Universidad Nacional de Chimborazo
NOVASINERGIA 2018, Vol. 1, No. 2, junio-noviembre (90-99)
ISSN: 2631-2654
https://doi.org/10.37135/unach.ns.001.02.10
Research Article
http://novasinergia.unach.edu.ec
Wet Adhesive Properties of Asian Green Mussel (Perna viridis) Foot
Protein Pvfp-5: An Underwater Adhesive Primer
Propiedades Adhesivas Húmedas de la Proteína del Pie de Mejillones Verdes
Asiático (Perna viridis) Pvfp-5: Una Base Adhesiva Subacuática
Navinkumar J Patil
1
*, Paola Gabriela Vinueza Naranjo
2
, Bruno Zappone
3
1
Dipartimento di Fisica, Università della Calabria, Arcavacata di Rende (CS), 87036, Italy.
2
DIET, Sapienza Università di Roma, Roma (RM), Italy.
3
Consiglio Nazionale delle Ricerche, Istituto di Nanotecnologia (CNR-Nanotec), Rende (CS) 87036, Italy;
paola.vinueza@uniroma1.it; bruno.zappone@cnr.it
* Correspondence: navinjpatil@gmail.com
Recibido 30 octubre 2018; Aceptado 28 noviembre 2018; Publicado 10 diciembre 2018
Abstract:
Asian green mussels (Perna viridis) are bivalves that attach firmly to rocks in wave-battered
intertidal seashores via a proteinaceous secretion. P. viridis mussels follow a precisely time-
regulated secretion of adhesive proteins where P. viridis foot protein-5 (Pvfp-5) was identified as
the first protein to be secreted during the formation of adhesive plaque. The high content of
catecholic amino acid 3,4-dihydroxy-L-phenylalanine (DOPA) (11 mol%) and cysteine (Cys)
(15 mol%) in Pvfp-5 and its localization near the plaque-substrate interface have prompted
speculation that the vanguard protein Pvfp-5 plays a key role as an adhesive primer in underwater
adhesion of P. viridis mussels. Surface Force Apparatus (SFA) was used to probe the adhesive
properties of Pvfp-5 at the nano-scale where pH dependent wet adhesion and antioxidant activity
of foot-extracted and purified Pvfp-5 were investigated. The study revealed that Pvfp-5 with its
high DOPA and CYS-content maintains adhesion even at higher pH by overcoming the
spontaneous oxidation of DOPA to quinone. SFA results are consistent with the apparent function
of Pvfp-5 acting as an adhesive primer, overcoming repulsive hydration forces by displacing
surface-bound water and generating strong and durable surface adhesion. Our findings reveal
molecular-scale insights that should prove relevant and impact material sciences to help the
development of new generation of wet-resistant adhesives, coatings and glues for biomedical,
therapeutic and antifouling applications.
Keywords:
Adhesion, Cysteine (CYS), Dihydroxy-L-phenylalanine (DOPA), Mussel foot proteins (Mfps),
Perna viridis foot protein (Pvfp), Surface Forces Apparatus (SFA)
Resumen:
Los mejillones verdes asiáticos (Perna viridis) son bivalvos que se adhieren firmemente a las rocas
en las costas intermareales golpeadas por las olas a través de una secreción proteica. Los
mejillones de P. viridis siguen una secreción de proteínas adhesivas regulada en el tiempo, donde
la proteína del pie-5 de P. viridis (Pvfp-5) fue identificada como la primera proteína que se secreta
durante la formación de la placa adhesiva. El alto contenido de aminoácido catecólico 3,4-
dihidroxi-L-fenilalanina (DOPA) (
11% en moles) y cisteína (Cys) (
15% en moles) en Pvfp-5 y
su localización cerca de la interfaz placa-sustrato tienen provocó la especulación de que la
proteína de vanguardia Pvfp-5 desempeña un papel clave como imprimación adhesiva en la
adhesión bajo el agua de los mejillones de P. viridis. El aparato de fuerza de superficie (SFA) se
usó para probar las propiedades adhesivas de Pvfp-5 a escala nanométrica, donde se investigó la
adhesión húmeda dependiente del pH y la actividad antioxidante de Pvfp-5 purificado y extraído
con el pie. El estudio reveló que Pvfp-5 con su alto contenido de DOPA y CYS mantiene la adhesión
incluso a un pH más alto al superar la oxidación espontánea de DOPA a quinona. Los resultados
de SFA son consistentes con la función aparente de que Pvfp-5 actúa como imprimación adhesiva,
superando las fuerzas de hidratación repulsivas al desplazar el agua unida a la superficie y
generando una adhesión superficial fuerte y duradera. Nuestros hallazgos revelan conocimientos
a escala molecular que deberían ser relevantes e impactar en las ciencias de los materiales para
ayudar al desarrollo de la nueva generación de adhesivos, recubrimientos y pegamentos resistentes
a la humedad para aplicaciones biomédicas, terapéuticas y antiincrustantes.
Palabras
clave:
Adhesión, Cisteína (CYS), Dihidroxi-L-fenilalanina (DOPA), Proteínas del pie de mejillón (Mfps),
Perna viridis proteína del pie (Pvfp), Aparato de fuerzas de superficie (SFA).
http://novasinergia.unach.edu.ec 91
1 Introduction
Animal attachment to a substrate is very different
in terrestrial and aquatic environments. In
terrestrial environment, gravity is considered as
the most important detachment force. However, in
submerged conditions gravity is nearly balanced
out by buoyancy but flow forces such as drag and
lift are of higher importance (Ditsche et al. 2014).
