Sunteți pe pagina 1din 23

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/263607497

A review on the mechanical methods for


evaluating coating adhesion

Article in Acta Mechanica February 2014


DOI: 10.1007/s00707-013-0979-y

CITATIONS READS

28 206

4 authors, including:

Kun Zhou
Nanyang Technological University
205 PUBLICATIONS 2,240 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Material development and optimisation of nanocomposite for selective laser sintering View project

Additive manufacturing via selective laser sintering for automobile application View project

All content following this page was uploaded by Kun Zhou on 03 July 2016.

The user has requested enhancement of the downloaded file.


Acta Mech 225, 431452 (2014)
DOI 10.1007/s00707-013-0979-y

Zhaoxiang Chen Kun Zhou Xuehong Lu


Yee Cheong Lam

A review on the mechanical methods for evaluating


coating adhesion

Received: 18 March 2013 / Revised: 5 August 2013 / Published online: 22 August 2013
Springer-Verlag Wien 2013

Abstract Coatings have been applied widely in aerospace, biomedical, electronic, and many other industries.
The performance of a coating is dictated by the adhesion between the coating and the underlying substrate.
Thus, the evaluation of coating adhesion is critical for the assessment of the quality of a coating and its fitness for
service. However, this evaluation is not straightforward, and various mechanical and non-mechanical evaluation
methods have been developed. This paper presents a comprehensive review of the currently most widely
employed mechanical methods. The interface fracture mechanics that provides the fundamental knowledge
for conducting adhesion assessment and analysis is briefly introduced. Detailed discussions are presented
on various sandwich specimen-based and bimaterial specimen-based testing methods, with emphasis on the
principles, merits, limitations, typical applications, and recent improvements of each method.

1 Introduction

Surface coating technology presents exciting opportunities for changing the surface properties of bulk materials
and imparting new and unique functions to engineering components underneath coatings, such as excellent
corrosion resistance, improved biological performance, and special magnetic or electrical properties [14].
It is worth noting that the performance and reliability of coatings depend on the mechanical integrity of
coating/substrate systems, i.e., the adhesion of coatings to their substrates. Adhesion is the most important
property of a coating/substrate system because it determines the durability and longevity of the whole system.
With delamination or spallation of coatings on the surface, the components underneath will be directly exposed
to environmental factors such as corrosive gases or liquids, abrasive particles, and high temperature, causing
the components to function improperly and even to fail prematurely [57]. Therefore, there is a need to evaluate
the coating adhesion with simple and reliable methods.
The mechanisms of coating adhesion can be categorized into: (a) interfacial adhesion in which the adhesive
forces are centered around a well-defined thin interface; (b) interdiffusion adhesion in which the coating and the
substrate diffuse into one another with a thicker interfacial region; (c) intermediate layer adhesion in which the
coating and the substrate are separated by one or more layers of materials of different chemical compositions;
(d) mechanical interlocking adhesion in which the coating/substrate interface is relatively rough [8]. Among

Z. Chen K. Zhou (B) Y. C. Lam


School of Mechanical and Aerospace Engineering, Nanyang Technological University,
50 Nanyang Avenue, Singapore 6389798, Singapore
E-mail: kzhou@ntu.edu.sg
Tel.:+65-6790-5499
Fax: +65-6792-4062

Z. Chen X. Lu
School of Materials Science and Engineering, Nanyang Technological University,
50 Nanyang Avenue, Singapore 6389798, Singapore
432 Z. Chen et al.

them, interdiffusion adhesion and mechanical interlocking adhesion usually exhibit relatively high strength.
In addition, the adhesion strength of a coating to a substrate is also a function of many other variables such
as the thermal and elastic mismatch between the coating and the substrate, the thickness of the coating, and
the presence of defects at or near the interface [911]. To quantify the effects of these variables on coating
adhesion, a good understanding of the adhesion and damage mechanisms of coating/substrate systems is of
advantage. To achieve the full potential of coatings, it is necessary to develop reliable techniques for evaluating
coating adhesion.
In recent years, various methods have been proposed and developed to characterize the adhesion of coatings
to the substrate [58]. These methods can mainly be categorized into nucleation methods, mechanical methods,
and miscellaneous techniques [5,6]. Nucleation methods are based on the observation of the formation kinetics
of films or coatings, and the early stages of coating formation are monitored. These methods cannot easily be
used to test the adhesion of coatings on practical tools or machine parts. In contrast, mechanical methods are
of practical interest and used widely in engineering and research [7,8]. According to the types of specimens,
mechanical methods can be broadly classified into sandwich specimen-based and bimaterial specimen-based
methods. A sandwich specimen is constructed by sandwiching the coating into the test structure, while a
bimaterial specimen includes only two types of materials: coating and substrate. Miscellaneous techniques
refer to the measurement techniques, which are still in a state of infancy but offer great potential, such as the
thermal method, X-ray method, and capacitance method [5,6].
The present review will focus on the mechanical methods used for adhesion evaluation of coatings, with
emphasis on the principles, merits, limitations, typical applications, and recent improvements of each method.

2 Interface fracture mechanics

Many mechanical techniques have been proposed to characterize and evaluate coating adhesion, such as
double cantilever beam test, four-point bend test, Brazil-nut test, and indentation test [1215]. Although these
techniques are very different in specimen shape or loading mode, their principles are fundamentally similar
from a fracture mechanics viewpoint. All these techniques establish an equilibrium between the elastic energy
provided by the external load and the propagation of an interfacial crack between the coating and the substrate.
The interfacial strain energy release rate G is derived from such energy balance and defined as
(W U )
G= , (1)
A
where W is the work done by external forces, U is the elastic energy stored in the system, and A is the crack
area. When G exceeds a critical value G c , interfacial crack propagation occurs, leading to the detachment of
the coating from the substrate.
For a crack in a homogeneous isotropic material, G can be simply related to the fracture toughness K by
K2
G= , (2)
E
where E = E for plane stress and E = E/(1 2 ) for plane strain, with E being Youngs modulus and
Poissons ratio.
However, the situation is more complicated for a crack located along the interface between two dissimilar
materials, such as an interfacial crack between the coating and the substrate. Fracture at the interface of two
dissimilar materials often exhibits a mixed mode. The crack tends to stay at the interface even when it is
subjected to a combination of shear and normal stresses because the interface is frequently the weakest path
for fracture [16]. Moreover, the elastic property difference across an interface also tends to disrupt the symmetry
even when the geometry and loading are symmetric about the crack [16,17]. Consequently, a complex interfacial
stress intensity factor (SIF) K has to be used to characterize the stress field ahead of an interfacial crack tip
[17,18].
For the interfacial crack shown in Fig. 1, the singular stresses directly ahead of the crack tip (along the
direction of = 0) are given by [18,19]
K i
yy + ixy = x , (3)
2 x
K = K I + iK II , (4)
Mechanical methods for evaluating coating adhesion 433

Fig. 1 An interfacial crack between two dissimilar materials

where is related to the mismatch between the properties of materials 1 and 2:


 
1 1 + 2 1
= ln , (5)
2 2 + 1 2

where j ( j = 1, 2) denotes the shear modulus of each material; j = 3 4 j with j being Poissons ratio
of the material for a plane strain problem, while j = (3 j )/(1 + j ) for plane stress.
For an interfacial crack, G is related to the complex interfacial SIF K by [18,19]
 
E 1 + E 2  2
G= K  , (6)
2 cosh2 () E 1 E 2

where E j = E j /(1 2j ) for plane strain, while E j = E j for plane stress.


In addition to the stated energy release rate, another useful parameter for a crack in a bimaterial interface is
the loading phase angle , which is related to the ratio of the normal to shear stresses ahead of the interfacial
crack tip. The phase angle is defined as [19,20]
  
i
1 Im  K L 
= tan , (7)
Re K L i

where L is a reference length used to characterize the remote field and is an arbitrary length parameter; the
functions Im (.) and Re (.) take the imaginary and the real part of a complex variable, respectively.

3 Mechanical methods for adhesion evaluation of coatings

In practice, two different ways are adopted to evaluate coating adhesion. First, the coating adhesion can be
quantified by strength based on stress analysis, more commonly known as either bond strength or adhesion
strength. The bond or adhesion strength of a coating is defined as the maximum stress that the coating/substrate
interface can sustain before the separation of the coating and the substrate. Such strength can be measured by a
number of simple techniques such as the pull-off test [5]. Alternatively, based on fracture mechanics analysis,
the coating adhesion can be quantified by interfacial fracture toughness. The interfacial fracture toughness
is defined as the resistance to interfacial crack initiation and propagation, which can be related to the amount
of energy required to initiate or propagate the crack. Since the interfacial fracture toughness usually depends
on the mode mixity, it is impossible to characterize it as a single quantity like the bond or adhesion strength.
For the measurement of interfacial fracture toughness of coating/substrate systems, tests such as the double
cantilever beam, four-point bend, and Brazil-nut have been employed.
434 Z. Chen et al.

Fig. 2 Specimen configuration for DCB testing of coating adhesion, modified with permission from Ref. [12]

3.1 Sandwich specimen-based methods

Sandwich specimens have been devised and employed for experimental evaluation of coating adhesion, and the
theoretical basis of these methods was mainly developed by Suo and Hutchinson [16]. This type of specimen
is constructed by bonding the coating into two halves of a substrate material, and the coating thickness is
very small compared to the crack length and other length scales of the specimen. Therefore, the bulk of the
specimen is homogeneous except for a thin layer of coating. For a homogeneous material, classical fracture
mechanics applies, and some solutions are already available for the evaluation of the fracture toughness or
strain energy release rate. Since the sandwiched coating has a relatively small thickness, the complication due
to inhomogeneity introduced by this thin coating is generally ignored. As such, it is a common practice to use
the fracture toughness or strain energy release rate determined for the homogeneous specimen without coating
to characterize the interfacial crack in the presence of coating [16]. Thus, the sandwich specimen with a thin
layer of coating in the middle has been widely employed for the quantitative analysis of coating adhesion. In
this Section, the sandwich specimens to be discussed include the double cantilever beam (DCB) specimen,
four-point bend specimen, Brazil-nut specimen, and double cleavage drilled compression specimen.

