You are on page 1of 7

Surface & Coatings Technology 205 (2010) 15–21

Contents lists available at ScienceDirect

Surface & Coatings Technology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / s u r f c o a t

Crack formation and its prevention in PVD films on epoxy coatings


Yuqing Bao, Jiming Gao, David T. Gawne ⁎
Department of Engineering and Design, Faculty of Engineering, Science and the Built Environment, London South Bank University, Borough Road, London, SE1 0AA, United Kingdom

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

Article history: Severe cracking was found to occur in PVD titanium films on epoxy powder coatings. After all baking treatments,
Received 28 January 2010 the epoxy coating had smooth, crack-free surfaces and the cracking of both the titanium film and the epoxy only
Accepted in revised form 20 May 2010 took place as a result of physical vapour deposition. Tensile cracks were observed in the titanium film and not the
Available online 27 May 2010
compressive cracks expected from the conventional two-layered theoretical model. An alternative model has
been developed for the prediction of thermal stress in a three-layered film–epoxy–substrate system. The model is
Keywords:
PVD films
consistent with the experimental trials and showed that cracking originated from thermal stresses developed in
Cracking the titanium–epoxy–aluminum system due to the PVD process. Tensile instability and cracking were initiated
Polymer powder coatings where pores intersected the film-coating interface. The results showed that crack formation could be prevented
Residual thermal stress by increasing the baking temperature to 210 °C. This critical temperature activates full crosslinking in the epoxy
Computer model structure and raises its strength sufficiently to avoid tensile instability due to residual stress. Crack-free and high-
Combined in-line PVD-powder coating gloss sputtered titanium films could therefore be produced on organic coatings. This offers the potential of a
technology combined in-line PVD-powder coating technology as an alternative to electroplating.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction alternative to electroplating in certain applications and, in some cases,


more environmentally compatible. This is particularly the case as many
Cracking of physical vapour deposited (PVD) films on dissimilar electroplating companies already possess powder coating facilities and
substrates can be a problem in many systems and renders products have the possibility of introducing in-line PVD equipment. Potential
unusable. There are other difficulties with PVD films. They are very thin property benefits would include high gloss, attractive colours and wear
and often relatively porous, which seriously impairs their corrosion resistance [2] together with levelling ability and corrosion resistance.
resistance. A further hindrance is the inability of PVD films to level a
substrate surface: a typical engineering surface will not provide a glossy
PVD surface, unless given an expensive polishing treatment before PVD. 2. Experimental details
This paper investigates the introduction of a polymer powder
coating in between the PVD film and metal substrate. The polymer 2.1. Film and coating deposition
coatings are much thicker than PVD films (60 µm compared with
0.5 µm), non-conducting and dense, and as a result, should provide a Aluminum (grade 6063) plate of dimensions 100 mm × 50 mm ×
corrosion resistant barrier layer for the underlying metal substrate. A 3 mm was used as the metal substrate. The plates were cleaned and
further benefit of the polymer is its potential for levelling a rough degreased with acetone immediately before applying the coating. An
metal surface, which obviates the need for polishing. However, the epoxy powder (Resicoat, Akzo Nobel) was deposited onto the
thermal expansion coefficient of polymers is much higher than that of aluminum plates by electrostatically spraying. The epoxy powder
metals, while the thermal conductivity and elastic modulus are much had a glass transition temperature (Tg) of 59 °C as measured by DSC.
lower. This property mismatch creates serious difficulties in produc- The powder coatings were subsequently cured in a baking oven under
ing high-quality inorganic films on organics. various temperature and time combinations in order to gain a range of
Overcoming these problems would provide a combined technology crosslinking in the polymer architecture.
of PVD-powder coatings [1] with significant advantages for metal Titanium films were deposited on both uncoated (bare) aluminum
products in manufacturing and properties. It could be a potential and epoxy-coated aluminum plates using a JLS MPS 500 Sputter
System. The epoxy undercoats were cleaned using dry compressed air
before entering the PVD chamber. Sputtering was carried out using
pure argon at a pressure of 15 mT, power of 350 W and without Vbias.
⁎ Corresponding author. A sputtering time of 2 h was used, unless stated otherwise, which gave
E-mail address: david.gawne@lsbu.ac.uk (D.T. Gawne). a film thickness of 0.5 µm.

