Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

ARTICLE IN PRESS

International Journal of Adhesion & Adhesives 30 (2010) 408–417

Contents lists available at ScienceDirect

International Journal of Adhesion & Adhesives


journal homepage: www.elsevier.com/locate/ijadhadh

Evaluation of mechanical interlock effect on adhesion strength of


polymer–metal interfaces using micro-patterned surface topography
Won-Seock Kim a, Il-Han Yun b, Jung-Ju Lee c,n, Hee-Tae Jung b
a
Satellite Structure Department, Korea Aerospace Research Institute, 45 Eoeun-dong, Yuseong-gu, Daejeon 305-333, South Korea
b
Department of Chemical and Biomolecular Engineering, KAIST, Science Town, Daejeon 305-701, South Korea
c
Department of Mechanical Engineering, KAIST, Science Town, Daejeon 305-701, South Korea

a r t i c l e in f o a b s t r a c t

Article history: This study concerns with the explanation of the wide range of adhesion strengths observed depending
Accepted 5 February 2010 on the nature of substrate surface topography by linking macroscopic adhesion strength to microscopic
Available online 27 May 2010 energy-expenditure mechanisms during fracture. The dominant factors to which the adhesion strength
Keywords: of polymer–metal interfaces is attributed are investigated theoretically and experimentally. In an
Surface roughness/morphology attempt to elucidate the effect of mechanical interlock on adhesion strength, micro-patterns were
Fracture Mechanics fabricated on metal surfaces as a designed surface topography. It was found that the molecular
Contact angles dissipation of the polymer in the vicinity of the interface is the major cause of the practical energy of
Adhesion by mechanical interlocking separation. Furthermore, it is shown that loading mode controls the mechanical interlock effect, which
Micro-pattern
is attributed to the fact that the stress distribution at the interface controls the deformation and failure
characteristics of the polymer resin near the interface. Therefore, mechanical interlock promoted by
adsorption provokes energy dissipation processes during fracture, which practically constitute the
adhesion strength of a polymer–metal interface. The contribution of mechanical interlock to adhesion
strength is systematically assessed by varying micro-pattern dimensions. The influence of the work of
adhesion, cohesion and other dissipation energy on adhesion strength is examined by measuring each
contribution to the total work of fracture.
& 2010 Elsevier Ltd. All rights reserved.

1. Introduction Among a number of mechanisms that account for the adhesion


phenomenon, it is well-known that adsorption and mechanical
Adhesive bonding is increasingly widely applied as a practical interlock are the dominant mechanisms that contribute to the
method for the joining of dissimilar materials in the automotive, adhesion strength of polymer–metal interfaces [1–4]. Adhesive
aircraft, microelectronic and medical devices industries, replacing forces induced by intermolecular attractions exist between a
conventional joining techniques such as welding, bolting and polymer and a metal provided molecular-scale intimate contact is
riveting [1–5]. Various types of polymers are used as adhesives, achieved, and the penetration of the polymer into the craters and
and they are selected according to the substrates to be bonded pores in rough metal substrates enhances the bond strength
and the operating conditions of the adhesive joint [6]. Metals, through mechanical interlock. Chemical bonding can also be
such as steel, aluminum, copper and gold are used in almost every introduced as a dominant adhesion strength enhancement factor
industry, and it is often necessary to join these metals with other by embedding a coupling agent at the interface and forming a
materials. Therefore, the polymer–metal interface becomes an metal/coupling agent/polymer system [7,8].
important design consideration in many structures that use In previous work, researchers have tried to correlate the
adhesive bonding. One of the predominant concerns for engineers measured joint strength with the theoretical work of adhesion
in using adhesive bonding is to optimize the adhesion strength. [9–17]. It is well-known that the thermodynamic work of
Thus, it is essential to understand the adhesion mechanisms adhesion is far less than the practical bond strength; however,
involved and to investigate the influence of each mechanism on the work of adhesion is regarded as playing a vital role in
adhesion strength since this understanding will allow the adhesion so that small changes in its value can cause large
formation of strong and reliable adhesive joints. changes in the practical adhesion strength. Many studies devoted
to measuring adhesion suggest that the greater values of the
practical work of separation compared to the work of adhesion
can be explained by taking the plastic or viscoelastic dissipation of
n
Corresponding author. Tel.: + 82 42 350 3033; fax: + 82 42 350 3210. the adhesive layer or substrates into consideration [9–17]. The
E-mail address: leejungju@kaist.ac.kr (J.-J. Lee). roughness of the interface has been presumed to be the major

