You are on page 1of 9

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

http://pubs.acs.org/journal/acsodf Article

Role of Hydrogen-Bonding and OH−π Interactions in the Adhesion


of Epoxy Resin on Hydrophilic Surfaces
Shin Nakamura, Yuta Tsuji, and Kazunari Yoshizawa*
Cite This: ACS Omega 2020, 5, 26211−26219 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Epoxy resin adhesives are widely used for joining metal alloys in
various industrial fields. To elucidate the adhesion mechanism microscopically, we
investigated the interfacial interactions of epoxy resin with hydroxylated silica (0 0 1)
and γ-alumina (0 0 1) surfaces using periodic density functional theory calculations as
well as density of states (DOS) and crystal orbital Hamilton population (COHP)
analyses. To better understand the interfacial interactions, we employed and analyzed
water and benzene molecules as hydrophilic and hydrophobic adsorbates,
respectively. Structural features and calculated adhesion energies reveal that these
small adsorbates have a higher affinity for the γ-alumina surface than that for the silica
surface, while a fragmentary model for the epoxy resin exhibits a strong interaction with the silica surface. This discrepancy suggests
that the structural features of the hydroxylated silica surface dictate its affinity to a specific species. Partial DOS and COHP curves
provide evidence for the presence of OH−π interactions between the OH groups on the surfaces and the benzene rings of the epoxy
resin fragments. The orbital interaction energies of the H-bonding and OH−π interactions evaluated from the integrated COHP
indicate that the OH−π interaction is a nonnegligible origin of the adhesion interaction, even when polymers with hydrophobic
benzene rings are adsorbed on hydroxylated surfaces.

■ INTRODUCTION
Adhesion between organic and inorganic materials is an
essential process in various industrial fields.1−4 Lightweight,
durable, and strong aircraft structural members are typically
made by bonding resin sheets to aluminum alloys. 5
Applications of adhesives include coatings preventing the
corrosion of various metal materials. The body of an
automobile is also a typical product, made of steel and
aluminum alloys coated with plastic paint. One notable
application of resin adhesion is fiber-reinforced plastics,
composites of glass and carbon fibers, and synthetic polymers
like epoxy resins, for use in the automotive, aerospace, and civil
engineering and construction sectors.6−12 For the development
of these adhesive and coating technologies, it is necessary to
understand the adhesion mechanism at the interface between
an adhesive and an adherend. Figure 1. Chemical structure of (a) bisphenol A-type epoxy resin and
Epoxy resin, synthesized from diglycidylether of bisphenol A (b) simplified model for the epoxy resin; the red, brown, and white
spheres are O, C, and H atoms, respectively.
(DGEBA), is one of the most frequently employed structural
adhesives (see Figure 1 for its structure), containing ether
(−O−) and hydroxyl (−OH) groups, which can act as an of graphite, while only dispersion interactions are substantially
acceptor and a donor−acceptor of H-bond, respectively. active on the hydrophobic surfaces of the basal face and the H-
Previous theoretical and experimental works have shown that
H-bonding interactions originating from these groups provide
a clue to the mechanism of adhesion under atmospheric Received: August 7, 2020
conditions, in which water molecules are adsorbed on Accepted: September 22, 2020
adherend surfaces.13−24 Published: October 1, 2020
Energy decomposition calculations have demonstrated that
both electrostatic and dispersion interactions are significant in
the OH-functionalized surfaces of silica and the armchair edge

© 2020 American Chemical Society https://dx.doi.org/10.1021/acsomega.0c03798


26211 ACS Omega 2020, 5, 26211−26219
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 2. Optimized geometries of (A) water molecule, (B) benzene molecule, and (C) epoxy fragment on (top) the silica and (bottom) alumina
(0 0 1) surfaces (red, oxygen; cyan, hydrogen; gray, carbon; yellow, silicon, and purple, aluminum atoms). Selected distances are shown in Å.

terminated armchair edge of graphite.16 The epoxy resin also We employed water and benzene molecules as a hydrophilic
has hydrophobic aromatic rings, which provide the resin with molecule and a hydrophobic ring, respectively; functional
good thermal property and corrosion resistance. A recent groups analogous to them are contained in the epoxy resin. A
computational study has proposed that the benzene rings in silica adherend with a low concentration of hydroxyl groups
the epoxy resin interact with hydrophobic surfaces like the and a γ-alumina surface with a high concentration of hydroxyl
basal planes of hexagonal boron nitride and graphite, via the groups were used. Their surface structures are well
π−π stacking.25 However, due to the lack of analysis of these established.29−33 Each of these molecules and the epoxy
behaviors and their interactions on hydrophilic surfaces, the resin fragment were placed on the surface separately, and
functional role of benzene rings in atmospheric adhesion periodic density functional theory (DFT) calculation, the
remains to be elucidated. density of states (DOS) analysis, and crystal orbital Hamilton
Hydrophilic surfaces can essentially be characterized by the population (COHP) analysis34−37 were performed. Based on
concentration of hydroxyl groups on the surface. McCafferty the optimized geometries and calculated adhesion energies,
and Wightman used X-ray photoelectron spectroscopy to one can observe that the small moleculeswater and
experimentally determine the densities of hydroxyl groups on benzeneare adsorbed in a recess surrounded by hydroxyl
several oxide films.26 Their results show that aluminum and groups on hydrophilic surfaces, indicating an energetic
chromium covered with oxides exhibit relatively high hydroxyl preference to the adhesion with the γ-alumina surface, while
densities (>13 nm−2), while tantalum and silicon oxide films the epoxy resin fragment would prefer the silica surface. The
exhibit low densities (∼8 nm−2). The hydroxyl density should DOS and COHP curves reveal that despite the hydrophobic
affect the affinity of the adherend to the epoxy adhesive, but nature of the benzene rings of the epoxy resin fragment,
the effect of different concentrations on adhesion has not been OH−π interactions between the epoxy resin fragment and the
adequately estimated. hydroxyl protons on the surfaces stabilize the π-electrons on
We have reported on the estimated adhesion strength of the aromatic ring.
epoxy resins adsorbed on hydrophilic silica and γ-alumina
surfaces.14,27,28 The importance of the H-bonding interaction
between the resin and the surfaces was revealed, but how the
■ RESULTS AND DISCUSSION
Adhesion Structures and Interactions. The stable
hydrophobic benzene ring is involved in the adhesion process structures of a water molecule, benzene molecule, and the
under atmospheric conditions remains to be clarified. Epoxy epoxy fragment on the silica (0 0 1) and γ-alumina (0 0 1)
resins contain a large number of benzene rings, seen from a surfaces were determined from geometry optimization at the
microscopic perspective, leading to hints for designing level of periodic DFT with plane-wave basis, as shown in
innovative adhesive polymers. Also, there have been a few Figure 2. The details of the calculation setup and the
systematic, comparative studies focusing on differences in construction of the surface slab models are given in the
molecular sizes and kinds of slab surfaces. Since the affinity of Computational Procedures section. The structure of the epoxy
each surface for various molecules is important when selecting fragment remains almost unchanged upon adsorption onto the
an adherend, these aspects not only provide a complete picture surfaces (see Figure S1 in the Supporting Information). The
of the adhesion mechanism but also facilitate the development water molecule in both slab models forms four H-bonds with
of the adherend in the industrial sector. Therefore, for starters hydroxyl groups and coordinatively unsaturated oxygen atoms
in this study, a comparative study of adhesion interactions in the surface, lying close to a surface recess. Strong H-bonds
brought about by small simple molecules on two surfaces with with a very short O−H distance of 1.66 Å are spotted. The
different hydrophilic properties was carried out. Then, we other H-bond distances are within the typical O−H distance
moved on to the scrutiny of the adhesion interaction of a range (1.8−2.0 Å). The average H-bond distances are 1.84 and
fragmentary model for the bisphenol A-type epoxy resin with 1.83 Å for the silica and alumina surfaces, respectively; no
the surfaces. significant difference can be found.
26212 https://dx.doi.org/10.1021/acsomega.0c03798
ACS Omega 2020, 5, 26211−26219
ACS Omega http://pubs.acs.org/journal/acsodf Article

