You are on page 1of 5

Materials Science and Engineering A319– 321 (2001) 919– 923

www.elsevier.com/locate/msea

Fatigue in thin films: lifetime and damage formation


O. Kraft *, R. Schwaiger, P. Wellner
Max-Planck-lnstitut für Metallforschung and Institut für Metallkunde of the Uni6ersity, Seestr. 92, 70714 Stuttgart, Germany

Abstract

Two new techniques, developed for studying the fatigue behavior of thin metal films on substrates, are presented: The first
technique involves deposition of Cu films onto elastic polyimide substrates. During cyclic tensile testing of the film–substrate
composite, the film is subjected to tension–compression cycles, since it is plastically deformed, while the substrate undergoes only
elastic deformation. Using this technique, it was found that, for 3 mm thick Cu films, the number of cycles to failure follows the
phenomenological Coffin–Manson relationship. For the other technique, thin films, here Ag films with thicknesses ranging from
0.2 to 1.5 mm, are deposited onto micromachined SiO2 cantilever beams. The beams are then deflected with a frequency of 45 Hz
using a nanoindentation system. A detailed investigation of the damage formation in both fatigued Cu and Ag films revealed
surface roughening prior to failure. Extrusions and cracks are formed inside large grains and between small grains, respectively.
© 2001 Elsevier Science B.V. All rights reserved.

Keywords: Polyimide; Fatigue; Thin films; Cu; Ag

1. Introduction amplitude does not exceed the yield strength of the


material and the number Nf of cycles to failure is
Fatigue of thin film materials has not yet been stud- typically above 103 –104. In the LCF regime, the mate-
ied systematically. Since the mechanical properties of rial is plastically strained in each cycle. Therefore, it is
thin films may be different from those of bulk materials common to report lifetimes of a material as a function
[1 –3], fatigue behavior cannot be extrapolated from of the plastic strain range Dmpl using the empirically
results of bulk studies. For ductile bulk materials, it is established Coffin –Manson relationship [8]:
well-known that fatigue crack initiation processes are
Dmpl
often related to the formation of persistent slip bands = m %(2N
f f)
c
(1)
[4]. These slip bands, which have a typical width of 2
several micrometers, lead to extrusions at the sample
where m %f and c are the fatigue ductility coefficient and
surface and stress concentrations, or to strain localiza-
exponent, respectively. In the HCF regime, fatigue tests
tion inside grains, and eventually to crack formation. It
are usually conducted under stress control and lifetimes
is uncertain, if similar processes can occur in thin film generally obey a similar equation:
materials, since the typical size of slip bands in bulk
materials exceeds the thickness and the grain size of D|a
most thin films. Indeed, in the few TEM studies, con- = | %(2N
f f)
b
(2)
2
ducted so far on fatigued metal foils and thin films, the
formation of dislocation structures was observed to be where D|a is the applied stress amplitude, | %f and b are
suppressed in fine grained Cu foils [5 – 7]. the fatigue strength coefficient and exponent, respec-
In fatigue testing of metals, it is common to distin- tively. It was found that Cu foils or films with a
guish between regimes of high cycle fatigue (HCF) and thickness of more than 10 mm follow, depending on the
low cycle fatigue (LCF). In the former, the stress fatigue test conducted, one of these empirical relation-
ships [5,6,9,10]. However, it is not obvious if the rela-
* Corresponding author. Tel./fax: + 49-711-2095123. tionships still hold for films with smaller thickness and
E-mail address: oliver.kraft@po.uni-stuttgart.de (O. Kraft). grain size.

0921-5093/01/$ - see front matter © 2001 Elsevier Science B.V. All rights reserved.
PII: S0921-5093(01)00990-X
920 O. Kraft et al. / Materials Science and Engineering A319–321 (2001) 919–923