0257-8972/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.surfcoat.2010.05.037
16 Y. Bao et al. / Surface & Coatings Technology 205 (2010) 15–21

2.2. Evaluation techniques

The extent of curing due to heat treatment of the epoxy powders


and coatings was investigated by differential scanning calorimetry
(DSC) with a linear heating rate of 20 K min− 1 in dry nitrogen. Fourier
transform infrared spectroscopy (FTIR) was used to investigate any
possible chemical changes caused by PVD. Each coating and powder
sample was subjected to 32 scans and the reflectance of the samples
recorded with a resolution of 4 cm− 1.
The gloss and surface roughness were used to quantify the surface
characteristics of the PVD films and undercoats. The gloss was measured
using a gloss meter (ETC GT60), which complies with ISO 2813, ASTM
D523, and DIN 67530. The surface topography of the coatings was
assessed by a Talysurf profilometer, optical and electron microscopy.
Adhesion was assessed by the pull-off and cross-cut tests. The pull-
off tensile test was used to assess the adhesion between the powder
coating and aluminum substrate. The test conforms to ASTM D4541 and
BS 3900-E10, and was carried out with a tensile machine. A test dolly
was bonded to the coating using an adhesive or glue and the tensile
force required to detach the coating from substrate was measured.
The PVD film was too thin for the pull-off test and the cross-cut test
was used to assess the adhesion between PVD film and the epoxy
undercoat. This test was performed using a multi-blade cutting tool in
accordance with ENISO2409:2007. The cross-cut test measures the area
percentage of the film detached as a result of cross-cutting and
subsequent tape stripping. The adhesion of the coatings is presented as
a six-step classification (Steps 0 to 5) in accordance with that defined in
ISO 2409:2007: Step 0 gives no detectable detachment of the film and the
highest adhesion performance, Step 1 gives less than 5% by area
detached, Step 2 gives 5–15% detached, Step 3 gives 15–35%, Step 4
35–65% and Step 5 more than 65% film detachment with the lowest
performance.
Vickers hardness of the coatings was measured using a Duramin-5
microhardness tester with a diamond indenter having 136° angle
Fig. 1. SEM micrographs of sputter-deposited Ti film on epoxy undercoat on aluminum
between opposite planes. The test was carried out under 0.5 N applied
substrate: (a) cross-section through film, (b) top surface of film.
load with a fixed loading time of 15 s.
The thickness of PVD Ti film was measured by a Dektak 6 M Stylus
Profiler. The standard deviation on successive 500 nm step-height roughness (Ra) of 0.40 μm and the titanium-coated aluminum of
measurements was typically 2–3%. A silicon sheet partly covered with 0.36 μm. However, using the interleaving layer of epoxy gave effective
Capron tape was placed in PVD chamber together with epoxy-coated levelling as shown by the micrograph in Fig. 2b and an Ra value of
substrate. After the sample was taken out of the PVD chamber, the 0.08 μm for the titanium film on the epoxy-coated aluminum. These
tape was removed and the difference between coated and uncoated results show that an epoxy powder coating can level a rough surface and
area was scanned by the stylus profiler. The step height was then has the potential of avoiding the need for polishing the substrate.
equated to the thickness of the film. In the fine-scale locality of the deposition zone, physical vapour
deposition will be expected to generate elevated temperatures due to
3. Results and discussion the impact and condensation at the substrate by gaseous atomic species
in the chamber. Infrared spectroscopy (FTIR) was used to assess
3.1. Characterization whether or not thermal degradation of the organic coating took place
during PVD. Fig. 3 shows the FTIR curves obtained: Curve A relates to the
Fig. 1 gives a through-thickness section and the top surface of the epoxy undercoat before PVD and Curve B to the epoxy surface
titanium film on the epoxy undercoat. The titanium film had a grain size immediately beneath the titanium film after sputter deposition. No
of the order of 100 nm and contained fine pores parallel to the through- detectable chemical change of the epoxy before and after sputter
thickness direction. The film structure was very similar to that deposited deposition is detected and so the FTIR results indicate that no significant
on uncoated aluminum. The fine-scale close contact between the film degradation of the epoxy takes place under the current sputter
and epoxy observed at the interface in Fig. 1a suggests strong adhesive deposition conditions. In support of this conclusion, the temperature
strength and this was confirmed by the high classification (Step 0) of epoxy was estimated at 140–170 °C (from its colour change) during
obtained in the cross-cut test. The adhesion of the cured epoxy powder the deposition and this is far below the epoxy decomposition
coating to the aluminum substrate was assessed by the pull-off tensile temperature. It may also be concluded that there will be no further
test and gave an average value of 8 ± 1 MPa. In fact, the tensile fracture crosslinking of the epoxy during physical vapour deposition, since the
was observed within the adhesive used, which indicates that the true latter estimated temperature is below the baking temperatures used.
epoxy–substrate adhesive strength was greater than 8 MPa. This
represents high adhesion between the epoxy and aluminum. 3.2. Cracking and crosslinking
Scanning electron microscopy under much lower magnifications
(Fig. 2a) showed that the PVD films replicated the original surface profile In the initial trials on sputtering the titanium films on the epoxy-
of aluminum and provided only slight levelling of the surface. This was coated substrate, the films were found to be cracked on removal from
confirmed by the profilometry results: the as-received aluminum had a the PVD chamber. Observations of the films in the chamber showed
Y. Bao et al. / Surface & Coatings Technology 205 (2010) 15–21 17