3.1.1 DCB specimen

The DCB specimen for the evaluation of coating adhesion was first proposed and developed in [12,21]. In the
DCB test, the coating is sandwiched between two beams of substrate materials with a preexisting crack along
the coating/substrate interface, as shown in Fig. 2. Under tensile loading, the crack begins to propagate along
the interface; the propagation can be monitored by an optical microscopy, and the crack opening displacement
be measured by an extensometer placed on the specimen. The experiment is performed in displacement control,
and the applied load is measured by a load cell. In the case that the load-point displacement increases at a
constant rate, the peak load is associated with the onset of the interfacial crack growth and the load drop is
associated with the interfacial crack jump. After a stable interfacial crack growth occurs, the displacement
ramp is interrupted and the DCB specimen is allowed to return to its original undeformed state. Subsequently,
the specimen may be re-loaded again with the constant displacement rate. Several loading/unloading cycles
are performed until complete failure of the specimen.
The fracture morphology of the specimen can be examined by optical microscopy or scanning electron
microscopy. The failure modes of the coating/substrate system such as adhesive (interfacial) failure, cohesive
failure, or mixed failure can be determined through the observation of the fracture morphology. More impor-
tantly, as a fracture mechanics test, the DCB test can be used to quantitatively determine the interfacial strain
energy release rate G of the coating/substrate system. When the coating thickness is small compared with
the length scale of the overall geometry of the DCB specimen and the ratio of the crack length to the beam
Mechanical methods for evaluating coating adhesion 435

half-height is large, the existence of this thin coating layer can be ignored. As such, G can be determined by
the following equation [22]:
 
12P 2 a 2 2h 2
G= 1 + , (8)
Eh 3 bbn 3a
where P is the applied tensile load, a the crack length, E the elastic modulus of the beam ( E = E/(1 2 )
with E being Youngs modulus and Poissons ratio for a plane strain problem, while E = E for plane
stress), h the beam half-height, b the beam thickness, and bn the crack-front thickness. With G determined,
the interfacial SIF can be calculated by Eq. (6). The phase angle shift caused by the thin coating layer is small.
Hence, the DCB specimen is dominantly subjected to mode I loading unless the elastic mismatch between the
coating and the substrate is significantly large.
The DCB test measures the interface fracture energy at a phase angle of about 0 (nearly mode I). Hence,
the measured energy is closely related to the intrinsic interfacial adhesion energy due to a minimum plastic
dissipation and asperity effects under mode I loading. In addition, compared with other simple techniques
such as the pull-off test and single-edge-notched specimen where the interfacial failure is often catastrophic,
the DCB test has stable and controllable crack growth, thus allowing multiple fracture toughness readings
to be taken from a single specimen. Therefore, the DCB test has become a popular method for evaluating
the adhesion of coating/substrate systems [12,21,2326]. For example, Guo and Wang [24] measured the
critical strain energy release rate G c of plasma-sprayed ceramic coatings using this test. They found that for
the sprayed ZrO2 5%CaO coating, G c showed an increase tendency (from about 30 to 50 J/m2 ) when the
crack length increased. Furthermore, they explored the failure mechanisms of the coatings by the adhesion test
and correlated the coating failure with the specific spraying processes, aiding the coating design and process
optimization. Mroz et al. [25] employed the DCB test to measure the adhesion properties of coatings on
cutting tools, and consistent adhesion data were obtained from independent optical and compliance techniques
used to measure the debond length. Furthermore, the results from the DCB test exhibited the same trends as
those obtained from the traditional scratch test technique. The compliance of the DCB specimen is important
for determining the fracture toughness. Troczynski and Camire [26] proposed a compliance model of DCB
specimen with coating and verified it experimentally. It was found that the compliance of the DCB specimen
increased significantly when the thickness of the coating was larger than 1 % of that of the beam and its
Youngs modulus was smaller than 50 % of that of the beam. This indicates that thin stiff coatings have no
influence on the compliance of the DCB specimen, while thick soft coatings increase the DCB compliance
significantly. This proposed compliance model allows the precise determination of coating adhesion by the
DCB test. It is of particular importance for thick soft coatings because they modify the compliance of the DCB
specimen significantly.
Despite its widespread application for coating adhesion evaluation, the conventional DCB test has its
shortcoming that the crack length has to be measured for the determination of the interfacial strain energy
release rate. In some cases, such as in an environmental chamber or during cyclic loading, it is difficult or even
impossible to accurately measure the crack length optically by a traveling microscope.
Mostovoy et al. [27] proposed a modified contoured DCB specimen, as shown in Fig. 3. The contour of
the specimen is designed so that its compliance is linearly proportional to the crack length. According to the
linear elastic beam theory, the applied tensile load is the only measured parameter required for establishing
the energy release rate. Thus, this specimen design allows the critical strain energy release rate to be measured
simply using the critical fracture load without the necessity of measuring the crack length. In addition, this test
method does not require the introduction of a sharp crack in the specimen. Specimen loading could commence
with a blunt crack. Once the crack starts propagating, the average load is used to calculate the energy release
rate.
Later, Li et al. [28] suggested that for the convenience of specimen preparation the tapered DCB specimen
with machined flat surfaces is preferable to the contoured DCB specimen (Fig. 3). When the beam is machined
into straight surface instead of curved one, although the beam height differed slightly from that of the curved
beam, the compliance of the tapered DCB specimen still yields experimentally a reasonable linear relation with
the crack length. They measured the interfacial fracture toughness of plasma-sprayed Al2 O3 coatings using the
tapered DCB specimen. The measured mean values of G C ranged from 13 to 27 J/m2 for the Al2 O3 coatings
at different spray distances. These values agreed well with those obtained by the conventional DCB specimen
test [28]. These experimental results and the dependency of the fracture toughness on the microstructure of
plasma-sprayed Al2 O3 coatings confirmed the feasibility of the tapered DCB method for the measurement of
fracture toughness of thermal sprayed coatings [28,29].
436 Z. Chen et al.

Fig. 3 Contoured and tapered DCB specimens, modified with permission from Ref. [28]

In addition to the widespread applications in coating adhesion evaluation, the DCB specimen has also been
used in the adhesion evaluation of other bimaterial interfaces. For example, in the microelectronic field, Dai
et al. [30] quantified the interfacial adhesion performance of a typical electronic packaging structure, namely
the flip-chip-on-organic-substrate. Their study provided the necessary data and inputs in optimizing materials
selections and process conditions for improving interfacial adhesion of an electronic packaging system. Loo
et al. [31] also demonstrated that the DCB specimen could be employed for the interfacial fracture toughness
evaluation of the solder joint with good repeatability.
The DCB test is well suited for measuring mode I interfacial fracture toughness. Multiple fracture toughness
readings can be taken from a single DCB specimen, making the test preferable when the specimen is expensive
or scarce. This test has no stringent restrictions on the mechanical properties of materials used for the coatings
and substrates and thus can be applied to a wide range of coating/substrate systems.

3.1.2 Four-point bend specimen

The four-point bend specimen was developed by Charalambides et al. [32,33] for the determination of the
critical energy release rate at the bimaterial interface under approximately equal shear and normal displacement
conditions. In this technique, a central notch through one layer provides the site for interfacial crack initiation
with the specimen loaded under four-point bending.
This technique has many advantages, they are as follows: (a) there is a closed-form solution for the interfacial
strain energy release rate; (b) the interfacial strain energy release rate exhibits steady-state characteristics and
thus it is not critical to monitor the crack length accurately during testing; (c) the specimen preparation is
relatively simple since a specimen with a rectangular cross-section is usually used for testing; (d) no special
specimen gripping is required and specimen mounting and dismounting are fairly straightforward.
However, the specimen developed by Charalambides et al. [32,33] is not readily applicable to a strongly
adhering coating/substrate system. This is because the strain energy stored in the bent coating may be insuf-
ficient to advance the interfacial crack, considering the vertical cracking and segmentation of bent coatings
during the test [34,35]. In addition, a critical thickness is required to store the energy by the applied load for
crack propagation at the interface. This critical thickness is usually much larger than the thickness of normal
coatings. Therefore, Hofinger et al. [35] proposed a simple modification to the four-point bend specimen for
the evaluation of the interfacial fracture toughness of thin brittle strongly adhering coatings. These coatings
usually tend to be separated by vertical cracking during the four-point bend loading. The modification is to
introduce the bonding of a stiffener on the top of the thin brittle surface coating, as shown in Fig. 4. This
stiffening plate suppresses the segmentation of the brittle coating and increases the stored energy in the coating
and therefore the driving force for coating delamination.
The bulk of the sandwich four-point bend specimen shown in Fig. 4 is homogeneous except for the thin
coating and adhesive layers in the structure, and a pre-crack lies at the coating/substrate interface. Suo and
Hutchinson [16] suggested that the presence of an interlayer (coating) can be ignored when its thickness is
small compared with the crack length and other in-plane dimensions of the specimen. This allows the use of
the strain energy release rate G of a homogeneous specimen without the interlayer for the characterization
Mechanical methods for evaluating coating adhesion 437

Fig. 4 Geometry of sandwich four-point bend specimen, modified with permission from Ref. [35]

Fig. 5 Modified loading arrangement using a floating platen, modified with permission from Ref. [36]

of the interfacial crack. Therefore, the expression for G of a four-point bend specimen without a sandwiched
layer can be adopted to characterize G in the sandwich four-point bend specimen. Thus,