As is well known, water or moisture significantly
compromises the performance of most man-made
adhesives, as water effectively competes for
surface bonding and eliminates contributions of
van der Waals interactions (Comyn, 1981; Lee et
al. 2011).
Despite these prevalent challenges and the
harshness of the physical environment, wave-
swept rocky shores and ship hulls are home to a
variety of organisms that have evolved to attach
themselves permanently to variety of substrates
that are wet, saline, corroded, and/or fouled by
biofilms. Figure 1 shows variety of invertebrate
animals attached underwater to wave-swept rocks
and ship hulls that have evolved separate and
distinct working strategies for their particular
underwater bonding requirements to protect
themselves against biological, chemical, and
mechanical stresses from the ocean (Wilker, 2010;
Stewart et al. 2011).
Figure 1: Mussels, barnacles, and tube worms sticking
to rocks (A-C); a starfish is shown adhering to a sheet of
glass (D). Figure adopted from (Wilker, 2010).
Mussels are prominent examples of sessile type
organisms attaching themselves permanently to
variety of underwater solid substrates. A detailed
study on Mytilus genus (blue mussels) during the
last decade has enhanced our understanding of
underlying biochemical and biophysical
adaptations of mussels for opportunistic and
durable adhesion to moist mineral and metal oxide
surfaces (Yu et al. 2011a; Wei et al. 2013; Yu et
al. 2013). Mussel adhesion is mediated by a
holdfast structure known as the byssus, essentially
a bundle of macroscopic extensible high-
performance fibers tipped by flattened adhesive
plaques that tether the mussel firmly to a variety of
hard surfaces and play a critical role in the ability
of mussels to dominate space on many temperate
shores worldwide (Waite et al. 2005; Lin et al.
2007; Priemel et al. 2017). Byssal thread consists
of several collagen-like proteins responsible for the
mechanical properties of the fibers while the plaque is
composed of number of protein-based adhesives
containing 3,4-dihydroxyphenylalanine (DOPA),
a post translationally modified amino acid that
forms hydrogen bonds with almost any surface
chemistry under wet conditions. Mussels adhere to
surfaces underwater through the secretion,
deposition, and complexation of these DOPA-
containing proteins, known as mussel foot proteins
(Mfps), to resist the forces produced by crashing
waves a feat unmatched by current industrial
glues (Priemel et al. 2017; Hamada et al. 2017).
Figure 2: Cartoon of mussel byssus and proposed model
of mussel’s adhesion mechanism. (A) schematic
overview of a mussel; (B) schematic of Mfps delivery
from different precursors under mussel’s foot; (C) Mfp
incremental secretion: step 1, acid; step 2, primers (mfp-
3 and mfp-5); step 3, reducing agent (mfp-6); step 4,
bulk adhesives (mfp-2 and mfp-4) and collagen; and step
5, coating proteins (mfp-1); (D) schematic showing the
distribution of major mfp’s and collagen in byssal
plaque. Figure adopted from (Kollbe Ahn, 2017).
Within the adhesive plaque of blue mussels
(Mytilus species), six major mussel foot proteins
(Mfp’s 1-6) have been identified with substantially
different amino acid sequences which are
responsible for adhesion and coating of the mussel
byssus (Kollbe Ahn et al. 2017; Lu et al. 2013).
Each Mfp is found at a different location in the foot
and byssus, and each one is believed to play a
distinct and important role, as shown in figure 2
(Ahn et al. 2017; Silverman at al., 2007; Hwang et
al. 2010). Mfps are polyelectrolytes distinguished
by their isoelectric point (pI 10) and mussels
appear to gain surface access for their adhesive
proteins by using their own protein-based ions to
outcompete the ions in their saltwater environment
http://novasinergia.unach.edu.ec 92
(Wilker, 2015). Also, the adhesive proteins found
near the plaque-substrate interface are the proteins
with the highest amount of DOPA, thus prompting
the important and involvement of DOPA in surface
adhesion. DOPA contains a chemical group called
catechol (3,4-dihydroxyphenyl), made from a
benzene ring bearing two adjacent hydroxyl (-OH)
groups, as shown in figure 3. Further studies on
catecholic amino acid revealed that the hydroxyl
groups of the catechol form chemical bonds with
rocks and other substrates that help to stick the
mussel in place (Ornes, 2013).
Figure 3: Dopamine and DOPA showing catechol
moiety.
Although DOPA-mediated wet adhesion is strong
and versatile on different surfaces, it has a
troubling tendency to readily oxidize to Dopa-
quinone and formation of Dopaquinone diminishes
adhesion of Mfps to surfaces by at least 80-95%.
There are several factors which may lead to
oxidation of DOPA to Dopa-quinone.
Spontaneous oxidation occurs at alkaline pH and
auto-oxidation which refers to spontaneous
oxidation of cathechol in the presence of oxygen at
neutral pH leads to loss of adhesion of Mfps (Lee
et al. 2006; Menyo et al. 2013). Regulation of
redox in the adhesive byssus of marine mussels is
an exotic and fascinating case in point of
extracellular redox where higher cysteine (CYS)
content in some Mfps reduces the risk of DOPA
oxidation.