0143-7496/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijadhadh.2010.05.004
ARTICLE IN PRESS
W.-S. Kim et al. / International Journal of Adhesion & Adhesives 30 (2010) 408–417 409

factor activating the energy dissipation processes in polymeric


adhesives during crack propagation. Various surface topographies
and characteristic surface parameters have been proposed to
improve adhesion strength and to find an influential parameter of
surface topography on adhesion strength [18–25]. However,
quantitative evaluation of mechanical interlock effect on adhesion
strength has not been reported since previous analysis was
performed based on randomly distributed surface roughness. The
topography of a surface is a kind of random data, which requires a
statistical analysis for the determination of its characteristic
parameters. Nevertheless, a surface can be completely character-
ized with several parameters if a surface topography is designed
and fabricated into a patterned structure. Designed surface
topography facilitates the investigation of how the morphological
change of the surface may influence adhesion strength.
The purpose of the present work is to systematically
investigate the influence of adhesion mechanisms on the adhesion
strength of a polymer–metal bond by incorporating periodic
surface patterns on the metal substrate. In order to link
microscopic adhesion phenomena to the macroscopic adhesion
strength of a bonded joint, the fracture mechanics concept has
Fig. 1. (a) Schematic diagram of a microline-patterned steel specimen. (b)
been utilized measuring adhesion strength as the required energy
Schematic diagram of microline-patterned metal–polymer interfaces; R: max-
for fracture. The interfacial fracture toughness of a polymer–metal imum etching depth, which corresponds to conventional definition of surface
bond was measured using the bi-material end-notched flexure roughness, w1: excavated line width, w2: preserved line width. Congruent pattern
(ENF) test for mode-II and the single-leg bend (SLB) test for design was utilized to investigate the roughness effect only. Other surface
parameters (w1/R, w2/R, Areal/Anom) could be maintained as a constant value.
mixed-mode. The cause of adhesion strength, rather than
adhesion strength enhancement, was targeted in this study by
elucidating the energy-expending process during fracture. Quan-
titative assessment of fracture energetics was achieved by characterized by these three independent parameters (Fig. 1(b)).
controlling the area fraction of adhesive and cohesive failure via The pattern parameters were varied by using different masks
systematic variation of surface topography. in the process of photolithography and by controlling etching
time. All of the pattern parameters (R, w1, w2) were in the range of
10–90 mm, which covers the minimum pattern sizes that can be
2. Experimental controlled using the isotropic wet-etching process. First, the three
surface design parameters were chosen to vary in a congruent
2.1. Materials and sample preparation manner as shown in Fig. 1(b) in order to investigate the roughness
effect only. Using a set of surfaces with congruent geometry, the
In order to investigate the fracture energetics of the polymer– aspect ratios (w1/R, w2/R) and roughness factor (the ratio of real
metal interface, composite/steel bonded joints were fabricated. contact area to nominal area, Areal/Anom), which have been
Steel (AISI 1045) as a conventional metal and carbon fiber- reported to have strong influence on adhesion strength as surface
reinforced polymer (CFRP) as an advanced structural material describing parameters [23–25], can be maintained as constants.
were chosen as the substrates to be bonded. The steel plates were Control joint specimens were also produced using steel plates
cut into 35  5  1 mm3 samples and were finely polished using a with polished surfaces and no patterns.
rotating disk-type polisher with diamond suspensions. Diamond
suspensions with particle sizes of 9, 6, 3, 1 and 0.25 mm were used
sequentially to attain smooth and flat surfaces. Afterwards, the 2.2. Contact angle measurement
polished steel surfaces were coated with a commercial photo-
resist (SU-8 from MicroChem, USA) and exposed to UV through a The work of adhesion of a polymer–steel interface was
transparency-printed mask. The UV-exposed photoresist film was determined by contact angle measurements. In general, contact
developed to have a series of line posts, whose width varied from angles are formed by the sessile drop method or annealing thin
10 to 90 mm using different masks. After the conventional films. In this study, a kind of annealing method was utilized to
photolithography, the steel surfaces not covered by the photo- obtain epoxy particles on a steel substrate. The epoxy resin which
resist films were chemically etched in an etchant called Nital composes the CFRP prepreg is in solid state at room temperature
solution (nitric acid and ethanol mixed solution, v/v ¼1/9) to and becomes a liquid as the temperature rises to the curing point
fabricate microline-patterns on the steel surface (Fig. 1(a)). After (125 1C), and then irreversible solidification is attained by the
the surface patterning process, CFRP prepregs (USN-150 from SK chemical reaction between the epoxide and the premixed cross-
Chemical, Korea) were laminated on the steel plate ([09/steel]) linking agent. The bisphenol-A based epoxy resin was extracted
and bonded through the standard curing process of the CFRP from the CFRP prepreg and shaped in the form of pellets with a
prepreg. In other words, composite/metal joint specimens were diameter of about 2 mm, weighing around 5 mg. The pellets were
fabricated using a co-cure bonding method. then placed on the polished steel surface and annealed using the
The micro-scale line-patterns, which were fabricated on the same temperature cycle as that of the curing process used for
steel plate prior to bonding, were created perpendicular to the bonding. During the annealing process, the epoxy pellets on the
direction of crack propagation. The effect of metal surface steel plate wet the substrate, forming small contact angles. Since
morphology on adhesion strength can be assessed by varying there was no significant change of the shape of the wetted
the width (w1), depth (R) and spacing (w2) of microline-patterns, particles after solidification, contact angles were measured at
since the line-patterned surface morphology can be completely room temperature. The contact angles of the epoxy particles were
ARTICLE IN PRESS
410 W.-S. Kim et al. / International Journal of Adhesion & Adhesives 30 (2010) 408–417

