Sunteți pe pagina 1din 7

Thin Solid Films 479 (2005) 130 136

www.elsevier.com/locate/tsf

Influence of growth temperature on microstructure and mechanical


properties of nanocrystalline zirconium carbide films
Cheng-Shi Chen*, Chuan-Pu Liu, C.-Y.A. Tsao
Department of Materials Science and Engineering, National Cheng-Kung University, Tainan 701, Taiwan, ROC

Received 18 September 2004; accepted in revised form 29 November 2004


Available online 30 December 2004

Abstract

ZrC films were grown on Si (100) substrates using magnetron sputtering where the growth temperature (Ts) was varied from 25 8C to 290
8C. Film/substrate practical adhesion of the ZrC films was determined by scratch testing while hardness, elastic modulus and fracture
toughness were measured by nanoindentation. Structures and morphologies of the ZrC films were analyzed using scanning electron
microscopy and X-ray diffraction. The results indicate that there exists an optimum growth temperature at Ts=120 8C, at which the film
exhibits the best adhesion. In addition, lower growth temperatures result in an increase in hardness and a decrease in modulus, while higher
growth temperatures degrade fracture toughness. The film structure reveals a change from columnar to equiaxed nanocrystalline at Ts=290
8C, which has a profound effect on some of the mechanical properties, such as hardness. The mechanism responsible for the nanocrystalline
structure is discussed.
D 2004 Elsevier B.V. All rights reserved.

PACS: 62.20.-x; 62.20.Qp; 68.35.-p


Keywords: Sputtering; Zirconium carbide; Adhesion; Hardness; Nano-crystalline

1. Introduction successful manufacture and applications of these devices.


Compared to the widely used ZrN materials, ZrC could be
ZrC is an advanced ceramic owing to both superior a good candidate because of 14% lower lattice mismatch
covalent properties such as high melting point (T m=3445 and of 3.4% lower expansion coefficient difference with Si.
8C) compared to TiN (T m=2950 8C), TaN (T m=2980 8C) ZrC thin films have been synthesized by e-beam
and ZrN (T m=2982 8C), great hardness, excellent mechan- bombardment [1], chemical vapor deposition [7], laser
ical stability, lower work function [1] and a metallic cladding [8], plasma-assisted metal organic chemical vapor
behavior in electrical and optical properties [2]. Therefore, deposition [9], and pulsed laser ablation-deposition [10].
ZrC films have been largely applied as refractory However, despite its technology interest, ZrC has not been
materials for cutting tools [3], crucibles in mechanical fabricated by simple magnetron sputter deposition. Here we
and nuclear industry [4,5] as well as in electronic devices report on a nanocrystalline ZrC film formed by sputtering at
[6]. 290 8C and the variations of film structure and mechanical
Although the primary function of ZrC deposited on Si is property with growth temperature are examined. Thin film
the electrical property for metallization applications, practical adhesion is an important property not only for
mechanical properties and interfaces are crucial to the microelectronics and magnetic recording industries, but also
for emerging technologies such as data transmission through
optical switches, which are typical of microelectromechan-
* Corresponding author. Tel.: +886 62 757 5756 2938; fax: +886 787
ical systems [11].
21967. ZrC coating on a roller bearing has a hardness of 11
E-mail address: fccs@ms48.hinet.net (C.-S. Chen). GPa and a critical load of 30 N [12]. Recently, Willmott
0040-6090/$ - see front matter D 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.tsf.2004.11.196
C.-S. Chen et al. / Thin Solid Films 479 (2005) 130136 131