A previous study using a cluster model reported how a water calculated using structures with one to three H-bonds between
molecule is bound via two H-bonds by an isolated silanol water and silanol, markedly different from the present study.
group, a model for the silica surface.38 Mian et al. reported a Therefore, we propose the idea that water molecules are
stable structure of a water molecule adsorbed on the silica (0 0 strongly stabilized when they penetrate the silanol layer of the
1) and (1 1 1) surfaces.20 These structures suggest the silica surface.
presence of strong H-bonds with O−H distances shorter than As shown in Figure 2, benzene weakly interacts with the
1.7 Å. Our data reasonably suggest the idea that a water surface hydroxyl groups, taking on a conformation almost
molecule can penetrate the silanol layer of the silica surface and perpendicular to the surface plane with angles of 75.0 and
approach the coordinatively unsaturated oxygen atoms that 88.7° on the silica and alumina surfaces, respectively. To
cross-link subsurface silicon atoms. visualize the effects of the adsorption on the electron densities
After the geometry optimization, we calculated the adhesion of adsorbed molecules, we calculated the charge density
energy, ΔEad, which represents the net energy change upon difference. For each model, the charge density difference, Δρ,
adhesion, expressed as upon adhesion was calculated using
ΔEad = (Eadhesive + Eadherend) − E(adhesive + adherend) Δρ = ρ(adhesive + adherend) − (ρadhesive + ρadherend )
where E(adhesive + adherend) represents the total energy of the
adhesive−adherend complex and Eadhesive and Eadherend, where ρ(adhesive+adherend), ρadhesive, and ρadherend are the charge
respectively, represent the energies of the adhesive and densities of the adhesive−adherend complex, adhesive, and
adherend calculated separately. Eadhesive and Eadherend are adherend, respectively. The charge densities of the adhesive
obtained by performing the geometry optimizations separately. and adherend were calculated separately, but their geometries
Positive values of ΔEad indicate exothermic adhesion, with were kept the same as those in the adhesive−adherend
higher values indicating a stronger interaction between the complex. The isosurfaces of charge density difference show an
adhesive and the adherend. The dispersion interaction upon increase in electron density in the region between the
adhesion was also estimated. The dispersion energy ΔEdisp interacting hydroxyl H atom and the carbon atom of the
working between the adhesive and the adherend is given by the benzene molecule (see Figure 3A), while a slight decrease in
following equation electron density on the lower H atoms of the benzene
disp disp disp molecule. These observations suggest that the main interaction
ΔEdisp = (Eadhesive + Eadherend ) − E(adhesive + adherend) for benzene adsorption is the OH−π interaction rather than
disp
where Eadhesive disp
and Eadherend represent, respectively, the the CH−O interaction. The adhesion energies of the benzene
dispersion energies of the adhesive and adherend calculated molecule are estimated to be 33.0 and 53.6 kJ/mol for the
separately and Edisp silica and alumina surfaces, respectively, as shown in Table 1. It
(adhesive+adherend) is the total dispersion energy of
the adhesive−adherend complex. These values were calculated is inferred that the OH−π interaction on the alumina surface
using the optimized geometry for each system. The calculated has a greater effect on the adhesion energy. The presence of
adhesion energies for the water molecule on the silica and γ- the OH−π interaction of a benzene molecule is consistent with
alumina (0 0 1) surfaces are 72.6 and 82.6 kJ/mol, respectively, the previous theoretical works by Rimola et al.41,42 The C−H
as summarized in Table 1. The results suggest that the alumina distance along the OH−π interaction in their model is in the
range of 2.2−2.6 Å, quite close to our results.
The benzene adhesion also denotes a high ratio of the
Table 1. Adhesion Energy, ΔEad (kJ/mol), Its Dispersion
dispersion energy, indicating that dispersion contribution to
Component, ΔEdisp (kJ/mol), and the Ratio between Them,
adhesion is crucial even on the hydrophilic surfaces. Our
ΔEdisp/ΔEad, Calculated for the Three Molecules (Water,
previous study has investigated the adhesion of an epoxy
Benzene, and Epoxy) on the Silica and Alumina (0 0 1)
fragment with the hydrophobic surface of hexagonal boron
Surfaces
nitride: the interaction exerted at the interface is predom-
silica (0 0 1) γ-alumina (0 0 1) inantly derived from dispersion forces.25 The large contribu-
ratio ratio tion of dispersion energy to the adhesion of benzene
(ΔEdisp/ (ΔEdisp/ investigated in the present study supports the idea that
model ΔEad ΔEdisp ΔEad) ΔEad ΔEdisp ΔEad) benzene rings are adsorbed on surfaces primarily through
water 72.6 22.0 0.303 82.6 35.1 0.425 dispersion interactions. Compared to benzene adsorption, the
benzene 33.0 40.3 1.223 53.6 67.3 1.256 ratios of ΔEdisp to ΔEad for the adsorption of water indicate a
epoxy 145.9 108.8 0.746 117.0 89.6 0.766 relatively low contribution of dispersion energy to adhesion
model
energy. The dispersion contribution seems less important for
the case of water adsorption.
surface has a slightly higher affinity for water than the silica The hydroxyl group of the epoxy resin model acts as an H-
surface. The ΔEad values of the water molecule adsorbed on bond donor−acceptor with the hydroxyl groups of both slab
the surfaces are comparable with those calculated using cluster model surfaces. These H-bond distances are within the typical
models.33,38−40 Mian et al. reported the water-binding energies distance range. The hydroxyl groups in the epoxy model, in
of 12.17 kcal/mol (50.9 kJ/mol) and 11.08 kcal/mol (46.4 kJ/ contrast to the water molecule, cannot enter the surface recess
mol) for the silica (0 0 1) and (1 1 1) surfaces, respectively.20 due to the steric hindrance caused by the two benzene rings.
Tielens et al. investigated the water-binding energy for various Compared to the hydroxyl groups, the interaction of the
silanols, finding that it ranges from 11.52 to 11.95 kcal/mol benzene rings of the epoxy fragment shows a different behavior
(48.2−50.0 kJ/mol).33 These binding energies are slightly between the silica and alumina surfaces. As shown in Figure 2,
lower than the water adhesion energies obtained in the present the benzene ring on the silica surface attracts hydroxyl protons
study. However, the energies reported in the literature were on the surface, but the benzene ring on the alumina surface
26213 https://dx.doi.org/10.1021/acsomega.0c03798
ACS Omega 2020, 5, 26211−26219
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 3. Isosurfaces of charge density difference calculated for (A) benzene adsorption and (B) epoxy adhesion on (left) the silica and (right)
alumina surfaces (yellow, charge accumulation; cyan, charge depletion). The isovalue is set to 0.00275 e/Å3.

