. R. Soc. a water droplet can easily roll off the surface of a Lotus leaf, it stays
pinned to the surface of these leaves. The different behavior of
wear depth on the nanostructure with CNT was not signi- wetting between the Lotus leaf and the rose petal can be explained
ficantly changed, but the morphology of the CNT differed by the different designs in the surface hierarchical micro- and nano-
slightly from that before the wear test. It can be interpreted that structure. Since the rose petal’s micro- and nanostructures have a
the individual CNT might be expected to slide or bend by the larger pitch value than the Lotus leaf, the liquid is allowed to impreg-
nate between the microstructure and partially penetrates into the
borosilicate ball applied by high normal load of 10 μN during
nanostructure. This is referred to as the Cassie impregnating wetting
the test process.
regime, in which the wetted surface area is less than that in the
Ebert and Bhushan26 used a mixture of microscale and nano- Wenzel regime but greater than that in the Cassie−Baxter regime.
scale SiO2 particles to create hierarchical structure patterns on Such an explanation implies that the extent of contact angle hys-
flat epoxy substrates. In addition, hierarchical structure was teresis increases with increasing wetted surface area, which is governed
created by depositing nanoscale SiO2 particles onto micro- by surface micro- and nanostructures. In the case of scallion and
patterned epoxy substrates. The data showed that the structures garlic leaves, contact angle hysteresis is high due to hydrophobic
are superhydrophobic and self-cleaning with low adhesion and defects responsible for contact line pinning. Superhydrophobic
high mechanical durability. surfaces with high adhesion have various potential applications, such
1703 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

Figure 7. Bar charts showing the static contact angle and contact angle hysteresis, and adhesive force and coefficient of friction measured using
15 μm radius of borosilicate tip measured on two rose petal surfaces with low and high adhesion (adapted from ref 29).

as the transport of liquid microdroplets over a surface without slid- the superhydrophobic surface with high adhesion due to the
ing or rolling, the analysis of very small volumes of liquid samples, lower real area of contact between the tip and petal sample.
and for the inside of an aircraft surface to minimize the falling of Table 1 presents the surface height parameters for the
condensed water droplets onto passengers. microbumps present on the two leaves. In order to understand
Bhushan and Her,29 for the first time, found that rose petals
can have either low or high adhesion. They measured the static Table 1. Microbump Map Statistics for Rose Petals
contact angle, contact angle hysteresis, adhesive force, and coef- with Superhydrophobicity and High Adhesion
ficient of friction. Based on the understanding of the relevant and Low Adhesion, Measured in Dried Leaves
mechanisms, artificial superhydrophobic surfaces with high and Using AFM29
low adhesions were fabricated.
Peak-to-
3.1. Characterization and Measurement of Two Kinds base (P−B) Peak
of Rose Petals and Their Underlying Mechanism. Figure 1 height Midwidth radius Bump density (1/10 000
(top right) shows optical micrographs of water droplets on the (μm) (μm) (μm) μm2)
Rosa, cv. Bairage petal. As a water droplet is deposited on its Rosa, cv. 6.8 16.7 5.8 23
Bairage
surface, a high static contact angle (152°) is observed on the (High
petal. When the petal is turned upside down, the water droplet adhesion)
does not drop down, which suggests high adhesion. In the case Rosa, cv. 8.4 15.3 4.8 34
of a droplet on the Rosa, cv. Showtime petal (not shown), it Showtime
(Low
also has a high static contact angle (167°), but the droplet easily adhesion)
rolls off the surface with a small tilt angle (6°).
Figure 7 shows the static contact angle, contact angle hysteresis, the mechanisms for the microstructures of the two super-
adhesive force, and coefficient of friction measured using a 15-μm- hydrophobic rose petals with different adhesive force, Figure
radius borosilicate tip in an AFM for two superhydrophobic rose 8 shows schematics of the petals’ hierarchical structures and
petals with high and low adhesion. Both leaves had high contact schematics of water droplet contact with petal surfaces. Pitch
angles, whereas contact angle hysteresis is extremely high for Rosa, value (or bump density) and peak-to-base (P−B) height of
cv. Bairage. microstructures are different in the two petals. On the
The adhesive force of Rosa, cv. Bairage is higher than that of superhydrophobic surface with low adhesion (Rosa, cv.
Rosa, cv. Showtime. Adhesive force arises from several sources: Showtime), its microstructure has a smaller pitch value and a
the presence of a thin liquid film, such as an adsorbed water larger P−B height compared to the superhydrophobic sur-
layer, that causes meniscus bridges to build up around the face with high adhesion (Rosa, cv. Bairage). A smaller value
contacting and near contacting bumps, and real area of contact of the ratio of pitch value (P) and P−B height (H) may lead
and surface energy effects.4 For fresh petals, there is moisture to the Cassie−Baxter regime. If the value of P/H is de-
within the plant material, which causes softening of the petal, creased, it leads to an increase in the propensity of air pocket
and so when the tip comes into contact with the petal sample, formation between microstructures, so the water droplet
the sample deforms, and a larger area of contact between the tip cannot touch its bottom and minimize the contact area be-
and sample causes an increase in the adhesive force. The coef- tween the droplet and the surface, resulting in high static
ficient of friction was measured on petal surfaces at a sliding contact angle, low contact angle hysteresis, and low adhesion.
velocity of 2 μm/s. The coefficient of friction for the super- In the case of the superhydrophobic surface with high
hydrophobic surface with low adhesion is lower than that for adhesion, its large pitch value and small P−B height leads
1704 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

Figure 8. Schematic illustrations of petals’ hierarchical structure and schematics of a water droplet contacting petal surfaces. Left column shows the
schematic for Rosa, cv. Bairage with a superhydrophobic and high adhesion surface, and right column shows Rosa, cv. Showtime with a
superhydrophobic and low adhesion.29 Copied with permission from Langmuir.