observed using a contact angle analyzer (Phoenix 300 from S.E.O. were miniaturized due to the technical difficulty of micro-
Corporation, Korea). patterning on a large area. Micro-patterns were located at the
uncracked ligament of 5 mm (from the initial crack tip to the
2.3. Fracture testing loading point). The thickness ratio of ENF specimens was held
constant at (composite thickness)/(steel thickness)¼1.25 to
The interfacial fracture toughness Gc, which is the released locate the neutral axis of the bonded beam at the interface. The
energy per unit interfacial area at which fracture occurs, was thickness ratio of SLB specimens was held constant at (composite
measured as the adhesion strength of the composite/steel inter- thickness)/(steel thickness)¼ 0.5 to induce a mode-I dominant
face. Mechanical testing methods for interfacial toughness loading condition. An interfacial pre-crack was introduced by
measurement include the double-cantilever beam (DCB) test for smearing mold release agent (ER-650 from Nabakem, Korea) on
mode-I, end-notched flexure (ENF) test for mode-II and single-leg the steel surface prior to bonding. A miniaturized three-point
bend (SLB) test for mixed-mode. Among various test methods, the bend fixture was used as shown in Fig. 2(c) and (d), and the
aforementioned beam-type specimen geometry has definite testing was performed under displacement control at a ramp rate
advantages. Beam-type specimens are simple to fabricate, amen- of 10 mm/s using a computer-controlled stepper motor.
able to miniaturization and easy to test. Most previous works Another factor that should be stated in adopting a test method
[9–15,22,26–29] performed pure mode-I tests using a peel test or for adhesion measurement is whether the measured adhesion
asymmetric DCB configuration to measure adhesion energy while strength is intrinsic to the interface. Engineering tests such as the
structural adhesive joints are designed to transfer loads by shear peel and blister tests are often sensitive to the global inelastic
stress distribution along the bond line. In this study, a bi-material deformation of the polymer occurring away from the interface.
ENF test, which provides a pure mode-II loading condition [30], Thus, the measured interfacial toughness is influenced by the
and a bi-material SLB test, which provides various mixed-mode thickness of a specimen and by the bulk properties of the
loading conditions [31], were employed. Fig. 2(a) and (b) show polymer. Therefore, inelastic deformations of the polymer should
schematic diagrams of the tests. The dimensions of the specimens be confined at the crack tip and in the vicinity of the interface to
propose the measured strength as a property of the interface. The
localization of inelastic deformation along the interface of the
specimen will be demonstrated in the results section.

3. Numerical modeling

A two-dimensional finite element (FE) analysis was conducted


in order to calculate the critical energy release rate, Gc using the
experimentally measured crack onset load, Pc. The numerical ENF
and SLB models consist of an upper CFRP layer and a lower steel
layer, and the interface was modeled as a straight line. Micro-
patterned interfacial roughness was not considered in the FE
model. For this reason, in this study, the measured interfacial
toughness is a macroscopic interfacial property which is a
function of micro- and nano-scale interfacial roughness. Second-
order eight-node quadrilateral elements with full integration
were used in the analysis. The interface between the two
adherends was modeled by sharing identical nodes, and the
interfacial crack was modeled by detaching nodes along some
portion of the interface. A frictionless contact option was applied
onto the crack surfaces. Near the interfacial crack tip, the FE mesh
was refined and quarter point singular elements were used to
describe the singular stress field efficiently. The models were
implemented in ABAQUS by performing a plane-strain analysis,
and the energy release rate was calculated through the domain
integration-based J-integral technique. Residual stresses caused
from high-temperature bonding of materials having disparate
thermal expansion coefficients were considered. Interfacial strain

Table 1
Material properties used in analyses.

Material properties of carbon fiber reinforced epoxy (USN 150)


Fiber direction modulus E1 (GPa) 131.6
Transverse direction modulus E2 ¼E3 (GPa) 8.2
In-plane shear modulus G12 (GPa) 4.5
Poisson’s ratio n12 0.28
Fiber direction thermal expansion coefficient a1 (10  6/1C)  0.9
Transverse direction thermal expansion coefficient a2 ¼ a3 (10  6/1C) 27

Material properties of steel (AISI 1045)