Recently identified byssal plaque proteins from the
Asian green mussel Perna viridis known as Pvfps
are all enriched with cysteine (CYS) and DOPA-
residues unlike Mfps, where only particular variant
of plaque protein (e.g. Mfp-6) has high CYS
content which acts as an anti-oxidant, reducing
dopamine back to DOPA. Also, the primary
sequences of Pvfps share very low homology with
earlier studied Mfps (Petrone et al. 2015). Our
current knowledge on the nature and properties of
mussel adhesion is based only on the limited set of
mussel species studied. Therefore, understanding
the molecular basis of adhesion in green mussels
can lead us to alternate mussel inspired
bioadhesives and targeted anti-fouling strategies.
Natural foot proteins from mussels of P. viridis
were extracted by artificially inducing the live
mussels to secrete plaque-forming proteins which
were collected from the groove at the tip of the foot
organ (Petrone et al. 2015; DeMartini et al. 2017).
It was identified that the byssal plaque proteins
were secreted in time-regulated manner where
Perna viridis foot protein-5 (Pvfp-5) was the first
protein secreted by Asian green mussels to initiate
interaction with the substrate, displacing
interfacial water molecules and forming adhesive
bonds with the substrate which was then followed
by other variants of Pvfps. Pvfp-5, which typically
appeared after 10 s of saline injection, had
molecular weight (MW) in the range 8-10 kDa and
pI 9.2. Amino acid sequence of Pvfp-5 revealed
that it contained 12 CYS residues, accounting for
15 mol% of its amino-acid content. It also revealed
21 mol% Tyr side chains in its primary sequence
and 9 DOPA residues accounting for 11 mol%
DOPA in Pvfp-5. Pvfp-5 predominantly exhibited
random coil structure and the dynamic light
scattering (DLS) revealed a hydrodynamic
diameter, d
H
= 9.500.26 nm (Petrone et al. 2015).
Here we report on our investigation of the
attractive forces and work of adhesion between
thin films of purified Pvfp-5 under changing pH
conditions. The results support the conclusion that
distinct mussel species adopt different strategies
for underwater adhesion, where Pvfp-5 unlike
Mfp-5 maintains adhesion even at higher pH due
to its high DOPA and CYS contain and CYS-based
redox activity.
2 Methodology
2.1 Extraction and Purification of
Mussel Foot Protein
P. Viridis foot protein (Pvfp-5) was extracted from
the foot organ of P. viridis mussels collected from
the northern coast of Singapore. Methodology for
extraction and purification of Pvfp-5 has been
previously reported (Petrone et al. 2015). The
secretion of adhesive foot proteins of P. Viridis
mussels was triggered by injecting 1 ml of KCl
phosphate buffer (0.56M, pH 7.2) into the mussels
pedal nerve located at the base of the foot.
Injecting KCl solution in the foot organ reasonably
mimics the natural byssus secretions of P. viridis,
and is referred as saline-induced mussel secretion.
These secretions of P. viridis were then collected
from the groove at the tip of the foot organ.
2.2 Surface Forces Apparatus
(SFA)
Adhesion and normal force measurements
between Pvfp-5 layers adsorbed on smooth and
chemically inert surfaces of mica (a common
alumino-silicate clay mineral) were obtained using
http://novasinergia.unach.edu.ec 93
the a SFA Mark III (Surforce LLC, Santa Barbara,
CA, USA). Figure 4 shows a detailed schematic
diagram of the SFA set-up used in this study. The
SFA is a technique to measure the interaction
forces between two macroscopic surfaces as a
function of absolute distance between them using
multiple beam interferometry (MBI) and fringes of
equal chromatic order (FECO). A detailed
description of the technique can be found in
Israelachvili & McGuiggan (Israelachvili et al.
1990).
Figure 4: Schematic of the SFA set-up: crossed cylinder
experimental set-up of the SFA-MK III (A); resulting
fringe pattern of fringes of equal chromatic order
(FECO) (B). Pvfp-5 was adsorbed on both mica surfaces
at pH 3 for 20 minutes followed by rinsing with protein-
free buffers of variable pH to study the effect of pH on
cohesive interaction between Pvfp-5 molecules.
All the SFA experiments were carried out using
mica as the base surface. Since the freshly cleaved
thin sheets of mica are atomically smooth,
optically transparent, and chemically inert, they act
as ideal substrates for SFA experiments to study
adhesive properties of Pvfp-5 proteins. The outer
opposing sides of mica sheets were coated with a
semitransparent layer of silver ( 55 nm) using
thermal evaporation or chemical sputtering
technique. Identically thick and flat back-slivered
mica sheets (thickness 1-5 µm) were glued with
the silver coated side on the cylindrical glass disks
(radius of curvature, R=1.82.2 cm) using
thermosetting glue (EPON 1007 by Shell). Two
semi-transparent curved disks were then mounted
orthogonally to each other in an inert nitrogen
atmosphere inside an airtight SFA chamber. The
top disk was fixed and the bottom disk was
attached to the free end of the double cantilever
spring where the stiffness kof the spring was in
the range of 250-950 N/m. The spring and
displacement mechanism allowed the motion of
opposing disks into a well-defined single asperity
contact at a given force.
During the experiment, collimated white light is
guided through these disks which act as semi-
transparent mirrors. When the surfaces are moved
close together, together with the intervening
medium it forms a symmetrical three-layer Fabry-
Perot interferometer employing MBI. The
constructive and destructive interference of the
white light at discrete wavelengths leads to the
generation of the FECO, which are an infinite
series of alternating sharp bright and dark bands
which are then detected by the spectrophotometer.