to a decrease in contact area, and water can penetrate to Schematics of the effect of nanostructure on the propensity
the bottom. This is responsible for a decrease in the static of air pocket formation are shown in Figure 10 (regime A and
contact angle and an increase in contact angle hysteresis B2 in Figure 9a). When a microstructure has the same pitch
and high adhesion. value and a nanostructure has low density (regime A), water
3.2. Fabrication of Surfaces with High and Low could impregnate between microstructures, but it is still not
Adhesion. From the understanding of real rose petals, artificial completely wetted into the nanostructure, resulting in
superhydrophobic surfaces with high and low adhesion were increasing static contact angle and adhesion. However, a high
fabricated by Bhushan and Her.29 The microstructure had 23 density of nanostructures (regime B2) prevents the transition
and 105 μm pitch values with the same diameter (14 μm) and from the Cassie−Baxter to the Wenzel regime and an increased
height (30 μm). To fabricate the nanostructure, various masses propensity of air pocket formation between micro- and
of n-hexatriacontane were coated on a microstructure. The nanostructures.
nanostructure was formed by three-dimensional platelets of 3.3. Summary. The fabrication of a superhydrophobic
n-hexatriacontane. Platelets were flat crystals grown perpendicular hierarchical surface with high and low adhesion to mimic the
to the surface. They were randomly distributed on the surface, rose petal consists of a microstructured substrate with two
and their shapes and sizes showed some variation. When different different pitch values (23 and 105 μm) coated with various
masses (0.1 and 0.2 μg/mm2) of wax were applied, the density of masses of wax (for nanostructure). In the surface with a 23 μm
pitch value, while the mass of n-hexatriacontane is changed,
the nanostructure changed. Different pitch values and masses of
there are only small changes in contact angle and contact angle
wax were used to provide hierarchical structured surfaces with
hysteresis values. In the surface with a 105 μm pitch value, high
high and low adhesion.
contact angle hysteresis (87°) with a superhydrophobic (static
The static contact angle and contact angle hysteresis were contact angle is 152°) state at 0.1 μg/mm2 mass of n-
studied as a function of the mass of n-hexatriacontane on hexatriacontane is found. When n-hexatriacontane (0.2 μg/
hierarchical structures with different pitch values is shown in mm2) is applied on the microstructure (105 μm pitch value), a
Figure 9a. In the surface with a 23 μm pitch value, while the droplet on the surface also reveals a high static contact angle
mass of n-hexatriacontane is changed, there are only small (168°) with low contact angle hysteresis (4°), and trapped air
changes in the static contact angle and contact angle hysteresis pockets can be seen.
values, which means that they are always in the Cassie−Baxter For a microstructure with large pitch values and a small peak-
wetting regime. Droplets did not adhere with any sample. to-base (P−B) height and a nanostructure with low density,
On the surface with a 105 μm pitch value, as the mass of water could impregnate between microstructures, but it is still
n-hexatriacontane was increased, the static angle increased, and not completely wetted into the nanostructure, resulting in high
the surface became superhydrophobic (static contact angle is contact angle hysteresis and high adhesion while maintaining a
152°) at 0.2 μg/mm2 (labeled as regime B2). Contact angle high static contact angle. However, the high density of a
hysteresis peaked out at 0.1 μg/mm2 to a value of 87° (labeled nanostructure even for a larger pitch value may prevent the
as regime A) with a minimum value at 0.2 μg/mm2. The transition from the Cassie−Baxter to the Wenzel regime and
comparison of regime A is made with regime B1 in the 23 μm may lead to an increased propensity of air pocket formation
pitch sample at the same mass of the wax. between micro- and nanostructures with low adhesion.
Figure 9b shows droplets on surfaces with regimes B1, B2,
and A indicated in Figure 9a. For 105 μm, at 0.1 μg/mm2, 4. GECKO FEET
there is complete wetting (regime A). The droplet adhered to The leg attachment pads of several animals including many
the surface at a tilt angle of 90° or turned upside down for insects, spiders, and lizards are capable of attaching to and
regime A. No air pockets were formed, which is believed to be detaching from a variety of surfaces and are used for
responsible for high contact angle hysteresis and high adhesion. locomotion, even on vertical walls or across the ceiling.30,31,9
At 0.2 μg/mm2, the proximity of features in the nanostructures Biological evolution over a long period of time has led to the
allow air pocket formation leading to low contact angle optimization of their leg attachment systems. This dynamic
hysteresis and low adhesion (regime B2). Air pockets were also attachment ability is referred to as reversible adhesion or smart
observed in regime B1, responsible for low contact angle adhesion.31 The attachment pads of geckos have been the most
hysteresis and low adhesion. widely studied due to the fact that they have the highest body
1705 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

Figure 9. (a) Static contact angle and contact angle hysteresis measured for hierarchical structures as a function of mass of n-hexatriacontane with
two different pitch values (23 and 105 μm). (b) The left set of images show droplets on hierarchical structures with 23 μm pitch and
n-hexatriacontane (0.1 μg/mm2) (regime B1), and on a hierarchical structure with 105 μm pitch and n-hexatriacontane (0.2 μg/mm2) (Regime B2).
In both cases, trapped air pockets can be observed. The right-hand set of images show a droplet on a hierarchical structure with 105 μm pitch and n-
hexatriacontane (0.1 μg/mm2) (regime A) which shows less trapped air with high adhesion since the droplet does not drop down when the substrate
is vertically inclined or turned upside down (adapted from ref 29).