It is the aim of this paper to elucidate the fatigue (O2 pressure 1× 10 − 2 mbar, power 100 W, bias
mechanisms in constrained volumes and to determine − 300 V). Without breaking vacuum, 3 mm thick Cu
fatigue lifetimes under these conditions. We present films were then sputter deposited at a substrate tem-
two methods, which we developed to study the fa- perature of 300°C, a rate of 60 nm per min with an
tigue behavior of thin metal films on substrates: The Ar pressure of 2× 10 − 3 mbar, a power of 150 W
first technique involves tensile testing of Cu films de- and a bias of −80 V. After deposition, the samples
posited onto elastic polyimide substrates. Tensile test- were annealed in a vacuum furnace (6× 10 − 6 mbar)
ing of these specimen results in tension– compression at a temperature of 400°C.
cycles in the plastically deformed thin film while the The tensile tests were performed in an electrome-
substrate is elastically deformed only. The second chanical tensile tester (Zwicki 1120, Zwick) under
technique utilizes the cyclic deflection of thin Ag films load control. Samples were cyclically loaded between
deposited onto micromachined SiO2 cantilever beam a minimum load of 2 N to maximum loads in the
using a nanoindentation system. range of 15–60 N. Under these conditions, the film is
strained with constant total strain ranges between 0.7
and 2.1% as determined from the crosshead displace-
2. Experimental ment and monitored as a function of the number of
cycles. It is not possible to determine the film stress
2.1. Tensile testing using this configuration because the externally applied
force is predominantly governed by the mechanical
For processing of the tensile specimen, shown in properties of the substrate. However, microtensile
Fig. 1a, the 125 mm thick polyimide substrates (Kap- tests on comparable samples using an X-ray diffrac-
ton, Du Pont) were initially cleaned by rinsing with tion technique enabled the measurements of stress–
ethanol and pressurized CO2. Then, the sample was strain behavior during such experiments as
mounted in the deposition system (base pressure of exemplified in Fig. 1b [11]. On loading and unload-
5×10 − 7 mbar) and activated by an oxygen plasma ing, the film is plastically deformed in tension and in
compression for a total and plastic strain range of 0.5
and 0.15%, respectively.

2.2. Microbeam bending

The microbeam samples are shown schematically in


Fig. 2a. They consist of a 2.83 mm thick SiO2 layer
and a thinner Ag film. The Ag films, ranging from
0.2 to 1.5 mm in thickness, were sputter deposited at
a temperature of about − 190°C and annealed at +
100°C. A detailed description of film deposition and
the testing procedure is given in [12]. The testing pro-
cedure is briefly described as follows, the beams are
deflected by a commercial nanoindentation system
(Nano II, MTS Corp.), which applies a cyclic load
P= Pmean + Pocos(2pnt), where Pmean is the mean
load, Po the load amplitude and n the frequency,
which was chosen to be 45 Hz. In this configuration,
the largest strain occurs at the fixed end of the beam
(as indicated in Fig. 2a). Based on elastic beam bend-
ing theory, this strain can be calculated from the
lever length L and the applied load as:
Fig. 1. Microtensile testing; (a) Sample geometry of the 125 mm thick P(L− x)
polyimide foil substrate and the Cu film sputtered in the middle m= (z− q) (3)
section of the dog-bone shaped sample. (b) Typical stress-strain
IE
behavior for a Cu film on a polyimide substrate during cyclic loading where I is the moment of inertia, E the Young’s
as measured by X-ray diffraction (second cycle of 0.7 mm thick Cu
film, from [11]). The total applied strain range of the cycle is 0.5%,
modulus of the substrate material, q the position of
whereas the plastic strain range amounts only to about 0.15% as the neutral axis, and x and z the coordinates along
indicated by the arrow within the hysteresis loop. the length and the thickness of the beam. The mo-
O. Kraft et al. / Materials Science and Engineering A319–321 (2001) 919–923 921

3. Results

3.1. Tensile testing

The 3 mm thick Cu films on the polyimide substrates


had grains which were found to be equiaxed rather than
columnar with a median grain size of 0.8 mm. Also, no
distinct texture was observed. As a typical example,
Fig. 3a shows the total strain range as a function of the
number of cycles for a bare substrate and a sample with
a Cu film. The maximum load was 25 N. It can be seen
that the strain range for the sample with the film is
initially smaller because the sample stiffness is higher
due the film. After about 4000 cycles, the strain range

Fig. 2. Microbeam bending; (a) Schematic of the sample and loading


geometry, the beam consists of a 2.83 mm thick SiO2 layer onto which
0.2– 1.5 mm thick Ag films were sputter deposited. The load P is
induced by a nanoindentation system with a wedge-shaped tip having
a wedge length of 10 mm and an opening angle of 90°. The tip
contacts the beam in the dark shaded area, where the Ag film has
been removed by focused ion beam milling. (b) Schematic of the
straining conditions: The load P is cycled with an amplitude of DP
around a mean value of Pmean. The resulting cyclic strain for the Ag
film at the fixed end of the beam (hatched area in (a)) is also
represented.