A typical FECO is shown in figure 4B, which
clearly depicts the flat section of FECO, indicating
an extended flat zone of contact. The FECO allows
the determination of the inter-mirror distance with
a resolution below 1Å. During the initial contact
mica-mica contact in dry nitrogen before
adsorbing proteins, the distance D=0 was defined
by recording the set of wavelengths. A change in
FECO fringe position was used to calculate the
shift in distance (D), of two opposing mirrors,
where discrete spectral wavelengths of mercury
were used to calibrate the spectrograph.
The SFA experiments were performed by precisely
approaching the surfaces until they come in
adhesive contact followed by retracting them from
one another. The adhesive forces were then
calculated by using Hooke’s law (F=k×ΔD) where
k is the known stiffness of the spring on which
lower surface was mounted and ΔD is the shift in
distance from the contact (of adhesive surfaces) to
their point of separation. Because of the curved
geometry of the surfaces, the measured force was
normalized by the radius of curvature of the
surfaces (R). By using the Derjaguin
approximation, the normalized force F(D)/R can
be directly transformed into the interaction energy
per unit area between the two flat surfaces, which
is given by E(D) = F(D)/2
R.
The error on F/R was smaller than 0.1mN/m (force
detection threshold) and the error on D about 0.5
nm.
2.3 Sample Preparation for SFA
Force Measurements
Mussel adhesive proteins like Pvfp-5 are
polyelectrolytes possessing ionization tendencies
which define their interactions. Therefore, careful
selection of pH was made during Pvfp-5
adsorption and for buffer solutions used for rinsing
the adsorbed proteins (Yu et al. 2011a; Danner et
al. 2012). Also, well defined salt concentration
was used to provide sufficient counterions to
prevent the electrostatic self-interaction from
dominating in the SFA experiments (Lin et al.
2007; Yu et al. 2011a; Danner et al. 2012; Lu et al.
2013).
Following mica-mica contact in dry air, Pvfp-5
proteins were adsorbed on both mica surfaces from
http://novasinergia.unach.edu.ec 94
0.02 mg/ml solution of protein in acid saline buffer
containing 10 mM acetic acid (AcOH) and 0.25M
potassium nitrate (KNO
3
) in purified water (Milli-
Q grade from Millipore) at pH 3. After a 20 min
adsorption, the surfaces were rinsed with protein-
free acidic buffer solution. Following positioning
into the SFA, the surfaces were brought almost
into contact with a droplet of buffer solution
providing a bridge between the two surfaces. To
keep the buffer droplet from evaporating, a second
droplet of buffer was placed in the box to maintain
the vapor pressure.
Figure 5: Representative force-distance (FD) profile: (A) Normalized force F/R measured between two Pvfp-5 coated
mica surfaces at pH 3. Positive forces are repulsive and negative forces are attractive. (B) Semi-logarithmic plot
showing the range of repulsion 2T, force detection threshold (dotted line) and exponential curves F/R=Ae
-D/d
. Solid
symbols (, , and ) indicate forces on approach and open symbols (, , and ) indicate forces on
retraction of surfaces. Each symbol (circle and square) corresponds to different contact position and different colors
indicate different surface approach/retraction cycles.
The influence of pH was investigated by flushing
the droplet of acidic buffer (~100 µL) with an
excess of the new treatment buffer (3ml) with
neutral and basic pH. Treatment buffers were
1) Neutral saline buffer, 0.25M KNO
3
, pH 7
(without HCl or NaOH)
2) Alkaline/basic saline buffer, 0.25M KNO
3
, pH
≈ 10 (with NaOH)
KNO
3
was used to replace sodium chloride in the
buffer solution to avoid corrosion of the thin silver
layers under the mica substrates. Having high
concentrations of chloride ions in solution, affect
the quality of the optical fringes during the SFA
experiments. Degassed buffer solutions were used
for dissolving Pvfp-5 as well as for rinsing.
Degassing helps removal of dissolved gases in
solution and minimizes the occurrence of bubbles
between two surfaces while performing SFA
experiments. Multiple force runs were recorded for
each pair of mica and protein adsorption and SFA
force measurements were done at a fixed
temperature of 25±1 C.
3 Results and Discussion
Pvfp-5 is the first protein to be secreted by Asian
green mussels (Perna viridis) in the precisely time
regulated secretion of a series of adhesive proteins.
Thus, Pvfp-5 is the first protein to initiate contact
with the substrate which acts as an adhesive
primer. Pvfp-5 is enriched with cysteine and
tyrosine (Tyr)/ DOPA residues and its localization
at plaque-substrate interface have provoked
interest in its adhesive properties. Using SFA, the
effects of changing pH on adhesion capability of
Pvfp-5 was explored. Since Pvfp-5 was adsorbed
onto both the opposing mica surfaces, the adhesion
forces reported below are proportional to the
cohesive interaction between Pvfp-5 molecules.