and Spillmann indicated that the hardness of ZrC0.6N0.4 is 3. Results and discussion
about 27 GPa [13]. Through a systematic study in this
paper, ZrC hardness and critical load were measured to 3.1. Microstructure characteristics of the ZrC thin films
be 30.217.5 GPa and 19.233.6 N, respectively, which
are significantly higher than reported. Interestingly, we Fig. 1 shows a series of XRD patterns from the as-
also found an inverse HallPetch [14,15] relationship as deposited ZrC films grown at various growth temperatures
approaching a nanocrystalline film due to nano-size along with a standard ZrC pattern in powder form from the
effect. JCPDS card [21]. From Fig. 1, all the as-deposited films
exhibited only ZrC cubic phase and the ZrC diffraction
peaks generally become narrower with growth temperature,
2. Experimental details indicating that there are larger crystallites and more order in
the lattice, except Ts=290 8C. The crystallite size was
The ZrC films were deposited on Si (100) substrates estimated by the Scherrer method as follows:
using a stoichiometric ZrC target (99.5% in purity) by dc
thkl kk=bcosh 1
magnetron sputtering. The distance between the target and
the substrate holder was fixed at 60 mm. A DC power of where t hkl is the thickness of crystallite perpendicular to
150 W was employed, while the base pressure of the the measuring surface, k is a constant (0.9 for the full
deposition chamber was 3106 Torr. During deposition, width at half maximum [22]); k is the wavelength of the
the working pressure was fixed at 8103 Torr with a incident beam; b is the half width of the peak and h is the
carrier gas of Ar for a total flux of 50 sccm. The sizes of diffraction angle of the peak. The calculated grain sizes are
the target and the substrate were 5 cm and 4 cm in 7.2, 12.1 and 3.5 nm for the growth temperature of 25 8C,
diameter, respectively, and the deposition time was about 6 180 8C and 290 8C, respectively, as shown in Fig. 2. In
min to maintain the ZrC film thickness in the range of fact, the peak broadening results not only from grain size
150200 nm. The growth temperature, Ts, was varied from effect but also from microstrain effect and crystal defects.
25 8C to 290 8C in order to study the effect of the
substrate temperature. The substrate temperature was
measured using a thermocouple directly on the backside
of the sample holder without water cooling. Therefore the
real substrate temperature would be slightly higher if ion
bombardment was taken into account. Subsequently, the
as-deposited thin films were characterized with X-ray
diffraction (XRD) for phase identification, scanning
electron microscopy (SEM, Hitachi S4100) for surface
morphology, and Auger electron spectroscopy (AES-310
D, VG) for composition.
Nanoindentation (Hysitron) is now widely accepted as an
important method for measuring the mechanical properties
of thin films [16] and thus is used to characterize elastic
properties such as modulus, E, and plastic properties such as
hardness, H [16]. Elastic modulus and hardness were
determined from a loaddisplacement data obtained by the
nanoindentation using a Berkovich tip. The indentation
technique for calculating fracture toughness was based on
the measurement of the length of a median-radial crack
produced by the indentation [17]. A Vickers indenter
(Matsuzaua, MXT 70) was used in making the Vickers
indentation.
Scratch test (Romulus IV, QUAD) was chosen for
measuring thin film adhesion for its simplicity and unique-
ness [1820]. During the test, a diamond stylus was drawn
over the film surface. The applied normal force, F n, was
increased progressively to 35 N. The normal force at which
thin films fail, as detected by a sudden increase in acoustic
emission (AE), is the measurement of the critical load value,
L c, corresponding to a complete detachment of the thin film Fig. 1. XRD patterns of the ZrC films deposited on Si at various growth
from the substrate surface. temperatures.
132 C.-S. Chen et al. / Thin Solid Films 479 (2005) 130136