does not interact with the hydroxyl group. The results of this view of this insight, the silica surface may be more suitable than
calculation are consistent with the charge density differences of the alumina surface for the formation of OH−π interactions.
the epoxy models on the surfaces, as shown in Figure 3. The This is due to the long distance (∼4.9 Å) between the convex,
characteristic features of the charge density differences around protruding oxygen atoms on the silica surface, providing
the benzene rings of the epoxy model on the silica surface can sufficient space to interact with the adhesive, as shown in
be observed: they are similar to those for the adsorption of Figure 4. Since the distance between the convex oxygen atoms
benzene. This observation strongly suggests that the OH−π
interactions contribute not only to the small molecule
adsorption but also to the adsorption of large systems (e.g.,
polymers and proteins). It should be noted that we obtained
the several optimized geometries of the epoxy fragment on the
silica and alumina surfaces by changing the initial structures for
the geometry optimizations, and then, we chose the most
stable conformation among them (see Table S1 in the
Supporting Information). For the epoxy−alumina complex,
we were not able to obtain the structure with both H-bonding
and OH−π interactions using geometry optimization. During
this process, we found that the OH−π interaction between the
epoxy fragment and the alumina surface is relatively weak in
comparison with that on the silica surface. The epoxy fragment
on the alumina surface does not show the charge density
difference on the benzene ring (see Figure S2 in the
Supporting Information).
The interaction between the epoxy model and the alumina
surface does not seem to bring about significant changes in the
Figure 4. Surface structures of silica (A) and γ-alumina (B) are shown
charge density difference. The benzene rings of the epoxy
along with distances between the surface convex, protruding oxygens
fragment on the alumina surface appear to be unable to (red, oxygen; yellow, silicon; and purple, aluminum atoms). The
interact effectively with the protons of the hydroxyl groups on selected distances are shown in Å. Hydrogen atoms are omitted for
the surface. This is partly because there is a lack of recessed clarity.
areas of the surface, which are necessary for the benzene rings
to take a conformation essential for the OH−π interaction on
the surface. The absence of dangling protons on the alumina
surface is also a major factor, in which the dangling proton on the alumina surface is about the same or shorter than the
means a proton that can move flexibly in response to changes typical H-bond distance, many of the hydroxyl protons prefer
in the electrostatic potential and the structure of hydrogen to form H-bonds with the nearby convex oxygen atoms rather
bonds on the surface. than with the adhesive. By contrast, many hydroxyl protons on
For the interaction of the benzene ring with a hydroxyl the silica surface prefer to interact with the adhesive. This
proton on the surface, the plane of the benzene ring needs to tendency seems to be critically important for the adhesion of
be slightly perpendicular to the O−H direction of the hydroxyl large molecules with a large steric hindrance (e.g., polymers
group on the surface, as in the case of benzene adsorption. In and proteins).
26214 https://dx.doi.org/10.1021/acsomega.0c03798
ACS Omega 2020, 5, 26211−26219
ACS Omega http://pubs.acs.org/journal/acsodf Article