Young’s modulus E (GPa) 200
Fig. 2. Adhesion fracture tests schematics and photographs: (a) ENF test schematics Poisson’s ratio n 0.3
(dimensions in mm); (b) SLB test schematics (dimensions in mm); (c) miniaturized Thermal expansion coefficient a (10  6/1C) 12
composite/metal ENF test and (d) miniaturized composite/metal SLB test.
ARTICLE IN PRESS
W.-S. Kim et al. / International Journal of Adhesion & Adhesives 30 (2010) 408–417 411

energy release rates in the layered specimens have non-zero images, presented in Fig. 4, show the steel/CFRP bonded
initial values before mechanical loading is applied due to the interface. The cross-section of the periodic grooves became
residual thermal stresses, which indicates that evaluation of semicircular due to the wet-etching process applied to the steel,
interfacial toughness ignoring residual thermal stresses will be an an isotropic material. Due to the isotropic wet-etching process,
underestimation compared to the true interfacial toughness. The surface parameters maintain a relationship of constant aspect
material properties used for calculations are listed in Table 1. ratio: R/w1 ¼0.5. The width of etched line w1 was varied from 10
to 90 mm with its maximum depth, R, ranging from 5 to 45 mm.
The epoxy resin, which was extracted from the CFRP prepreg
4. Results and discussion during its curing process, completely filled the micro-grooves on
the steel surface. Therefore, mechanical keying, or interlocking, of
the polymer into the steel surface cavities was completely
4.1. Surface and interface observation
attained at the microline-patterned steel/CFRP interface.
The submicroscopic morphology of the etched metal surface
Surfaces and cross-sections of interfaces were observed using
also has effect on adhesion strength though it was not under
both optical and scanning electron microscopes (SEM). The SEM
control in this study [32]. Exactly speaking, micro-scale surface
photographs, presented in Fig. 3, show the micro-scale line-
roughness was designed, but the overspread nano-scale surface
patterned steel surface. Optical microscope cross-sectional
roughness was naturally formed during the micro-patterning
process. Consequently, the naturally formed nano-morphology
was uniformly applied to all etched steel surfaces.

4.2. Work of adhesion

From the thermodynamic point of view, the work of adhesion


of a bi-material bond is the amount of energy required to create
free surfaces by separating the interface. The work of adhesion is
calculated by the Dupre equation:
Wa ¼ g1 þ g2 g12 ð1Þ

where Wa is the reversible energy of adhesion, g1 and g2 are the


surface free energies of the contacting solids denoted by 1 and 2,
and g12 is the interface free energy between the two solids.
Therefore, the thermodynamic work of adhesion, Wa, is a constant
for a specific pair of solids as an intrinsic property of the interface,
and it does not depend on loading mode. In the thermal
equilibrium condition, the interface free energy gi between a
solid particle and a solid substrate is given by the Young equation:
gi ¼ gS gP cos y ð2Þ

where gS and gP are the surface free energies of the substrate and
particle, respectively, and y is the contact angle (Fig. 5(a)).
According to the Young–Dupre equation, the adhesion energy
between the particle and substrate is calculated as follows:
Wa ¼ gS þ gP gi ¼ gP ð1 þ cos yÞ ð3Þ
Therefore, using the surface free energy of the epoxy and the
measured contact angle between the epoxy particle and the steel
substrate, the adhesion energy of the epoxy/steel bond can be
calculated. It is necessary to remember that the surface free
energy, g, refers to the energy required to create new surfaces of
the material in a vacuum. The simplification of zero spreading
Fig. 3. Scanning electron micrographs of a microline-patterned steel surface. Line pressure underlies using Eq. (3) for calculating the work of
widths were maintained as a constant ratio of w1/w2 ¼ 1. adhesion in the presence of ambient air.

Fig. 4. Optical micrographs of the polished cross-section of a steel/composite bonded interface. Patterned surface parameters were w1 ¼w2 ¼30 mm and R¼ 15 mm.
ARTICLE IN PRESS
412 W.-S. Kim et al. / International Journal of Adhesion & Adhesives 30 (2010) 408–417

Fig. 5. (a) Schematic diagram of contact angle measurement. An epoxy particle on


the steel substrate: (b) before annealing and (c) contact angle formed after
annealing.

Fig. 6. A typical load–displacement curve from the bi-material ENF test.

As the sample was heated near the curing temperature of the


epoxy resin (125 1C), the epoxy pellets on the steel plate wetted
the substrate by forming small contact angles (Fig. 5(b) and (c)).
The temperature cycle for annealing was identical to the
manufacturer’s recommended curing cycle of the prepreg, which
was also used as a bonding process in this study. The measured
contact angle of the epoxy particles on the steel substrate is
24.21 72.31 (the average of ten individual determinations). The
room-temperature surface energy of typical amine-cured epoxy
resin is reported to be 46.2 mJ/m2 ðgS ¼ gD P
S þ gS ¼ 41:2mJ=m þ
2