mass and exhibit the most versatile and effective adhesive


known in nature.
There are two kinds of attachment padsrelatively smooth
and hairy. The first kind are relatively smooth pads, so-called
arolia and euplantulae, which are soft and deformable and are
found in some amphibians, such as tree frogs, torrent frogs,
cockroaches, grasshoppers, and bugs. They are able to attach to
and move over wet or even flooded environments without
falling.32 Tree frog toe attachment pads consist of a hexagonal
array of flat-topped epidermal cells about 10 μm in size separat-
ed by approximately 1-μm-wide mucus-filled channels; the
flattened surface of each cell consists of submicrometer array of
nanopillars or pegs of approximately 100−400 nm diameter.
The toe pads are made of an extremely soft, inhomogeneous
material; epithelium itself has an effective elastic modulus of
about 15 MPa, equivalent to silicone rubber. The pads are per-
manently wetted by mucus secreted from glands that open into
the channels between epidermal cells. They attach to mating
surfaces by wet adhesion.32 They are capable of climbing on
wet rocks even when water is flowing over the surface.33 During
walking, the squeezing is expected to occur rapidly. Torrent
Figure 10. Schematic illustrations of droplets on hierarchical structure
frogs can resist sliding even on flooded surfaces. The surface of with two nanostructures with 0.1 μg/mm2 (top) and 0.2 μg/mm2
their toe pads is similar to that of tree frogs with some changes (bottom) with 105 μm pitch. Mass of n-hexatriacontane changes the
in the structure to handle the large flow of water.34 density of nanostructure (Regime A and B2 in Figure 9(a)). Nano-
The second kind are the hairy types which consist of long structures play an important role in contact area between water and
deformable setae and are found in many insects (e.g., beetles, underlying substrate (adapted from ref 29).

1706 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714


Langmuir Invited Feature Article

flies), spiders, and lizards. The microstructures utilized by The attachment pads on two feet of the Tokay gecko have an
beetles, flies, spiders, and geckos have similar structures. As the area of about 220 mm2. About three million setae on their toes
size (mass) of the creature increases, the radius of the terminal that branch off into about three billion spatula on two feet can
attachment elements decreases. This allows a greater number of produce a clinging ability of about 20 N (vertical force required
setae to be packed into an area, hence increasing the linear to pull a lizard down a nearly vertical (85°) surface)40 and allow
dimension of contact and the adhesion strength. them to climb vertical surfaces at speeds over 1 m/s with the
The attachment pads of geckos have been the most widely capability to attach and detach their toes in milliseconds.
studied due to the fact that they have the higher body mass and In isolated setae, a 2.5 μN preload yielded adhesion of 20 to
exhibit the most versatile and effective adhesion in nature.35−41 40 μN, and thus the adhesion coefficient, which represents
The explanation for the adhesion properties of gecko feet can the strength of adhesion as a function of preload, ranges from 8
be found in the surface morphology of the skin on the toes of to 16.43
the gecko. The skin is composed of a complex hierarchical 4.1. Adhesion Mechanisms. It is known that van der
structure of lamellae, setae, branches, and spatulae. Figure 1 Waals attraction is the dominant adhesion mechanism. Typical
(bottom left) shows a SEM micrograph of a gecko foot, show- rough, rigid surfaces are able to make intimate contact with a
ing the hierarchical structure down to the nanoscale.42 Figure 11 mating surface only over a very small portion of the per-
ceived apparent area of contact. In fact, the real area of
shows a schematic of the structure. The gecko attachment
contact is typically 2 to 6 orders of magnitude less than the
system consists of an intricate hierarchy of structures beginning
apparent area of contact.44,45,4 Autumn et al.46 proposed that
with lamellae, soft ridges 1−2 mm in length that are located on
divided contacts serve as a means for increasing adhesion.
the attachment pads (toes) that compress easily so that contact The surface energy approach can be used to calculate adhe-
can be made with rough bumpy surfaces. Tiny curved hairs sion force in the dry environment in order to calculate the
known as setae extend from the lamellae with a density of effect of division of contacts. Spatula is considered as a hemi-
approximately 14 000 per square millimeter. These setae are sphere with radius, R, adhesion force of a single contact, Fad,
typically 30−130 μm in length, 5−10 μm in diameter, and are based on the so-called JKR (Johnson-Kendall-Roberts) theory47
composed primarily of β-keratin with some α-keratin is given as
component. At the end of each seta, 100 to 1000 spatulae
with typically 2−5 μm in length and a diameter of 0.1−0.2 μm 3
branch out and form the points of contact with the surface. The Fad = πWadR
2 (6)
tips of the spatulae are approximately 0.2−0.3 μm in width, 0.5
μm in length, and 0.01 μm in thickness and garner their name where Wad is the work of adhesion (units of energy per unit
from their resemblance to a spatula. area). Equation 6 shows that adhesion force of a single contact

Figure 11. Schematic of the three layer hierarchy of gecko seta with three levels of branches: seta level, middle level, and spatula level. ρ represents
the number of spatulae.31 Copied with permission from J. Adhesion Sci. Technol.