3.1 Adhesion between Pvfp-5
Layers at Acidic pH
Interaction forces (F/R) between the thin layers of
Pvfp-5 adsorbed symmetrically to both the mica
surfaces at acidic pH were measured as a function
of separation distance (D) (figure 5A). Seven
distinct force runs were performed at three
different contact positions where the surfaces were
brought into contact in the acid saline buffer and
then separated after a brief contact time. At pH 3
and 0.25M ionic strength, the maximum adhesion
force after separation was measured to be F
a
/R =
5.42 mN/m, which can be converted to adhesion
energy, E = (2/3
) F
a
/R 1.14 mJ/m
2
. Different
force runs displayed variability in adhesion force
since factors like contact time or maximum load
applied during compression were not explicitly
controlled. However, a clear trend emerged that
adhesion decreased with time as the surfaces were
repeatedly approached and retracted at the same
contact position and as different contact positions
were tested. This was most likely due to oxidation
of DOPA (Ahn et al. 2014; Nicklisch et al. 2012).
http://novasinergia.unach.edu.ec 95
The average value of the adhesion force was F/R =
3.63 mN/m. The repulsive force was recorded for
all the force runs during surface approach, showing
that adhesion was generated by the formation of
adhesive bonds between interacting Pvfp-5
molecules. The force consistently exceeded the
detection threshold of 0.1 mN/m when the surface
separation distance D became smaller than 2T =
24.530.8 nm (figure 5B), where T was the
thickness of adsorbed protein layer on one surface.
Since the Debye length of the saline buffer was
smaller than 1 nm, the recorded long repulsion
range (2T) can be attributed to the overlap between
protein layers adsorbed on opposite surfaces, each
having an approximate thickness T = 12.215.4
nm. The value of the thickness of adsorbed protein
layer on mica was larger than the size of Pvfp-5
molecule determined by DLS (Petrone et al. 2015),
indicating aggregation behavior of Pvfp-5
molecules at pH 3. The aggregation of Pvfp-5
molecules in solution was also confirmed by
dynamic light scattering (DLS) results indicating
the hydrodynamic diameter of Pvfp-5 to be 8.56
nm at pH 4, which was most likely via DOPA
DOPA crosslinks.
The maximum value thickness (T) of a Pvfp-5
layer on mica substrate was obtained during the
first approach, indicating that Pvfp-5 was adsorbed
as a soft hydrated layer, as opposed to a compact
“hard-wall” coating, allowing the protein
molecules and water to rearrange and move as the
compressive force was increased. This also
indicates that the protein layers were flattened,
perhaps due to the formation of additional DOPA-
DOPA crosslinks or DOPA-mica hydrogen bonds
on compression. Besides, the determination of
thickness (T) of the adsorbed protein layer on mica
substrates, a “hard wall” distance (D
HW
) was also
measured using SFA.
Hard wall distance is defined as the asymptotic
thickness of confined Pvfp-5 under increasing
normal load or pressure. Hard wall distance was
determined by compressing the adsorbed Pvfp-5
layer until FECO fringes cease to move towards
the lower wavelength side of the signal. The
measured hard wall distance for Pvfp-5 at pH 3
was reported to be D
HW
≈ 2.6 nm.
Figure 6: Representative force-distance (FD) profile: (A) Normalized force F/R measured between two Pvfp-5 coated
mica surfaces at neutral pH (=7). (B) Semi-logarithmic plot showing the range of repulsion 2T, force detection threshold
(dotted line) and exponential curves F/R=Ae
-D/d
. Solid symbols (●, , and ) indicate forces on approach and open
symbols (, , and ) indicate forces on retraction of surfaces. Each symbol (circle and square) corresponds to
different contact position and different colors indicate different surface approach/retraction cycles.
3.2 Adhesion between Pvfp-5
Layers at Neutral pH
Interaction forces (F/R) between the thin layers of
Pvfp-5 at neutral pH were measured by first
adsorbing Pvfp-5 to both mica surfaces at acidic
pH, followed by rinsing the surfaces with a
protein-free neutral saline solution. At pH 7 and
0.25M ionic strength, the surfaces were brought
together and the maximum adhesion force
measured after separation was F
a
/R = 1.28 mN/m
(figure 6A), corresponding to the work of
adhesion, E = (2/3
) F
a
/R 0.27 mJ/m
2
. Maximum
adhesion was measured during separation of the
first approach-retraction cycle which consequently
reduced for the next three force runs recorded on
the same contact position.
This trend at pH 7 was similar to that observed at
pH 3 earlier, which was most likely due to DOPA
oxidation. The average value of the adhesion force
at pH 7 was F/R = 0.84 mN/m. Hysteresis was
present for all the force runs recorded and
repulsive forces were observed during surface
approach with the onset of repulsion at 2T = 36
45.8 nm (figure 6B). Thin hard-wall distance, D
HW
2.2 nm was measured at neutral pH.
http://novasinergia.unach.edu.ec 96
3.3 Adhesion between Pvfp-5
Layers at Basic pH
To assess the interaction forces (F/R) between
Pvfp-5 layers at basic pH, thin layers of Pvfp-5
were initially adsorbed on both mica surfaces at
acidic pH, followed by rinsing with protein-free
alkaline buffer solution. Two force runs were
recorded at pH 10 and 0.25M ionic strength and
small adhesion were measured for both the force
runs. The maximum adhesion force measured at
pH 10 was 0.68 mN/m (figure 7A), corresponding
to a work of adhesion, E = 0.14 mJ/m
2
.
The average value of the adhesion force at pH 10
was F/R = 0.6 mN/m. The approaching branch of
force curves revealed a repulsion setting in at a
comparably smaller surface separation of 2T =
11.429.4 nm than that recorded for acidic and
neutral pH (figure 7B). The measured hard wall
distance for Pvfp-5 at pH 10 was reported to be
D
HW
≈ 3.85 nm.