Fig. 4. Optical micrograph of top surface of titanium film sputter-deposited on the


epoxy coating cured at 200 °C for 20 min.

time was only 5 min and 10 min is required at this temperature to


suppress crack formation. There is, therefore, a critical cure temper-
ature and time necessary to avoid cracking in the epoxy coatings and
both parameters must be specified. Fig. 5 shows the trials results
plotted on a predominance diagram: combinations of temperature
and time specifying coordinates above the curve give uncracked films,
whereas those below provide cracked films. The curve is approximate
and only applies to the particular epoxy and PVD conditions used in
this investigation.
Differential scanning calorimetry examination indicated that
raising the baking temperature resulted in, as expected, an increased
degree of curing by thermally activating crosslinking between the
molecular chains of the epoxy structure. This provides strengthening
of the epoxy as confirmed by Fig. 6, which shows that the measured
microhardness of the epoxy coating increases with baking tempera-
tures up to 240 °C. The error bars in the diagram refer to the standard
deviations of the measurements. However, most of the strengthening
Fig. 2. SEM micrographs of sputter-deposited Ti film on: (a) uncoated aluminum plate;
takes place by 210 °C, which is consistent with the above results on
(b) epoxy undercoat on aluminum plate. the suppression of cracking. The research shows therefore that the
underlying mechanism for the avoidance of cracking is the enhance-
ment of fracture toughness by crosslinking of the epoxy.
that there were no cracks during deposition and the cracks formed
during cooling in the chamber. The cracking of the film was extensive 3.3. Origin of residual stress
and a crack network formed, as shown in Fig. 4.
Experimental trials were undertaken in which the epoxy coating There are three main contributions to residual stress in PVD films:
was baked at temperatures in the range 180 to 240 °C for times of 5 to (i) thermal stress, (ii) intrinsic stress, (iii) extrinsic stress [3–5]. The
20 min. The results reveal a minimum baking temperature of 210 °C at total stress in the film, σ, may thus be described as:
which crack-free coatings could be obtained. Cure time was also
important. The latter critical baking temperature of 210 °C related to a σ = σthermal + σintrinsic + σextrinsic : ð1Þ
time of 20 min, and reducing the time, raised the critical temperature.
For instance, cracking was seen to occur at 240 °C because the cure Thermal stress is due to the mismatch of thermal expansion
between the film and the substrate upon cooling from the deposition
temperature to ambient. The fact that the cracks in the film formed on

Fig. 3. FTIR spectra of epoxy undercoat. A: before PVD; B: epoxy surface immediately Fig. 5. Fail-safe diagram for sputtered titanium film on epoxy coating. Cracking occurs
beneath the Ti film after sputter deposition. below the curve but not above it.
18 Y. Bao et al. / Surface & Coatings Technology 205 (2010) 15–21

