ArticlePDF Available

Factors that influence elastomeric coating performance: The effect of coating thickness on basal plate morphology, growth and critical removal stress of the barnacle Balanus amphitrite

Authors:
  • United States Naval Research Laboratory (retired)

Abstract and Figures

Silicone coatings are currently the most effective non-toxic fouling release surfaces. Understanding the mechanisms that contribute to the performance of silicone coatings is necessary to further improve their design. The objective of this study was to examine the effect of coating thickness on basal plate morphology, growth, and critical removal stress of the barnacle Balanus amphitrite. Barnacles were grown on silicone coatings of three thicknesses (0.2, 0.5 and 2 mm). Atypical ("cupped") basal plate morphology was observed on all surfaces, although there was no relationship between coating thickness and i) the proportion of individuals with the atypical morphology, or ii) the growth rate of individuals. Critical removal stress was inversely proportional to coating thickness. Furthermore, individuals with atypical basal plate morphology had a significantly lower critical removal stress than individuals with the typical ("flat") morphology. The data demonstrate that coating thickness is a fundamental factor governing removal of barnacles from silicone coatings.
Content may be subject to copyright.
Factors that influence elastomeric coating performance: the effect
of coating thickness on basal plate morphology, growth and critical
removal stress of the barnacle Balanus amphitrite
D. E. WENDT
1
, G. L. KOWALKE
1
, J. KIM
2
& I. L. SINGER
2
1
Biological Sciences Department and Center for Coastal Marine Science, California Polytechnic State University, San Luis
Obispo and
2
Code 6176 US Naval Research Laboratory, Washington, DC, USA
Abstract
Silicone coatings are currently the most effective non-toxic fouling release surfaces. Understanding the mechanisms that
contribute to the performance of silicone coatings is necessary to further improve their design. The objective of this study was
to examine the effect of coating thickness on basal plate morphology, growth, and critical removal stress of the barnacle
Balanus amphitrite. Barnacles were grown on silicone coatings of three thicknesses (0.2, 0.5 and 2 mm). Atypical (‘‘cupped’’)
basal plate morphology was observed on all surfaces, although there was no relationship between coating thickness and i) the
proportion of individuals with the atypical morphology, or ii) the growth rate of individuals. Critical removal stress was
inversely proportional to coating thickness. Furthermore, individuals with atypical basal plate morphology had a significantly
lower critical removal stress than individuals with the typical (‘‘flat’’) morphology. The data demonstrate that coating
thickness is a fundamental factor governing removal of barnacles from silicone coatings.
current understanding of the mechanisms by which
Introduction
elastomeric coatings provide easy release is at best
There has been considerable effort over the past 25 incomplete. It has been demonstrated that a variety of
years to develop non-toxic, foul-release coatings to factors influences the performance of elastomeric
aid in the control and prevention of biofouling coatings. For example, silicone fluids or the in-
(Swain, 2004). This effort has been driven mainly corporation of oils enhance the performance of
by environmental regulations that have reduced or in elastomeric coatings (e.g. Truby et al. 2000;
some cases eliminated the use of highly effective Kavanagh et al. 2001; 2003; Stein et al. 2003). It is
toxic paints (e.g. triorganotin-based paints) (Walker, also known that certain factors are important such as
1998). In contrast to toxic paints, which prevent the surface energy (e.g. Baier et al. 1968; Finlay et al.
recruitment and growth of organisms on surfaces, 2002), coating modulus (e.g. Brady & Singer, 2000;
non-toxic coatings allow biofouling to occur and Singer et al. 2000; Berglin et al. 2003; Chaudhury
instead rely on the inability of organisms to adhere et al. 2004), frictional slippage (Newby et al. 1995;
well to surfaces. Weak adhesion by organisms Newby & Chaudhury, 1997), and coating thickness
facilitates their removal through factors such as biotic (Kohl & Singer, 1999; Brady & Singer, 2000).
disturbance or hydrodynamic forces (e.g. Swain et al. However, the practical importance of the aforemen-
1998; Schultz et al. 1999). tioned factors has not been thoroughly examined in
The most effective non-toxic surfaces to date are situ using live organisms. It is thus not clear whether
silicone-based elastomeric coatings. Much of the such factors are applicable to the release of hard-
knowledge of best performing coatings has been fouling from elastomeric coatings under natural
derived empirically through extensive laboratory conditions. The primary objective of this study was
and field testing of emerging coatings (e.g. Swain to examine using live barnacles the effect of coating
et al. 1992; Swain & Schultz, 1996). However, thickness on the performance of elastomeric coatings.
The notion that coating thickness is important to
performance is based on a fracture mechanics model
developed by Kendall (1971). Kendall showed that
the force required to pull off a rigid cylinder attached
to a thick elastomer is given by
1=2
2
8Ew
a
P
c
¼ pa ð1Þ
pað1 n
2
Þ
where P
c
, E,w
a
, and a are crack initiation force,
elastic modulus, Dupre’s work of adhesion between
epoxy and elastomer and contact radius, respectively.
However, if the contact radius is much larger than
the elastomer thickness, i.e. a h, where h is the
thickness of the elastomer, then the crack initiation
force is given by
1=2
2
2w
a
K
P
c
¼ pa ð2Þ
h
where K, the bulk modulus of the elastomeric film is
related to the elastic modulus by
K ¼ E=½3ð1 2nÞ ð3Þ
Since K E, because n approaches 0.5 for an
elastomer, a thin coating has a higher P
c
than a
more compliant thick coating.
Indeed, pull-off tests of epoxy bonded to silicone
coatings (often called pseudobarnacle tests) have
verified the behavior predicted by Kendall’s model
(Kohl & Singer, 1999; Brady & Singer, 2000; Singer
et al. 2000). However, the only two published reports
to date using live animals (in contrast to pseudo-
barnacles) failed to find an inverse relationship
between coating thickness and pull-off force (Singer
et al. 2000; Sun et al. 2004). The absence of a
thickness dependence found by Singer et al. (2000)
was likely to be due to the fact that most of the
barnacle basal plates broke during removal. In this
case the mechanics of detachment should not follow
Kendall’s model, which describes a fracture process
where a basal plate would peel from an elastomer.
A proportion of animals growing on silicone
coatings have been shown to have atypical basal
plate morphology (sometimes referred to as
‘‘cupped’’) that consists of a thick callus of cement
present between the calcareous basal plate and the
substratum; the basal plate often forms a cup over
this thick callus (Watermann et al. 1997; Berglin &
Gatenholm, 2003; Wiegemann & Watermann,
2003). Moreover, Holm et al. (2005) have shown
that the occurrence of the atypical basal plate
morphology has both an environmental and a genetic
underpinning and that there is a significant interac-
tion between these factors. Sun et al. (2004) suggest
that atypical basal plate morphology may form as a
result of increased production of adhesive as animals
try to maintain contact with a PDMS substratum. In
the present study tests were carried out to see if the
atypical morphology may be in part related to the
effective compliance of the substratum, and there-
fore, it was predicted that the frequency of the
atypical morphology may increase on the thicker,
more compliant coatings.