The arrangement of the convex oxygen atoms depends


predominantly on the concentration of hydroxyl groups on the
surface. Differences in the coordination structure of silicon and
aluminum atoms in the slabs seem to be the main cause of the
difference in the concentration of surface hydroxyl groups.
Aluminum atoms in alumina basically take an octahedral
geometry consisting of six ligand oxygen atoms, whereas silicon
atoms in silica take a tetrahedral geometry consisting of four
ligand oxygen atoms. The combination of the octahedral
geometry and the (0 0 1) facet induces a flat surface and many
surface aluminum atoms are left coordinatively unsaturated
immediately after cleavage (before passivation). This view is
consistent with the fact that the distances between the convex
oxygen atoms on the alumina surface are almost the same as
the nearest Al−Al distance, which ranges from 2.8 to 2.9 Å as
measured for the optimized bulk structure of alumina. For
example, chromium in chromium(III) oxide (i.e., Cr2O3),
which has a six-coordinated structure similar to alumina, also
has a high concentration of hydroxyl groups on its surface.26
Owing to the tetrahedral geometry of silicon, concave and
convex silicon sites are generated on the silica surface: the
former are hydroxylated, while the latter are not. This leads to
a reduced number of hydration sites. Since a relatively low
density of hydroxyl groups is reported for the silica (0 0 1) and
(1 1 1) surfaces, which are derived respectively from α- and β-
cristobalites,20 the choice of lattice planes to be cleaved may Figure 5. Partial density of states (PDOS) for the 2p orbitals of the C
not significantly affect the hydration concentration of the atoms in (A) benzene and the aromatic rings of the epoxy fragment
(B) adsorbed on the surface (red lines) and isolated from the surface
surface at least as far as silica goes. This suggests that the (blue lines). The dashed line indicates the Fermi level. In the
coordination structure of the cationic element in the bulk may calculation of the isolated epoxy molecule, the slab surface remains in
have a significant effect on the concentration and structure of the periodic boundary box but they are separated so far that the
surface hydroxyl groups. surface−molecule interaction can be negligible. Therefore, the Fermi
Adhesion energies calculated for the epoxy fragment on the energy of the system including the isolated molecule matches that of
silica and alumina surfaces are 145.9 and 117.0 kJ/mol, the surface.
respectively, as shown in Table 1. This result seems consistent
with the absence of the OH−π interaction on the alumina adsorption on the silica surface does not appear to affect the
surface. The dispersion contribution to the adhesion energy is peak positions. The larger adhesion energy of benzene on the
estimated to be 108.8 and 89.6 kJ/mol on the silica and alumina surface than on the silica surface (see Table 1) looks
alumina surfaces, respectively. The ratio of ΔEdisp to ΔEad in consistent with the occurrence of the downward peak shift in
epoxy adhesion lies between those for water and benzene the PDOS for the alumina system and its absence in the silica
adsorption. This suggests that the epoxy fragment adheres to system.
the hydrophilic surfaces through the H-bonding interaction The PDOS spectrum for the carbon 2p orbitals of the
due to the hydroxyl group and the dispersion interaction due benzene rings of the epoxy fragment on the alumina surface
to the benzene rings. looks almost identical to that calculated for the isolated epoxy
We have previously reported that the adhesion energies of fragment. On the silica surface, the PDOS peaks for the carbon
an epoxy resin fragment on the silica (0 0 1) surface and the γ- 2p orbitals of one of the benzene rings are downshifted by
alumina (0 0 1) surface are 145.727 and 64.7−65.8 kJ/mol,13 about 0.08 eV after the adsorption of the epoxy fragment on
respectively. The value of ΔEad calculated for the silica surface the surface. This suggests that the electrons on the benzene
in this study is close to that of the previous study, while the ring are somewhat stabilized by the OH−π interaction, which
value of ΔEad for the γ-alumina surface in this study is may be one of the important interactions for the adhesion of
significantly larger than those calculated in the previous polymers including benzene rings.
studies. However, the previous studies on the alumina surface As shown in Figure 6, in contrast to the OH−π interaction,
did not apply dispersion correction to the periodic DFT the H-bonding interaction leads to the broadening of the
calculations. The mismatch in the adhesion energy of the PDOS peaks for the hydroxyl O atom of the epoxy fragment
epoxy fragment to the alumina surface may be due to the lack and those for the oxygen atom of the adsorbed water molecule.
of dispersion correction in the previous studies, as dispersion Also, the H-bonding interaction lowers the energy levels of the
interactions contribute significantly to adhesion. PDOS peaks. These features are indicative of electronic
Effects on Electronic Structures and Orbital Inter- interactions induced by H-bonds between the adherend and
actions. To estimate the effect of adhesion on the electronic the adhesive.
structure, we calculated the partial DOS (PDOS) for the 2p In addition to the PDOS calculations, we also performed
orbitals of the carbon atoms of the benzene molecule and the COHP analyses based on the wave function obtained from the
benzene rings in the epoxy model, as shown in Figure 5. Most periodic DFT calculations. The wave function is expanded
peaks of the PDOS are found downshifted by 0.34−0.39 eV using plane waves, so a projection scheme is required for
when benzene is adsorbed on the alumina surface, while its moving from the plane-wave to a local atomic-orbital basis set.
26215 https://dx.doi.org/10.1021/acsomega.0c03798
ACS Omega 2020, 5, 26211−26219
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 6. Partial density of states (PDOS) for the O atom of the


water molecule and that for the O atom of the hydroxyl group in the
epoxy fragment. They were calculated for the state of adsorption on
the silica and γ-alumina surfaces (red) and the state in which the
adsorbate is separated from the surface (blue). The dashed line
indicates the Fermi level.