3000 µm thick aluminum substrate. The dimensions of the system are


100 mm × 50 mm × 3.06 mm. It is assumed that the titanium film,
epoxy interlayer and aluminum substrate have homogenous struc-
tures and properties, and undergo uniform biaxial thermal strain in
response to changes in temperature (the thermal strain along the
thickness direction is neglected). Due to the differences in thermal
expansion coefficients, the strain of the epoxy (which has the largest
free thermal expansion) will be restricted by the film and the
substrate, as they both have smaller free expansions. During cooling
from the deposition temperature to room temperature, therefore, the
epoxy will be under a biaxial tensile force Fe, while the titanium film
and aluminum substrate will be under biaxial compressive forces Ff
and Fa respectively. As there are no external forces acting upon the
cross-section of the combined three-layer system, the following force
equilibrium is established:
Fig. 6. Effect of baking temperature on measured hardness and estimated tensile
strength of epoxy coatings. Fe − Ff − Fa = 0 ð2Þ

cooling and not during deposition suggests that thermal stress has a where subscripts f, e and a denote, the film, the epoxy and aluminum
major effect. Other relevant results are: (a) no cracks were observed respectively.
to form in titanium films deposited under the same PVD conditions on Referring to Fig. 7, Fe is considered as the resultant force of Fe1
bare aluminum during deposition or during cooling, and (b) the (caused by Ti film) and Fe2 (caused by aluminum substrate) and so
results in Section 3.2 show that cracking is strongly influenced by the Eq. (2) can be rewritten as:
strength of the epoxy. These results confirm that thermal stress is
expected to be a major contributor to residual stress. Fe1 + Fe2 −Ff −Fa = ðFe1 −Ff Þ + ðFe2 −Fa Þ = 0: ð3Þ
Intrinsic stress is formed by the atom species impacting the
substrate or growing surface layer in the chamber during deposition. Eq. (3) indicates that the three-layer force system of the titanium
The Thornton structure zone model provides a very approximate film on the epoxy undercoat on the aluminum substrate can be
prediction of microstructure from PVD process conditions [5–8]. In the considered as the superposition of two two-layer sub-systems:
present investigation, the substrate temperature normalized to the titanium film on epoxy (Sub-system 1) and epoxy on aluminum
melting point of titanium (Ts/Tm = 428 K/1998 K) was 0.21, the (Sub-system 2). This superposition is shown schematically in Fig. 7.
chamber pressure was 15 mT and no bias was applied. For these For the film on epoxy
conditions, the Thornton model predicts a Zone 1 type of structure,
which is a fairly porous columnar structure under low-tensile intrinsic Fe1 −Ff = 0 ⇒Fe1 = Ff = F1 : ð4Þ
stress. The observed structure in Fig. 1 is consistent with this
prediction and so it is possible that this relatively open structure For the epoxy on aluminum
may also have been under a tensile intrinsic stress during deposition
of the film. However, no cracks were observed in the films during Fe2 −Fa = 0 ⇒Fe2 = Fa = F2 : ð5Þ
deposition (before cooling in the chamber began) on uncoated
aluminum or on epoxy-coated aluminum. This suggests that intrinsic The above treatment can simplify the stress calculation. The forces
stress might not be the predominant contribution, although it might in Eq. (4) produce a couple that must be resisted by internal moments
still have been present (at the end of deposition before cooling) but at of the film and epoxy, hence [13]:
a level below the fracture stress. Extrinsic stress is due to various
molecules, particularly gaseous polar species, penetrating open voids 1 F
or pores, adsorbing on pore walls and creating residual stress. ðE I + Ee I1e Þ = 1 ðtf + te Þ: ð6Þ
R1 f f 2
However, the cracks formed during cooling under low pressure in
the chamber and no change was observed on exposing the samples to
Likewise, the forces in Eq. (5) produce a couple that must be
the atmosphere. This suggests that extrinsic stress is not a major
resisted by internal moments of the epoxy and aluminum, hence:
contributor.
1 F
3.4. Model for thermal stress ðE I + Ea Ia Þ = 2 ðte + ta Þ ð7Þ
R2 e 2e 2