21P 2 L 2 (1 2 )
G= , (9)
16Eb2 h 3
where P is the total load; L is the spacing between the inner and outer loading lines; b and h are the width and
half-height of the specimen, respectively; E and are Youngs modulus and Poissons ratio of the substrate
material, respectively. Equation (9) applies to the propagation of interfacial cracks located within the inner
loading lines of the specimen.
In the four-point bend test, the interfacial cracks are assumed to propagate symmetrically with respect to the
central notch. However, such an ideal scenario is rarely achieved in actual experiments. Instead, asymmetric
crack propagation is more common. For asymmetric crack propagation, the calculated G may be incorrect if
symmetric crack advance is assumed. There are two main causes for the asymmetric crack propagation [36,37].
Firstly, for a very stiff four-point bend beam specimen under a small load, even a very small shape distortion of
the specimen (100 m) or misalignments of the test rig can generate significant loading asymmetries. This
will have a substantial effect on the load at which the interfacial cracking occurs [36]. Secondly, the presence
of imperfect pre-cracks or the change of material properties along the interface can also result in asymmetrical
crack propagation [37]. To avoid the asymmetry crack propagation, Howard et al. [36] designed a modified
bend loading arrangement with a floating platen, as shown in Fig. 5. This modified four-point bend rig is
self-aligned and flexible to use. It ensures that a uniform moment is applied to the specimen between the inner
loading rollers and makes possible the generation and propagation of two cracks at both sides of the specimen
symmetrically along the interface. This modified four-point bend fixture was applied by Wang and Siegmund
[37] to obtain improved fracture toughness results. They also developed equations for calculating the fracture
toughness for the case in which the two cracks are asymmetrical. The obtained two fracture toughness values
correspond to the two load plateaus caused by asymmetrical crack propagation.
The sandwich four-point bend test has been found to be one of the most useful and simple methods in the
evaluation of coating adhesion. Yamazaki et al. [38] evaluated the delamination resistance of a thermal barrier
coating system with this method. They found that the average critical energy release rate G C was independent
438 Z. Chen et al.

of the loading rate and the interface roughness. The measured G C increased from about 150 to 250 J/m2
with increasing aging time at 1,000 C. These results indicated that the sandwich four-point bend test was an
effective method to measure the delamination resistance of thermal barrier coatings and to evaluate the effects
of thermal aging on coating adhesion. A previous study by one of the present authors and his co-workers [13]
also indicated that a sandwich four-point bend specimen could be employed to evaluate the interfacial crack
growth resistance in plasma-sprayed thermal barrier coatings. Well-controlled, stable and measurable crack
propagation could be achieved, and the strain energy release rate G was found to increase with crack extension
until a steady value of about 170 J/m2 , i.e., the specimen exhibited a rising R-curve behavior during the test.
Through in situ observation of crack propagation and subsequent finite element analysis, it was concluded that
such a rising R-curve was caused by crack bridging in the wake zone, which shielded the crack tip from the
far-field loading. This rising R-curve means an increasing crack growth resistance in thermal barrier coatings,
and this feature is beneficial against coating segmentation and delamination during service.
The sandwich four-point bend specimen can also be employed for adhesion measurement of multilayered
thin-film structures and composite structures. For example, Gan et al. [39] quantified the interfacial adhe-
sion energy of multilayered thin-film structures. They successfully determined the adhesion energy of the
SiC/polyarylene ether (PAE) interface, and their measured value was 26.2 J/m2 higher than that of SiN/PAE
interface reported by others. Zou et al. [40] characterized the interfacial toughness of Si3 N4 /BN composites.
Compared with the conventional four-point bend test in which large specimens were usually adopted and
pre-cracks were created by three-point bending, their test was improved by adopting a sandwich specimen of
small size and introducing a weak crack path directly connecting to the interphase. Their measured results
showed that the interfacial toughness of Si3 N4 /BN composites increased from 37.16 to 117.76 J/m2 as the
volume fraction of the Si3 N4 interphase modifier increased. The improved specimen was found to be very
effective in measuring the interfacial toughness of Si3 N4 /BN composites.
Unlike the DCB test that mainly measures mode I interfacial fracture toughness, the sandwich four-point
bend test can be used to determine the mixed-mode interfacial fracture toughness. Furthermore, it can be
widely applied to many types of coatings, from soft flexible coatings to hard brittle coatings. It is not only
applicable to a bilayer system but also to a multilayer system. The main disadvantage lies in the complexity
of the preparation of pre-cracks.

3.1.3 Brazil-nut specimen

The Brazil-nut specimen was originally introduced by Carneiro in Brazil [41]. It was initially used for measuring
the tensile strength of concrete. Atkinson et al. [42] adopted this specimen to study the mixed-mode fracture
of homogeneous material. A uniform disk-shaped specimen containing a pre-crack along a diameter is loaded
in compression along another diameter at an angle to the crack plane, as shown in Fig. 6a. With different
orientation of the pre-crack with respect to the loading direction, this specimen provides a wide range of phase
angles. For example, for = 0 , the crack deformation is in mode I and for = 25 , the crack deformation
is in mode II. The real and imaginary parts of the stress intensity factor are given by [43]

P fI a
KI = , (10)
br

P f II a
K II = , (11)
br
where P is the load applied in compression, a is the semi-crack length, b and r are the disk thickness and
radius, respectively. The non-dimensional factors f I and f II are functions of the compression angle and the
crack length a. They are formulated in fitting polynomial forms in [42]. The energy release rate G can be
calculated from
1  2    P 2a
G= K I + K II2 = f I2 + f II2 , (12)
E Eb2 r 2
where E = E for plane stress and E = E/(1 2 ) for plane strain. The corresponding mode mixity is
given by
   
K II f II
= tan1 = tan1 . (13)
KI fI
Mechanical methods for evaluating coating adhesion 439

Fig. 6 a Brazil-nut specimen geometry, modified with permission from Ref. [42]. b Sandwich Brazil-nut specimen geometry,
modified with permission from Ref. [44]

Subsequently, Wang and Suo [44] developed a cylindrical sandwich specimen, namely the Brazil-nut-sandwich
or sandwich Brazil-nut. In this specimen, the coating is bonded in between two identical halves, and a pre-crack
is introduced into the interface (see Fig. 6b). Assuming that the thickness of the coating layer is much smaller
than the other characteristic dimensions of the specimen, the SIF solution for a homogeneous disk shown in
Fig. 6a can be used for a sandwich Brazil-nut specimen shown in Fig. 6b. In the test, the sandwich Brazil-nut
specimen can be loaded in pure mode I, pure mode II, or in any mode mixity between modes I and II. Thus, a
wealth of interface fracture information at different modes can be provided by a simple compression loading.
In addition, this specimen is compact, and the test does not require sophisticated loading fixtures.
The sandwich Brazil-nut specimen has been a valuable tool in practice for the study of interfacial fracture
behavior in bimaterials. Tong et al. [45] studied the interfacial fracture toughness of a synthetic bonecement
interface using the sandwich Brazil-nut specimen with polyurethane (cancellous bone analogous materials) as
substrates and bone cement as the interlayer. Selected loading angles were employed for a mixture of mode I
and mode II loading conditions. The solution for the critical strain energy release rate G C at the chosen phase
angle was obtained using finite element analysis. At 45 and | | < 18 , the measured values of G C
were 47 and 20 J/m2 on average, respectively. Their results successfully demonstrated the determination of
interfacial fracture toughness of cancellous bonecement interface using the sandwich Brazil-nut specimen.
Rahbar et al. [46] studied the mode mixity dependence of interfacial fracture toughness between marble and
adhesive. It was found that the measured interfacial fracture toughness increased with increasing mode mixity.
Tan et al. [47] studied the effects of silane on the adhesion and mixed-mode interfacial fracture between the
parylene layer and 316L stainless steel with a combined experimental and theoretical approach. Their results
showed that silane had a modest effect on the adhesion of parylene to 316L steel, and the measured interfacial
fracture energies by the Brazil-nut specimens were consistent with those theoretical predictions over a range
of mode mixities.
The Brazil-nut specimen has a key advantage that it can conveniently evaluate the interfacial fracture
toughness as function of the loading mode. This is because this specimen can be easily loaded in pure mode
I, pure mode II, or in any mode mixity between modes I and II. Furthermore, there is no stringent restriction
on mechanical properties of the materials used for the coatings and substrates. The main disadvantage is that
the interfacial crack tends to propagate in an unstable manner.

3.1.4 Double cleavage drilled compression specimen

The double cleavage drilled compression (DCDC) specimen was developed by Janssen [48]. The specimen is
a rectangular bar with a circular hole at the center, as shown in Fig. 7a. Under uniform axial compression, the
Poisson effect leads to a concentrated tensile stress at the crowns of the central hole. Once the tensile stress
reaches the fracture strength of the material under increased compression, a mode I crack is generated at each
crown. The generated crack subsequently propagates along the mid-plane of the specimen. Finally, it will be
arrested by frictionally constrained end conditions [4951].
The DCDC specimen provides a long stretch of stable crack growth [52]. This feature allows the crack to be
monitored by optical microscopy, thus making the in situ observation and video recordings of crack extension
440 Z. Chen et al.