The LOBSTER program allows us to do this. The COHP Figure 7. Negative of projected crystal orbital Hamilton population
(−pCOHP) calculated for the interactions found to be related to the
analysis is a method of the decomposition of the DOS into the
OH−π interactions between the aromatic rings of the epoxy fragment
bonding and antibonding interactions for a specific pair of and the hydroxyl protons on the silica surface (A). −pCOHP curves
atoms using the corresponding Hamiltonian matrix element for the H-bonding interactions between the hydroxyl group of the
and the crystal orbital expansion coefficients. Since the epoxy fragment and the hydroxyl groups on the silica and γ-alumina
projection is done, one may call the COHP generated from surfaces (B). The dashed line indicates the Fermi level. (C) Top: the
LOBSTER the projected COHP (pCOHP). Bonding inter- atomic pairs associated with the OH−π interactions at the silica−
actions are concomitant with negative Hamiltonian matrix epoxy interface are indicated by colored lines. Bottom: the interface
elements, so negative pCOHP values ensue. The values of structure between silica/alumina and epoxy are shown in an
pCOHP plots to be presented below have already been abstracted fashion, where the atomic pairs associated with the H-
multiplied by −1 such that positive −pCOHP values bonding interactions are indicated by colored lines. The interaction
lines are color-coded with the colors of the lines in the −pCOHP
correspond to bonding interactions.
plots.
Figure 7A shows the −pCOHP curves calculated for the
atomic pairs found to be related to the OH−π interaction at
Table 2. Distances (Å) and Integrated pCOHP (ICOHP)
the silica−epoxy interface. In these spectra, a manifold of
Values for the Atomic Pairs Shown in Figure 7C
positive peaks are observed below the Fermi level, a relatively
large positive peak at around 4 eV, and a couple of negative silica (0 0 1) γ-alumina (0 0 1)
peaks way above the Fermi level (>5 eV). Although the large distance (Å) ICOHP (eV) distance (Å) ICOHP (eV)
positive band around 4 eV contains no electrons and does not
Hslab−Oa 1.85 −0.78 1.97 −0.64
contribute to bonding, the peaks below the Fermi level indicate
Ha−Oslab 1.86 −0.67 1.79 −0.95
that the OH−π interactions between the aromatic ring and the
Hslab−Oc 1.93 −0.53
hydroxyl groups of the silica surface can be characterized by
Hslab−C1f 2.34 −0.16
the presence of bonding-type orbital interactions between
Hslab−C2b 2.08 −0.32
them over a relatively wide energy range.
Hslab−C2d 2.45 −0.14
Figure 7B shows the −pCOHP curves calculated for the
atomic pairs found to be related to the H-bonding interactions
at the silica−epoxy and alumina−epoxy interfaces. The the atomic pairs. Note that the unit of ICOHP is eV. This
calculated O−H pairs with H-bonding interactions are found value can be read as bonding energy of a certain kind.
to be characterized by bonding-type orbital interactions, The ICOHP value for the H-bonding interaction falls in the
although one can see a few negative peaks (antibonding range from −0.53 to −0.95 eV, while that for the OH−π
interactions) at around −20 and −3 eV for the alumina interaction falls in the range from −0.14 to −0.32 eV. These
interface and in the range of −20 to −18 eV and at around −6 computed values imply that the H-bonding interactions are
eV for the silica interface. We visualized the OH−π orbital stronger than the OH−π interactions. Thus, the H-bonding
interactions and the H-bonding orbital interactions by interaction is likely to have a dominant effect on adhesion as
calculating the partial charge density maps, as shown in Figure expected and as reported in the preceding studies. On top of
S2 in the Supporting Information. that, since epoxy resins have more benzene rings than hydroxyl
To compare the orbital interaction energies between the H- groups, the interaction energy due to the OH−π interaction
bonding and OH−π interactions, the pCOHP values were cannot be regarded as low as a whole. This suggests that the
integrated up to the Fermi level for each atomic pair. They are OH−π interaction should not be ignored even though
called ICOHP, as summarized in Table 2. A negative ICOHP polymers containing hydrophobic benzene rings are used for
value suggests that there is a net bonding interaction between adhesion to hydrophilic surfaces.
26216 https://dx.doi.org/10.1021/acsomega.0c03798
ACS Omega 2020, 5, 26211−26219
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 8. Periodic slab models of (A) 2 × 3 supercell of the hydroxylated silica (0 0 1) surface and (B) 3 × 1 supercell of the hydroxylated γ-
alumina (0 0 1) surface (red, oxygen; cyan, hydrogen; yellow, silicon; and purple, aluminum atoms).

■ CONCLUSIONS
To elucidate the adhesion mechanism of the DGEBA epoxy
with the benzene ring stabilizes the π-electrons on the ring.
The COHP values were integrated up to the Fermi level to
obtain the ICOHP, which is a good measure of the bonding
resin to the hydrophilic surfaces of alumina and silica from a
energy due to orbital interactions. Calculated ICOHP values
microscopic point of view, periodic DFT calculations and
indicate that the H-bonding interaction is stronger than the
orbital analysis have been used to explore the interactions at
OH−π interaction. However, considering that the number of
the interfaces. Water and benzene molecules were selected as a
the benzene rings in epoxy resins is greater than that of the
simple model for the hydroxyl group and the benzene ring in
hydroxyl groups, it can be concluded that OH−π interactions
the epoxy resin. A fragmentary molecular structure of the
cannot be ignored as one of the origins of the adhesion force.