carbon covers the surface of a growing crystal, acting as


new nucleation sites. Thus, the film growth is blocked by a
surface covering layer of an impurity phase [23] and the
film growth proceeds by repeated renucleation. Therefore,
the grain structure reaches an equiaxed grain morphology
in nanometer range, as evidenced from the cross-sectional
SEM image in Fig. 4(a), compared to a typical columnar
grain morphology at Ts=180 8C in Fig. 4(b). Petrov et al.
[24] and Barna and Adamik [25] have reported similar
results on grain refiners.
The preferred orientation of the grains can be expressed
by a (111) texture coefficient, defined as I ZrC (111)/(I ZrC (111)+
I ZrC (200)), as shown in Fig. 2. The ZrC film changes the (111)
texture coefficient from 0.64 (Ts=25 8C) to 1 (Ts=290 8C),
Fig. 2. Texture coefficient and grain size of the ZrC films as a function of while the theoretical texture coefficient of 0.57 corresponds
growth temperature. to a polycrystalline microstructure of randomly oriented ZrC
grains. It implies that ZrC films grown at lower temperatures
However, grain size effect predominates the peak broad- tend to be random, while the preferred orientation is (111)
ening in this case and the accuracy is estimated to be about when grown at even higher temperatures. More (111)
20% when compared with SEM images in Fig. 3. The increased orientation of grains at higher temperatures is
trend of the grain size variation can also be confirmed by related to more greater thermodynamic favorability. The
SEM images as shown in Fig. 3. The commonly observed highly (111) oriented equiaxed grain morphology for Ts=290
increase of grain size with growth temperature up to 180 8C implies that the renucleation process is also highly
8C obviously results from higher atomic mobility and grain orientation dependent. Texture phenomena in equiaxed grain
boundary mobility at elevated temperatures. However, the structures have been reported [26].
grain size is abruptly decreased to 3.5 nm at Ts=290 8C, ZrC lattice parameter decreases toward the ideal ZrC
probably for the following reasons. When the temperature constant with increasing growth temperature as indicated
rises to a temperature where carbon diffusion is many by the shift to larger angles in Fig. 1. The dilation in the
times faster than Zr atom, carbon easily diffuses along the growth direction is generally the result of in-plane
lattice because a large number of vacant octahedral sites compressive stress. Thus, increasing temperature results
are available. In this case, Zr metal acts as a bcatalystQ and in compressive strain relaxation. The best-fitted formula
relating the ZrC lattice parameter to the growth temper-
ature is given by
a 0:4732  2  105 T nm; 25 8CbT b290 8C 2
where a is the lattice parameter of ZrC in nm and T (8C) is
the growth temperature.

Fig. 3. SEM images of the surface of the as-deposited ZrC films at different Fig. 4. Cross-sectional SEM images of the ZrC/Si for (a) Ts=290 8C, (b)
growth temperatures. Ts=180 8C.
C.-S. Chen et al. / Thin Solid Films 479 (2005) 130136 133

pyramidal tip (Berkovich), the projected contact area


(A) in relation to the contact depth (h c) is given by
A 24:5h2c 5
The reduced modulus is given by
 
1=Er 1  m2f =Ef 1  m2i =Ei 6
where E and m are elastic modulus and Poissons ratio and
the subscripts, f and i, represent the film and the indenter,
respectively (for a diamond indenter, E i=1141 GPa, and
m i=0.07).
Doerner and Nix [19] have employed the unloading data
for measuring contact hardness and stiffness values, which
Fig. 5. Schematic representation of loaddisplacement data from a have been proved to be reliable and accurate [19]. By curve
nanoindentation experiment, where P max is the maximum applied load, fitting using a power law to the unloading curve for a range
h max being the maximum displacement, h f being the final penetration depth from 95% to 40%, stiffness can be calculated. Therefore,
and dP/dh is the slope of the unloading curve.
elastic modulus E is obtained from Eqs. (4)(6), and
hardness H is given by the well-known relation:
The X-ray peak shift in Fig. 1 conforms to the in- H Pmax =A 7
troduction of a residual stress, which can be calculated as
Hardness represents a complex combination of the
follows [27].
deformation characteristics of thin films. Fig. 6 shows the
rm  E=2ma0  a0 4=a0  3 variations of hardness and elastic modulus as a function of
the growth temperature for the ZrC films. The hardness
where a 0 and a 0* denote stress-free and stressed lattice
values of the ZrC films range from 17.5 to 30.2 GPa. The
parameters, respectively, while m ZrC is the Poissons ratio of
ZrC hardness decreases with increasing growth temperature
the layer. The values of a 0* can be determined by Braggs
from Ts=25 8C to Ts=180 8C, and then increases slightly to
law from the shift of the ZrC (111) peaks. The residual
Ts=290 8C. The decreased hardness for the films deposited
stress was calculated to be 5.84, 2.44 and 2.25 GPa
in the temperature range of 25 8C to 180 8C is due to stress
for the ZrC films deposited at Ts=25, 180 and 290 8C,
relaxation and larger grain size for higher growth temper-
respectively.
atures. Possible explanations for the larger hardness in the
ZrC films at Ts=290 8C are a combination of smaller grain
3.2. Hardness, elastic modulus and fracture toughness
size, texture strengthening [29] and a nano-size effect,
which is discussed in the following. In the case of lower
Since the film thickness is only about 200 nm,
growth temperatures from 25 8C to 180 8C, the dependence
hardness was determined by nanoindentation to avoid
of hardness on grain size agrees well with the well-known
contributions from the substrate. The indentation depth
HallPetch [14,15] relation as
performed ranged from 5 to 20 nm to ensure the
correctness of this method for measuring mechanical H H0 KH d 1=2 8
properties of the thin films. For measuring hardness and
modulus using the nanoindentation method with a where H is the hardness, d being the average grain diameter,
Berkovich indenter, indentation load, P, with displace- K H being the slope of the straight line drawn through the
ment, h, is continuously recorded during one complete data, and H 0 is the intercept of the line with the coordinate
cycle of loading and unloading [28]. A typical set of axis and corresponds to the hardness expected at a hypo-
loaddisplacement data is shown in Fig. 5, which
illustrates some of the experimental quantities involved
in the analysis. The key measured quantities are the peak
load, P max, the displacement at the peak load, h max and
the initial unloading contact stiffness, S=dP/dh, (i.e., the
slope of the initial portion of the unloading curve), which
is given by
S dP=dh 2=p1=2 Er A1=2 4
where the reduced modulus, E r, accounts for the fact that
the measured displacement includes contributions from Fig. 6. Hardness and elastic modulus of the ZrC films as a function of
both the specimen and the indenter. For an ideal growth temperature.
134 C.-S. Chen et al. / Thin Solid Films 479 (2005) 130136