1707 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714


Langmuir Invited Feature Article

is proportional to a linear dimension of the contact. For a


contact area divided into a large number of contacts or setae, n,
the radius of a divided contact, R1, is given by R1 = R/√n (self-
similar scaling).48 Therefore, the adhesion force of eq 6 can be
modified for multiple contacts such that

3 ⎛ R ⎞
F ′ad = πWad⎜ ⎟n = n Fad
2 ⎝ n⎠ (7)

where F′ad is the total adhesion force from the divided contacts.
Thus, the total adhesive force is simply the adhesion force of a
single contact multiplied by the square root of the number of
contacts.
For contact in a humid environment, the meniscus (or capil-
lary) forces further increase the adhesion force.4 The attractive
meniscus force (Fm) consists of a contribution by both Laplace
pressure and surface tension. The contribution by Laplace pres-
sure is directly proportional to the meniscus area. The other
contribution is from the vertical component of surface tension
around the circumference. This force is proportional to the
circumference, as is the case for the work of adhesion.4
4.2. Peeling. Although geckos are capable of producing
large adhesion forces, they retain the ability to remove their feet
from an attachment surface at will by peeling action. The orien-
tation of the spatulae facilitates peeling. Autumn et al.43 were
the first to experimentally show that the adhesion force of
gecko setae is dependent on the three-dimensional orientation
as well as the preload applied during attachment. Due to this
fact, geckos have developed a complex foot motion during Figure 12. One-, two-, and three-level hierarchical spring models for
walking.49,50 First, the toes are carefully uncurled during detach- simulating the effect of hierarchical morphology on interaction of a
ment. The maximum adhesion occurs at an attachment angle of seta with a rough surface. In this figure, lI, II, III are lengths of
30°the angle between a seta and the mating surface. The structures, sI is space between spatulae, kI, II, III are stiffnesses of structu-
gecko is then able to peel its foot from surfaces one row of setae res, I, II, and III are level indexes, R is radius of tip, and h is distance
at a time by changing the angle at which its setae contact a between upper spring base of each model and mean line of the rough
surface. At an attachment angle greater than 30°, the gecko will profile.31 Copied with permission from J. Adhesion Sci. Technol.
detach from the surface.
4.3. Adhesion Enhancement during Contact with
Rough Surfaces. Geckos are able to climb and remain attached mechanism utilized by gecko and spider attachment systems
to surfaces with different roughnesses ranging from smooth glass appears to be van der Waals forces. The hierarchical struc-
windows to painted walls and tree bark. Hierarchical structure pro- ture involving complex divisions of the gecko skin (lamellae-
vides adhesion enhancement during contact with rough surfaces. setae-branches-spatulae) enable a large number of contacts between
In order to study the effect of the number of hierarchical levels in the gecko skin and the mating surface. These hierarchical fibril-
the attachment system on attachment ability, models with one, lar microstructured surfaces would be capable of reusable dry
two, and three levels of hierarchy were simulated (Figure 12).31 adhesion and would have uses in a wide range of applications
Figure 13 shows the calculated spring force−distance curves for from everyday objects such as adhesive tapes, fasteners, toys,
the one-, two-, and three-level hierarchical models in contact microelectronic and space applications, and treads of wall-
with rough surfaces of different values of root-mean-square (rms) climbing robots. In the design of fibrillar structures, it is neces-
amplitude σ ranging from σ = 0.01 μm to σ = 30 μm at an applied sary to ensure that the fibrils are compliant enough to easily
load of 1.6 μN, which was derived from the gecko’s weight. When deform to the mating surface’s roughness profile, yet rigid
the spring model is pressed against the rough surface, contact enough to not collapse under their own weight. Spacing
between the spring and the rough surface occurs at point A; as the between the individual fibrils is also important. If the spacing is
spring tip presses into the contacting surface, the force increases too small, adjacent fibrils can attract each other through inter-
up to point B, B′, or B″. During pull-off, the spring relaxes, and the molecular forces which will lead to bunching. The development
spring force passes an equilibrium state (0 N); tips break free of of nanofabricated surfaces capable of replicating this adhesion
adhesion forces at point C, C′, or C″ as the spring moves away force developed in nature is limited by current fabrication
from the surface. The perpendicular distance from C, C′, or C″ to methods. Many different techniques have been used in an
zero is the adhesion force. The data show that adhesion forces for attempt to create and characterize bioinspired adhesive tapes.
various levels of hierarchy are comparable when in contact with a Attempts are being made to develop climbing robots using
smooth surface. However, the adhesion force for the three level is gecko inspired structures.51
the largest when in contact with a rough surface. Thus, multilevel Lee and Bhushan52 fabricated hierarchical-structured super-
hierarchy facilitates adaptability to various rough surfaces. hydrophobic surfaces using a porous membrane as a template.
4.4. Fabrication of Biomimetic Gecko Skin. Based on By changing the density and diameter of nanofibers, surfaces
the studies reported in the literature, the dominant adhesion with either Lotus effect or gecko effect could be fabricated.
1708 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

Figure 14. SEM micrographs and measured contact angles, contact


angle hysteresis, and adhesion forces of various samples with micro-,
nano-, and hierarchical structures made of polypropylene (adapted
from ref 52).

superhydrophobic governed by air pocket formation on the


Figure 13. Force−distance curves of one-, two-, and three-level
surfaces. The oriented fibers of 100 and 600 nm diameter
models in contact with rough surfaces with different σ values for an exhibited the gecko effect (high adhesion) due to their high
applied load of 1.6·μN.31 Copied with permission from J. Adhesion Sci. fiber densities and large contact areas. Whereas the oriented
Technol. fibers of 5 and 14 μm exhibited the Lotus effect due to their
smaller fiber density.
Figure 14 shows the SEM micrographs of selected micro-, nano-, 4.5. Summary. The adhesive properties of geckos and other
and hierarchical structures made from polypropylene. The figure creatures such as flies, beetles, and spiders are due to the
also shows the contact angle, contact angle hysteresis, and hierarchical structures present on each creature’s hairy
adhesive force data. It can be seen that all structured surfaces are attachment pads. Geckos have developed the most intricate
1709 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