epoxy resin was also used to study the adhesion mechanism,
including the effects of resin-specific steric hindrances.
The alumina and silica slab surfaces were modeled on COMPUTATIONAL PROCEDURES
hydroxylated γ-alumina (0 0 1) and silica (0 0 1) surfaces. The We performed all of the ab initio periodic calculations using
water molecule was found adsorbed into a recess surrounded VASP program package version 5.4.4.43−46 We carried out
by hydroxyl groups on both surfaces to form four H-bonds, geometry optimizations using DFT with the Perdew−Burke−
while the benzene molecule stands almost perpendicular to the Ernzerhof functional47 with a dispersion correction of the
surface in the recess, and their carbon atoms attract hydroxyl DFT-D2 method by Grimme.48 The plane-wave basis set
protons through OH−π interactions. cutoff and self-consistent field tolerance were set to 500 and
In contrast to the small adsorbates, the optimized geometry 10−5 eV, respectively. The threshold of atomic forces for the
of the epoxy resin fragment differs from surface to surface. The convergence of the geometry optimization was set to 0.03 eV/
hydroxyl group in the epoxy fragment forms H-bonds with Å. We employed a mesh of k-points of 2 × 2 × 1 for sampling
both silica and alumina surfaces, whereas the resin’s benzene the Brillouin zone. We calculated DOS and COHP using
rings interact only with the silica surface by attracting the LOBSTER software version 3.1.0,34−36 utilizing the wave
protons of surface hydroxyl groups thereon. Calculated functions generated from the VASP periodic DFT calculations.
adhesion energies indicate that the alumina surface has a Structures were visualized using the VESTA software pack-
higher affinity for water and benzene molecules than the silica age.49
surface. By contrast, the epoxy fragment shows a stronger A periodic silica (0 0 1) surface was created by cleaving the
interaction with the silica surface than with the alumina bulk structure of α-cristobalite obtained from the Materials
surface. Studio 6.1 database.50 The unit cell of silica contains four
Although this contrast seems to be a contradiction, this may silicon and eight oxygen atoms with lattice constants of a =
be because of the steric hindrance of the epoxy fragment. The 4.931 Å, b = 4.931 Å, c = 6.802 Å, and β = 90.00°, as
interactions of the fragment with protons wobbling on the determined from DFT calculations. By repeating the unit cell
alumina surface and limited recessed space available for twice in the direction of the c-axis, a supercell containing eight
interaction may also be responsible for the difference. This atomic layers was constructed, and an empty space with a
indicates that by controlling the size of the recess on the thickness of 20 Å was added as a vacuum layer on the surface.
surface, which is derived from surface hydroxyl groups, one can Two water molecules were dissociatively adsorbed to passivate
ameliorate the adhesion affinity of the surface to the polymer. coordinatively unsaturated silicon atoms on the surface. The
Orbital interactions responsible for the H-bonding and hydroxylated surface thus generated was optimized using the
OH−π interactions at the adhesion interfaces were also periodic DFT method with the atomic positions in the middle
investigated using PDOS and COHP analyses. The results layers (i.e., the fourth and fifth layers) kept fixed. The slab
show that the benzene molecule on the alumina surface and model obtained was virtually identical to the one investigated
one of the benzene rings of the epoxy fragment on the silica in our previous study,27 as shown in Figure 8A.
surface cause a downshift of the PDOS peaks corresponding to A periodic γ-alumina (0 0 1) surface slab model was
the 2p orbitals of the carbon atoms therein when the adhesion constructed in the same way as used for the construction of the
interaction sets in. This suggests that the OH−π interaction silica (0 0 1) surface model. The bulk structure of γ-alumina
26217 https://dx.doi.org/10.1021/acsomega.0c03798
ACS Omega 2020, 5, 26211−26219
ACS Omega http://pubs.acs.org/journal/acsodf Article

was obtained from the literature, where a model based on DFT Notes
calculations by Toulhoat et al. is reported.51,52 The unit cell of The authors declare no competing financial interest.


γ-alumina contains 16 aluminum atoms and 24 oxygen atoms
with lattice constants of a = 5.528 Å, b = 8.333 Å, c = 8.022 Å, ACKNOWLEDGMENTS
and β = 90.68°, as determined from DFT calculations. A 20 Å
thick empty space was placed on the surface as a vacuum layer This work was supported by KAKENHI grants (numbers
and eight water molecules were dissociatively adsorbed on the JP17K14440 and JP17H03117) from the Japan Society for the
surface to passivate coordinatively unsaturated aluminum Promotion of Science (JSPS) and the Ministry of Education,
atoms on the surface. The surface structure created was Culture, Sports, Science and Technology of Japan (MEXT)
optimized with the atomic positions of the middle layers (i.e., through the MEXT projects Integrated Research Consortium
the second and third layers of the four layers) kept fixed. As on Chemical Sciences (IRCCS), Cooperative Research
can be seen from Figure 8B, the optimized surface structure for Program of Network Joint Research Center for Materials and
γ-alumina is in good agreement with what was calculated in Devices, and Elements Strategy Initiative to Form Core
previous studies.13,28 Research Center and by JST-CREST JPMJCR15P5 and JST-
Owing to the different molecular sizes of the three adhesive Mirai JPMJMI18A2. The computations in this work were
molecules, a 2 × 2 supercell of the silica slab was used for the primarily performed using the computer facilities at the
calculations of water and benzene adsorption and a 2 × 3 Research Institute for Information Technology, Kyushu
supercell for epoxy adhesion. A 2 × 1 supercell of the γ- University. Y.T. is grateful for a JSPS Grant-in-Aid for
alumina slab was used for the calculations of benzene Scientific Research on Innovative Areas (Discrete Geometric
adsorption and a 3 × 1 supercell for epoxy adhesion. The Analysis for Materials Design, grant number JP20H04643, and
longer dimensions of the a- and b-directions of the periodic Mixed Anion, grant number JP19H04700).
boundary box are ∼15 and ∼17 Å for the silica and γ-alumina
surfaces, respectively, both of which are longer than the end-to-
end distance of the epoxy fragment (∼13 Å). The unit cell of
■ REFERENCES
(1) Kinloch, A. J. The science of adhesion. J. Mater. Sci. 1980, 15,
the γ-alumina slab was used for the calculations of water 2141−2166.
adsorption. In the optimizations of the slab structures with the (2) Edward, P. Epoxy Adhesive Formulations; McGraw-Hill
addition of each adhesive model, the atomic positions of the Professional, 2005.
lower half of the silica and alumina slabs were fixed and those (3) Fourche, G. An overview of the basic aspects of polymer
of the upper half were fully relaxed. Each of the adsorbate adhesion. Part II: Application to surface treatments. Polym. Eng. Sci.
molecules was placed randomly on the surface and geometry 1995, 35, 968−975.
(4) Fourche, G. An overview of the basic aspects of polymer
optimization was carried out several times to determine a adhesion. Part I: Fundamentals. Polym. Eng. Sci. 1995, 35, 957−967.
stable interfacial structure.