An initial study was conducted to determine which
of two commercially available silicones, Sylgard
184
TM
or a cross-linked polydimethylsiloxane elas-
tomer by Gelest
TM
, hereafter referred to as
PDMSdp125, gave the more effective foul-release
surface. After determining that the PDMSdp125 was
the more effective of the two, the earlier work with
pseudobarnacles was then extended by using live
barnacles on PDMSdp125 to investigate the effect of
coating thickness on i) the rate of growth and size of
individuals, ii) the frequency of atypical basal plate
and adult cement morphology, and iii) critical release
stress (sometimes referred to as adhesion strength).
Attempts were also made to determine whether the
critical removal stress differed between animals with
typical (‘‘flat’’) and atypical (‘‘cupped’’) basal plate
and adult cement morphology.
Methods
Silicone materials
Sylgard 184
TM
(Dow Corning Corp.) is a two-
component, high-temperature-vulcanized (HTV) si-
licone elastomer. Following the product instructions,
the base resin and curing agent provided in two
separate containers were thoroughly mixed in a ratio
of 10:1 by weight and cured at 858C for 2 h. The
PDMSdp125 coatings were prepared from a base
resin (vinyldimethylsiloxy-terminated polydimethyl-
siloxane, Gelest
TM
catalog DMS-V22, 200 cSt, 9400
g mole
71
(125 dp), 0.4 0.6 wt% vinyl) using a
crosslinker (25 30% methylhydrosiloxane-dimethyl-
siloxane copolymer, Gelest
TM
catalog HMS-301,
25 35 cSt), a catalyst (platinum-divinyltetramethyl-
disiloxane complex, Gelest
TM
catalog SIP6830.0,
3 3.5% platinum concentration in vinyl terminated
polydimethylsiloxane) and a curing control agent
(Maleate, from Dow Corning). The mixture was
cured at 758Cfor 1h.
Coating preparation
Two sets of coatings were prepared at NRL for
barnacle adhesion and removal studies. The first set
was 10 slides each with 1 mm thick coatings of
Sylgard 184
TM
and of PDMSdp125. The second set
was PDMSdp125 coatings at thicknesses of 0.1 mm,
0.5 mm and 2 mm. To achieve the correct thickness,
the silicone mixtures were poured into a cast
consisting of a microscope glass slide at the base and
spacers on four sides of fixed height. The surface of
the glass slide was coated with a silane coupling agent
to promote bonding to the coating. After pouring the
mixture, it was carefully covered by a hydrophobic
glass slide to constrain the thickness of the mixture
during curing; then the cast assembly was secured
with six clamps. The assembly was placed in an oven
where the mixture cured. After curing, the hydro-
phobic glass slide was removed. It should be noted
that the process of using a hydrophobic glass slide
can affect the surface properties of a cross-linked
PDMS, and can differ from an air cured PDMS. The
coatings were air-cured for several weeks before
leaching. Measurements via indentation showed that
the cured coatings acted as true elastomeric surfaces,
indicating complete curing.
Ten slides of each thickness were sent to Cal Poly
for barnacle studies and two slides remained at NRL
for pseudobarnacle testing (to be reported in a
separate paper). Both sets of coatings were leached
in 0.2 mm filtered natural sea water for 6 d.
Larval settlement and growth conditions
Balanus amphitrite cypris larvae were obtained from
the Duke University Marine Laboratory. A 500 ml
drop of seawater containing 20 50 cypris larvae was
placed on each treated microscope slide in a covered
Petri dish. Cypris larvae were allowed to settle in
these conditions for 72 h at 258C in a constant
temperature incubator under a 12 h light/dark cycle.
After that time, each slide was transferred to a
separate 100 mm 6 20 mm Petri dish and immersed
in 40 ml of natural, 0.2 mm filtered seawater. For the
pilot study on two different commercially available
PDMS formulations, settlement was performed on
28 April 2003. For the coating thickness experiment
this was done on two separate occasions: 27
December 2003 and 27 February 2004.
Newly metamorphosed barnacles were fed the
unicellular alga Duneliella tertiolecta and the diatom
Skeletonema costatum. Nauplius larvae of the brine
shrimp, Artemia sp., were added to their diet 2 weeks
following settlement. Feeding was done every
Monday, Wednesday and Friday, and involved a
complete replacement of the seawater and food
suspension within each Petri dish. Animals were
reared at 258C in a constant temperature incubator
under a 12 h light/dark cycle.
Growth of basal plate
Barnacles growing on 5 slides of each thickness were
monitored monthly from 31 December 2003 to 31
March 2004 to determine if there were any differ-
ences in growth rate of the basal plates as a function
of coating thickness. Photography was performed
using a Canon
TM
EOS 10D digital camera attached
to an Olympus
TM
SZX12 dissecting microscope.
The basal plates of the developing barnacles were
photographed through the transparent substratum,
and the area of the basal plate was calculated using
NIH’s ImageJ image analysis software. Basal plate
areas were compared by analysing data from 31
March 2005 in one- and two-factor ANOVAs.
Shear testing
The shear test apparatus consisted of an IMADA
TM
AXT 70 digital force gauge (2 kg) mounted on an
IMADA
TM
SV-5 motorised stand. The force gauge
was attached to a motorised stand that moved a
shearing head parallel to the coating surface at an
average speed of approximately 67 um s
71
. The glass
slides were clamped into a custom-built Plexiglas
chamber that allowed complete submersion of coat-
ings during release tests. Prior to each test, basal
plates were photographed using a Canon
TM
EOS
10D digital camera attached to an Olympus
TM
SZX12 dissecting microscope, and areas were
calculated using NIH’s ImageJ. As the coating
surface was flat, barnacle adhesive was assumed to
be in contact with the substratum over the entire
basal plate surface; therefore the size of the basal
plate was measured to determine area. Only barna-
cles with a diameter 43 mm were shear tested. The
force at which the barnacle detached from the
coating, the maximum force measured, is hereafter
called the critical removal force; the critical removal
stress is obtained by dividing the critical removal
force by the measured basal area. The shear test was
discarded in cases where any fraction of the basal
plate remained attached to the surface. Procedures
for performing shear tests differed from ASTM D
5618 in that the shear force was applied by an
automated test stand and the barnacle release was
performed in water. In June 2003, shear tests were
performed on the first set of PDMSdp125 and
Sylgard 184
TM
(1 mm thick) coatings. The second
set (0.2, 0.5 and 2 mm thick) of coatings was shear
tested on two occasions: mid-February and mid-June
2004. Data from shear testing satisfied the assump-
tions for an ANOVA, and were analysed using one-
and two-factor ANOVAs and Fisher’s PLSD post-hoc
tests.
Photographs were also used to determine the
morphology of the barnacle basal plates. The atypical
or altered cement produced by barnacles growing on
silicone surfaces could be observed visually as a white
mass that obscured the radial structures charac-
teristic of the barnacle’s basal plate (Berglin &
Gatenholm, 2003; Wiegemann & Watermann,
2003). Barnacles exhibiting any ‘‘clouding’’ of the
basal plate when viewed from below were designated
as having the atypical or ‘‘cupped’’ morphology, with
all others being designated ‘‘flat’’. The rate of
occurrence of atypical basal plates was determined
for each slide, and compared via a one-way ANOVA.