Fig. 7 a DCDC specimen geometry, modified with permission from Ref. [51]. b Sandwich DCDC specimen geometry, modified
with permission from Ref. [56]

possible. Other experimental advantages of the DCDC specimen include the experimental simplicity of axial
compression, auto-pre-cracking, and controllable crack growth [53]. Although this specimen is usually used
for the determination of mode I fracture toughness, a wide range of mode mixities (from = 0 to 64 )
can be achieved by varying the locations of the hole and longitudinal crack (Fig. 7a) [54,55]. An analytical
solution was also developed by He et al. [54] for calculating the strain energy release rate G under mode I
loading condition:

  2
2 R W W a
G= + 0.235 0.259 , (14)
E R R R

where is the applied compressive stress; E is the elastic modulus ( E = E/(1 2 ) with E being Youngs
modulus and Poissons ratio for a plane strain problem, while E = E for plane stress); R is the hole radius;
W is the half-width of the specimen; a is the crack length. Equation (14) is applicable when the specimen
geometries are varied within the ranges 2 W/R 4 and W/R a/R 15.
The DCDC specimen can be extended to study interface fracture by sandwiching a thin layer of coating
or film between the two beams of the substrate material (Fig. 7b). The interfacial crack has almost the same
energy release rate as that for the monolithic material as long as the sandwiched layer is thin compared with
the in-plane dimensions of the specimen. Therefore, the interfacial energy release rate of a sandwich DCDC
specimen can be evaluated using Eq. (14). A notable advantage of this specimen for the study of interfacial
fracture is that the cracks tend to propagate along the interface even when the interfacial fracture toughness
significantly exceeds the toughness of the adjoining materials [56]. During the test, when cracks attempt to
propagate into the adjoining materials rather than being confined at the interface, cracks are subjected to a
substantial mode II loading; this loading will deflect the cracks back toward the interface. Thus, a strong
preference for interface fracture is achieved in the DCDC specimen.
The sandwich DCDC specimen was originally used by Turner et al. [56]. Their specimen was validated
by a model system consisting of a thin polymer layer between two glass beams. Their experimental results
demonstrated the unique capability of the DCDC specimen for measuring mode I interfacial fracture toughness.
Subsequently, Turner and Evans [57] investigated the mechanism of mode I crack propagation along the
Au/sapphire interface in both humid and dry environments. Through in situ experimental observations, the
intrinsic mechanisms that allow debonding to occur along the oxide/metal interfaces had been identified. Lipkin
et al. [58] investigated the effect of interfacial carbon on adhesion and toughness of gold/sapphire interfaces.
Mechanical methods for evaluating coating adhesion 441

The full details of the crack extension along the interface were recorded by a real-time video recording
system and a linear variable differential transformer for position tracking. The variation in each quantity with
interfacial carbon was independently measured, and the possible origins of the observed sensitivity of adhesion
to carbon heat treatment were discussed. Results showed that the adhesion was reduced by the carbon heat
treatment. Ritter et al. [59] employed the sandwich DCDC specimen to measure the critical energy release rate,
the moisture-assisted crack growth, and the fatigue threshold of epoxy/glass interfaces. They discussed the
mechanisms of moisture-assisted crack growth at different epoxy/glass interfaces, based on the testing results
of epoxy/glass interfaces.
A unique and notable advantage of the sandwich DCDC specimen is that the crack always tends to propagate
along the interface, making this specimen a most suitable choice when the interfacial fracture toughness
exceeds the fracture toughness of the adjoining materials. In addition, the feature that the specimen can be
automatically pre-cracked is attractive as it greatly simplifies the experimental requirements. There is no
particular requirement for the mechanical properties of the materials used for the coatings and substrates. This
specimen is usually used for the determination of mode I interfacial fracture toughness.
In summary, the reviewed sandwich specimens can be employed to perform a quantitative analysis of
coating adhesion using fracture mechanics. Such analysis is important not only for understanding the fac-
tors contributing to coating adhesion, but also for further numerical simulations and lifetime predictions.
However, the sandwich specimens still have some shortcomings such as relatively sophisticated experimental
techniques and complicated specimen preparation with possible undesirable effects on experimental results.
The sandwich specimen is usually made by sandwiching a thin layer of coating into the test structure. This
is typically done through diffusion bonding, which may alter both the coating microstructure and interfa-
cial adhesion. This is because the bonding process usually takes a long time and sometimes at a relatively
high temperature. For example, Yamazaki et al. [38] used a heat curing type of epoxy adhesive to sand-
wich the thermal barrier coatings into two steel beams and observed that the adhesive infiltrated into the
top coating for about 100 m. Thus, sandwich specimens should not be employed to coatings with inter-
connected pores or coatings that are temperature-sensitive. In addition, sandwich specimens cannot satis-
factorily simulate the stress/strain loading environment that real manufactured parts must endure. The sand-
wich structure could well present a different stress state at the interface from the expected service stress
state.

3.2 Bimaterial specimen-based methods

The adhesion measurement techniques based on bimaterial specimens are relatively simple to perform and
could provide good qualitative or quantitative information. Furthermore, the adhesion evaluation of coatings
with the bimaterial specimens does not usually involve the attachment of new devices to the coating. This avoids
the possible epoxy penetration into the coating and thus minimizes the distortion on interfacial adhesion as
compared to what will be expected in service. In this Section, the bimaterial specimen-based test methods
discussed include indentation, scratch, peel, blister, and tensile tests.

3.2.1 Indentation test

Microindentation and nanoindentation techniques have commonly been used for measuring the mechanical
properties of bulk materials, such as hardness, Youngs modulus, and fracture toughness [6062]. Recently,
these techniques have been further developed and used for evaluating the adhesion strength of coatings or films
to the substrate [6366]. An indentation test for coating adhesion evaluation can be performed in two ways: the
indentation perpendicular to the coating surface (Fig. 8) and the indentation perpendicular to the cross-section
of the coating/substrate system (Fig. 10).
When the indentation test is performed at the coating surface (Fig. 8), the applied load drives the indenter
into the coating/substrate system and induces a compressive radial stress in the coating, which provides the
driving force for coating detachment [67]. According to Drory and Hutchinson [63], three types of coating
delamination have been observed during the indentation test (Fig. 9) such as : (a) delamination leaving behind
a very narrow annular plate of coating; (b) delamination with an unbuckled annular plate of coating left
behind the crack tip; and (c) delamination with a buckled annular plate of coating left behind the crack tip.
They developed solutions to evaluate the interfacial energy release rate from the measurement of the applied
indentation load, delamination radius, coating thickness, and coating/substrate material properties. Considering
the most common model shown in Fig. 9b as an example, the interfacial energy release rate value is given by
442 Z. Chen et al.

Fig. 8 Schematic of indentation test at the surface of a coating/substrate system

Fig. 9 Coating delamination caused by indentation, modified with permission from Ref. [63]

(1 c2 )h
G= [r (R) r (R )]2 , (15)
2E c
where vc , h, and E c are Poissons ratio, thickness, and Youngs modulus of the coating, respectively, and r (R)
and r (R ) are the radial stress components at the inner and outer edges of the annular plate, respectively.
The indentation test at the surface of a coating/substrate system has demonstrated to be a relatively simple
and economic measurement method for providing a good estimate of interfacial adhesion. For example, Xie
and Hawthorne [68] measured the adhesion of solgel-derived coatings to a 316 stainless steel substrate and
found that this method is sensitive enough to differentiate the adhesions of the coatings processed by different
conditions. The interfacial fracture toughness of the coatings studied ranged between 25 and 73 J/m2 for
various coating processing temperatures and substrate treatment methods. However, it should be pointed out
that this type of indentation technique is mainly applied to the systems of brittle coatings weakly bonded to
ductile substrates. A ductile and strongly adhered coating often deforms before delaminating from the substrate.
Furthermore, the calculation of the radial stress in Eq. (15) is rather complicated due to the complexity of the
triaxial stress state of the indentation, the different elastic and plastic properties of the coating and substrate,
and the existence of the interface.
The indentation test performed perpendicularly to the cross-section of the coating/substrate system (Fig. 10),
which is called the interfacial indentation test, was developed by Chicot et al. [64]. In this test, a mechanically
stable crack is introduced into the coating/substrate interface using the conventional Vickers indentation [64
66]. One diagonal line of the Vickers indenter is first aligned along the coating/substrate interface and then an
increasing indentation load is applied. With the penetration of the indenter into the coating/substrate system,
a crack with a semicircular shape is generated and localized in the interface plane. Subsequently, the critical
Mechanical methods for evaluating coating adhesion 443

Fig. 10 The interfacial indentation test, modified with permission from Ref. [64]

load Pc required to initiate the interface debonding and the measured interfacial crack length ac is used to
calculate the apparent interfacial fracture toughness [69],
 1/2
Pc E
K = 0.015 3/2 . (16)
ac H I
In Eq. (16), the ratio (E/H )I characterizes the global behavior of the coating/substrate system and is defined
by the following equation:
 1/2  E 1/2  E 1/2
E H S H C
= 1/2 + 1/2 , (17)
H I HS
1 + HC 1+ H C
HS

where E is Youngs modulus, H is the hardness, and the subscripts I, S, and C stand for interface, substrate,
and coating, respectively.
The interfacial indentation test offers a technique for quick measurement since only simple experimental
manipulation and short data acquisition time are required. Furthermore, the measured interface fracture resis-
tance is uniquely related to the interfacial bonding because the load is applied directly on the interface. Marcano
et al. [70] studied the adhesion of vacuum plasma-sprayed Cr3 C2 -NiCr coatings to plain carbon steel. They
found that the interfacial toughness increased with coating thickness, achieving values from 7.6 to 10.1 MPa
m1/2 when the thickness increased from 150 to 450 m. Similarly, Sadeghi-Fadaki et al. [71] investigated
the interfacial fracture toughness of plasma-sprayed yttria-stabilized zirconia (YSZ) coatings. The adhesion
behaviors of the YSZ coatings were found to be related to the microstructural characteristics of powders and
coatings.
However, for this method, there are difficulties in interfacial crack length measurement: (a) the crack
nucleated may deviate across the coating instead of being confined at the interface, making an accurate
measurement of the crack length difficult; (b) the existence of complex microstructure around the interface,
e.g., pores and intersplat microcracks in thermal sprayed coatings, hinders the optical observation of the
interface and the accurate determination of the crack length [65]. Furthermore, the coating thickness in the
interfacial indentation test usually has to be higher than 100150 m to avoid the occurrence of cracking
within the coating instead of being localized at the interface. However, this coating thickness requirement has
been reduced with the widespread use of nanoindentation in scientific and technical laboratories in the recent
years. The nanoindentation technique introduces a highly localized interface debonding and thus allows the
measurement of adhesion strength of a thin coating with this test [72].
The advantages of the indentation test include simplicity of operation, small quantity of specimen materials,
and ready availability of commercial equipments. As such, this test is easily implemented both in the laboratory
and in the manufacturing line. The conventional indentation test is often used for the adhesion measurement
of a brittle coating on a ductile substrate, while the interfacial indentation test has a wider applicability to a
variety of coating/substrate systems. Although the interfacial indentation test requires careful preparation of
the specimen cross-section, it is especially suitable for the adhesion measurement of thick coatings. The main
limitation of the interfacial indentation test lies in the difficulty of crack length measurement.