The discussion in the previous section indicates that thermal stress


where R1 and R2 are the radii of curvature of the neutral axis in Sub-
plays a major role in the generation of residual stress in the coating
system 1 (film on epoxy) and Sub-system 2 (epoxy on aluminum)
system under investigation. A quantitative analysis based on biaxial
respectively. E is the elastic modulus of the materials, I is the second
forces, bending moment equilibrium and classical beam bending
moment area of the materials about the neutral axis of the cross-
theory has therefore been carried out to predict thermal stress in
sections, and t is the thickness.
order to aid understanding the observed behaviour. Analytical models
In the film–epoxy sub-system, the total strain equals the sum of
using force and moment equilibrium theory have been reported in the
the direct strain caused by the direct force F1, the bending strain due
literature and their solutions applied to the calculation of residual
to the bending moment and a thermal strain due to temperature
stress in metal and ceramic coatings produced by PVD and thermal
variation. These must be compatible at the film–epoxy interface [13].
spray deposition for single or multi-layered coating systems [5,9–12].
Hence, the force F1 can be derived as:
This paper considers a three-component system but uses a simpler
approach of the superposition of two two-component sub-systems. ðαe −αf ÞΔT
The three-layer system consists of a top layer of a 0.5 µm titanium F1 = : ð8Þ
1−νf 1−νe ðtf + te Þ2
+ +
film, an interlayer of a 60 µm epoxy coating and a bottom layer of a Af Ef Ae Ee 4ðEf If + Ee I1e Þ
Y. Bao et al. / Surface & Coatings Technology 205 (2010) 15–21 19

Fig. 7. Schematic representation of a three-layer force system (titanium–epoxy–aluminum) by the superposition of two two-layer sub-systems: titanium on epoxy (Sub-system 1)
and epoxy on aluminum (Sub-system 2).

Similarly, force F2 can be calculated as: still under a compressive stress but at a smaller magnitude. This is
mainly attributed to a reduced thermal strain mismatch between the
ðαe −αa ÞΔT titanium and aluminum (Table 1). Fig. 9b indicates that the aluminum
F2 = ð9Þ
1−νe 1−νa ðte + ta Þ2
Ae Ee + Aa Ea + 4ðEe I2e + Ea Ia Þ
substrate in the latter system is under a tensile stress but the
magnitude is very small (0.3 MPa) owing to its large thickness. In all
where ν is the Poisson ratio of the materials, α is the linear coefficient systems (Figs. 8 and 9), the stresses are predicted to increase with the
of thermal expansion, ΔT is the temperature difference and A is the PVD deposition temperature.
cross-sectional area of the materials.
The distribution of thermal stress across the cross-section of the 3.5. Mechanism of crack formation
film and epoxy can therefore be calculated by considering the
combined effect of the direct stress and bending stresses: Figs. 4 and 10 are progressively higher magnifications of the cracks
observed in the titanium film on the epoxy cured at 200 °C for 20 min.
F1 E
σf = + f y1f ð10Þ This baking treatment is below the critical temperature of 210 °C and
Af R1
so gives rise to cracking. The tensile strength of the epoxy, estimated
as one third of its hardness [14], is 69 MPa at 210 °C, 59 MPa at 200 °C
F1 + F2 y y
σe = − FEe 1e −Ee 2e ð11Þ
Ae R1 R2

σf and σe are stresses within the film and epoxy respectively, y1


and y2 are the distances to the neutral axis of the Sub-system 1 (film–
epoxy) and Sub-system 2 (epoxy–aluminum) respectively.
Eqs. (2) and (11) were used to calculate thermal stress generated
within the film and epoxy-interlayer. The material properties used in
the calculation are given in Table 1. The film stress calculated from
Eq. (10) considers only interactions between the film and epoxy-
interlayer (Sub-system 1). As the thickness of film is far smaller than
that of epoxy (0.5 μm compared with 60 μm) and the crack propagation
was observed to end within the epoxy interlayer rather than pass
through it, it is reasonable to neglect the effect of the aluminum
substrate. However, the stress of the epoxy interlayer is calculated as the
consequence of the superposition of two sub-systems.
Fig. 8a was derived using the above equations for the film–epoxy–
substrate system and shows that the titanium film is under an almost
uniform compressive stress due to its lower thermal expansion
coefficient than epoxy. Across the film–epoxy interface, the stress in
the epoxy undercoat is tensile: Fig. 8b gives a magnified view to show
how it decreases linearly with the distance from the film–epoxy
interface. The highest tensile stresses occur at the film–epoxy
interface. In comparison, Fig. 9 relates to the titanium film–aluminum
substrate system (without the epoxy interlayer) and shows the film is

Table 1
Material properties at room temperature used in analytical modelling [15]; [16,17].