from AES. The results indicate that the composition changes


very little, implying that the composition effect is insignif-
icant in this experiment.
To summarize the film properties, we have obtained the
maximum hardness of 30.2F0.5 GPa and the highest elastic
modulus of 325F0.5 GPa for the ZrC film deposited at 25
8C and 290 8C, respectively. To compare the properties, the
hardness is higher than 17.5 GPa for ZrN [10] and 27 GPa
for ZrC0.6N0.4 [13]. The results obtained so far imply that
Fig. 7. Hardness of the ZrC as a function of the reciprocal of the square root increasing growth temperature can enhance stiffness, while
of grain size. degrading hardness.
SEM images of the Vickers indentation made of a normal
thetical infinite grain size. By fitting into the experimental load of 0.5 N held for 15 s on the ZrC films are shown in
data based on this relation, the hardness extrapolated to a Fig. 9. It is interesting to note that the cracking profile is like
grain size of 3.5 nm for Ts=290 8C is 53.5 GPa, while the a bloopQ at Ts=25 8C, while the others exhibit similar four
experimental hardness is only 23.2 GPa. We thus found a bradiatedQ lines. This implies that the increase in the
grain size softening effect, which exhibited a negative Hall hardness is accompanied by high brittleness and residual
Petch relation, as shown in Fig. 7. Conrad and Narayan and stresses [36], and the cracking mechanism for Ts=25 8C is
Schiotz et al. [30,31] indicated that when grain size is different from the other samples.
reduced, a larger fraction of atoms belong to grain The measurement of fracture toughness in thin films was
boundaries, and grain boundary sliding becomes easier, developed by Lawn et al. [17], who related the length (c) of
which then leads to material softening. an indentation crack to K Ic as follows,
Elastic modulus is a measure of the stiffness of a  
material. From Fig. 6, the elastic modulus of the ZrC films KIc a E=H 1=2 P=c3=2 ; 9
increases rapidly with increasing growth temperature from
216.2 GPa at Ts=25 8C to 325 GPa at Ts=290 8C. There are where a is an empirical constant depending on the geometry
conflicting results in the literature regarding the variation of of the indenter and P is the indentation load. For the Vickers
elastic modulus E with grain size in nanocrystalline indenter, a is 0.016. Both E and H values have been
materials. Some experimental results indicate a reduction obtained previously. The fracture toughness for the ZrC
in E when grain size is reduced to nanoscale [32,33], while films with various growth temperatures is shown in Fig. 10.
others show almost no grain size effect [34]. However, our It shows that the fracture toughness of the films increases
results tend to support that elastic modulus is less dependent with increasing growth temperature from 25 8C to 180 8C,
on grain size than other effects, presumably preferred then decreases to 290 8C, which is analogous to the trend of
orientation in this case. Elastic modulus is anisotropic in the grain size and just opposite to the hardness behavior.
nature, and thus sample orientation dependent. From Fig. 1, Therefore, the grain size effect is responsible for the
apparently, (111) and (200) peaks coexist, but the relative toughness.
intensity of (111) to (200) monotonically increases with
substrate temperature, suggesting that the elastic modulus of 3.3. Practical adhesion
the ZrC (111) planes might be larger than that of the (200)
planes. The dependence of elastic modulus on crystal The ZrC practical adhesion strength analyzed by
orientation has been reported [35]. scratch testing is characterized by the critical load, L c,
Composition change in the ZrC films with growth
temperature is another factor to consider. Fig. 8 shows the
C/Zr ratio of the films as a function of growth temperature