Results
Critical removal stress for PDMSdp125 vs Sylgard
184
TM
Only 44% of the barnacles removed from Sylgard
184 exhibited complete release from the surface,
without leaving all or part of the basal plate on the
surface of the coating. By contrast, 87% of the
barnacles removed from PDMSdp125 exhibited
complete release from the surface. Mean critical
removal stresses for barnacles removed from 1 mm
thick coatings of Sylgard 184
TM
and PDMSdp125
were 0.10 + 0.02 MPa and 0.069 + 0.007 MPa,
respectively; although technically not significant
(one-factor ANOVA; n ¼ 21, F ¼ 4.11, p ¼ 0.057)
the p-value was only marginally higher than a 0.050
critical value.
Occurrence of atypical basal plate and adult cement
morphology
The formation of atypical basal plates and adult
cement were observed on all coatings tested (see
Figure 1). The proportion of individuals with the
atypical morphology was significantly different
among coatings of different thickness (one-factor
ANOVA; n ¼ 15, F ¼ 3.9, p 5 0.05) (Figure 2). The
lowest frequency of occurrence was on the 0.5 mm
coating. The atypical morphology was observed in
essentially equal frequency on the thinnest (0.1 mm)
and thickest (2.0 mm) samples. Many of the
individuals displayed an ‘‘intermediate’’ morphol-
ogy; that is, the basal plate had the cupped
appearance in the centre and a flat morphology
closer to the perimeter (Figure 1C). The reverse
situation was not observed.
Growth rate and size as a function of coating thickness
Growth rate, as measured by the increase in basal
plate area, did not differ among coating thicknesses
(see Figure 3). Likewise, the areas of basal plates on
the three thicknesses were statistically indistinguish-
able throughout the three months of the experiment
(one-factor ANOVA; n ¼ 13, F ¼ 0.52, p ¼ 0.61)
(Figure 4A). The size of individuals with ‘‘cupped’’
basal plates was not significantly different from
individuals with ‘‘flat’’ basal plates (two-factor
ANOVA; n ¼ 37, thickness F ¼ 1.23, p ¼ 0.30; mor-
phology F ¼ 0.04, p ¼ 0.83) (Figure 4B).
Critical removal stress as a function of coating thickness
Critical removal stress was significantly different and
inversely proportional to coating thickness (one-factor
ANOVA; n ¼ 44, F ¼ 6.681, p ¼ 0.0031) (Figure 5).
The average critical removal stress was 0.093 + 0.008
MPa, 0.074 + 0.005 MPa, and 0.055 + 0.006 MPa
on 0.1 mm, 0.5 mm, and 2 mm thick coatings,
respectively. A post hoc test showed a significant
difference in critical removal stress between 2 mm
and both 0.5 mm and 0.1 mm, but not between
0.1 mm and 0.5 mm (Figure 5A). The data also show a
lower critical removal force for barnacles with an
atypical basal plate than for barnacles with a typical
basal plate (two-factor ANOVA: n ¼ 44, thickness,
F ¼ 7.437, p ¼ 0.0019; basal plate morphology, F ¼
4.182, p ¼ 0.0478; interaction, F ¼ 0.478, p ¼ 0.6338)
(Figure 5B).
Discussion
Growth rate and size as a function of coating thickness
The hypothesis that the growth rate of the basal
plate might differ significantly among coating
thickness was not supported (Figures 3 and 4).
However, using the basal plate area as a proxy for
growth does not take into account the actual mass
of the animal. Thus differences among barnacles
growing on different thicknesses may not have been
detected.
Occurrence of atypical basal plate and adult cement
morphology
The appearance of atypical basal plates and adult
cement has been observed when barnacles grow on
silicone coatings (Watermann et al. 1997, Berglin &
Gatenholm, 2003; Wiegemann & Watermann,
2003; Holm et al. 2005; Kavanagh & Swain,
personal communication) (Figure 1). Individuals
exhibiting the atypical morphology were observed to
synthesise a thick, paste-like cement that was
granular in nature, which was similar to the atypical
morphology reported by Berglin and Gatenholm
(2003). The exact mechanism by which silicone
coatings disrupt the typical process of basal plate
growth and adult cement formation is not under-
stood. Sun et al. (2004) suggest that atypical basal
plate morphology may form as a result of increased
production of adhesive as animals try to maintain
contact with a PDMS substratum. Berglin and
Gatenholm (2003) showed using electron probe
microanalysis and infrared spectroscopy that cal-
cium was incorporated as calcite in barnacles with
typical basal plate and adult cement morphology
grown on polymethlymethacrylate; in contrast, no
Figure 1. Examples of the various morphologies of a barnacle basal plate grown on an elastomeric surface. A ¼ typical or ‘‘flat’’ morphology;
B ¼ atypical or ‘‘cupped’’ morphology; C ¼ transitional basal plate exhibiting both ‘‘cupped’’ and ‘‘flat’’ characteristics.
calcium could be detected in the adult cement of
animals with the atypical morphology grown on a
polydimethylsiloxane (silicone) coating. In the same
study the authors did not report on the frequency of
occurrence of the atypical morphology in their
experiments. The present results showed that some
individuals grown on silicone coatings exhibit the
typical basal plate and adult cement morphology.
It was also observed that individuals may initially
have an atypical (‘‘cupped’’) morphology, but
then change to the typical (‘‘flat’’) morphology
as the animal grows. Holm et al. (2005) have
demonstrated genetic underpinnings that at least in
part determine the frequency of occurrence of the
atypical morphology. In their experiments different
genetic families of the barnacle B. amphitrite ex-
hibited different rates of occurrence of the atypical
morphology when grown on the same silicone
coatings. That a monotonic relationship between
coating thickness and occurrence of the atypical
morphology was not found suggests that coating
thickness (and therefore effective compliance) does
not directly influence the frequency of occurrence of
the atypical morphology (Figure 2).
Figure 2. The proportion of atypical, or ‘‘cupped’’ basal plates occurring on three different thicknesses of PDMS.
Figure 3. Growth rate of barnacle basal plates grown on three thicknesses of PDMS: 0.1 mm, 0.5 mm, and 2 mm.
Critical removal stress as a function of coating thickness
The results showed that critical removal force
decreased as a function of coating thickness,
although not as 1/h
1/2
, where h is coating thickness,
as predicted by Kendall (1971) for pull-off removal.
There are several reasons why this particular power-
law behaviour might not apply to shear of barnacles.
First, no theoretical model of shear currently exists
that predicts a 1/h
1/2
behaviour. Secondly, the
deviation could be an experimental artifact, e.g.
due to the application of non-shearing forces (e.g.
torque). In addition, Chaudhury has calculated the
thickness dependence of the shear removal stress for
both solid and flexible plates attached to elastomeric
coatings (Chaudhury, personal communication,
2005). He found that the thickness dependence
weakened as the flexibility of the plate decreased to
that of the coating. The present results also show that
the removal stress was lower for barnacles with
atypical basal plates than for barnacles with typical
basal plates (Figure 5B). These data suggest that the
functionality of the adhesive has been compromised
in some capacity, which is consistent with the
observations and data of Berglin and Gatenholm
(2003) and Sun et al. (2004).