adhesive structures of any of the aforementioned creatures. magnitude of pressure drag can be reduced by creating strea-
The attachment system consists of ridges called lamellae that mlined shapes. Friction or viscous drag is caused by the interac-
are covered in microscale setae that branch off into nanoscale tions between the fluid and a surface parallel to the flow, as well
spatulae, about three billion spatulae on two feet. The so-called as the attraction between molecules of the fluid. Friction drag is
division of contacts provides high dry adhesion. Multiple-level similar to the motion of a deck of cards sliding across a table.
hierarchically structured surface construction plays an impor- The above discussion of friction drag assumes all neighboring
tant role in adapting to surface roughness bringing the spatulae fluid molecules move in the same relative direction and
in close proximity with the mating surface. These structures, as momentum transfer occurs between layers of fluid flowing at
well as material properties, allow the gecko to obtain a much different velocities. Fully developed turbulent flow is commonly
larger real area of contact between its feet and a mating surface said to exhibit complete randomness in its velocity distribution,
than is possible with a nonfibrillar material. Two feet of a Tokay but there exist distinct regions within fully developed turbulent
gecko have about 220 mm2 of attachment pad area on which flow that exhibit different patterns and flow characteristics.57
the gecko is able to generate approximately 20 N of adhesion While organization is evident in the viscous sublayer, the layer
force. Although capable of generating high adhesion forces, a closest to the surface, the outer layers of the turbulent boundary
gecko is able to detach from a surface at willan ability known layer are chaotic and disorganized. Much of this chaotic motion
as smart adhesion. Detachment is achieved by a peeling motion above the viscous sublayer is caused by the outward bursting of
of the gecko’s feet from a surface. the streamwise vortices that form at the surface in the viscous
There is a great interest among the scientific community to sublayer. Streamwise vortices (vortices which rotate about axes
create surfaces that replicate the adhesion strength of gecko in the direction of mean velocity) dominate the viscous sublayer.
feet. These hierarchical fibrillar microstructured surfaces would As these vortices rotate and flow along the surface, they naturally
be capable of reusable dry adhesion and would have uses in a translate across the surface in the cross-flow direction. The
wide range of applications from everyday objects such as interaction between the vortices and the surface, as well as be-
adhesive tapes, fasteners, toys, microelectronic and space
tween neighboring vortices that collide during translation initiate
applications, and treads of wall-climbing robots.
bursting motions where vortices are rapidly ejected from the sur-
face and into the outer boundary layers. As vortices are ejected,
5. SHARK SKIN
they tangle with other vortices and twist such that transient
Nature has created ways of reducing drag in fluid flow, evident velocity vectors in the cross-stream direction can become as large
in the efficient movement of fish, dolphins, and sharks. The as those in the average flow direction.57 The translation, bursting
mucus secreted by fish causes a reduction in drag as they move of vortices out of the viscous sublayer, and chaotic flow in the
through water, and also protects the fish from abrasion by mak- outer layers of the turbulent boundary layer flow are all forms of
ing the fish slide across objects rather than scrape and disease momentum transfer and are large factors in fluid drag. Reducing
by making the surface of the fish difficult for microscopic the bursting behavior of the streamwise vortices is a critical goal
organisms to adhere to.53 It has been known for many years of drag reduction, as the drag reduction possibilities presented by
that, by adding as little as a few hundred parts per million guar, this are sizable.
a naturally occurring polymer, friction in pipe flow can be A flow visualization technique was used to capture cross-
reduced by up to two-thirds. The compliant skin of the dolphin
sectional images, shown in Figure 15, of the streamwise vortex
has also been studied for drag reducing properties. By
formations above both flat-plate and riblet surfaces.58 The streaky
responding to the pressure fluctuations across the surface, a
structure that was seen in the horizontal cross section was
compliant material on the surface of an object in a fluid flow
representative of local average velocity flows and is caused by the
has been shown to be beneficial. Though early studies showed
interactions between the neighboring vortices. The average cross-
dramatic drag reduction benefits, later studies have only been
able to confirm 7% drag reduction.54 stream wavelength of these high- and low-speed streaks, the added
Another set of aquatic animals which possesses multipurpose widths of one high speed streak and one low speed streak, is equal
skin is fast-swimming sharks. The skin of fast-swimming sharks to the added diameters of two neighboring vortices and has been
protects against biofouling and reduces the drag experienced by measured at 70−100 wall units.57,59 It is useful to use nondimen-
sharks as they swim through water.55 The tiny scales covering sional length values to better compare studies performed in dif-
the skin of fast-swimming sharks, known as dermal denticles ferent flow conditions. Dimensionless wall units, marked +, are
(skin teeth), are shaped like small riblets and aligned in the used for all length scales, which are calculated by multiplying the
direction of fluid flow (Figure 1, bottom right). Shark skin in- dimensional length by Vτ/ν. For example, s+ = sVτ/ν, where s+ is
spired riblets have been shown to provide a drag reduction the nondimensional riblet spacing, s is the dimensional riblet
benefit up to 9.9%.56 The spacing between these dermal den- spacing, ν is the kinematic viscosity, and Vτ = (τ0/ρ)0.5 is the wall
ticles is such that microscopic aquatic organisms have difficulty stress velocity, for which ρ is the fluid density and τ0 is the wall
adhering to the surface. Slower sharks are covered in dermal shear stress. The 70−100 wall units data just reported correspond
denticles as well, but not those which are shaped like riblets or to a vortex diameter of 35−50 wall units. Flow visualizations
provide any drag reduction benefit. shown in Figure 15 show vortex cross sections and relative length
5.1. Mechanisms of Fluid Drag. Fluid drag comes in scales demonstrating vortex diameters smaller than 40 wall units.58
several forms, the most basic of which are pressure drag and The small riblets that cover the skin of fast-swimming sharks
friction drag. Pressure or form drag is the drag associated with work by impeding the cross-stream translation of the
the energy required to move fluid out from in front of an object streamwise vortices in the viscous sublayer. The mechanism
in the flow, and then back in place behind the object. Much of by which the riblets interact with and impede vortex translation
the drag associated with walking through water is pressure drag, is complex, and the entirety of the phenomena is not yet fully
as the water directly in front of a body must be moved out and understood. On a practical level, impeding the translation of
around the body before the body can move forward. The vortices reduces the occurrence of vortex ejection into the outer
1710 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