(5) Brockmann, W.; Hennemann, O. D.; Kollek, H.; Matz, C.
Adhesion in bonded aluminium joints for aircraft construction. Int. J.
ASSOCIATED CONTENT Adhes. Adhes. 1986, 6, 115−143.
* Supporting Information
sı (6) Simon, R.; Prosen, S. P.; Duffy, J. Carbon fibre composites.
The Supporting Information is available free of charge at Nature 1967, 213, 1113−1114.
https://pubs.acs.org/doi/10.1021/acsomega.0c03798. (7) Ishida, H.; Koenig, J. L. The reinforcement mechanism of fiber-
glass reinforced plastics under wet conditions: a review. Polym. Eng.
Relative energies for the optimized geometries of the Sci. 1978, 18, 128−145.
epoxy fragment on the hydroxylated surfaces, structural (8) Fitzer, E.; Geigl, K. H.; Hüttner, W.; Weiss, R. Chemical
changes of the epoxy fragments upon adsorption, interactions between the carbon fibre surface and epoxy resins.
isosurfaces of charge density difference calculated for Carbon 1980, 18, 389−393.
epoxy adhesion on the alumina surface without H- (9) Park, S. J.; Kim, B. J. Roles of acidic functional groups of carbon
bonding interactions, and partial charge density maps for fiber surfaces in enhancing interfacial adhesion behavior. Mater. Sci.
Eng., A 2005, 408, 269−273.
the epoxy fragment adsorbed onto the hydroxylated
(10) Dai, Z.; Shi, F.; Zhang, B.; Li, M.; Zhang, Z. Effect of sizing on
surfaces (PDF) carbon fiber surface properties and fibers/epoxy interfacial adhesion.

■ AUTHOR INFORMATION
Corresponding Author
Appl. Surf. Sci. 2011, 257, 6980−6985.
(11) Deng, S. H.; Zhou, X. D.; Fan, C. J.; Lin, Q. F.; Zhou, X. G.
Release of interfacial thermal stress and accompanying improvement
of interfacial adhesion in carbon fiber reinforced epoxy resin
Kazunari Yoshizawa − Institute for Materials Chemistry and composites: induced by diblock copolymers. Composites, Part A
Engineering and IRCCS, Kyushu University, Fukuoka 819- 2012, 43, 990−996.
0395, Japan; orcid.org/0000-0002-6279-9722; (12) Sathishkumar, T. P.; Satheeshkumar, S.; Naveen, J. Glass fiber-
Email: kazunari@ms.ifoc.kyushu-u.ac.jp reinforced polymer composites - a review. J. Reinf. Plast. Compos.
2014, 33, 1258−1275.
Authors (13) Semoto, T.; Tsuji, Y.; Yoshizawa, K. Molecular understanding
Shin Nakamura − Institute for Materials Chemistry and of the adhesive force between a metal oxide surface and an epoxy
Engineering and IRCCS, Kyushu University, Fukuoka 819- resin. J. Phys. Chem. C 2011, 115, 11701−11708.
0395, Japan; orcid.org/0000-0001-6223-2323 (14) Semoto, T.; Tsuji, Y.; Yoshizawa, K. Molecular understanding
Yuta Tsuji − Institute for Materials Chemistry and Engineering of the adhesive force between a metal oxide surface and an epoxy
and IRCCS, Kyushu University, Fukuoka 819-0395, Japan; resin: effects of surface water. Bull. Chem. Soc. Jpn. 2012, 85, 672−
orcid.org/0000-0003-4224-4532 678.
(15) Semoto, T.; Tsuji, Y.; Tanaka, H.; Yoshizawa, K. Role of edge
Complete contact information is available at: oxygen atoms on the adhesive interaction between carbon fiber and
https://pubs.acs.org/10.1021/acsomega.0c03798 epoxy resin. J. Phys. Chem. C 2013, 117, 24830−24835.

26218 https://dx.doi.org/10.1021/acsomega.0c03798
ACS Omega 2020, 5, 26211−26219
ACS Omega http://pubs.acs.org/journal/acsodf Article