Basal plate breakage during removal in shear
It is known that on non-easy release surfaces the
basal plates of barnacles will often fracture as animals
are dislodged. If the integrity of the basal plate is
compromised during removal, then it is likely that
the adhesion between the surface and the barnacle
cement is so great that the basal plate cannot support
the forces needed to dislodge the animal. As the
animal grows, the basal plate structural integrity
increases (e.g. Berglin et al. 2001). Berglin et al.
(2001) showed a gradual transition during barnacle
growth in failure mode, viz. i) the smallest barnacles
showed a total cohesive failure leaving the entire
basal plate on the surface, ii) as the animals grew they
shifted to a mixed failure mode where a portion of
the basal plate is removed (i.e. breaking of the basal
plate), and iii) complete removal of the barnacle
Figure 4. Mean area of the basal plates of barnacles grown on three thicknesses of PDMS, as measured during shear testing. A ¼ mean area
for all barnacles on each of three thicknesses; B ¼ mean area of barnacles exhibiting the typical (‘‘flat’’) basal plate and the atypical
(‘‘cupped’’) basal plate.
including the entire basal plate, indicating failure of
the adhesive bond to the surface. This transition,
which they suggest is a measure of the balance
between the cohesive strength of the barnacle basal
plate and the adhesion bond to the surface, occurs
earlier (i.e. for smaller barnacles) on better perform-
ing foul-release surfaces. Earlier observations of
Singer et al. (2000) showed that the barnacle B.
improvisus often broke during removal in tensile
from Sylgard 184
TM
. In the present study with
B. amphitrite a similar situation was observed;
only 44% of the barnacles grown on Sylgard 184
TM
were removed from the surface without cohesive
failure, whereas animals of the same size grown
on PDMSdp125 showed 87% complete removal.
In the coating thickness experiment using only
PDMSdp125 coatings total cohesive failure was
never observed and greater than 90% of the animals
showed failure of the adhesive bond to the surface
(i.e. complete basal plate removal). These data
clearly demonstrate i) that not all silicone coatings
are easy release, and ii) that a pure PDMS such as
PDMSdp125 can be easy release without additives
such as oils.
Figure 5. Mean critical removal stress of barnacles removed in shear from three thicknesses of PDMS. A ¼ overall mean critical removal
stress; B ¼ mean critical removal stress of barnacles exhibiting the two different basal plate morphologies: typical (‘‘flat’’) and atypical
(‘‘cupped’’).
Summary
No relationship was found between coating thickness
and the rate of growth or the size of barnacle
basal plates. There was additionally no discernable
relationship between coating thickness and the
occurrence of atypical basal plate and adult cement
morphology (i.e. ‘‘cupped’’ vs ‘‘flat’’). Critical
release stress qualitatively obeyed fracture mechanics
model, i.e. release force decreased with increasing
thickness in confinement regime. Indeed, critical
removal stress was significantly lower on 2 mm
coatings than on either 0.1 mm or 0.5 mm coatings.
Moreover, critical removal stress was lower for
barnacles with atypical basal plates than for barnacles
with typical basal plates. Although 2 mm is thicker
than is practical for most commercial coatings, the
data demonstrate that coating thickness is an
important parameter governing removal of barnacles
from elastomeric coatings; coating thickness should
be considered when optimising the design of
elastomeric foul-release surfaces.
Acknowledgements
We thank Dan Rittschoff and Beatriz Orihuela-Diaz,
both of Duke University, for providing us with
barnacle cypris larvae, and Emily Wilson, Ruth
Armour, Danielle Castle, and Lisa Needles for their
assistance with the culture of juvenile barnacles at
Cal Poly. The US Office of Naval Research is
gratefully acknowledged for financial support
(Grants no. N00014-02-0935 to DEW; N00014-
04-WX-2-0311 to ILS). Jongsoo Kim gratefully
acknowledges funding support of Lehigh University
and ONR (2002 2004) and North Dakota State
University (2004 2005). We also thank two anon-
ymous reviewers for their helpful comments and
suggestions.
References
Anonymous. 1997. ASTM D 5618-94. Standard test method for
measurement of barnacle adhesion strength in shear. Annual
book of ASTM standards Vol. 06.02. West Conshohocken, PA:
American Society for Testing and Materials.
Baier RE, Shafrin EG, Zisman WA. 1968. Adhesion: mechanisms
that assist or impede it. Science 162:1360 1368.
Berglin M, Gatenholm P. 2003. The barnacle adhesive plaque:
morphological and chemical differences as a response to
substrate properties. Colloids Surf B: Biointerfaces 28:107
117.
Berglin M, Lonn N, Gatenholm P. 2003. Coating modulus and
barnacle bioadhesion. Biofouling 19:63 69.
Berglin M, Larsson A, Jonsson PR, Gatenholm P. 2001. The
adhesion of the barnacle, Balanus improvisus, to poly(dimethyl-
siloxane) fouling-release coatings and poly(methylmethacrylate)
panels: the effect of barnacle size on strength and failure mode.
J Adhesion Sci Technol 15:1485 1502.
Brady RF, Singer IL. 2000. Mechanical factors favoring release
from fouling release coatings. Biofouling 15:73 81.
Chaudhury MK, Finlay JA, Chung JY, Callow ME, Callow JA.
2004. The influence of elastic modulus and thickness on the
release of the soft-fouling green alga Ulva linza (syn.
Enteromorpha linza) from poly(dimethylsiloxane) (PDMS)
model networks. Biofouling 21:41 48.
Finlay JA, Callow ME, Ista LK, Lopez GP, Callow JA. 2002. The
influence of surface wettability on the adhesion strength of
settled spores of the green alga Enteromorpha and the diatom
Amphora. Integr Comp Biol 42:1116 1122.
Holm ER, Orihuela B, Kavanagh CJ, Rittschof D. 2005. Variation
among families for characteristics of the adhesive plaque in the
barnacle Balanus amphitrite. Biofouling 21:121 126.
Kavanagh CJ, Swain GW, Schultz MP, Stein J, Truby K,
Darkangelo-Wood C. 2001. Variation in adhesion strength of
Balanus eburneus, Crassostrea virginica and Hydroides dianthus to
fouling release coatings. Biofouling 17:155 167.
Kavanagh CJ, Swain GW, Kovach BS, Stein J, Darkangelo-Wood
C, Truby K, Holm ER, Monetemarano J, Meyer A, Wiebe D.
2003. The effects of silicone fluid additives and silicone
elastomer matrices on barnacle adhesion strength. Biofouling
19:381 390.
Kendall K. 1971. The adhesion and surface energy of elastic
solids. J Physics D: Appl Physics 4:1186 1195.
Kohl JG, Singer IL. 1999. Pull-off behavior of epoxy bonded to
silicone duplex coatings. Prog Organic Coatings 36:15 20.
Newby BZ, Chaudhury MK. 1997. Effect of interfacial slippage on
viscoelastic adhesion. Langmuir 13:1805 1809.
Newby BZ, Chaudhury MK, Brown HR. 1995. Macroscopic
evidence
of the effect of interfacial slippage on adhesion.
Science 269:1407 1409.