5:0 mJ=m2 Þ [1,4]. Using the surface free energy value of a typical
epoxy resin (46.2 mJ/m2) and the measured value of the contact
angle (24.21), the work of adhesion between the epoxy resin and
steel was calculated to be 88.3 mJ/m2 via Eq. (3).
The adhesion energy, Wa, between two materials influences
how a liquid (adhesive) will wet a solid (adherend). Thus, the
adhesion energy plays a significant role in the spreading of hot-
melt adhesives. Penetration of a liquid into the irregularities of a
substrate surface is a prerequisite for the functioning of the Fig. 7. Interfacial fracture toughness, GIIc for different pattern depths, R. The
patterns with different pattern widths and depth maintained congruent shapes.
mechanical interlock mechanism. Therefore, adsorption promotes
mechanical interlock. The work of adhesion is an important
parameter as a promotion factor for other bond strength
enhancement mechanisms though its value is far below that of line-based critical load, Pc, as the crack onset load. The critical
the practical work of separation. In the epoxy/steel bond, the load, Pc was transformed into the critical energy release rate Gc,
measured value of the adhesion energy, Wa ¼88.3 mJ/m2, was the energy dissipated during fracture per unit fracture area by
enough to form a mechanical interlock-activating interface by using the numerical method explained in Section 3. The deter-
completely filling the epoxy resin into the micro-scale roughened mined values of interfacial fracture toughness from ENF test, GIIc
steel surfaces. The strengthening mechanism of the polymer/ were plotted against the pattern depth, R, in Fig. 7. The width and
metal bond caused by mechanical interlock and the further role of depth of the patterns were varied in a congruent manner by
the work of adhesion during the fracture process will be discussed maintaining the width ratio as w1:w2 ¼1:1 to isolate the influence
in the following sections. of the surface roughness (Fig. 1(b)). The lack of correlation in Fig. 7
suggests that the surface roughness, R, does not control adhesion
strength. In order to investigate the locus of failure, fracture
4.3. Interfacial fracture toughness and failure locus surfaces were examined using a scanning electron microscope.
As seen in Fig. 8, the excavated region on the steel surfaces shows
A typical load–displacement curve obtained from an ENF test is pure cohesive failure and the preserved region on the steel
presented in Fig. 6. Although the epoxy resin in the fabricated surfaces shows pure interfacial failure regardless of the variation
epoxy/steel bond is brittle, there is a gradual transition from of pattern dimensions. As long as the width ratio is maintained as
linear to non-linear behavior due to the stable crack propagation, w1:w2 ¼ 1:1, the total areas of cohesive and adhesive failure
and the point at which crack propagation began is hard to regions are the same. This observation reveals that the total area
determine. The critical point at which the compliance has fraction of the cohesive and adhesive failure regions is directly
increased by 5% was interpreted as the onset of crack related to adhesion strength while the surface roughness, R, is an
propagation and used for the calculation of the critical energy indirect parameter that contributes to failure mode transition
release rate, Gc [33,34]. The load at the critical point Pc always fell from adhesive to cohesive or vice versa.
between the proportional limit, PNL, and maximum load, Pmax In order to increase adhesion strength by increasing the area
(Fig. 6), which validates the employment of the 95% secant fraction of cohesive failure, the width ratio of line patterns was
ARTICLE IN PRESS
W.-S. Kim et al. / International Journal of Adhesion & Adhesives 30 (2010) 408–417 413

Fig. 8. Fracture surfaces of ENF specimens on the patterned steel side. Preserved steel surface region failed adhesively showing the clear steel adherend, and excavated
steel surface region failed cohesively showing the epoxy resin.

mode (Fig. 10(c)). The observed partial cohesive failure mode


implies that the secondary bonds across the interface are strong
enough to cause crack deflection into the polymer resin, resulting
in large energy dissipation during the gradual crack sliding.
In the case of mode-I dominant loading, both mirror-polished
and micro-patterned steel substrates showed poor adhesion
strength compared to the mode-II test result. The failure loads
reached only one-tenth of those under the mode-II loading
condition (Fig. 11). The SLB test of a mirror-polished steel
substrate resulted in almost pure interfacial failure (Fig. 12(a))
while the micro-patterned steel substrates resulted in a partial
cohesive failure (Fig. 12(b)). In a magnified view of the interfacial
failure region on a patterned substrate (Fig. 12(c)), many scraps of
epoxy resin could be found on the fracture surface, which
indicates frictional resistance during pull out of the interlocked
epoxy resins. Therefore, although adhesion strength can be
enhanced under crack opening-mode loading, a mechanically
interlocked steel/epoxy interface under mode-I dominant loading
does not provide sufficiently reliable bond strength to be used for
Fig. 9. Typical load–displacement curves of bi-material ENF specimens with load-bearing joints.
different pattern width ratios (w1/w2). Scanning electron micrographs of the CFRP/steel interface after
crack propagation show that crack growth under mode-II loading
causes plastic damages in the polymer resin for both adhesive
varied as w1:w2 ¼ 1:1; 2:1; 4:1; 6:1; 8:1. Typical load–displace- and cohesive failure (Fig. 13(a)). A long slender damage zone
ment curves obtained from ENF tests with different pattern width containing numerous micro-cracks developed during the sliding-
ratios are presented in Fig. 9. The failure load of adhesive joints mode crack propagation (Fig. 13(b)). The failure in the interfacial
increases according to the increment of the ratio of the excavated failure mode region, which showed the clean metal adherend
width, w1, to the preserved width, w2, of the line-patterns. As seen on the fracture surfaces, was not mere adhesive failure at the
in Fig. 10(a) and (b), the excavated region on the steel surfaces interface; it created a damage layer of about 5 mm in the polymer
always shows pure cohesive failure and the preserved region on neighboring the interface. The damage layer in the cohesive
the steel surfaces always shows pure interfacial failure regardless failure region is also thin and parallel with the interface. Micro-
of the variation of the pattern width ratio (w1/w2). The cause of cracks are oriented approximately 451 from the main crack
the high interfacial toughness of specimens having a large pattern propagation direction, which means that the micro-cracks are
width ratio (w1/w2) can be explained by the increase of the area oriented perpendicular to the maximum normal stress as is usual
fraction of cohesive failure. Cohesive failure, caused by the for brittle materials (Fig. 13(b)). As loading increases, these micro-
molecular decohesion of the polymer resin, expends larger cracks coalesce with each other, forming the main crack growth.
energy for crack propagation than interfacial failure, caused by The high interfacial toughness in mode-II results from energy
the debonding of secondary bonds across the interface. In the case dissipation in this damage zone. The experimental observation of
of mirror-polished surfaces without any pattern, the load– enhanced interfacial toughness by the increment of the area
displacement curve of an ENF test shows that the initial slope fraction of cohesive failure region suggests that viscoelastic and
starts to decrease earlier than those of patterned specimens plastic dissipation in the polymer material is more effectively
(Fig. 9). This indicates that the crack started to propagate at the induced during the crack growth within the polymer resin than
moment when the initial slope of the load–displacement curve during the crack growth along the interface, though the interfacial
started to decrease, though the maximum load is much higher debonding also induces a damage zone in the polymer. The steel/
than the crack onset load. Since there is no mechanical interlock composite interface after mode-I dominant loading fracture
effect on mirror-polished steel surfaces, the epoxy resin attached shows that the epoxy resin that had penetrated into the cavities
to the metal substrate can easily slide, and the crack can of the patterned steel substrates was pulled out without breaking
propagate along the interface without mechanical obstacles. the internal bonds of the polymer, and plastic damage zones were
However, the fracture surface shows a partial cohesive failure not found in the polymer (Fig. 13(c)). The epoxy resin near the
ARTICLE IN PRESS
414 W.-S. Kim et al. / International Journal of Adhesion & Adhesives 30 (2010) 408–417