(16) Yoshizawa, K.; Semoto, T.; Hitaoka, S.; Higuchi, C.; Shiota, Y.; (36) Maintz, S.; Deringer, V. L.; Tchougreeff, A. L.; Dronskowski, R.
Tanaka, H. Synergy of electrostatic and van der Waals interactions in LOBSTER: A tool to extract chemical bonding from plane-wave
the adhesion of epoxy resin with carbon-fiber and glass surfaces. Bull. based DFT. J. Comput. Chem. 2016, 37, 1030−1035.
Chem. Soc. Jpn. 2017, 90, 500−505. (37) Dronskowski, R.; Bloechl, P. E. Crystal orbital Hamilton
(17) Henry, D. J.; Yiapanis, G.; Evans, E.; Yarovsky, I. Adhesion populations (COHP): energy-resolved visualization of chemical
between graphite and modified polyester surfaces: a theoretical study. bonding in solids based on density-functional calculations. J. Phys.
J. Phys. Chem. B 2005, 109, 17224−17231. Chem. B 1993, 97, 8617−8624.
(18) Köppen, S.; Bronkalla, O.; Langel, W. Adsorption config- (38) Pelmenschikov, A. G.; Morosi, G.; Gamba, A. Adsorption of
urations and energies of amino acids on anatase and rutile surfaces. J. water and methanol on silica hydroxyls: ab initio energy and
Phys. Chem. C 2008, 112, 13600−13606. frequency calculations. J. Phys. Chem. A 1997, 101, 1178−1187.
(19) Koppen, S.; Langel, W. Simulation of adhesion forces and (39) Civalleri, B.; Garrone, E.; Ugliengo, P. Cagelike clusters as
energies of peptides on titanium dioxide surfaces. Langmuir 2010, 26, models for the isolated hydroxyls of silica: ab initio B3-LYP
calculations of the interaction with ammonia. Langmuir 1999, 15,
15248−15256.
5829−5835.
(20) Mian, S. A.; Saha, L. C.; Jang, J.; Wang, L.; Gao, X.; Nagase, S.
(40) Civalleri, B.; Garrone, E.; Ugliengo, P. Cage-like clusters as
Density functional theory study of catechol adhesion on silica models for the hydroxyls of silica: ab initio calculation of 1H and 29Si
surfaces. J. Phys. Chem. C 2010, 114, 20793−20800. NMR chemical shifts. Chem. Phys. Lett. 1999, 299, 443−450.
(21) Mian, S. A.; Yang, L. M.; Saha, L. C.; Ahmed, E.; Ajmal, M.; (41) Rimola, A.; Civalleri, B.; Ugliengo, P. Physisorption of aromatic
Ganz, E. A fundamental understanding of catechol and water organic contaminants at the surface of hydrophobic/hydrophilic silica
adsorption on a hydrophilic silica surface: exploring the underwater geosorbents: a B3LYP-D modeling study. Phys. Chem. Chem. Phys.
adhesion mechanism of mussels on an atomic scale. Langmuir 2014, 2010, 12, 6357−6366.
30, 6906−6914. (42) Rimola, A.; Sodupe, M.; Ugliengo, P. Affinity Scale for the
(22) Bahlakeh, G.; Ghaffari, M.; Saeb, M. R.; Ramezanzadeh, B.; De Interaction of Amino Acids with Silica Surfaces. J. Phys. Chem. C
Proft, F.; Terryn, H. A close-up of the effect of iron oxide type on the 2009, 113, 5741−5750.
interfacial interaction between epoxy and carbon steel: combined (43) Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid
molecular dynamics simulations and quantum mechanics. J. Phys. metals. Phys. Rev. B 1993, 47, 558−561.
Chem. C 2016, 120, 11014−11026. (44) Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation
(23) Ogata, S.; Takahashi, Y. Moisture-induced reduction of of the liquid-metal-amorphous-semiconductor transition in germa-
adhesion strength between surface oxidized Al and epoxy resin: nium. Phys. Rev. B 1994, 49, 14251−14269.
dynamics simulation with electronic structure calculation. J. Phys. (45) Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy
Chem. C 2016, 120, 13630−13637. calculations for metals and semiconductors using a plane-wave basis
(24) Ogata, S.; Uranagase, M. Unveiling the chemical reactions set. Comput. Mater. Sci. 1996, 6, 15−50.
involved in moisture-induced weakening of adhesion between (46) Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab
aluminum and epoxy resin. J. Phys. Chem. C 2018, 122, 17748−17755. initio total-energy calculations using a plane-wave basis set. Phys. Rev.
(25) Tsuji, Y.; Kitamura, Y.; Someya, M.; Takano, T.; Yaginuma, M.; B 1996, 54, 11169−11186.
Nakanishi, K.; Yoshizawa, K. Adhesion of epoxy resin with hexagonal (47) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient
boron nitride and graphite. ACS Omega 2019, 4, 4491−4504. approximation made simple. Phys. Rev. Lett. 1996, 77, 3865−3868.
(26) McCafferty, E.; Wightman, J. P. Determination of the (48) Grimme, S. Semiempirical GGA-type density functional
constructed with a long-range dispersion correction. J. Comput.
concentration of surface hydroxyl groups on metal oxide films by a
Chem. 2006, 27, 1787−1799.
quantitative XPS method. Surf. Interface Anal. 1998, 26, 549−564.
(49) Momma, K.; Izumi, F. VESTA 3for three-dimensional
(27) Higuchi, C.; Tanaka, H.; Yoshizawa, K. Molecular under- visualization of crystal, volumetric and morphology data. J. Appl.
standing of the adhesive interactions between silica surface and epoxy Crystallogr. 2011, 44, 1272−1276.
resin: effects of interfacial water. J. Comput. Chem. 2019, 40, 164−171. (50) Materials Studio 6.1; Accelrys, Inc.: San Diego, CA, 2012.
(28) Yoshizawa, K.; Murata, H.; Tanaka, H. Density-functional tight- (51) Krokidis, X.; Raybaud, P.; Gobichon, A.-E.; Rebours, B.; Euzen,
binding study on the effects of interfacial water in the adhesion force P.; Toulhoat, H. Theoretical study of the dehydration process of
between epoxy resin and alumina surface. Langmuir 2018, 34, 14428− boehmite to γ-alumina. J. Phys. Chem. B 2001, 105, 5121−5130.
14438. (52) Digne, M.; Sautet, P.; Raybaud, P.; Euzen, P.; Toulhoat, H. Use
(29) Liu, Z.; Ma, L.; Junaid, A. S. M. NO and NO2 adsorption on of DFT to achieve a rational understanding of acid-basic properties of
Al2O3 and Ga modified Al2O3 surfaces: a density functional theory γ-alumina surfaces. J. Catal. 2004, 226, 54−68.
study. J. Phys. Chem. C 2010, 114, 4445−4450.
(30) Kwak, J. H.; Hu, J.; Mei, D.; Yi, C. W.; Kim, D. H.; Peden, C.
H.; Allard, L. F.; Szanyi, J. Coordinatively unsaturated Al3+ centers as
binding sites for active catalyst phases of platinum on γ-Al2O3. Science
2009, 325, 1670−1673.
(31) Peri, J. B. A model for the surface of γ-alumina1. J. Phys. Chem.
A 1965, 69, 220−230.
(32) Zhuravlev, L. T. Concentration of hydroxyl groups on the
surface of amorphous silicas. Langmuir 1987, 3, 316−318.
(33) Tielens, F.; Gervais, C.; Lambert, J. F. o.; Mauri, F.; Costa, D.
Ab initio study of the hydroxylated surface of amorphous silica: a
representative model. Chem. Mater. 2008, 20, 3336−3344.
(34) Deringer, V. L.; Tchougreeff, A. L.; Dronskowski, R. Crystal
orbital Hamilton population (COHP) analysis as projected from
plane-wave basis sets. J. Phys. Chem. A 2011, 115, 5461−5466.
(35) Maintz, S.; Deringer, V. L.; Tchougreeff, A. L.; Dronskowski, R.
Analytic projection from plane-wave and PAW wavefunctions and
application to chemical-bonding analysis in solids. J. Comput. Chem.
2013, 34, 2557−2567.

26219 https://dx.doi.org/10.1021/acsomega.0c03798
ACS Omega 2020, 5, 26211−26219

You might also like