Schultz MP, Kavanagh CJ, Swain GW. 1999. Hydrodynamic
forces on barnacles: implications on detachment from fouling
release surfaces. Biofouling 13:323 335.
Singer IL, Kohl JG, Patterson M. 2000. Mechanical aspects of
silicone coating for hard foulant control. Biofouling 16:301
309.
Stein JK, Truby K, Darkangelo-Wood C, Stein J, Gardner M,
Swain G, Kavanagh C, Kovach B, Schultz M, Wiebe D, Holm
E, Montemarano J, Wendt D, Smith C, Meyer A. 2003.
Silicone foul release coatings: effect of the interaction of oil and
coating functionalities on the magnitude of macrofouling
attachment strengths. Biofouling 19 (Suppl):71 82.
Sun Y, Guo S, Walker GC, Kavanagh CJ, Swain GW. 2004.
Surface elastic modulus of barnacle adhesive and release
characteristics from silicone surfaces. Biofouling 20:279 289.
Swain GW. 2004. In search of the perfect ship hull coating.
Plenary Presentation, 12th International Congress on Marine
Corrosion and Fouling, University of Southampton, South-
ampton, UK.
Swain GW, Schultz MP. 1996. The testing and evaluation of non-
toxic antifouling coatings. Biofouling 10:187 197.
Swain GW, Nelson WG, Preedeenkanit S. 1998. The influence of
biofouling adhesion and biotic disturbance on the development
of fouling communities on non-toxic surfaces. Biofouling
12:257 269.
Swain GW, Griffith JR, Bultman JD, Vincent HL. 1992. The use
of barnacle adhesion measurement for the field evaluation of
non-toxic foul release surfaces. Biofouling 6:105 114.
Truby K, Wood C, Stein J, Cella J, Carpenter J, Kavanagh C,
Swain G, Wiebe D, Lapota D, Meyer A, Holm E, Wendt D,
Smith C, Montemarano J. 2000. Evaluation of the performance
enhancement of silicone biofouling-release coatings by oil
incorporation. Biofouling 15:141 150.
Walker I. 1998. Non-toxic fouling control systems. Pitture Vernici
Eur 13:17 22.
Watermann B, Berger HD, Sonnichsen H, Willemsen P. 1997.
Performance and effectiveness of non-stick coatings in sea-
water. Biofouling 11:101 118.
Wiegemann M, Watermann B. 2003. Peculiarities of barnacle
adhesive cured on non-stick surfaces. J Adhesion Sci Technol
17:1957 1977.
... 1,2 Secondly, a coating has been sought that would provide easy release of foulants from its surface. While the roles of surface chemistry and mechanics guided scientific and technological developments in this field, and laboratory tests [3][4][5] have been carried out to validate various models [6][7][8][9] , identification of a coating that meets the above goals is still far from our sight. An ideal combination 10 of these two approaches would be such that not only fouling would be deterred as much as possible, but also that the adhesion of foulants to the coating would be so weak that the weight of the foulant or the hydrodynamic forces created by the ship's motion would dislodge the biofoulants. ...
... Secondly, the adhesive released by the fouling organism can spread on the coating forming a thin film ahead of the thicker central region of the foulant. [3][4][5] Here, crack does not initiate from the edge of contact, so that the only venue to release is the central region, where crack initiates as bubbles via interfacial cavitation. Following cavitation, the bubbles would nucleate and a crack would be formed that then would eventually propagate from the center to the outer region of contact as a run away instability. ...
... The 3D amplitude of the instability on these coatings was also estimated using this method by setting the cross-section of the micrograph of the coating as a reference. After fixing the top horizontal position of the pre-soaked coating as origin (0,0,0), the amplitude scan was performed using the extract profiles program 38 images [4]. The advantage with this method of pre-processing the image by removing the background noise and enhancing the contrast in imageJ helped to obtain a 8-bit scale of the final images. ...
Preprint
Full-text available
Recent studies demonstrated that an elastomer containing hygroscopic inclusions absorbs moisture and swell. Here we show that a thin film of such an elastomer bonded to a rigid substrate undergoes morphological instability upon absorption of water, the wavelength of which increases linearly with its thickness. As the driving force for such a morphological instability arises from the difference of the chemical potential of water between its source and that in the film, its development is slowed down as the salinity of the water increases. Nonetheless, the wavelength of the fully developed morphology, but not its amplitude, is independent of the salinity. We also demonstrate that if a domed disk shaped adherent is attached to the hygro-elastomeric film before moisture absorption, the elastic force generated during the morphological transition is able to dislodge it completely without the need of any external force. These patterns, once developed in pure water, is subdued when the salinity of water increases or if it is exposed to dry air. They re-emerge when the film is immersed in water again. Such an active response could be important in fouling release when a ship coated with such a hygro-elastomer changes its location during its long travel through sea, where salinity varies from place to place.
... Quantitative techniques to assess the transport of species through hull-fouling are not very widespread, although some data have been used to predict the abundance and diversity of organisms transported by ocean-going yachts in New Zealand and Scotland [15,16]. The expiration date of paint efficiency and long periods of inactivity/reduced sailing increase the risk of hull macrofouling development [15][16][17][18]. ...
... In terms of density, the barnacles were the most numerically abundant on all surveyed vessels and were dominated by the suspension-or filter-feeding barnacles: A. amphitrite, B. trigonus, and A. improvisus. In particular, barnacles are greatly resistant to dislodgement by currents and waves as their adhesion strength is very high (0.02 MPa in juveniles to over 2.00 MPa in adults), although variable depending on the substratum, species, and age/size of the individuals [17,18,47]. These results are believed to be one of the reasons why barnacles, along with attached organisms, take over ship hulls and spread worldwide. ...
Preprint
Full-text available
Global ecological concern regarding the transfer of fouling organisms to ship hulls is increasing. This study investigated the species composition, dominant species, distribution patterns, community structure, and life-cycle differences of hull-fouling macroinvertebrates on five research vessels (R/Vs: Isabu, Onnuri, Eardo, Jangmok 1, and Jangmok 2) operated by the Korea Institute of Ocean Science and Technology (KIOST). Hull-fouling macroinvertebrates were collected three to five times on quadrats from the upper and middle sectors of the hull sides, bottom, and niche areas (the propellers and shafts and the thrusters). A total of 47 macroinvertebrate species were identified, represented by 8,519 individuals(ind.)/m2 and a biomass of 1,967 gWWt/m2 on the five vessels. The number of species, density, and biomass were greater on the coastal vessels the Eardo, Jangmok 1, and Jangmok 2 than on the ocean-going vessels the Isabu and Onnuri. Among the coastal vessels, barnacles were the most abundant and had the greatest density, while mollusks had the highest biomass. Differences between hull sectors showed that the highest species abundance and density appeared on all hulls in ports and bays where the Jangmok 1 operated, while the highest species abundance, density, and biomass were identified in the niche areas of the Eardo, which operated farther from the coast. The hull-fouling macroinvertebrates that exceeded 1% of all organisms were the barnacles Amphibalanus amphitrite, Balanus trigonus, and Amphibalanus improvisus; the polychaete Hydroides ezoensis; the bivalves Magallana gigas and Mytilus galloprovincialis; and the amphipod Jassa slatteryi. The dominant species were cosmopolitan and globally distributed, and many of them were cryptogenic. Six native species were identified: M. gigas, H. ezoensis, the amphipod Melita koreana, the isopod Cirolana ko-reana, and the barnacle B. trigonus and F. kondakovi. Eight non-indigenous species (NIS) were de-tected: barnacles A. amphitrite and A. improvisus, the bivalve M. galloprovincialis, the polychaete Perinereis nuntia, amphipods J. slatteryi and Caprella californica, and bryozoans Bugulina californica and Bugula neritina. Of the fouling macroinvertebrates found on the vessel hulls, 13% were native and 17% were NIS. The nMDS ordination analysis of the five vessels showed four clusters of fouling macroinvertebrate communities: two ocean clusters (the Isabu and Onnuri), one coastal cluster (the Eardo and Jangmok 2), and one local cluster (the Jangmok 1). More diverse communities developed on the hulls of vessels that operated locally rather than globally or in deep oceans. Spearman rank correlation showed that biological indices such as number of species, density, and biomass correlated highly and positively with the numbers of days on coastal operations and negatively with the number days on ocean operations. The species diversity index correlated positively with the total number of anchoring days and coastal operation days and negatively with the total number of operation days and ocean operation days. The macroinvertebrates differed by the area of operation, the port of anchorage, the number of days in operation and at anchor, and the hull sectors. There is no previous research data on hull-fouling macroinvertebrates in the Republic of Korea, and this study provides a basis for future studies to identify introduced species and their differences based on operation area.