Fig. 11. Typical load–displacement curves of bi-material SLB specimens with and
without micro-pattern.

main source of the adhesion strength of polymer–metal interfaces


by provoking energy dissipation processes in the polymer resin
during joint failure. Furthermore, different loading modes near
the interfacial crack tip cause different interactions between the
material property and interface geometry, resulting in different
adhesion strengths.

4.4. Energetics of the fracture process

Many researchers have suggested that the adhesion fracture


energy, or interfacial toughness, Gc, comprises two principal
components; one is the intrinsic fracture energy, G0, and the other
is the dissipated energy for the deformation of materials during
fracture, Gdiss:
Gc ¼ G0 þ Gdiss ð4Þ
The intrinsic fracture energy G0 is the energy required to
propagate a crack in the absence of any energy losses in the
material and is only related to thermodynamic surface energy
terms. The energy dissipated in the deformation of the material
during fracture, Gdiss, is usually much larger than G0 and is
responsible for the dependence of Gc on loading mode, tempera-
ture and crack propagation rate. Multiplicative relationships of
the two effects have been sought on the hypothesis that G0 and
Gdiss are coupled
Gc ¼ G0  fðc, T, cÞ ð5Þ
where j (c, T, c) is a mode mixity, c, temperature, T and loading
rate, c dependent energy loss term. The theoretical derivation of
the relationship between G0 and Gdiss was also proposed by
Andrews and Kinlock [10]. The hypothesis that the value of Gdiss is
Fig. 10. Fracture surface on steel side of ENF specimens: (a) pattern width ratio of connected with G0 so that Gdiss is proportional to G0 is now
w1:w2 ¼ 2:1; (b) pattern width ratio of w1:w2 ¼8:1 and (c) mirror-polished surface generally accepted [35–37].
without patterning.
In the present work, quantitative assessment of the role of the
thermodynamic work of adhesion and cohesion in adhesion
epoxy/steel interface does not behave in an elastic manner when strength was facilitated by the adhesive and cohesive failure
loaded to the point of fracture. Chain pullout, microvoid mode control using microline-patterned surface design. Inter-
formation and crazing are common failure mechanisms in facial toughness values, GIIc obtained from ENF tests using various
polymers, and the energy dissipation during these damage pattern dimensions are plotted in Fig. 14 as a function of the line
formations constitutes the practical fracture energy. The width ratio, w1/(w1 + w2) on steel surfaces. Since the pattern width
variation of stress distribution at the polymer–metal interface ratios w1/(w1 +w2) and w2/(w1 +w2) were always identical to the
caused from the increased crack opening-mode loading changed area fractions of the cohesive and interfacial failure regions, the
the fracture process at the interface from a more to a less energy- measured joint failure energy per unit area, Gc, can be
consuming mode. Therefore, mechanical interlock provides the decomposed into cohesive and adhesive failure energies by an
ARTICLE IN PRESS
W.-S. Kim et al. / International Journal of Adhesion & Adhesives 30 (2010) 408–417 415