3.4 Discussion
Mytilus edulis foot protein-5 (Mefp-5) is one of the
adhesive proteins in the byssal adhesive plaque of
the mussels of M. edulis that has been reported to
have highest molecular DOPA concentration
which is localized near the plaque substrate
interface similar to Pvfp-5 (Danner et al. 2012,
Petrone et al. 2015). The adhesion capabilities of
Pvfp-5 were assessed by SFA experiments using
mica as the substrate, which enabled direct
comparison of Pvfp-5 with extensively studied
Mefp-5. Strong adhesion force was consistently
measured between symmetrically adsorbed Pvfp-5
films at pH 3 (maximum adhesion force, F
a
/R =
5.42 mN/m). Previous studies on Mefp-5 showed
adhesion force, F
a
/R = 6.0 mN/m, which slightly
exceeded the adhesion force of Pvfp-5.
Considering the 30 mol% DOPA content of Mefp-
5 compared to 11 mol% DOPA in Pvfp-5, the
adhesion force between interacting proteins films
of both the adhesive proteins were not significantly
different (Danner et al. 2012).
It has been reported that exposure of Mefp-5 to pH
above 7.5 reduces its adhesion by 95% due to rapid
oxidation of DOPA to Dopaquinone (Kan et al.
2014). However, the CYS-enriched Pvfp-5 was
adhesive at both neutral and basic pH with the
maximum adhesion force, F
a
/R = 1.28 mN/m at pH
7.0 and F
a
/R = 0.68 mN/m at pH 10 (table 1).
Thus, increasing the pH of gap solution from pH3
to pH7 and pH10, reduced the adhesion force by
approximately 75% and 87%, respectively,
compared to initial adhesion force measured at
acidic pH. Therefore, CYS residues in the amino
acid sequence of Pvfp-5 participate in the
reduction of quinones by providing a pH-
dependent antioxidant activity to Pvfp-5, the first
protein secreted by P. viridis, justifying its role as
a vanguard protein that acts as an adhesive primer
which initiates first adhesive interactions with the
substrate.
Besides adhesion, table 1 summarizes the variation
of other parameters with respect to change in pH
conditions which reveals an important trend: the
hard wall and the thickness (T) of Pvfp-5 layer on
mica, changed with pH.
Table 1: Summary of various parameters for Pvfp-5
interactions in symmetric configuration measured by
SFA.
pH
Pvfp-5 film
thickness,
T (nm)
Hardwall
distance,
D
HW
(nm)
pH 3
12.2 15.4
2.6
pH 7
18 22.9
2.2
pH 10
5.7 14.7
3.8
The film thickness (T) of Pvfp-5 on mica at pH 3
was in the range 12.2-15.4 nm with hard wall, D
HW
(pH 3) 2.6 nm. After increasing the pH of the gap
solution to 7.0, repulsion at comparatively higher
distance was noticed during approach. The film
thickness (T) at pH 7 was measured to in the range
of 18-22.9 nm with hard wall, D
HW
(pH 7) 2.2
nm. The increase in film thickness at pH 7 suggests
that the Pvfp-5 film expands at neutral pH;
however, the hard wall distance measured on
compression remained almost unchanged. When
the pH of the gap solution was changed from acidic
to alkaline, significantly low repulsion range
corresponding to film thickness (T) in the range
5.7-14.7 nm was recorded. Surprisingly, the hard
wall distance measured upon compressing the
Pvfp-5 layer under higher loads was D
HW
(pH 10)
= 3.8 nm, which was higher than D
HW
recorded at
pH 3 and pH 7, respectively. Lower film thickness
and higher hard wall at pH 10 suggests a more
compact and rigid conformation of Pvfp-5, which
may prevent the exposure of DOPA residues to
higher pH, thus preventing DOPA from
spontaneous oxidation.
Surprisingly, SFA experiments on Pvfp-5 also
showed that adsorption of thicker layers of protein
on mica resulted in diminished adhesion (results
not shown). Series of experiments performed with
varying adsorption time and protein concentration
in solution revealed that thinner layer have more
exposed DOPA molecules available to interact and
form adhesive bonds whereas, thicker layers
screen DOPA molecules as well as enhances cross-
linking between DOPA molecules making them
unavailable for subsequent adhesive interactions.
http://novasinergia.unach.edu.ec 97
Figure 7: Representative force-distance (FD) profile: (A) Normalized force F/R measured between two Pvfp-5 coated
mica surfaces at pH 10. (B) Semi-logarithmic plot showing the range of repulsion 2T, force detection threshold (dotted
line) and exponential curves F/R=Ae
-D/d
. Solid symbols ( and ) indicate forces on approach and open symbols (
and ) indicate forces on retraction of surfaces. Different colors indicate different surface approach/retraction cycles.
The film thickness (T) of Pvfp-5 on mica at pH 3
was in the range 12.2-15.4 nm with hard wall, D
HW
(pH 3) 2.6 nm. After increasing the pH of the gap
solution to 7.0, repulsion at comparatively higher
distance was noticed during approach. The film
thickness (T) at pH 7 was measured to in the range
of 18-22.9 nm with hard wall, D
HW
(pH 7) 2.2
nm. The increase in film thickness at pH 7 suggests
that the Pvfp-5 film expands at neutral pH;
however, the hard wall distance measured on
compression remained almost unchanged. When
the pH of the gap solution was changed from acidic
to alkaline, significantly low repulsion range
corresponding to film thickness (T) in the range
5.7-14.7 nm was recorded. Surprisingly, the hard
wall distance measured upon compressing the
Pvfp-5 layer under higher loads was D
HW
(pH 10)
= 3.8 nm, which was higher than D
HW
recorded at
pH 3 and pH 7, respectively. Lower film thickness
and higher hard wall at pH 10 suggests a more
compact and rigid conformation of Pvfp-5, which
may prevent the exposure of DOPA residues to
higher pH, thus preventing DOPA from
spontaneous oxidation.