... In terms of density, the barnacles were the most numerically abundant on all surveyed vessels and were dominated by the suspension-or filter-feeding barnacles A. amphitrite, B. trigonus, and A. improvisus. In particular, barnacles are greatly resistant to dislodgement by currents and waves, as their adhesion strength is very high (0.02 MPa in juveniles to over 2.00 MPa in adults), although variable depending on the substratum, species, and age/size of the individuals [17,18,47]. These results are believed to be one of the reasons why barnacles, along with attached organisms, take over ship hulls and spread worldwide. ...
Article
Full-text available
Global ecological concern regarding the transfer of fouling organisms to ship hulls is increasing. This study investigated the species composition, dominant species, distribution patterns, community structure, and life-cycle differences of hull-fouling macroinvertebrates on five research vessels (R/Vs: Isabu, Onnuri, Eardo, Jangmok 1, and Jangmok 2) operated by the Korea Institute of Ocean Science and Technology (KIOST). Hull-fouling macroinvertebrates were collected three to five times on quadrats from the upper and middle sectors of the hull sides, bottom, and niche areas (the propellers, shafts, and thrusters). A total of 47 macroinvertebrate species were identified, represented by 8519 individuals (ind.)/m2 and a biomass of 1967 gWWt/m2 on the five vessels. The number of species, density, and biomass were greater on the coastal vessels Eardo, Jangmok 1, and Jangmok 2 than on the ocean-going vessels the Isabu and Onnuri. Among the coastal vessels, barnacles were the most abundant and had the greatest density, while mollusks had the highest biomass. Differences between hull sectors showed that the highest species abundance and density appeared on all hulls in ports and bays where the Jangmok 1 operated, while the highest species abundance, density, and biomass were identified in the niche areas of the Eardo, which operated farther from the coast. The hull-fouling macroinvertebrates that exceeded 1% of all organisms were the barnacles Amphibalanus amphitrite, Balanus trigonus, and Amphibalanus improvisus; the polychaete Hydroides ezoensis; the bivalves Magallana gigas and Mytilus galloprovincialis; and the amphipod Jassa slatteryi. The dominant species were cosmopolitan and globally distributed, and many of them were cryptogenic. Six native species were identified: M. gigas, H. ezoensis, the amphipod Melita koreana, the isopod Cirolana koreana, and the barnacles B. trigonus and F. kondakovi. Eight non-indigenous species (NIS) were detected: the barnacles A. amphitrite and A. improvisus, the bivalve M. galloprovincialis, the polychaete Perinereis nuntia, the amphipods J. slatteryi and Caprella californica, and the bryozoans Bugulina californica and Bugula neritina. Of the fouling macroinvertebrates found on the vessel hulls, 13% were native, and 17% were NIS. More diverse communities developed on the hulls of vessels that operated locally rather than globally or in deep oceans. The species diversity index correlated positively with the total number of anchoring days and coastal operation days and negatively with the total number of operation days and ocean operation days. The macroinvertebrates differed by the area of operation, the port of anchorage, the number of days in operation and at anchor, and the hull sectors. There is no previous research data on hull-fouling macroinvertebrates in the Republic of Korea, and this study provides a basis for future studies to identify introduced species and their differences based on operation area.
... While some of the barnacle adhesion strength difference between i-PDMS and o-PDMS (Figs. 1D, 6, 7) could also be due to the swelling-induced thickness differences (~ 100 µm for the o-PDMS coating vs. ~ 150 µm for the i-PDMS coating) 38 , with thicker coating requiring less force to detach adhered barnacles, the difference in thickness is too small to explain the magnitude of the differences in de-adhesion forces 31 . ...
Article
Full-text available
For many decades, silicone elastomers with oil incorporated have served as fouling-release coating for marine applications. In a comprehensive study involving a series of laboratory-based marine fouling assays and extensive global field studies of up to 2-year duration, we compare polydimethylsiloxane (PDMS) coatings of the same composition loaded with oil via two different methods. One method used a traditional, one-pot pre-cure oil addition approach (o-PDMS) and another method used a newer post-cure infusion approach (i-PDMS). The latter displays a substantial improvement in biofouling prevention performance that exceeds established commercial silicone-based fouling-release coating standards. We interpret the differences in performance between one-pot and infused PDMS by developing a mechanistic model based on the Flory–Rehner theory of swollen polymer networks. Using this model, we propose that the chemical potential of the incorporated oil is a key consideration for the design of future fouling-release coatings, as the improved performance is driven by the formation and stabilization of an anti-adhesion oil overlayer on the polymer surface.
... While some of the barnacle adhesion strength difference between i-PDMS and o-PDMS (Fig 1D, 6,7) could also be due to the swelling-induced thickness differences (~100 µm for the o-PDMS coating vs ~150 µm for the i-PDMS coating), with thicker coating requiring less force to detach adhered barnacles, the difference in thickness is too small to explain the magnitude of the differences in de-adhesion forces 31 . ...
Preprint
Full-text available
For many decades, silicone elastomers with oil incorporated have served as fouling-release coating for marine applications. In a comprehensive study involving a series of laboratory-based marine fouling assays and extensive global field studies of up to 2-year duration, we compare polydimethylsiloxane (PDMS) coatings of the same composition loaded with oil. Methods using a traditional, one-pot pre-cure oil addition approach (o-PDMS) and a newer post-cure infusion approach (i-PDMS) were evaluated. The latter display a substantial increase in biofouling prevention performance that exceeds an established commercial silicone-based fouling-release coating standard. We interpret the differences in performance between one-pot and infused PDMS by developing a mechanistic model based on the Flory-Rehner theory of swollen polymer networks. Using this model, we propose that the chemical potential of the incorporated oil is a key consideration for the design of future fouling-release coatings, as the improved performance is driven by the formation and stabilization of an anti-adhesion oil overlayer on the polymer surface.