Properties Titanium Epoxy Aluminum


−3
Density (kg m ) 4510 1200 2700
Linear coefficient of thermal expansion 8.6 81–117 23.6
(10− 6 K− 1)
Fig. 8. Calculated stress in a titanium–epoxy–aluminum coating system for specified
Young's modulus (GPa) 103 2.41 69
deposition temperatures: (a) stress through thickness; (b) expanded scale showing
Poisson's ratio 0.34 0.34 0.33
stress through epoxy thickness.
20 Y. Bao et al. / Surface & Coatings Technology 205 (2010) 15–21

The crack morphology in the titanium film clearly indicates tensile


cracking, whereas the thermal stress in this film is predicted to be
highly compressive. On the other hand, the thermal stress in the
epoxy is tensile. Examination of the epoxy within the crack in Fig. 10
shows extensive plastic work or plasticity has taken place. The
plasticity is in the form of stretching, twisting and tearing, which is
indicative of tensile, ductile fracture. This plastic deformation implies
that significant strain hardening must occur in the epoxy, so as to
delay tensile instability and enable the observed ductility. The
greatest plasticity is seen to take place at the interface with the film
and this is consistent with the model predictions in Fig. 8, in which the
stress is a maximum at the interface and then declines with depth into
the epoxy. Fig. 10 also shows the crack stops within the epoxy and
does not extend down to the aluminum interface, which is again
consistent with the declining stress in the model.
Close inspection of Fig. 10 shows the presence of holes or voids,
particularly in the epoxy close to the titanium–epoxy interface, and
the tearing often appears to have initiated from these holes. The
diameter of the holes is approximately 100–500 nm, which is of a
similar order of magnitude as the pores observed in the titanium film
(Fig. 1). Where a pore intersects the interface, there will clearly be no
bonding between the titanium film and the epoxy. This will represent
a position of weakness where the epoxy is not constrained or
strengthened by the film or alternatively, a position of stress
concentration near which the stress is raised locally. This causes
instability in the epoxy side of the interface next to the pore, where
strain hardening can no longer support thinning with the result that
local deformation of the epoxy takes place. This leads to the formation
of voids or holes in the epoxy followed by crack propagation and
Fig. 9. Calculated stress in a titanium–aluminum coating system for specified deposition ductile failure of the epoxy coating. The fracture of the epoxy is
temperatures: (a) stress through thickness; (b) expanded scale showing stress through
expected to initiate crack propagation from the pores through the
aluminum substrate thickness.
thickness of the titanium film. Close inspection of the film fracture
surface in Fig. 10 shows stretching and tearing on a fine scale, which
indicates some ductility in the titanium and suggests the imposition of
and 49 MPa at 180 °C as given in Fig. 6. In broad terms, these baking a net tensile stress system at the point of failure.
temperatures produce approximately fully cured (at 210 °C) and
partially cured epoxies. In comparison with these strength levels, the 4. Conclusions
imposed thermal stresses as calculated by the model (Section 3.4 and
Fig. 8) are tensile and 56, 70 and 85 MPa corresponding to a 1. The topography of the PVD films is governed by that of the epoxy
deposition temperature of 140 °C, 170 °C and 200 °C at the film– undercoat. An epoxy powder coating can enable the deposition of a
epoxy interface. These tensile stresses are above the estimated tensile smooth PVD film on a rough substrate, which avoids the
strengths (Fig. 6) of the epoxies baked below 210 °C (49–59 MPa) and requirement for substrate polishing.
coincide with the appearance of cracks. The value of the critical 2. A quantitative model based on uniform biaxial strain has been
temperature therefore relates to that providing maximum strength developed for the calculation of the thermal stress in a three-layer
and sufficient crosslinking in the epoxy. The calculated results are PVD film–epoxy–metal substrate system. The calculated results
therefore consistent with the experimental observations. have been used to analyze the mechanism of crack formation.