Surprisingly, SFA experiments on Pvfp-5 also
showed that adsorption of thicker layers of protein
on mica resulted in diminished adhesion (results
not shown). Series of experiments performed with
varying adsorption time and protein concentration
in solution revealed that thinner layer have more
exposed DOPA molecules available to interact and
form adhesive bonds whereas, thicker layers
screen DOPA molecules as well as enhances cross-
linking between DOPA molecules making them
unavailable for subsequent adhesive interactions.
4 Conclusion
In summary, the objective of this work was
understanding the molecular interactions between
Pvfp-5 proteins adsorbed on mica substrates at
different pH conditions using the SFA technique.
The results provide important insights into the wet
adhesion mechanism of Asian green mussels (P.
Viridis), whereby a DOPA-rich primer (Pvfp-5) is
initially secreted because of its superior adsorptive
and adhesive abilities to interact with foreign
surfaces. Other experimental studies of Pvfp-5
shows the ability of Pvfp-5 to act as an adhesive
primer to displace surface-bound water from
hydrophilic surfaces allows the formation of strong
and durable bonds via its adhesive DOPA residues
exposed on the protein surface, hence creating an
environment conducive to the assembly of other
plaque components (Petrone et al. 2015).
By studying the effects of changing pH on
adhesive properties of Pvfp-5, the emerging
picture is that Pvfp-5 with its high DOPA and
CYS-content, maintains adhesion even at higher
pH and acts as a self-antioxidant, overcoming the
spontaneous oxidation of DOPA to quinone at
higher seawater pH ( 8) with saturating levels of
dissolved oxygen. Also, tyrosine, DOPA and
cysteine motifs in Pvfp-5 play a key role in
enabling the robust underwater adhesion exhibited
by P. viridis. More broadly, the study provided a
molecular basis for understanding the impressive
underwater adhesion of marine organisms, an
insight that should prove relevant for diverse areas,
from designing of mussel inspired bioadhesives to
engineering targeted anti-fouling strategies.
DOPA- and CYS- containing proteins like Pvfp-5
are crucial for wet adhesion in mussels that retards
oxidation by shielding the amino acids from the
solvent and endowing the protein with the ability
to maintain adhesion at neutral as well as basic pH
levels (Kaushik et al. 2015). Recent studies on the
underwater adhesion mechanisms of marine
fouling organisms have been explored by SFA,
where three key mechanisms of successful
underwater adhesion, i.e. DOPA and REDOX
chemistry, cation-π interactions and complex
coacervation, have been suggested (Hwang et al.
2010; Yu et al. 2011; Gebbie et al. 2017). SFA and
other techniques have not yet fully elucidated all
the adhesion mechanisms of the marine fouling
organisms like mussels, but the key mechanisms
http://novasinergia.unach.edu.ec 98
exploited should be effective strategies for the
development of bio-inspired medical adhesives
and tissue sealants, antifouling surfaces for
medical devices and environmentally friendly
antifouling paints for use in the marine industry.
Further, understanding the aspects of natural redox
control can provide fundamentally important
insights for adhesive polymer engineering, surface
modifications and antifouling strategies.
Interest Conflict
The authors declare that they have no conflict of
interest.
Acknowledgment
We would like to thank Prof. Ali Miserez (NTU,
Singapore) for providing the isolated and purified
mussel foot proteins. This research was supported
by the University of Calabria (Unical), Italy
through its doctoral fellowship to N. J. Patil under
Bernardino Telesio School of Science and
Technique. We thank CNR-Nanotec and
Department of Physics, Unical for various
experimental and computing facilities.
References
Comyn, J. (1981). The relationship between joint
durability and water diffusion. In Developments
in Adhesives; Kinloch, A. J., Edition, U.K.:
Applied Science Publishers: Barking, Vol. 2,
279-313.
Danner, E. W., Kan, Y., Hammer, M. U., Israelachvili,
J. & Waite, H. (2012). Adhesion of Mussel Foot
Protein Mefp-5 to Mica: An Underwater
Superglue. Biochem, 51(33), 65116518.
DeMartini, D. G., Errico, J. M., Sjoestroem, S., Fenster,
A. & Waite, H. J. (2017). A cohort of new
adhesive proteins identified from transcriptomic
analysis of mussel foot glands. J R Soc Interface,
14(131), 0151.
Ditsche, P. & Summers, A. P. (2014). Aquatic versus
terrestrial attachment: Water makes a difference.
Beilstein J Nanotechnol, 5, 24242439.
Gebbie, M. A., Wei, W., Schrader, A. M., Cristiani, T.
R., Dobbs, H. A., Idso, M., Chmelka, B. F.,
Waite, H. & Israelachvili, J. N. (2017). Tuning
underwater adhesion with cation–π interactions.
Nat Chem, 9, 473-479.
Hamada, N. A., Roman, V. A., Howell, S. M. & Wilker,
J. J. (2017). Examining Potential Active
Tempering of Adhesive Curing by Marine
Mussels. Biomim, 2(16), doi:
10.3390/biomimetics2030016.