Article
Acorn barnacles are efficient colonizers on a wide variety of marine surfaces. As they proliferate on critical infrastructure, their settlement and growth have deleterious effects on performance. To address acorn barnacle biofouling, research has focused on the settlement and adhesion processes with the goal of informing the development of novel coatings. This effort has resulted in the discovery and characterization of several proteins found at the adhesive substrate interface, i.e. cement proteins, and a deepened understanding of the function and composition of the biomaterials within this region. While the adhesive properties at the interface are affected by the interaction between the proteins, substrate and mechanics of the calcified base plate, little attention has been given to the interaction between the proteins and the cuticular material present at the substrate interface. Here, the proteome of the organic matrix isolated from the base plate of the acorn barnacle Amphibalanus amphitrite is compared with the chitinous and proteinaceous matrix embedded within A. amphitrite parietal plates. The objective was to gain an understanding of how the basal organic matrix may be specialized for adhesion via an in-depth comparative proteome analysis. In general, the majority of proteins identified in the parietal matrix were also found in the basal organic matrix, including nearly all those grouped in classes of cement proteins, enzymes and pheromones. However, the parietal organic matrix was enriched with cuticle-associated proteins, of which ca 30% of those identified were unique to the parietal region. In contrast, ca 30–40% of the protease inhibitors, enzymes and pheromones identified in the basal organic matrix were unique to this region. Not unexpectedly, nearly 50% of the cement proteins identified in the basal region were significantly distinct from those found in the parietal region. The wider variety of identified proteins in the basal organic matrix indicates a greater diversity of biological function in the vicinity of the substrate interface where several processes related to adhesion, cuticle formation and expansion of the base synchronize to play a key role in organism survival.
Article
The aim of this work is to study the adhesion strength of Amphibalanus amphitrite in the İzmir Bay and compare the results with the pseudobarnacle adhesion test. Normally, adhesion tests are performed to evaluate the performance of the antifouling coatings, but the test results can also be used to predict biofouling cleaning process efficacy. The biofouling process is highly dependent on environmental conditions. For this reason, laboratory tests are required to perform the performance tests on self-polishing coatings in cases where living organisms cannot be reached. For this purpose, different self-polishing antifouling coatings have been formulated. Field tests for the coatings were carried out in the Aegean Sea for 10 weeks. After 10 weeks, barnacle and pseudobarnacle adhesion tests were conducted on coatings. When the results were compared, similarity was observed between the adhesion strength of barnacles and pseudobarnacles with 10 mm diameter on coating with the rosin/xylene/BaSO4 (40:40:20 w/w %). The adhesion strength of barnacles and pseudobarnacles on the coating 12 was found to be 0.46 and 0.45 MPa, respectively. In conclusion, the present study exhibits the first data related to the adhesion strength of A. amphitrite on rosin-based self-polishing coatings in the Aegean Sea. Moreover, based on field tests, a pseudobarnacle adhesion test methodology was developed to mimic barnacles and the correlation between barnacle and pseudobarnacle tests was examined.
Chapter
The non-stick & foul release property of silicones was first reported in the early 1970s, with surface free energy of 22 – 24 dynes/cm offering a minimally adhesive surface to biological organisms. The superior antifouling performance of tri-butyl tin- self-polishing coatings TBT-SPC systems outshone all other antifouling formulations from 1970 to 1980s until environmental regulations warranted a total ban on the use of the TBT-SPC system. Foul release coatings (FRC’s) use hydrodynamic stress during navigation to minimize adhesion between fouling organisms and coating surfaces so that fouling can be removed. Addition of hydrophobic silicone oils along with other properties like low surface energy, elasticity and low glass transition temperature, low micro-roughness, attributed to the foul release property of siloxane polymers. Inhibition of fouling on FRC is dependent on several factors like chemical bonding of marine bio adhesives, electrostatic interactions, physical adsorptions between coatings and secreted bio adhesives, diffusion, penetration and interlocking of bio adhesives within the coating matrix. Foul release coatings are prone to biofouling and their fouling load decreases with an increase in hydrodynamic stress due to water flow. Fouling release occurs due to weak interfacial bond created by the organism’s cement and the coating surfaces as a result of low surface free energy (SFE) and cohesive failure of bio adhesives occurs due to shear forces created by flowing water across the coatings. Even though FRC has been shown to be eco-friendly & reduce drag they have many drawbacks viz: weak adhesion strength between coating and substrate, weak mechanical properties, poor AF performance under static conditions, inefficient against diatom and bacterial slimes. Bacterial and diatom biofilms on FRC’s increase frictional resistance reduce drag reduction and fuel savings. To improve the biofouling resistance of FRC, several approaches like amphiphiles, zwitterions, quaternary ammonium salts (QAs), and metal oxide nanoparticles have been investigated. PEG-based amphiphiles is one such example where findings have translated into a commercial paint Intersleek 1100SR and HempasilX3 formulations which have been reported to offer better fouling release of barnacles and diatoms. Surface chemistry, mechanical property, binding to substrates, and durability are vital factors in designing modern-day antifouling coatings and fouling resistance is a ubiquitous parameter in consideration. This review reports the advancements and modifications to the siloxane backbone by each of these parameters which have enabled in development of superior and environmentally benign foul release coatings.KeywordsAntifouling coatingsFouling release coatingsPolydimethylsiloxanePolymerNanofillersPolymer brushesHydrogelBarnaclesAmphiphilicZwitterion
Article
This paper characterizes the effect of barnacle settlement on the durability and microstructure of concrete. Twenty-seven concrete samples with different coverage areas of barnacles were extracted from a breakwater in the tidal zone of Huangdao district in the Yellow Sea. Durability properties, including the water absorption, chloride ion resistance, chloride ion content, and pH profiles of the concrete were determined, and the microstructure and pore structure of the concrete were explored to elucidate these durability properties. Results confirm that the settlement of the barnacles could enhance the resistance of the concrete to water absorption, chloride ion penetration and neutralization, and this positive effect is increased as the coverage area of the barnacles expands. The barnacle adhesive at the interface between the concrete and the barnacles has a dense structure and significantly improves the pore structure of the concrete. The shells of the barnacles, barnacle adhesive, and the penetration of the adhesive collectively act as three lines of defense, aid in enhancing the durability of the concrete and then protecting the concrete from the severe marine environment.
Article
Full-text available
Several non‐stick coatings mainly based on silicones were tested under laboratory and field conditions in regard to their toxicity and their influence on the settlement of cyprid larvae (Balanus amphitrite). The toxicity was very low for most of the coatings. The antifouling properties were evaluated in exposure trials over 6 months at several stations along the German coast. The field trials were performed by stationary exposure and with periodically moved catamarans. The silicon coatings exhibited the best performance and 4 out of 9 of these were highly effective. The adhesion force of barnacles on silicones was substantially reduced in relation to control substrata. The leaching of organotin compounds derived from some of the silicone coatings used was less than, or in the range of, the induction leaching‐rates found in organotin‐stabilized Formica tubes.