3.2.2 Scratch test

The scratch test has been widely used as a technique for the evaluation of coating adhesion since its invention
by Heavens in 1950 [73]. In this method, a spherically tipped diamond indenter is drawn across the coated
444 Z. Chen et al.

Fig. 11 Schematic of scratch adhesion test, modified with permission from Ref. [68]

surface with the applied normal load to the indenter increasing continuously or stepwise until the detachment
of the coating occurs [74], as illustrated in Fig. 11. The normal force required to produce coating detachment,
also called the critical load, is considered as a qualitative measure of the coating/substrate adhesion and
can be used to compare the adherence of coating/substrate systems of similar nature. Depending on the
coating/substrate system and the scratch tester, the coating detachment and the corresponding critical load can
be determined by optical microcopy, scanning electron microscopy, or acoustic emission detectors [75]. In
addition, in situ measurement of tangential friction and normal forces has also proved to be a reliable means
of coating detachment detection [76]. Generally speaking, the acoustic emission is often used to characterize
the occurrence of fragile coating failure; the in situ measurement is often used to determine the critical failure
load of ductile coatings; the microscopic observation and analysis are suitable for the failure observation of
all coating/substrate systems [7577].
A key consideration of the scratch test for coating adhesion evaluation is to establish correct relation-
ship between the measured critical load and the actual adhesion strength. In the scratch test, there is a range
of different failure modes of coatings [78]. The most common and important failure mode is the coating
detachment ahead of the indenter. Such adhesive failure of the coating is necessary for coating adhesion eval-
uation. This failure occurs when the induced compressive stress in the coating exceeds a critical value [79].
However, the normal load applied to the indenter not only causes adhesive failure of the coating but also
its cohesive failure [80]. Other failure modes (e.g., transverse cracking that occurs when the coating is still
adhered to the substrate) have also been observed [78]. Therefore, various non-adhesive failures may occur
before or along with adhesive failure during the scratch test. If the scratch test is to be used for the assess-
ment of coating/substrate adhesion, adhesion-related failure modes are the most important. In addition to the
failure modes, the critical load is also strongly affected by several parameters related to testing conditions
and the coating/substrate system [75]. For testing conditions, common factors affecting the measurements
include scratching speed, loading rate, indenter tip radius, and indenter wear; for the coating/substrate sys-
tem, the factors include substrate hardness, coating thickness, friction coefficient, and residual stress in the
coating.
The mechanics of the scratch test was first analyzed by Benjamin and Weaver [81]. They used the plasticity
theory to arrive at an expression for the critical shearing force for coating removal. However, their analysis is
insufficient for most practical coating systems because the actual process of coating removal during the scratch
test is more complex than the general assumption adopted in their analysis. Later, Laugier [82] proposed that
the adhesion behavior can be modeled in terms of the strain energy released during coating removal. In his
model, the compressive stressed coating in the region ahead of the indenter is considered to reduce its stored
energy by detaching from the substrate at the critical applied load. This energy description is applicable to both
ductile and brittle coatings, and it is not necessary to consider the details of coating removal in determining
the adhesion energies provided that the work for ductile deformation or brittle fragmentation of the coatings is
small. More recently, Xie and Hawthorne [79,80] developed a mathematical model to calculate the distribution
of compressive stresses in a thin coating induced by the scratch indenter. They derived a simple but useful
equation from the model to rank the adhesion of different coating systems.
Nowadays, the scratch adhesion testing has increasingly been used in engineering and research. For exam-
ple, Jaworski et al. [83] characterized the adhesion of suspension plasma-sprayed TiO2 coatings using this
test. The critical force necessary to peel off the coatings from the substrate was determined and considered
as a measure of the coating/substrate adhesion. Their measured results indicated that the spray distance has
Mechanical methods for evaluating coating adhesion 445

Fig. 12 Schematic drawing of the 90 peel test around a mandrel, modified with permission from Ref. [89]

a significant effect on the adhesion of coatings. Similarly, Barnes et al. [84] used the scratch test to measure
the adhesion strength of calcium phosphate coatings applied to poly (carbonate urethane) substrates. They
demonstrated that the scratch test is an effective technique for obtaining comparative, rather than absolute,
values of adhesion strength for hard coatings on a compliant substrate.
Like the indentation test, the scratch test is fairly reliable, easy to use, and no special specimen shape or
preparation is required. This test is most often used for measuring the adhesion of a thin, hard coating to a
relatively soft substrate. The main challenge is to interpret the value of the critical load in terms of adhesion
because this value is affected by a number of factors that are not adhesion-related [85]. Thus, the measured
critical load is only a semiquantitative measure of the coating/substrate adhesion. A careful analysis of the
failure modes of coatings is necessary and important for a proper understanding of the implication of the test
results.

3.2.3 Peel test

The peel test is originated from the so-called tape test [86]. In the simplest version, a piece of adhesive tape is
pressed against a film or coating and then removed to observe if the coating is peeled off. Through the years,
several standard peeling methods were developed, such as multiangle fixed arm peel, T-peel, and roller-assisted
peel [8790]. Figure 12 illustrates the common 90 peeling around a mandrel and it is suitable for flexible
coatings on rigid substrates [89]. Indeed, the peel test can be performed at any angle between 0 and 180 ,
and the varying peel angles are used to investigate the effect of mode mixity on the peel strength. This is
important because the adhesion strength of many coatings is sensitive to mode mixity. In the peel test, the
interfacial crack propagates in a stable manner, and the peel force versus displacement curve is recorded to
represent the adhesion of the coating. However, the applied peel force not only separates the coating from
the substrate, but also plastically deforms the coating and overcomes friction. Therefore, understanding and
analyzing this combination are essential for a proper evaluation of the adhesion strength of coatings. The peel
force is significantly affected by the peel parameters. Nase et al. [91] have showed that the peel force increased
exponentially with an increase in the peel rate. In addition, adhesion strengths at different positions can be
measured when the crack continuously propagates along the interface. Therefore, this test can exhibit the effect
of the inhomogeneity of the substrate surface on coating adhesion. Horgnies et al. [92] used this test to study
the influence of the surface properties of concrete on the adhesion of polyurea coatings. It was demonstrated
that the adhesion of the polyurea coatings is mainly conditioned by the mechanical anchorage at the concrete
surface.
The principal limitation of the peel test is that it is only suitable for tough and flexible coatings. However,
brittle coatings such as thermal sprayed ceramic coatings are increasingly used in engineering and research.
Sexsmith and Troczynski [93] had developed a novel peel test to allow adhesion evaluation of brittle coatings.
In this test, a thin metal foil is attached to a massive copper block. The surface of the metal foil is partly sand-
blasted, as shown in Fig. 13a. Subsequently, the coating is deposited onto the metal foil. As such, the copper
block underneath provides the necessary mechanical support and also serves as a heat sink (Fig. 13b). The
assembly of the block, foil, and coating is glued to a stiff aluminum plate using a strong epoxy, and the copper
block is subsequently removed. The metal foil, which serves as the substrate in the remaining coating/substrate
system, is then peeled from the coating at a constant speed, causing a crack to propagate precisely along the
interface (Fig. 13c). With this arrangement, the ductile substrate rather than the brittle coating is being peeled
in the test. Sexsmith and Troczynski [93] demonstrated the application of this modified peel test for adhesion
446 Z. Chen et al.

Fig. 13 Schematic of the novel peel test, modified with permission from Ref. [93]

measurement of brittle thermal sprayed coatings. Kurzweg and Heimann [94] also employed this modified peel
test to evaluate the adhesion of thermally sprayed hydroxyapatite (HAP) coatings to the Ti6Al4V substrate.
They found that an appropriate bond interlayer between the substrate and the HAP coatings improved the
adhesion of HAP coatings to the metallic substrate.
The most important advantage of the peel test is that it allows the study of adhesion as a function of the
interfacial crack position. Thus, this test can be used to investigate the inhomogeneity effect of the substrate
surface on coating adhesion. Other advantages include simplicity, minimal cost, and controllable crack propa-
gation. The main limitation of the conventional peel test is that it is only suitable for tough and flexible coatings.
Although the discussed novel peel test here can be used for brittle ceramic coatings, it requires a very com-
plex specimen preparation. A major difficulty of both peel tests is that the majority of the mechanical energy
consumed in the tests serves to bend and deform the specimen, and contributes relatively little to interface
fracture. Consequently, it is difficult to extract adhesion energy from peel test data.