Fig. 13. SEM micrographs of the composite/steel interface after crack propagation:
Fig. 12. Fracture surfaces on steel side of SLB specimens: (a) mirror-polished
(a) crack growth under mode-II loading; (b) magnified view of the damage zone
surface without patterning; (b) pattern width ratio of w1:w2 ¼ 1:1 and
and (c) crack growth under mode-I dominant loading.
(c) magnified view of patterned fracture surface.

equation of the form are related with thermodynamic surface energy terms as:
w1 w2 Ea ¼ Wa  a ð7Þ
Gc ¼ Ec þ Ea
w1 þ w2 w1 þ w2
w1 Ec ¼ Wc  b ð8Þ
¼ ðEc Ea Þ þ Ea ð6Þ
w1 þ w2
where a and b are factors of viscoelastic and plastic energy
where Ec and Ea are the energies required for unit area separation dissipation for the polymer deformation during joint failure, and
in the pure cohesive and adhesive modes, respectively. Ea and Ec Wa and Wc are the intrinsic work of adhesion and cohesion,
ARTICLE IN PRESS
416 W.-S. Kim et al. / International Journal of Adhesion & Adhesives 30 (2010) 408–417

at the interface controls the deformation and failure character-


istics of the polymer resin near the interface, strong adhesive
joints can be attained not only through consideration of the
surface morphology of the metal substrate and the mechanical
properties of the polymer, but also by joint design optimization to
transfer loads via shear stresses across the bond line.
This study attempts to explain why the adhesion strength of
polymer–metal bonded joints shows great variation according to
the surface topography of the metal substrates and the loading
mode. It is important to emphasize that mechanical interlock is
not just one aspect governing the adhesion phenomenon.
Mechanical interlock promoted by adsorption provokes energy
expenditure during fracture, which practically constitutes the
adhesion strength.