Hwang, D. S., Zeng, H., Masic, A., Harrington, M. J.,
Israelachvili, J. N. & Waite, J. H. (2010).
Protein- and metal-dependent interactions of a
prominent protein in mussel adhesive plaques. J
Biol Chem, 285(33), 2585025858.
Hwang, D. S., Zeng, H., Srivastava, A., Krogstad, D. V.,
Tirrell, M., Israelachvili, J. N. & Waite, H. J.
(2010). Viscosity and interfacial properties in a
mussel-inspired adhesive coacervate. Soft
Matter, 6, 32323236.
Israelachvili, J. N. & McGuiggan, P. M. (1990).
Adhesion and short-range forces between
surfaces. Part I: new apparatus for surface force
measurements. J Mater Res, 5, 22232231.
Kan, Y., Danner, E. W., Israelachvili, J. N. & Waite, H.
(2014). Boronate Complex Formation with Dopa
Containing Mussel Adhesive Protein Retards
pH-Induced Oxidation and Enables Adhesion to
Mica. PLoSONE, 9(10), doi:
10.1371/journal.pone.0108869.
Kaushik, N. K., Kaushik, N., Pardeshi, S., Sharma, J. G.,
Lee, S. H. & Choi, E. H. (2015). Biomedical and
Clinical Importance of Mussel-Inspired
Polymers and Materials. Mar Drugs, 13, 6792-
6817.
Kollbe Ahn, B. (2017). Perspectives on Mussel-Inspired
Wet Adhesion. J Am Chem Soc, 139,
10166−10171.
Lee, H. S., Scherer, N. F. & Messersmith, P. B. (2006).
Single-molecule mechanics of mussel adhesion.
Proc Natl Acad Sci U S A, 103, 1299913003.
Lee, B. P., Messersmith, P. B., Israelachvili, J. N.
& Waite, J. H. (2011). Mussel-Inspired
Adhesives and Coatings. Annu Rev Mater Res,
41, 99 132.
Lin, Q., Gourdon, D., Sun, C., Holten-Andersen, N.,
Anderson, T. H., Waite, J. H. & Israelachvili, J.
N. (2007). Adhesion mechanisms of the mussel
foot proteins mfp-1 and mfp-3. Proc Natl Acad
Sci U S A, 104(10), 3782-3786.
Lu, Q., Danner, E., Waite, J. H., Israelachvili, J. N.,
Zeng, H. & Hwang, D. S. (2013). Adhesion of
mussel foot proteins to different substrate
surfaces. J Roy Soc Interf, 10(79), 111.
Menyo, M. S., Hawker, C. J. & Waite. J. H. (2013).
Versatile tuning of supramolecular hydrogels
through metal complexation of oxidation-
resistant catechol-inspired ligands. Soft Matter,
9(43), 10314-10323.
Ornes, S. (2013). Mussels' sticky feet lead to
applications. Proc Natl Acad Sci U S A, 110(42),
16697-9.
Petrone, P., Kumar, A., Sutanto, C. N., Patil, N. J.,
Kannan, S., Palaniappan, A., Amini, S.,
Zappone, B., Verma, C. & Miserez, A. (2015).
Mussel adhesion is dictated by time-regulated
http://novasinergia.unach.edu.ec 99
secretion and molecular conformation of mussel
adhesive proteins. Nat comm, 6, 8737.
Priemel, T., Degtyar, E., Dean, M. N. & Harrington, M.
J. (2017). Rapid self-assembly of complex
biomolecular architectures during mussel byssus
biofabrication. Nat Comm, 8, 14539.
Silverman, H. G. & Roberto, F. F. (2007).
Understanding marine mussel adhesion. Mar
Biotech, 9(6), 661681.
Stewart, R. J., Ransom, T. C. & Hlady, V. (2011).
Natural Underwater Adhesives. J Polym Sci B
Polym Phys, 49(11), 757-771.
Waite, J. H., Andersen, N. H., Jewhurst, S. & Sun, C. J.
(2005). Mussel Adhesion: Finding the Tricks
Worth Mimicking. J Adhes, 81, 297317.
Wei, W., Yu, J., Broomell, C., Israelachvili, J. N. &
Waite, J. H. (2013). Hydrophobic enhancement
of Dopa-mediate adhesion in a mussel foot
protein. J Am Chem Soc, 135, 377383.
Wilker, J. J. (2010). Marine bioinorganic materials:
mussels pumping iron. Curr Opin Chem Biol, 14,
276283.
Wilker, J.J. (2015). BIOLOGICAL ADHESIVES.
Positive charges and underwater adhesion.
Science, 349(6248), 582-3.
Yu, J., Kan, Y., Rapp, M., Danner, E., Das, S., Miller,
D. R., Chen, Y., Waite, J. H & Israelachvili, J.
N. (2013). Adaptive hydrophobic and
hydrophilic interactions of mussel foot proteins
with organic thin films. Proc Natl Acad Sci U S
A, 110, 1568015685.
Yu, J., Wei, W., Danner, E., Ashley, R. K., Israelachvili,
J. N. & Waite, J. H. (2011a). Mussel protein
adhesion depends on thiol-mediated redox
modulation. Nat Chem Biol, 7, 588590.
Yu, J., Wei, W., Danner, E., Israelachcili, J. N. & Waite,
H. J. (2011b). Effects of Interfacial Redox in
Mussel Adhesive Protein Films on Mica. Adv
Mater, 23(20), 23622366.
doi:10.1002/adma.201003580.