3.2.4 Blister test

The blister test was initially proposed by Dannenberg [95] to measure the adhesion of organic coatings to
metals. In a standard blister test, the testing geometry consists of an overlying coating against a rigid substrate
with a circular hole constructed at the center, as shown in Fig. 14a. Either liquid or gas is pressurized through
the hole against the underside of the coating, causing the coating to debond from the substrate and form a blister
[96]. The coating/substrate interfacial adhesion energy can be quantitatively determined from the geometry of
the blister and the applied pressure.
The main limitation of the standard pressurized blister test is that the coating may rupture prior to debonding
from the substrate. Furthermore, catastrophic debonding often occurs once a critical pressure is reached.
Therefore, many improvements and refinements have been made. Chang et al. [97] and Napolitano et al. [98]
developed a constrained blister test to limit the vertical displacement and rupture of the coating by placing
a flat constraint above the blister, achieving a nearly constant interfacial strain energy release rate. Allen
and Senturia [99] developed an island blister test to reduce stresses on the coating. Dillard and Bao [100]
Mechanical methods for evaluating coating adhesion 447

Fig. 14 a Schematic of the pressurized blister test, modified with permission from Ref. [96]. b Schematic of the shaft-loaded
blister test, modified with permission from Ref. [104]

later proposed a peninsular blister test, which has a steadier delamination front than the island blister test. A
significant improvement was made by Wan and his co-workers [101103]. They developed a unique blister
test driven by a fixed amount of gas trapped at the coating/substrate interface. Unlike the conventional blister
test loaded by constant pressure, which usually leads to catastrophic crack propagation, this improved blister
test is driven by an internal thermal expansion of trapped gas for achieving stable crack growth. Furthermore,
this novel test allows coating adhesion investigation at an elevated temperature.
Despite the above-mentioned improvements, the application of these pressurized blister tests is limited
because they usually require a complicated pressure control system and a sophisticated experimental setup
to continuously monitor changes in pressure and blister dimension. Therefore, a more convenient method
named the shaft-loaded blister test, first proposed by Malyshev and Salganik [104], has attracted attention in
recent years. In this version of blister test, the mechanical load is applied to the center of the overhanging
coating with a spherically capped shaft to drive an axisymmetric conical delamination at the coating/substrate
interface, as shown in Fig. 14b. The shaft-loaded test is an attractive alternative to pressurized tests due to
its simpler experimental setup. Furthermore, the load-shaft displacement and the blister central deflection are
easily obtained and controlled. For this test, Wan and Mai [105,106] have developed an analytic solution to
derive the strain energy release rate for a pure stretching film or coating in a conical geometry. During stable
crack growth, the applied load P, the central deflection w0 , and the radius of debonding a are predicted to
increase linearly. From the slope of P versus w0 , the strain energy release rate G can be calculated by
 2
1 P
G= 2 (18)
Eh w0
where E and h are elastic modulus and thickness of the coating, respectively. Alternatively, G also can be
calculated from the slope of P versus a:
 1/3  4/3
1 P
G= , (19)
16 4 Eh a
or from the slope of w0 versus a:
Eh w0 4
G= . (20)
16 a
The blister test has found increasing use in engineering and research. For instance, OBrien et al. [107] used the
shaft-loaded blister test to measure interfacial adhesion energy G of a pressure-sensitive adhesive tape bonded
to a rigid substrate. They calculated G using Eqs. (18)(20), respectively. It was found that using Eq. (19)
accurate values of G were obtained for a thin adhesive coating independent of the coating stiffness and plastic
deformation at the shaft tip. Recently, Na et al. [108] used the same test method to measure the adhesion
work of an electrospun polymer membrane on a commercial cardboard substrate. The average adhesion work
448 Z. Chen et al.

was found to be 206 26 mJ/m2 . Their work shows that the shaft-loaded blister test is a viable method for
evaluating the adhesion energy of electrospun polymer fabrics.
Wang [109] used the pressurized blister test to measure the interfacial adhesion energy G of an elastic
tape bonded to a poly(methyl methacrylate) substrate. The effect of the volume flow rate of the injected fluid
was also investigated. It was found that G was about 3.0 0.5 J/m2 and the flow rate of the injected fluid did
not have a pronounced effect on the measured interfacial strength. Koenig et al. [110] measured the adhesion
energy of graphene sheets with a silicon oxide substrate by the novel pressurized blister test developed by Wan
and his co-workers [101103]. They found the adhesion energy of 0.45 0.02 J/m2 for monolayer graphene
and 0.31 0.03J/m2 for specimens containing two to five graphene sheets. They attributed such a strong
adhesion to the extreme flexibility of graphene. This flexibility allows graphene to conform to the topography
of the very smooth substrates.
The notable advantages of the blister test include: (a) the axisymmetric fracture surface minimizes the
effect of specimen non-uniformity; (b) the small detachment angle minimizes the dissipative effects of the
coating; (c) a fully quantitative analysis of interfacial adhesion energy is available; (d) testing the environmental
degradation of the coating/substrate system is possible because the pressurizing medium can be chosen as the
degrading fluid or gas. The blister test is usually limited to soft flexible polymeric or metallic coatings because
hard brittle coatings will tend to crack before forming a blister. The main disadvantages are the relatively
complicated specimen preparation and high experimental requirements.

3.2.5 Tensile test

A tensile test with circumferentially notched specimen was recently proposed and developed by Elambasseril
and Ibrahim [111] to determine the interfacial fracture toughness of coatings. The circumferentially notched
tensile (CNT) specimen is of cylindrical shape with the notch machined in the middle, as illustrated in Fig. 15.
The notch angle can be varied to give the interfacial crack various orientations with respect to the tensile load,
thus making it possible to investigate the mixed-mode failure mechanisms. Prior to the coating deposition,
the cylindrical substrate is polished, and then a thin layer of gold is electroplated on the masked substrate
to form a pre-crack at different locations on the notch. Next, the coating is deposited on the middle part of
the U-notched cylindrical substrate (Fig. 15). Afterward, a uniaxial tensile load is applied on the coated CNT
specimen until the failure of the coating. The corresponding applied force is recorded as the critical tensile load
and determined by comparing the acoustic events due to the interfacial fracture and the loaddisplacement
curves. Finally, the measured critical load can be applied to the finite element model to calculate the critical
J integral JC . The complex critical SIF K C can be evaluated by

 
 K  =  E JC  . (21)
C
1 2

In Eq. (21), and E are given as


1 (2 1) 2 (1 1) 2E 1 E 2
= , E = , (22)
1 (2 + 1) + 2 (1 + 1) E1 + E2
where j ( j = 1, 2 designating
  materials
  1 and 2) is the material shear modulus and E j is Youngs modulus.
For plane stress, j = 3 j / 1 + j with j being Poissons ratio, while for plane strain j = 3 4 j .
As a newly developed method, the CNT technique is attractive because it is simple, repeatable, easy for
modeling, and widely applicable to various types of coatings. The CNT specimen can simulate the service
conditions of coatings more closely than the sandwich specimen. Furthermore, the CNT technique can be used
to investigate the mixed-mode failure mechanism by varying the notch angle. For example, Elambasseril and

Fig. 15 Schematic of the tensile test with circumferentially notched specimen, modified with permission from Ref. [111]
Mechanical methods for evaluating coating adhesion 449

Ibrahim [112,113] performed a quantitative evaluation of the adhesion of nickel coatings to a mild steel sub-
strate. Their measured interfacial fracture toughness exhibited good agreement with those obtained by others.
Nevertheless, it should be pointed out that this technique still has some limitations. Firstly, the determination
of the critical load is challenging. An acoustic emission system is needed during the test, but it is difficult to
collect the acoustic signals for the analysis because both the cohesive failure in the coating and the adhesive
failure at the interface may occur. Secondly, the calculation of interfacial fracture toughness has to rely on
finite element modeling. This could be tedious and thus an impediment for the technique implementation.
Thirdly, in situ observation of interfacial crack initiation and propagation is almost impossible during the test.
Therefore, further improvement to address these deficiencies will be desirable.

4 Conclusions

The currently widely used mechanical methods for evaluating coating adhesion are discussed. Their advantages
and limitations have been reviewed from the perspective of interfacial fracture mechanics, their underlying
principles, typical applications, and specimen preparations. This review shows that there is no single method that
is totally satisfactory and/or of general applicability for all coating/substrate systems. Each method has its own
merits and limitations, depending on the targeted applications and the characteristics of the coating/substrate
system. By properly selecting the testing method or a combination of methods, useful and reliable data could
be obtained. Indeed, the deficiencies highlighted by this review indicate that there are rooms for improvement
for the development of more efficient, reliable, and widely applicable methods for evaluating coating adhesion.

Acknowledgments The authors acknowledge the support by the Agency for Science, Technology and Research (A*STAR),
Singapore (Grant No: SERC 1123004027).

References

1. Padture, N.P., Gell, M., Jordan, E.H.: Thermal barrier coatings for gas-turbine engine applications. Science 296,
280284 (2002)
2. Lima, R.S., Marple, B.R.: Thermal spray coatings engineered from nanostructured ceramic agglomerated powders for
structural, thermal barrier and biomedical applications: a review. J. Therm. Spray Technol. 16, 4063 (2007)
3. Wood, R.J.K.: Tribo-corrosion of coatings: a review. J. Phys. D: Appl. Phys. 40, 55025521 (2007)
4. PalDey, S., Deevi, S.C.: Single layer and multilayer wear resistant coatings of (Ti, Al) N: a review. Mater. Sci. Eng.
A 342, 5879 (2003)
5. Valli, J.: A review of adhesion test methods for thin hard coatings. J. Vac. Sci. Technol. A 4, 30073014 (1986)
6. Mittal, K.L.: Adhesion measurement of thin films. Electrocompon. Sci. Technol. 3, 2142 (1976)
7. Lin, C.K., Berndt, C.C.: Measurement and analysis of adhesion strength for thermally sprayed coatings. J. Therm. Spray
Technol. 3, 75104 (1994)
8. Hull, T.R., Colligon, J.S., Hill, A.E.: Measurement of thin film adhesion. Vacuum 37, 327330 (1987)
9. Mellali, M., Fauchais, P., Grimaud, A.: Influence of substrate roughness and temperature on the adhesion/cohesion of
alumina coatings. Surf. Coat. Technol. 81, 275286 (1996)
10. Singh, R.K., Gilbert, D.R., Fitz-Gerald, J., Harkness, S., Lee, D.G.: Engineered interfaces for adherent diamond coatings
on large thermal-expansion coefficient mismatched substrates. Science 272, 393398 (1996)
11. Takadoum, J., Bennani, H.H.: Influence of substrate roughness and coating thickness on adhesion, friction and wear of
TiN films. Surf. Coat. Technol. 96, 272282 (1997)
12. Ostojic, P., McPherson, R.: Determining the critical strain energy release rate of plasma-sprayed coatings using a double-
cantilever-beam technique. J. Am. Ceram. Soc. 71, 891899 (1988)
13. Chen, Z.X., Qian, L.H., Zhu, S.J.: Determination and analysis of crack growth resistance in plasma-sprayed thermal barrier
coatings. Eng. Fract. Mech. 77, 21362144 (2010)
14. Tong, J., Wong, K.Y., Lupton, C.: Determination of interfacial fracture toughness of bone-cement interface using sandwich
Brazilian disks. Eng. Fract. Mech. 74, 19041916 (2007)
15. Zhang, J., Lewandowski, J.J.: Interfacial fracture toughness measurement using indentation. J. Mater. Sci. 29,
40224026 (1994)
16. Suo, Z., Hutchinson, J.W.: Sandwich test specimens for measuring interface crack toughness. Mater. Sci. Eng. A 107,
135143 (1989)
17. Rice, J.R., Sih, G.C.: Plane problems of cracks in dissimilar media. J. Appl. Mech. 32, 418423 (1965)
18. Rice, J.R.: Elastic fracture mechanics concepts for interfacial cracks. J. Appl. Mech. 55, 98103 (1988)
19. Klingbeil, N.W., Beuth, J.L.: Interfacial fracture testing of deposited metal layers under four-point bending. Eng. Fract.
Mech. 56, 113126 (1997)
20. Hutchinson, J.W., Suo, Z.: Mixed-mode cracking in layered materials. Adv. Appl. Mech. 29, 63191 (1992)
21. Berndt, C.C., McPherson, R.: A fracture mechanics approach to the adhesion of flame and plasma sprayed coatings. Trans.
Inst. Eng. Aust. 6, 5358 (1981)
450 Z. Chen et al.