Figure 15. Turbulent flow visualization of streamwise vortices in a vertical cross-section over flat-plate and riblet surfaces. All experiments were
performed using atomized olive oil in air (adapted from ref 58).

boundary layers as well as the momentum transfer caused by created for a riblet array with a specific set of characteristic
tangling and twisting of vortices in the outer boundary layers. dimensions. The use of nondimensional characteristic dimen-
In the turbulent flow regime, fluid drag typically increases sions for riblet studies, namely nondimensional spacing, s+, is
dramatically with an increase in surface area due to the shear important for comparison between studies performed under
stresses at the surface acting across the new, larger surface area. different flow conditions. Nondimensionalization accounts for
However, as vortices form above a riblet surface, they remain the change in size of flow structures like vortex diameter, which
above the riblets, interacting with the tips only and rarely is the critical value to which riblets must be matched. Experi-
causing any high-velocity flow in the valleys of the riblets. Since ments have been carried out with various types of riblets, surface
the higher velocity vortices interact only with a small surface materials and in air, oil, and water.60−62,56,59 Under the same
area at the riblet tips, only this localized area experiences high nondimensionalized flow conditions, riblet arrays sharing charac-
shear stresses. The low velocity fluid flow in the valleys of the teristic dimension ratios create similar performance curves whether
riblets produces very low shear stresses across the majority of they are made of different materials, are tested in different fluids, or
the surface of the riblet. By keeping the vortices above the riblet fabricated at a different scale.
tips, the cross-stream velocity fluctuations inside the riblet Sawtooth, scalloped, and blade riblets shown in Figure 16
valleys are much lower than the cross-stream velocity fluctua- have been studied. When comparing the optimal drag reduction
tions above a flat plate.58 This difference in cross-stream geometries for sawtooth, scalloped, and blade riblets shown in
velocity fluctuations is evidence of a reduction in shear stress Figure 16, it is clear that blade riblets provide the highest level
and momentum transfer near the surface, which minimizes the of drag reduction, scalloped riblets provide the second most,
effect of the increased surface area. and sawtooth riblets provide the least benefit.
5.2. Optimization of Riblet Geometry. Most studies are More recently there have been studies which have investigated
done by changing the nondimensionalized riblet spacing, s+, by the drag reduction properties of riblet-topped shark scales as
varying only fluid velocity and collecting shear stress data from both a static structure63 and a flexiblepossibly controllable
a shear stress balance in a wind tunnel or open flow channel member.64 Using the scales molded in epoxy resin from the skin
with oil and water. Measured shear stress is compared to shear of the Spiny Dogfish (Squalus acanthias), shown in Figure 17a,
stress over a flat plate and plotted against the calculated s+ value Jung and Bhushan63 have reported a decrease in pressure drop
for the flow conditions. In this manner, a performance curve is corresponding to a decrease in fluid dragversus a smooth
1711 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714
Langmuir Invited Feature Article

Figure 16. Drag reduction comparison for sawtooth, scalloped, and


blade riblets with optimum h/s values. All experiments were performed
using oil as the test fluid in an open channel (adapted from ref 56).

surface in a rectangular flow cell flow experiment. Pressure drop


from inlet to outlet of a channel is a measure of drag with a large
pressure drop occurring as a result of high drag (Figure 17b). In
addition, some decrease in pressure drop was realized in a similar
experiment using segmented aligned riblets fabricated on acrylic
compared to a smooth acrylic test section (Figure 17b).
5.3. Application of Riblets for Drag Reduction. The
dominant and perhaps only commercial market where riblet tech-
nology for drag reduction is commercially sold is competitive swim- Figure 17. (a) SEM micrographs of shark skin replica patterned in
wear. The general population became aware of shark skin’s drag epoxy and segmented blade-style riblets fabricated from acrylic, and
reduction benefits with the introduction of the FastSkin suits by (b) comparison of pressure drop in a closed rectangular channel
flow over a flat epoxy surface and epoxy with shark skin replica
Speedo in 2004. Speedo claimed a drag reduction of several percent surface and over flat acrylic surface and segmented blade riblets. All
in a static test compared to other race suits. However, given the experiments were performed using water. Data are compared with
compromises of riblet geometry made during manufacturing, it is the predicted pressure drop function for a hydrophilic surface
hard to believe the full extent of the drag reduction. (adapted from ref 63).

1712 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714


Langmuir


Invited Feature Article

It is clear that creating surface structures by weaving threads AUTHOR INFORMATION


is difficult. As a result, riblet geometries woven from thread Corresponding Author
have limited options of feasible riblet shapes. By the pattern *E-mail: bhushan.2@osu.edu.