Acknowledgements

Fig. 14. Interfacial fracture toughness, GIIc as a function of pattern width ratio. This research was supported by Basic Science Research
Program through the National Research Foundation of Korea
respectively. If viscoelastic and plastic energy dissipations are (NRF) funded by the Ministry of Education, Science and Technol-
excluded, interfacial failure requires the work of adhesion, ogy (No. 2009-0083647).
Wa ¼ g1 + g2  g12, and cohesive failure requires the work of
cohesion, Wc ¼ 2g1 (or 2g2). Energy dissipation factors a and b References
caused from mechanical interlock effects can be inferred using the
known intrinsic adhesion strengths and the measured adhesion
[1] Kinloch AJ. Adhesion and adhesives: science and technology. New York:
strength variation according to surface parameter variation Chapman and Hall; 1987.
(Fig. 14). By the interpolation and extrapolation of the [2] Packham DE. Handbook of adhesion. UK: Longman; 1992.
[3] Pizzi A, Mittal KL, editors. Handbook of adhesive technology. New York:
interfacial toughness data in Fig. 14, the quantities Ea and Ec can
Marcel Dekker; 1994.
be evaluated as Ea E210 and Ec E570 J/m2. From the results [4] Pocius AV. Adhesion and adhesives technology: an introduction. New York:
obtained in the previous sections (Wa ¼88.3 mJ/m2 and Hanser; 1997.
Wc ¼46.2 mJ/m2), a E2380 and b E12 340. This means that the [5] Yacobia BG, Martin S, Davis K, Hudson A, Hubertb M. Adhesive bonding in
microelectronics and photonics. J Appl Phys 2002;91:6227–62.
intrinsic energies were increased by the 3rd order of magnitude in [6] Wake WC. Theories of adhesion and uses of adhesives: a review. Polymer
adhesive failure mode and by the 4th order of magnitude in 1978;19:291–308.
cohesive failure mode. The dissipation factors, a and b, will be [7] Plueddemann EP. Silane coupling agents, 2nd ed. New York: Plenum Press;
1991.
functions of the mechanical properties of the polymer, which can [8] Mittal KL, editor. Silanes and other coupling agents. Utrecht: VSP; 1992.
be modified by altering its formulation. The upper limit of the [9] Kaelble DH. Peel adhesion: influence of surface energies and adhesive
adhesion strength of a polymer–metal bond will be rheology. J Adhes 1969;1:102–23.
[10] Andrews EH, Kinloch AJ. Mechanics of adhesive failure. I. Proc Roy Soc A
Gc ¼Ec ¼Wc  b in pure cohesive failure. Therefore, in order to 1973;332:385–99.
increase the failure energy of a bonded joint, the interaction [11] Andrews EH, Kinloch AJ. Mechanics of adhesive failure. II. Proc Roy Soc A
between surface morphology and the mechanical properties of 1973;332:401–14.
[12] Gent AN, Kinloch AJ. Adhesion of viscoelastic materials to rigid substrates. III.
the polymer resin should be optimized to cause increased energy
Energy criterion for failure. J Polym Sci, Part A: Polym Chem 1971;9:
expenditure during crack propagation. 659–68.
[13] Mittal KL. The role of the interface in adhesion phenomena. Polym Eng Sci
1977;17:467–73.
[14] Maugis D, Barquins M. Fracture mechanics and the adherence of viscoelastic
5. Conclusions bodies. J Phys D: Appl Phys 1978;11:1989–2023.
[15] Carre A, Schultz J. Polymer–aluminum adhesion II. Role of the adhesive and
The surface topography effect on the adhesion strength of an cohesive properties of the polymer. J Adhes 1984;17:135–56.
[16] Allen KW. Some reflections on contemporary views of theories of adhesion.
epoxy/steel bond was investigated. Experimental results show Int J Adhes Adhes 1993;13:67–72.
that the major source of strength enhancement caused by metal [17] Seshadri M, Saigal S, Jagota A, Bennison SJ. Scaling of fracture energy in
surface topography modification in polymer–metal bonded joints tensile debonding of viscoelastic films. J Appl Phys 2007;101:093504.
[18] Gent AN, Lin CW. Model studies of the effect of surface roughness and
is the transition from interfacial to cohesive failure, which is also mechanical interlocking on adhesion. J Adhes 1990;32:113–25.
controlled by loading mode. The adhesion strength is increased by [19] Chen Y, Kalyon DM, Bayramli E. Effects of surface roughness and the chemical
the increased surface roughness of a substrate only if the structure of materials of construction on wall slip behavior of linear low
density polyethylene in capillary flow. J Appl Polym Sci 1993;50:1169–77.
increased roughness causes the transition of failure mode from [20] Shanahan MR, Morris AP. Adhesion to sintered substrates. J Adhes
adhesive to cohesive. The high fracture toughness of microline- 1996;59:51–9.
structured interfaces is attributed to the mechanical interlock [21] Janarthanan V, Garrett PD, Stein RS, Srinivasarao M. Adhesion enhancement
in immiscible polymer bilayer using oriented macroscopic roughness.
mechanism, which provides the energy expending processes
Polymer 1997;38:105–11.
whereby viscoelastic and plastic energy dissipation and corre- [22] Arola DD, Yang DT, Stoffel KA. Fatigue of the cement/bone interface: surface
sponding damage zone development in the polymer are required texture and wear debris. J Biomed Mater Res, Part B 2001;58:519–24.
[23] Packham DE. Surface energy, surface topography and adhesion. Int J Adhes
for crack propagation. Therefore, once the adhesion between a
Adhes 2003;23:437–48.
specific metal with a certain surface morphology and a specific [24] Prolongo SG, Rosario G, Urena A. Study of the effect of substrate roughness on
polymer is achieved by adsorption, the fracture characteristics adhesive joints by SEM image analysis. J Adhes Sci Technol 2006;20:
will depend on the interaction between the surface topography of 457–70.
[25] Jiang ZX, Huang YD, Liu L, Long J. Effects of roughness on interfacial
the metal substrate and the mechanical properties of the polymer by performances of silica glass and non-polar polyarylacetylene resin compo-
mechanical interlock. Furthermore, since the stress distribution sites. Appl Surf Sci 2007;253:9357–64.
ARTICLE IN PRESS
W.-S. Kim et al. / International Journal of Adhesion & Adhesives 30 (2010) 408–417 417

[26] Gent AN, Petrich RP. Adhesion of viscoelastic materials to rigid substrates. [32] Brockmann W, Geiss P, Klingen J, Schroder K. Adhesive bonding: materials,
Proc Roy Soc A 1969;310:433–48. applications and technology. Darmstadt: Wiley-VCH; 2009. [Chapter 7].
[27] Kendall K. Peel adhesion of solid films—the surface and bulk effects. J Adhes [33] ASTME 1820-05. Standard test method for measurement of fracture
1973;5:179–202. toughness.
[28] Cao HC, Evans AG. An experimental study of fracture resistance of bimaterial [34] ASTMD 5045-99. Standard test methods for plane-strain fracture toughness
interfaces. Mech Mater 1989;7:295–305. and strain energy release rate of plastic materials.
[29] Akisanya AR, Fleck NA. Brittle fracture of adhesive joints. Int J Fract [35] Jokl ML, Vitek V, McMahon CJ. A microscopic theory of brittle fracture in
1992;58:93–114. deformable solids: a relation between ideal work to fracture and plastic
[30] Kim WS, Lee JJ. Fracture characterization of interfacial cracks with frictional work. Acta Metall 1980;28:1479–88.
contact of the crack surfaces to predict failures in adhesive-bonded joints. [36] Volinsky AA, Moody NR, Gerberich WW. Interfacial toughness measurements
Eng. Fract Mech 2009;76:1785–99. for thin films on substrates. Acta Mater 2002;50:441–66.
[31] Davidson BD, Sundararaman V. A single leg bending test for interfacial [37] Je H, Was GS, Thouless MD. Measurement of the noibium/sapphire interface
fracture toughness determination. Int J Fract 1996;78:193–210. toughness via delamination. Int J Fract 2003;119/120:441–8.

You might also like