22. Cannon, R.M., Dalgleish, B.J., Dauskardt, R.H., Oh, T.S., Ritchie, R.O.: Cyclic fatigue-crack propagation along
ceramic/metal interfaces. Acta Metall. Mater. 39, 21452156 (1991)
23. Heintze, G.N., McPherson, R.: Fracture toughness of plasma-sprayed zirconia coatings. Surf. Coat. Technol. 34,
1523 (1988)
24. Guo, D.Z., Wang, L.J.: Measurement of the critical strain energy release rate of plasma-sprayed coatings. Surf. Coat.
Technol. 56, 1925 (1992)
25. Mroz, J., Daukardt, R.H., Schleinkofer, U.: New adhesion measurement technique for coated cutting tool materials. Int. J.
Refract. Met. Hard Mater. 16, 395402 (1998)
26. Troczynski, T., Camire, J.: On use of double cantilever beam for coatings and adhesion tests. Eng. Fract. Mech. 51,
327332 (1995)
27. Mostovoy, S., Crosley, P.B., Ripling, E.J.: Use of crack-line-loaded specimens for measuring plane-strain fracture tough-
ness. J. Mater. 2, 661681 (1967)
28. Li, C.J., Wang, W.Z., He, Y.: Measurement of fracture toughness of plasma-sprayed Al2 O3 coatings using a tapered double
cantilever beam method. J. Am. Ceram. Soc. 86, 14371439 (2003)
29. Wang, W.Z., Li, C.J., Ye, F.X.: Effect of specimen geometry on fracture toughness measurement of plasma-sprayed ceramic
coatings by the tapered double cantilever beam approach. Vacuum 73, 649654 (2004)
30. Dai, X., Brillhart, M.V., Ho, P.S.: Adhesion measurement for electronic packaging applications using double cantilever
beam method. IEEE Trans. Compon. Packag. Technol. 23, 101116 (2000)
31. Loo, S.Z.Y., Lee, P.C., Lim, Z.X., Yantara, N., Tee, T.Y., Tan, C.M., Chen, Z.: Interface fracture toughness assessment of
solder joints using double cantilever beam test. Int. J. Mod. Phys. B. 24, 164174 (2010)
32. Charalambides, P.G., Lund, J., Evans, A.G., McMeeking, R.M.: A test specimen for determining the fracture resistance of
bimaterial interfaces. J. Appl. Mech. 56, 7782 (1989)
33. Charalambides, P.G., Cao, H.C., Lund, J., Evans, A.G.: Development of a test method for measuring the mixed mode
fracture resistance of bimaterial interfaces. Mech. Mater. 8, 269283 (1990)
34. Chandra, L., Clyne, T.W.: Characterization of the strength and adhesion of diamond films on metallic substrates using a
substrate plastic straining technique. Diam. Relat. Mater. 3, 791798 (1994)
35. Hofinger, I., Oechsner, M., Bahr, H.A., Swain, M.V.: Modified four-point bending specimen for determining the interface
fracture energy for thin, brittle layers. Int. J. Fract. 92, 213220 (1998)
36. Howard, S.J., Phillipps, A.J., Clyne, T.W.: The interpretation of data from the four-point bend delamination test to measure
interfacial fracture toughness. Composite 24, 103112 (1993)
37. Wang, B., Siegmund, T.: A modified 4-point bend delamination test. Microelectron. Eng. 85, 477485 (2008)
38. Yamazaki, Y., Schmidt, A., Scholz, A.: The determination of the delamination resistance in thermal barrier coating system
by four-point bending tests. Surf. Coat. Technol. 201, 744754 (2006)
39. Gan, Z., Mhaisalkar, S.G., Chen, Z., Zhang, S., Chen, Z., Prasad, K.: Study of interfacial adhesion energy of multilayered
ULSI thin film structures using four-point bending test. Surf. Coat. Technol. 198, 8589 (2005)
40. Zou, L., Huang, Y., Wang, C.: The characterization and measurement of interfacial toughness for Si3 N4 /BN composite by
the four-point bend test. J. Eur. Ceram. Soc. 24, 28612868 (2004)
41. Carneiro, F.L.L.B.: Concrete tensile strength. RILEM Bull. Paris 13, 103107 (1953)
42. Atkinson, C., Smelser, R.E., Sanchez, J.: Combined mode fracture via the cracked Brazilian disk test. Int. J. Fract. 18,
279291 (1982)
43. Ayatollahi, M.R., Aliha, M.R.M.: Cracked Brazilian disc specimen subjected to mode II deformation. Eng. Fract.
Mech. 72, 493503 (2005)
44. Wang, J.S., Suo, Z.: Experimental determination of interfacial toughness curves using Brazil-nut-sandwiches. Acta Metall.
Mater. 38, 12791290 (1990)
45. Tong, J., Wong, K.Y., Lupton, C.: Determination of interfacial fracture toughness of bone-cement interface using sandwich
Brazilian disks. Eng. Fract. Mech. 74, 19041916 (2007)
46. Rahbar, N., Jorjani, M., Riccardelli, C., Wheeler, G., Yakub, I., Tan, T., Soboyejo, W.O.: Mixed mode fracture of
marble/adhesive interfaces. Mater. Sci. Eng. A 527, 49394946 (2010)
47. Tan, T., Meng, J., Rahbar, N., Li, H., Papandreou, G., Maryanoff, C.A., Soboyejo, W.O.: Effects of silane on the interfacial
fracture of a parylene film over a stainless steel substrate. Mater. Sci. Eng. C 32, 550557 (2012)
48. Janssen, C.: Sample for fracture mechanics studies on glass. In: 10th Int. Congr. on Glass: Kyoto, Japan. Ceram. Soc. Jpn.
(1974)
49. Pallares, G., Ponson, L., Grimaldi, A., George, M., Prevot, G., Ciccotti, M.: Crack opening profile in DCDC specimen. Int.
J. Fract. 156, 1120 (2009)
50. Fett, T., Rizzi, G., Guin, J.P., Lpez-Cepero, J.M., Wiederhorn, S.M.: A fracture mechanics analysis of the double cleavage
drilled compression test specimen. Eng. Fract. Mech. 76, 921934 (2009)
51. Plaisted, T.A., Amirkhizi, A.V., Nemat-Nasser, S.: Compression-induced axial crack propagation in DCDC polymer sam-
ples: experiments and modeling. Int. J. Fract. 141, 447457 (2006)
52. Jenne, T.A., Keat, W.D., Larson, M.C.: Limits of crack growth stability in the double cleavage drilled compression
specimen. Eng. Fract. Mech. 70, 16971719 (2003)
53. Ayatollahi, M.R., Bagherifard, S.: Numerical analysis of an improved DCDC specimen for investigating mixed mode
fracture in ceramic materials. Comput. Mater. Sci. 46, 180185 (2009)
54. He, M.Y., Turner, M.R., Evans, A.G.: Analysis of the double cleavage drilled compression specimen for interface fracture
energy measurements over a range of mode mixities. Acta Metall. Mater. 43, 34533458 (1995)
55. Olvera, D., Zimmermann, E.A., Ritchie, R.O.: Mixed-mode toughness of human cortical bone containing a longitudinal
crack in far-field compression. Bone 50, 331336 (2012)
56. Turner, M.R., Dalgleish, B.J., He, M.Y., Evans, A.G.: A fracture resistance measurement method for bimaterial interfaces
having large debond energy. Acta Metall. Mater. 43, 34593465 (1995)
Mechanical methods for evaluating coating adhesion 451