woven into the FastSkin swimsuits, riblets are formed which
resemble wide blade riblets with small grooves on top. The
REFERENCES
larger riblets are formed by the macro weaving pattern, and the
smaller riblets are created by the individual weaves of thread (1) Bhushan, B. Phil. Trans. R. Soc. A 2009, 367, 1445.
(2) Bar-Cohen, Y. Biomimetics: Nature-Based Innovation; CRC Press:
aligned with the macro riblets. Dean and Bhushan55 reported Boca Raton, FL, 2011.
that unstretched riblets are tightly packed. As the fabric stret- (3) Nosonovsky, M. Bhushan, B. Multiscale Dissipative Mechanisms
ches, the riblet width and spacing increase. The associated de- and Hierarchical Surfaces: Friction, Superhydrophobicity, and Biomi-
crease in h/s ratio depends on the dimensions of each swimmer’s metics; Springer-Verlag: Heidelberg, Germany, 2008.
body, which is another compromising factor in the design. (4) Bhushan, B. Introduction to Tribology; Wiley: New York, 2002.
Riblet thickness is also a factor considered in the design. (5) Barthlott, W.; Neinhuis, C. Planta 1997, 202, 1.
Aside from the limitations imposed by the weaving patterns (6) Neinhuis, C.; Barthlott, W. Annals of Botany 1997, 79, 667.
(7) Wagner, P.; Furstner, R; Barthlott, W.; Neinhuis, C. J. Exper.
available, flexibility in the riblet tips will hinder the fabric’s Botany 2003, 54, 1295.
ability to impede the cross-stream translation of streamwise (8) Bhushan, B.; Jung, Y. C. Nanotechnology 2006, 17, 2758.
vortices. Thicker riblets are probably needed, used for strength, and (9) Bhushan, B. Biomimetics: Bioinspired Hierarchical-Structured
cause a decrease in the peak drag reduction capability compared to Surfaces for Green Science and Technology; Springer-Verlag: Heidelberg,
thinner riblets. Germany, 2012.
5.4. Summary. Fluid drag in the turbulent boundary layer (10) Wenzel, R. N. Indust. Eng. Chem. 1936, 28, 988.
(11) Cassie, A.; Baxter, S. Trans. Faraday Soc. 1944, 40, 546.
is in large part due to the effects of the streamwise vortices (12) Jung, Y. C.; Bhushan, B. Nanotechnology 2006, 17, 4970.
formed in the fluid closest to the surface. Turbulence and associa- (13) Bhushan, B.; Jung, Y. C. Ultramicroscopy 2007, 107, 1033.
ted momentum transfer in the outer boundary layers is in (14) Checco, A.; Guenoun, P.; Daillant, J. Phys. Rev. Lett. 2003, 91,
large part due to the translation, ejection, and twisting of these 186101.
vortices. Additionally, the vortices also cause high velocities at (15) Lafuma, A.; Quėrė, D. Nat. Mater. 2003, 2, 457.
the surface which create large shear stresses on the surface. Two (16) Marmur, A. Langmuir 2003, 19, 8343.
mechanisms of riblet drag reduction are believed to be (17) Patankar, N. A. Langmuir 2004, 20, 7097.
(18) Bhushan, B.; Jung, Y. C.; Koch, K. Phil. Trans. R. Soc. A 2009,
dominant. First, riblets impede the translation of the stream- 367, 1631.
wise vortices, which causes a reduction in vortex ejection and (19) Liu, M.; Wang, S.; Wei, Z.; Song, Y.; Jiang, L. Adv. Mater. 2009,
outer layer turbulence. Second, riblets lift the vortices off of the 21, 665.
surface and reduce the amount of surface area exposed to the (20) Jung, Y. C.; Bhushan, B. Langmuir 2009, 25, 14165.
high velocity flow. By modifying the velocity distribution, riblets (21) Lau, K. K. S.; Bico, J.; Teo, K. B. K.; Chhowalla, M.;
facilitate a net reduction in shear stress at the surface. Amaratunga, G. A. J.; Milne, W. I.; McKinley, G. H.; Gleason, K. K.
Various riblet shapes have been studied for their drag reduc- Nano Lett. 2003, 3, 1701.
(22) Huang, L.; Lau, S. P.; Yang, H. Y.; Leong, E. S. P.; Yu, S. F.
ing capabilities, but sawtooth, scalloped, and blade riblets are J. Phys. Chem. 2005, 109, 7746.
most common. By varying flow properties or riblet geometries, (23) Zhu, L.; Xiu, Y.; Xu, J.; Tamirisa, P. A.; Hess, D. W.; Wong, C.
optimization studies have been performed. Drag reduction by Langmuir 2005, 21, 11208.
riblet surfaces has been shown to be as high as nearly 10% given (24) Hong, Y. C.; Uhm, H. S. Appl. Phys. Lett. 2006, 88, 244101.
an optimal geometry of h/s ∼0.5 for blade riblets. The (25) Jung, Y. C.;; Bhushan, B. ACS Nano 2009, 3, 4155−4163.
maximum reliable drag reduction provided by scalloped (26) Ebert, D.; Bhushan, B. J. Colloid Interface Sci. Online early
access; http://dx.doi.org/10.1016/j.jcis.2011.09.049.
riblets and sawtooth riblets is about 6% at h/s ∼0.7 and 5% (27) Jin, M.; Feng, X.; Feng, L.; Sun, T.; Zhai, J.; Li, T.; Jiang, L. Adv.
at α ∼60°, respectively. Commercial applications of riblets Mater. 2005, 17, 1977.
include competition swimsuits, which use a thread-based (28) Feng, L.; Zhang, Y.; Xi, J.; Zhu, Y.; Wang, N.; Xia, F.; Jiang, L.
riblet geometry. Langmuir 2008, 24, 4114.
(29) Bhushan, B.; Her, E. K. Langmuir 2010, 26, 8207.
6. OUTLOOK (30) Gorb, S. Attachment Devices of Insect Cuticles; Kluwer:
Dordrecht, Netherlands, 2001.
The emerging field of biomimetics is already gaining a foothold (31) Bhushan, B. J. Adhesion Sci. Technol. 2007, 21, 1213.
in the scientific and technical arena. It is clear that nature has (32) Federle, W.; Barnes, W. J. P.; Baumgartner, W.; Drechsler, P.;
evolved and optimized a large number of materials and struc- Smith, J. M. J. R. Soc. Interface 2006, 3, 689.
(33) Barnes, W. J. P.; Smith, J.; Oines, C.; Mundl, R. Tire Technol. Int.
tured surfaces with rather unique characteristics. Nature capi- 2002, 56.
talizes on hierarchical structure using basic materials to provide (34) Ohler, A. Asiatic Herpetological Res. 1995, 6, 85.
functionality of interest. As we understand the underlying mech- (35) Ruibal, R.; Ernst, V. J. Morphol. 1965, 117, 271.
anisms, we can begin to exploit them for commercial appli- (36) Russell, A. P. J. Zool. Lond. 1975, 176, 437.
cations. Significant advancements in nanofabrication allow one (37) Russell, A. P. Can. J. Zool. 1986, 64, 948.
to replicate structures of interest in biomimetics using smart (38) Williams, E. E.; Peterson, J. A. Science 1982, 215, 1509.
materials. The commercial applications include nanomaterials, (39) Schleich, H. H.; Kästle, W. Amphibia Reptilia 1986, 7, 141.
(40) Irschick, D. J.; Austin, C. C.; Petren, K.; Fisher, R. N.; Losos, J. B.;
nanodevices, and processes. These include surfaces with Ellers, O. Biol. J. Linn. Soc. 1996, 59, 21.
superhydrophobicity, self-cleaning, and low or high adhe- (41) Autumn, K.; Peattie, A. M. Integr. Comp. Biol. 2002, 42, 1081.
sion, with reversible adhesion, and low drag surfaces, to (42) Gao, H.; Wang, X.; Yao, H.; Gorb, S.; Arzt, E. Mech. Mater.
name a few. 2005, 37, 275.