ment of inertia and the position of the neutral axis


depend on the thicknesses and the elastic properties of
the beam and film [13]. As a result of the applied cyclic
load, the metal film at the fixed end of the microbeam
is subjected to a cyclic strain as schematically shown in
Fig. 2b. For lifetime studies, the lever length, the mean
load and the load amplitude were varied, resulting in
values for mmean and Dm in the range of 0.2– 2.3% and
0.2 –0.6%, respectively. We report our results as a func-
tion of mmax =mmean +1/2Dm. The film is yielding at such
strains and the corresponding stress range in the exper-
iment can, therefore, not be calculated. Fig. 2b also
indicates that under the dynamic loading conditions,
the cyclic strain is phase shifted relative to the cyclic
load. This phase shift was measured throughout the
experiment and used to monitor the stiffness of the
beam as a function of the number of cycles as described
elsewhere [12].
All films were investigated by focused ion beam Fig. 3. (a) Total strain range for a 3 mm thick Cu film on a polyimide
microscopy (FEI 200) and X-ray diffraction to deter- substrate and a bare substrate. Both experiments were conducted
under load control with a maximum load of 25 N and a minimum
mine grain size and texture, respectively. After testing, load of 2 N. The lifetime was defined as the number of cycles at
fatigue damage was characterized by scanning electron which the strain range increased. (b) Number of cycles to failure as a
and focused ion beam microscopy. function of total strain range.
922 O. Kraft et al. / Materials Science and Engineering A319–321 (2001) 919–923

3.2. Microbeam bending

The grains in the Ag films on the SiO2 microbeams


were found to be columnar and had a median grain size
of about 0.9 mm independent of the film thickness. The
majority of grains was (111)-oriented with a smaller
fraction of (100)-oriented grains. The latter increased
with increasing film thickness. As an example of a
typical microbeam bending experiment, Fig. 4a shows
the beam stiffness as function of the number of cycles
for a beam with a 0.8 mm thick film and a maximum
strain mmax of 2.2%. Initially, the beam stiffness is nearly
constant. The dramatic decrease in stiffness after ap-
proximately 6× 10 − 5 cycles indicates the onset of
severe damage formation. Similar experiments con-
ducted on samples with different film thicknesses are
summarized in Fig. 4b. The filled symbols denote the
experiments in which a distinct decrease in beam stiff-
ness was observed within 3.9× 106 cycles accompanied
by crack formation. Only films with a thickness of 0.6
mm or more suffered from fatigue damage.

3.3. Damage morphology

Typical micrographs of fatigue damage are shown in


Fig. 5. For both Cu and Ag films, the damage morphol-
ogy can be described as follows, extrusions (marked as
E) are formed within large grains, whereas intergranu-
lar cracks (C) were found in fine grained regions of the
film. The extrusion height is comparable to the film
thickness. Underneath the extrusions, large voids are
found at the interface between the film and the sub-
strate. Some more detailed observations of the fatigue
damage in the Ag films are described in [12].

Fig. 4. (a) Stiffness of a microbeam with 0.8 mm thick film cycled to


mmax of 2.2% at a frequency of 45 Hz. The sudden decrease of stiffness 4. Discussion
indicates the onset of damage formation. (b) Map for damage forma-
tion as a function of Ag film thickness and maximum strain. Thick Two techniques have been developed to study the
films showed fatigue damage within 3.9 × 106 cycles (filled symbols) fatigue behavior of thin films on substrates. Central to
whereas thin films did not show structural features of fatigue (open
symbols). The transition occurs for 0.6 and 0.8 mm thick films as
both methods is the use of an elastic substrate, which
function of maximum strain. acts as an antagonist to an external mechanical loading.
As a result, films are deformed in tension on loading
and in compression on unloading. Another important
aspect of these experiments is that film fatigue damage
is not catastrophic because the undamaged substrate
increases, i.e. the sample stiffness decreases. After maintains some mechanical integrity of the sample.
10000 cycles, the strain range has reached the level of However, the degradation of the film, i.e. crack forma-
the bare substrate indicating that the film is no longer tion and propagation, is manifested as a decrease in
contributing to the sample stiffness. The number Nf of stiffness of the film–substrate composite. Based on this
cycles to failure in these experiments was defined as the distinct decrease in stiffness, the lifetime of Cu films on
interception of two linear curve fits as demonstrated in polyimide substrates was determined as a function of
Fig. 3a. In Fig. 3b, this number is plotted versus the applied total strain range. It was not possible to deter-
total strain range. As expected, the lifetime decreases mine the plastic strain range, since the film stress was
with increasing strain range following a relationship as not measurable during the experiments. As a result,
given by Eq. (1). Fig. 3b is plotted as the initial total strain range as a
O. Kraft et al. / Materials Science and Engineering A319–321 (2001) 919–923 923