Fig. 10. SEM micrographs showing the cracks viewed from the top surface of titanium film deposited on the epoxy coating cured at 200 °C for 20 min.
Y. Bao et al. / Surface & Coatings Technology 205 (2010) 15–21 21

3. The model predicts that major compressive stresses will be Acknowledgements


generated in the PVD titanium film simultaneously with tensile
stresses in the epoxy undercoat due to the mismatch in properties The authors wish to thank the European Commission (Sixth
between the epoxy and titanium. The model also predicts that the Framework Programme COLL-CT-2006-030409 Flexicoat) for finan-
highest tensile stress is at the film–epoxy interface at a magnitude cial support of this work. They would also like to thank Akzo Nobel
that exceeds the tensile strength of epoxy cured below 210 °C. This Powder Coatings Ltd., Gateshead, UK, for providing the epoxy powder.
is in close agreement with the observed critical temperature of
210 °C below which cracking occurs in titanium films deposited on References
epoxy.
[1] Guide to Exploitation of PVD, Editors: David Gawne, Cor Schrauwen and Arjan de
4. The tensile stress is predicted by the model to decrease with depth Bruin, Kennistransfer Centrum Bouw & Industrie, Zeist, The Netherlands, ISBN:
into the epoxy undercoat and this is consistent with the 978-90-81 5273-1-6, 2010.
observation that cracks end within the epoxy and do not extend [2] D.M. Mattox, Handbook of Physical Vapor Deposition (PVD) processing: film
formation, adhesion, Surface Preparation and Contamination Control, Noyes
down to the aluminum interface. The magnitude of the stresses in Publications, Westwood, N.J, 1998.
the film and epoxy increases with the deposition temperature. [3] Y. Pauleau, Vacuum 61 (2001) 175.
5. A network of tensile cracks was observed in the titanium PVD film, [4] M.M.M. Bilek, D.R. McKenzie, R.N. Tarrant, S.H.M. Lim, D.G. McCulloch, Surf. Coating
Technol. 156 (2002) 136.
despite the expected compressive stresses in the film. The research [5] V. Teixeira, Thin Solid Films 392 (2001) 276.
shows that tensile cracking and ductile fracture initiates in the [6] J.A. Thorton, Ann. Rev. Mater. Sci. 7 (1977) 239.
epoxy at the interface opposite pores in the titanium film. This [7] E. Mirica, G. Kowach, H. Du, Cryst. Growth Des. 4 (2004) 157.
[8] A.J. Hermann, Surf. Coating Technol. 112 (1999) 210.
fracture also causes crack propagation from the pores through the
[9] V. Teixeira, Thin Solid Films 64 (2002) 393.
thickness of the titanium and produces a tensile crack network on [10] X.C. Zhang, B.S. Xu, H.D. Wang, Y. Jiang, Thin Solid Films 497 (2006) 223.
the film surface. [11] Y.C. Tsui, T.W. Clyne, Thin Solid Films 306 (1997) 23.
6. Increasing the baking temperature to 210 °C and above raises the [12] X.C. Zhang, B.S. Xu, H.D. Wang, Y.X.Wu. Thin Solid, Films 488 (2005) 274.
[13] Mechanics of solids and structures, in: D.W.A. Rees (Ed.), McGraw-Hill Book
tensile strength of the epoxy by crosslinking sufficiently to resist Company (UK) Ltd, London, 1990, ISBN 0-07-708222-7.
failure by the residual stress and produce crack-free PVD titanium [14] J.R. Cahoon, W.H. Broughton, A.R. Katzak, Metall. Mater. Trans. B 2 (1971) 1979.
films. [15] Modern Plastics Encyclopaedia 96, McGraw-Hill, New York, NY.
[16] Shiqiang Deng, Meng Hou, Lin Ye, Polym. Test. 26 (2007) 803.
7. No degradation of the epoxy was detected in this work during the [17] Metals Handbook: Properties and Selection: Non-ferrous Alloys and Pure Metals,
physical vapour deposition of titanium films. Vol 2, 9th Edition.

You might also like