1713 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714


Langmuir Invited Feature Article

(43) Autumn, K.; Liang, Y. A.; Hsieh, S. T.; Zesch, W.; Chan, W. P.;
Kenny, T. W.; Fearing, R.; Full, R. J. Adhesive Force of a Single Gecko
Foot-Hair. Nature 2000, 405, 681−685.
(44) Bowden, F. P. ;Tabor, D. The Friction and Lubrication of Solids,
Part I; Clarendon Press: Oxford, UK, 1950; revised ed., 1954;
paperback ed., 1964.
(45) Bowden, F. P.; Tabor, D. The Friction and Lubrication of Solids,
Part II; Clarendon Press, Oxford, UK, 1964.
(46) Autumn, K.; Sitti, M.; Liang, Y. A.; Peattie, A. M.; Hansen, W. R.;
Sponberg, S.; Kenny, T. W.; Fearing, R.; Israelachvili, J. N.; Full, R. J. Proc.
Natl. Acad. Sci. U.S.A. 2002, 99, 12252.
(47) Johnson, K. L.; Kendall, K.; Roberts, A. D. Proc. R. Soc. London,
Ser. A 1971, 324, 301.
(48) Arzt, E.; Gorb, S.; Spolenak, R. Proc. Natl. Acad. Sci. U.S.A. 2003,
100, 10603.
(49) Tian, Y.; Pesika, N.; Zeng, H.; Rosenberg, K.; Zhao, B.; McGuiggan,
P.; Autumn, K.; Israelachvili, J. Proc. Nat. Acad. Sci. U. S. A. 2006, 103,
19320.
(50) Pesika, N. S.; Tian, Y.; Zhao, B.; Rosenberg, K.; Zeng, H.;
McGuiggen, P.; Autumn, K.; Israelachvili, J. N. J. Adhes. 2007, 83, 383.
(51) Cutkosky, M.; Kim, S. Phil. Trans. R. Soc. A 2009, 367, 1799−
1813.
(52) Lee, H.; Bhushan, B. J. Colloid Interface Sci., in press.
(53) Shephard, K. L. Rev. Fish Biol. Fisheries 1994, 4, 401.
(54) Choi, K. S.; Yang, X.; Clayton, B. R.; Glover, E. J.; Altar, M.;
Semenov, B. N.; Kulik, V. M. Proc. R. Soc. London, Ser. A 1997, 453,
2229.
(55) Dean, B.; Bhushan, B. Phil. Trans. R. Soc. A 2010, 368, 4775;
2010, 368, 5737.
(56) Bechert, D. W.; Bruse, M.; Hage, W.; Van Der Hoeven, J. G. T.;
Hoppe, G. J. Fluid Mech. 1997, 338, 59.
(57) Kline, S. J.; Reynolds, W. C.; Schraub, F. A.; Runstadler, P. W.
J. Fluid Mech. 1967, 30, 741.
(58) Lee, S.-J.; Lee, S.-H. Exp. Fluids 2011, 30, 153.
(59) Bechert, D. W.; Bruse, M.; Hage, W. Exper. Fluids 2000, 28, 403.
(60) Walsh, M. J. Turbulent Boundary Layer Drag Reduction using
Riblets; Paper # AIAA-82−0169; AIAA 20th Aerospace Sciences
Meeting, Orlando, FL, Jan 11−14, 1982.
(61) Walsh, M. J. Lindemann, A. M. Optimization and Application of
Riblets for Turbulent Drag Reduction; Paper # AIAA-84−0347; AIAA
22nd Aerospace Sciences Meeting, Reno, NV, Jan 9−12, 1984.
(62) Bechert, D. W., Bartenwerfer, M., Hoppe, G., Reif, W.-E. Drag
Reduction Mechanisms Derived from Shark Skin, Paper # ICAS-86−
1.8.3; Proceedings 15th ICAS Congress, Vol. 2 (A86−48−97624−01);
AIAA: New York, 1986; pp 1044−1068.
(63) Jung, Y. C.; Bhushan, B. J. Phys.: Condens. Matter 2010, 22,
035104.
(64) Lang, A. W.; Motta, P.; Hidalgo, P.; Westcott, M. Bioinspir.
Biomim. 2008, 3, 1.

1714 dx.doi.org/10.1021/la2043729 | Langmuir 2012, 28, 1698−1714

You might also like