significantly slower or suppressed in thinner films,


which would be qualitatively in agreement with the fact
that the strength of metal films increases with decreas-
ing film thickness [1,2,14]. However, it is currently not
clear, if the formation of the extrusions is related to the
occurrence of dislocation structures, such as persistent
slip bands, as in bulk materials. More systematic stud-
ies, including TEM investigations, are currently pur-
sued to obtain a deeper understanding of fatigue
mechanisms in thin films.

5. Summary

In an attempt to systematically study the fatigue


behavior of thin metal films as a function of film
thickness, grain size, and straining conditions, we have
developed two new methods. Central to both methods
is the use of an elastic substrate, which acts as an
antagonist to an external mechanical loading. First
results can be summarized as follows; fatigue damage
includes the formation of large transgranular extrusions
and intergranular cracks. The crack formation is associ-
ated with a distinct decrease in stiffness of the film –
substrate composite. The lifetime of 3 mm thick Cu
films on polyimide substrates follows a Coffin –
Manson-type relationship. No fatigue damage was
found in Ag films with thicknesses below 0.6 mm as
tested by dynamic microbeam deflection.
Fig. 5. Damage morphology after fatigue testing; (a) Focused ion
beam micrograph of a 3 mm thick Cu film on a polyimide substrate,
large extrusions and cracks are marked by E and C, respectively. (b)
SEM micrograph showing the 0.8 Ag film, from the test described in References
Fig. 4a, at the fixed end of a microbeam. Large extrusions (marked
by E) are connected by a crack (C). [1] W.D. Nix, Met. Trans. A 20A (1989) 2217 – 2245.
[2] R.-M. Keller, S.P. Baker, E. Arzt, J. Mater. Res. 13 (1998)
function of Nf. The data follow Eq. (1) with an expo-
1307 – 1317.
nent of − 0.4, which is somewhat higher than typically [3] E. Arzt, Acta Mater. 46 (1998) 5611 – 5626.
observed values of − 0.5 to −0.7 for many bulk metal [4] U. Essmann, H. Mughrabi, Phil. Mag. A 40 (1979) 731 –756.
materials [8]. This is partially accounted for by the use [5] S. Hong, R. Weil, Thin Solid Films 283 (1996) 175 –181.
of the total strain range instead of the plastic strain [6] M. Judelewicz, H.U. Künzi, N. Merk, B. Ilschner, Mat. Sci. Eng.
A186 (1994) 135 – 142.
range, because the difference between the two is more
[7] D.T. Read, Int. J. Fatigue 20 (1998) 203 – 209.
important at small strain ranges. As a result of this [8] S. Suresh, Fatigue of Materials, Second ed, Cambridge Univer-
consideration, the exponent of −0.4 can be regarded sity Press, Cambridge, 1999, pp. 137 – 139.
as an upper bound. [9] Y. Oshida, P.C. Chen, J. Electr. Packaging 113 (1991) 58 –62.
The microbeam bending experiments reveal that the [10] H.D. Merchant, M.G. Minor, Y.L. Liu, J. Electron Mater. 28
(1999) 998 – 1007.
fatigue behavior changes significantly, when the film
[11] M. Hommel, O. Kraft, and E. Arzt. J. Mater. Res., 14 (1999).
thickness is reduced below 1 mm. Fig. 4b shows that [12] R. Schwaiger, O. Kraft, Scr. Mat. 41 (1999) 823 – 829.
films thinner than 0.6 mm did not fatigue within 3.9× [13] S.P. Baker, W.D. Nix, J. Mater. Res. 9 (1994) 3131 –3144.
106 cycles. The formation of extrusions appears to be [14] R. Venkatraman, J.C. Bravman, J. Mater. Res. 7 (1992) 2040.

View publication stats

You might also like