57. Turner, M.R., Evans, A.G.: An experimental study of the mechanisms of crack extension along an oxide/metal inter-
face. Acta Mater. 44, 863871 (1996)
58. Lipkin, D.M., Clarke, D.R., Evans, A.G.: Effect of interfacial carbon on adhesion and toughness of gold-sapphire inter-
faces. Acta Mater. 46, 48354850 (1998)
59. Ritter, J.E., Fox, J.R., Hutko, D.I., Lardner, T.J.: Moisture-assisted crack growth at epoxy-glass interfaces. J. Mater.
Sci. 33, 45814588 (1998)
60. Oliver, W.C., Pharr, G.M.: Measurement of hardness and elastic modulus by instrumented indentation: advances in under-
standing and refinements to methodology. J. Mater. Res. 19, 320 (2004)
61. Kruzic, J.J., Kim, D.K., Koester, K.J., Richie, R.O.: Indentation techniques for evaluating the fracture toughness of
biomaterials and hard tissues. J. Mech. Behav. Biomed. Mater. 2, 384395 (2009)
62. Ranade, A.N., Krishna, L.R., Li, Z., Wang, J., Korach, C.S., Chung, Y.W.: Relationship between hardness and fracture
toughness in Ti-TiB2 nanocomposite coatings. Surf. Coat. Technol. 213, 2632 (2012)
63. Drory, M.D., Hutchinson, J.W.: Measurement of the adhesion of a brittle film on a ductile substrate by indentation. Proc.
R. Soc. Lond. A 452, 23192341 (1996)
64. Chicot, D., Dmarcaux, Ph., Lesage, J.: Apparent interface toughness of substrate and coating couples from indentation
tests. Thin Solid Films 283, 151157 (1996)
65. Marot, G., Lesage, J., Dmarcaux, Ph., Hadad, M., Siegmann, St., Staia, M.H.: Interfacial indentation and shear tests to
determine the adhesion of thermal spray coatings. Surf. Coat. Technol. 201, 20802085 (2006)
66. Qi, H., Yang, X., Wang, Y.: Interfacial fracture toughness of APS bond coat/substrate under high temperature. Int. J.
Fract. 157, 7180 (2009)
67. Xie, Y., Zhang, X., Robertson, M., Maric, R., Ghosh, D.: Measurement of the interface adhesion of solid oxide fuel cells
by indentation. J. Power Sources 162, 436443 (2006)
68. Xie, Y., Hawthorne, H.M.: Measuring the adhesion of solgel derived coatings to a ductile substrate by an indentation-based
method. Surf. Coat. Technol. 172, 4250 (2003)
69. Lesage, J., Chicot, D.: Models for hardness and adhesion of coatings. Surf. Eng. 15, 447453 (1999)
70. Marcano, Z., Lesage, J., Chicot, D., Mesmacque, G., Puchi-Cabrera, E.S., Staia, M.H.: Microstructure and adhesion of
Cr3 C2 -NiCr vacuum plasma sprayed coatings. Surf. Coat. Technol. 202, 44064410 (2008)
71. Sadeghi-Fadaki, S.A., Zangeneh-Madar, K., Valefi, Z.: The adhesion strength and indentation toughness of plasma-sprayed
yttria stabilized zirconia coatings. Surf. Coat. Technol. 204, 21362141 (2010)
72. Lee, C.Y., Dupeux, M., Tuan, W.H.: Cross-sectional indentation technique for thick film interfacial adhesion characterisa-
tion. Surf. Eng. 25, 97100 (2009)
73. Heavens, O.S.: Some factors influencing the adhesion of films produced by vacuum evaporation. J. Phys. Radium 11,
355360 (1950)
74. Laugier, M.: The development of the scratch test technique for the determination of the adhesion of coatings. Thin Solid
Films 76, 289294 (1981)
75. Steinmann, P.A., Tardy, Y., Hintermann, H.E.: Adhesion testing by the scratch test method: the influence of intrinsic and
extrinsic parameters on the critical load. Thin Solid Films 154, 333349 (1987)
76. Valli, J., Mkel, U.: Applications of the scratch test method for coating adhesion assessment. Wear 115, 215221 (1987)
77. Richard, P., Thomas, J., Landolt, D., Gremaud, G.: Combination of scratch-test and acoustic microscopy imaging for the
study of coating adhesion. Surf. Coat. Technol. 91, 8390 (1997)
78. Bull, S.J.: Failure mode maps in the thin film scratch adhesion test. Tribol. Int. 30, 491498 (1997)
79. Xie, Y., Hawthorne, H.M.: A model for compressive coating stresses in the scratch adhesion test. Surf. Coat.
Technol. 141, 1525 (2001)
80. Xie, Y., Hawthorne, H.M.: Effect of contact geometry on the failure modes of thin coatings in the scratch adhesion test. Surf.
Coat. Technol. 155, 121129 (2002)
81. Benjamin, P., Weaver, C.: Measurement of adhesion of thin films. Proc. R. Soc. Lond. A 254, 163176 (1960)
82. Laugier, M.T.: An energy approach to the adhesion of coatings using the scratch test. Thin Solid Films 117, 243249 (1984)
83. Jaworski, R., Pawlowski, L., Roudet, F., Kozerski, S., Petit, F.: Characterization of mechanical properties of suspension
plasma sprayed TiO2 coatings using scratch test. Surf. Coat. Technol. 202, 26442653 (2008)
84. Barnes, D., Johnson, S., Snell, R., Best, S.: Using scratch testing to measure the adhesion strength of calcium phosphate
coatings applied to poly (carbonate urethane) substrates. J. Mech. Behav. Biomed. Mater. 6, 128138 (2012)
85. Bull, S.J., Berasetegui, E.G.: An overview of the potential of quantitative coating adhesion measurement by scratch
testing. Tribol. Int. 39, 99114 (2006)
86. Jacobsson, R.: Measurement of the adhesion of thin films. Thin Solid Films 34, 191199 (1976)
87. Roy, D., Simon, G.P., Forsyth, M., Mardel, J.: Modification of thermoplastic coatings for improved cathodic disbondment
performance on a steel substrate: a study on failure mechanisms. Int. J. Adhes. Adhes. 22, 395403 (2002)
88. Guermazi, N., Haddar, N., Elleuch, K., Ayedi, H.F.: On the peel behavior of polymer coating-steel system: effect of
hygrothermal aging. Adv. Polym. Technol. 29, 185196 (2010)
89. Sexsmith, M., Troczynski, T.: Peel strength of thermal sprayed coatings. J. Therm. Spray Technol. 5, 196206 (1996)
90. Nase, M., Zankel, A., Langer, B., Baumann, H.J., Grellmann, W., Poelt, P.: Investigation of the peel behavior
of polyethylene/polybutene-1 peel films using in situ peel tests with environmental scanning electron microscopy.
Polymer 49, 54585466 (2008)
91. Nase, M., Langer, B., Baumann, H.J., Grellmann, W., Geiler, G., Kaliske, M.: Evaluation and simulation of the peel
behavior of polyethylene/polybutene-1 peel systems. J. Appl. Polym. Sci. 111, 363370 (2009)
92. Horgnies, M., Willieme, P., Gabet, O.: Influence of the surface properties of concrete on the adhesion of coating:
characterization of the interface by peel test and FT-IR spectroscopy. Prog. Org. Coat. 72, 360379 (2011)
93. Sexsmith, M., Troczynski, T.: Peel adhesion test for thermal sprayed coatings. J. Therm. Spray Technol. 3, 404411 (1994)
94. Kurzweg, H., Heimann, R.B.: Adhesion of thermally sprayed hydroxyapatite-bond-coat systems measured by a novel peel
test. J. Mater. Sci. Mater. Med. 9, 916 (1998)
452 Z. Chen et al.

95. Dannenberg, H.: Measurement of adhesion by a blister method. J. Appl. Polym. Sci. 5, 125134 (1961)
96. Taheri, N., Mohammadi, N., Shahidi, N.: An automatic instrument for measurement of interfacial adhesion of polymeric
coatings. Polym. Test. 19, 959966 (2000)
97. Chang, Y.S., Lai, Y.H., Dillard, D.A.: The constrained blistera nearly constant strain energy release rate test for adhe-
sives. J. Adhes. 27, 197211 (1989)
98. Napolitano, M.J., Chudnovsky, A., Moet, A.: The constrained blister test for the energy of interfacial adhesion. J. Adhes.
Sci. Technol. 2, 311323 (1988)
99. Allen, M.G., Senturia, S.D.: Application of the island blister test for thin film adhesion measurement. J. Adhes. 29,
219231 (1989)
100. Dillard, D.A., Bao, Y.: The peninsula blister test: a high and constant strain energy release rate fracture specimen for
adhesives. J. Adhes. 33, 253271 (1991)
101. Wan, K.T., Mai, Y.W.: Fracture mechanics of a new blister test with stable crack growth. Acta Metall. Mater. 43,
41094115 (1995)
102. Wan, K.T., Breach, C.D.: Thermodynamics of a stable blister delamination at elevated temperature. J. Adhes. 66,
183202 (1998)
103. Wan, K.T.: A novel blister test to investigate thin film delamination at elevated temperature. Int. J. Adhes. Adhes. 20,
141143 (2000)
104. Malyshev, B.M., Salganik, R.L.: The strength of adhesive joints using the theory of cracks. Int. J. Fract. Mech. 1,
114128 (1965)
105. Wan, K.T., Mai, Y.W.: Fracture mechanics of a shaft-loaded blister of thin flexible membrane on rigid substrate. Int. J.
Fract. 74, 181197 (1995)
106. Wan, K.T., Prima, A.D., Ye, L., Mai, Y.W.: Adhesion of nylon-6 on surface treated aluminium substrates. J. Mater.
Sci. 31, 21092116 (1996)
107. OBrien, E.P., Ward, T.C., Guo, S., Dillard, D.A.: Strain energy release rates of a pressure sensitive adhesive measured by
the shaft-loaded blister test. J. Adhes. 79, 6997 (2003)
108. Na, H., Chen, P., Wan, K.T., Wong, S.C., Li, Q., Ma, Z.: Measurement of adhesion work of electrospun polymer membrane
by shaft-loaded blister test. Langmuir 28, 66776683 (2012)
109. Wang, C.: Measurements of interfacial strength from the blister test. J. Appl. Polym. Sci. 73, 18991912 (1999)
110. Koenig, S.P., Boddeti, N.G., Dunn, M.L., Bunch, J.S.: Ultrastrong adhesion of graphene membranes. Nat. Nanotech-
nol. 6, 543546 (2011)
111. Elambasseril, J., Ibrahim, R.N.: Determination of interfacial fracture toughness of coatings using circumferentially notched
cylindrical substrate. Mater. Sci. Eng. A 529, 406416 (2011)
112. Elambasseril, J., Ibrahim, R.N.: Quantitative evaluation of the adhesion of metallic coatings using cylindrical
substrate. Mater. Des. 33, 641651 (2012)
113. Elambasseril, J., Ibrahim, R.N.: Validity requirements of circumferentially notched tensile specimens for the determination
of the interfacial fracture toughness of coatings. Compos. Part B: Eng. 43, 24152422 (2012)

View publication stats

S-ar putea să vă placă și