Article
Full-text available
Concern over the environmental impact of traditional biocide‐containing antifouling formulations has led to the development of non‐toxic, fouling‐release coatings. An “ideal”; fouling‐release surface for a ship is one that reduces the tenacity of biofouling to a point that hydrodynamic self‐cleaning occurs under normal operating conditions. In order to address the feasibility of designing such a surface, an understanding of the relationship between the hydro‐dynamic forces and the strength of adhesion of the organism to the substrate is needed. This paper presents the results of an experimental investigation to measure the hydrodynamic lift and drag of an acorn barnacle (Balanus eburneus) attached to an instrumented foil. The results are used to generate a model of the tensile and shear stresses at the base of the barnacle, as a function of the water velocity. Predictions of the maximum barnacle adhesion strength in shear and tension that will allow detachment at operational speeds are offered.
Article
Full-text available
Barnacle release mechanisms and the durability of silicone coatings have been studied. Release studies were performed on both transparent, single‐layer silicone coatings and duplex silicone coatings. The release forces of pseudobarnacles (epoxied studs) and Chesapeake Bay barnacles (Balanus improvisus) were measured with a pull‐off (tension) tester; modes of release were revealed in video recordings of the separation process from transparent coatings on glass. Scratch tests with 0.8 mm spherically‐tipped diamond provided a measure of durability (tear resistance). Release forces from both coatings decreased as coating thickness increased. Both pseudobarnacles and barnacles separated by a peeling process, although differences in peeling modes were seen. The durability of coatings increased with increasing coating thickness. Release behavior is discussed in terms of a fracture mechanic's model for pull‐off separation, and the differences in adhesion between barnacles and pseudobarnacles are described.
Article
Measurements were made of the bond strength of cyprids and barnacles (Balanus improvisus) attached to poly(dimethylsiloxane) (PDMS) fouling-release coatings and poly(methyl methacrylate) (PMMA) panels as a function of barnacle base plate size (0.05-90 mm). The vertical forces necessary to dislodge cyprids and newly metamorphosed barnacles (base plate < 0.5 mm) were found to be equal for the two different substrates. This unexpected result was explained by the occurrence of cohesive failure in the cyprid/barnacle part. A significantly higher detachment force was observed for larger barnacles (base plate > 0.5 mm) when dislodged from the PMMA compared to the PDMS. Analysis of the failure surfaces with light microscopy and scanning electron microscopy (SEM) showed a gradual transition in failure mode from a total cohesive within the barnacle to a mixed failure mode during barnacle growth. This transition, which is a measure of the balance between the cohesive strength of the barnacle base plate and the adhesion bond to the surface, occurs earlier or with smaller barnacles when detached from the PDMS. The quantification of the remaining fraction of the base plate at the polymeric failure surfaces appears to be a function of barnacle bioadhesive bond strength and is, therefore, suggested to be used as a new parameter for evaluating the release properties of new coatings formulations.
Article
The interfaces between barnacle adhesive and a diverse range of materials were studied using light microscopy, scanning electron microscopy and atomic force microscopy. Barnacles grown on non-stick surfaces and those grown on easy-to-attach surfaces revealed differences in the calcified part of the barnacle base and the adhesive's ultrastructure. The formation of highly hydrated adhesive plaques was considered as the repair mechanism to bridge the gap between the barnacle base and the substratum. The formation of a loose web and water uptake results in higher elasticity at the expense of reduced cross-linking and low cohesive strength.
Article
This study compared the shear adhesion strength of barnacles, oysters and tubeworms on eight RTV 11‐based silicone fouling‐release coatings containing different silicone oil additives. It was found that adhesion strength differed among species and coating types. In most cases, oysters and tubeworms had higher adhesion strengths than barnacles. Barnacle adhesion strength was reduced on all coatings containing oil additives; however, this was not generally true for oysters and tubeworms. The difference in the adhesion strength among the three organisms tested in this study emphasizes the importance of understanding the fundamental interaction between marine invertebrate adhesives and the substratum.
Article
Flat-head studs, 7.2mm diameter, were epoxied to 0.1–0.9mm thick silicone layers attached to a rigid substrate. Pull-off tests were performed at nominally 1.5 N/s to establish critical pull-off forces and detachment modes at the epoxy–silicone interface. Most tests were performed on a two-layer coating consisting of a silicone top coat and a stiffer silicone bond coat. A few tests were performed with a single layer of transparent silicone on a glass substrate, which allowed video analysis of the interface during pull-off. Critical pull-off force, Pc, decreased as the thickness of the coated layers increased. The thickness dependence of Pc was shown to be in good agreement with a model presented by Kendall for pulling a rigid cylinder from a elastomeric glue layer, but modified to account for a two-layer elastomeric coating; Pc varied as t*1/2, where t* is an effective thickness, which depends on the bulk moduli of the coatings as well as their thicknesses. The observed pull-off mode was peeling, which occurred by nucleation, growth and coalescence of multiple voids initiated inside the contact area.
Article
Field studies were designed to examine the effect of surface type (with respect to bioadhesion) and biotic disturbance (with respect to predation and grazing) on the percentage cover by biofouling communities and on barnacle adhesion strength. Non‐toxic surfaces consisting of epoxy, Teflon™, and silicone were chosen to represent surfaces with high, medium, and low bioadhesion. Substrata were compared in a fully crossed experimental design; samples in fully caged treatments assessed biotic disturbance effects, those in partial cages assessed cage effects and uncaged panels acted as controls. The caged panels for all the surface types developed the greatest fouling cover and had the highest diversity of organisms. The number of taxa and survivorship were reduced on partially caged and uncaged panels. The uncaged and partially caged silicone surfaces remained almost free of fouling. The mean barnacle adhesion strengths for the caged treatments of epoxy and Teflon™ were less than their uncaged and partially caged equivalents. This may be a consequence of biological disturbance. The results have important implications for the interpretation of the performance of non‐toxic foul‐release coatings.
Article
Barnacle adhesion measurements provide an excellent method of assessing the foul release characteristics of non‐toxic surfaces. This paper describes both tensile and shear force methods that have been developed for field applications. Selected data are presented and the results discussed with respect to the variables that affect the observed adhesion values.
Article
Peeling of a viscoelastic adhesive from a solid substrate poses a wonderful problem of polymer flow. When the adhesive is peeled, it is also stretched in a direction normal to the substrate. The concomitant Poisson contraction creates a pressure gradient and thus induces a shear flow in the adhesive close to the delamination front. Earlier it was pointed out that the strength of viscoelastic adhesion would decrease if the shear stress in the adhesive is relaxed by a slip process at the interface. Here we report experimental results which confirm that viscoelastic adhesives do indeed slip on segmentally mobile organic surfaces at and near the crack tip regions. Evidence of slip was obtained from the interfacial displacements of small fluorescent particles when the adhesive was peeled from various substrates. While on most surfaces the slip distances were about 1−2 μm, a large slip (13 μm) was observed on segmentally mobile tethered chains of polydimethylsiloxanes (silicones). On the latter surface, slippage is so extensive that the adhesive flow pattern near the delamination zone is like plug flow. We believe it is due to the propensity of huge slippage that the silicone-containing polymers exhibit their unusually low adhesion to most materials.