Fig. 9. SEM images of the Vickers indentation made at a normal load of 0.5
Fig. 8. C/Zr ratio of the films as a function of growth temperature. N on the ZrC films.
C.-S. Chen et al. / Thin Solid Films 479 (2005) 130136 135

which is shown in Fig. 11 as a function of growth


temperature. Each data point in Fig. 11 was averaged over
three samples and five independent measurements were
acquired from each sample. The critical load, and thus
practical adhesion strength, increases initially, reaching a
plateau at Ts=50 8C, and then decreases above Ts=120 8C,
which shows the complex nature of this property.
Practical adhesion is a macroscopic property that
depends on chemical and mechanical bonding, intrinsic
stress, stress gradients, and adhesive failure modes across an
interfacial region. The adhesion failure is composed of crack
initiation and propagation. The crack initiation begins at a
flaw that allows stress concentration or bond weakening,
while the crack propagation occurs by repeated bond
breaking.
Figs. 2 and 11 indicate that the initial increase in L c Fig. 11. Critical load and thermal stress of the ZrC films as a function of
growth temperature.
coincides with the trend on grain size and fracture
toughness, implying that the initial increase is due to less
stress concentration from dislocation pile-ups on grain grain size and thermal stress, respectively, as indicated in
boundary with larger grain size. However, while L c starts to Fig. 11.
decrease at Ts=120 8C, both grain size and fracture
toughness still continue to increase to Ts=180 8C, suggest-
ing that the adhesion may also be degraded by thermal 4. Conclusion
stress. In other words, bond breaking occurs when the
thermal stress exceeds the fracture stress. The creation of ZrC films were deposited by sputtering as a function of
the thermal stress r th in the as-deposited ZrC films is growth temperature and the resulting mechanical properties
attributed to the difference between the thermal expansion were systematically examined in the paper. The results
coefficients of ZrC, a ZrC, and the Si substrate, a Si, as the show that the mechanical properties of the ZrC films are
film/substrate assembly is cooled from the deposition greatly influenced by the growth temperature and can be
temperature, Ts, to the ambient temperature, Tam. The closely related to the microstructure evolution during
thermal stress developed in the ZrC film is approximated sputtering. For the texture characteristics, the relative
by intensity of (111) to (200) from the XRD spectra monotoni-
cally increases with increasing growth temperature, accom-
rth EZrC Ts  Tam aZrC  aSi =1  mZrC 10 panied by an increase of grain size until Ts=290 8C, where
the film structure changes from columnar to equiaxed nano-
where E ZrC is the Youngs modulus of the ZrC films and size grains, which is caused by a renucleation process. The
m ZrC is the Poissons ratio of the layer. The estimated nanocrystalline effect then causes an inverse HallPetch
thermal stresses for Ts=25 8C, 120 8C and 290 8C are 0.02 relation in hardness due to easier grain boundary sliding.
GPa, 1.23 GPa and 4.64 GPa, respectively. It clearly shows For films grown at lower growth temperatures, the elastic
that the thermal stress rises to a significant level compared modulus is degraded, probably due to the preferred
with the residual stress at Ts=120 8C. Therefore, the critical orientation toward more randomly oriented grains with
load for the ZrC films deposited over the temperature range decreasing growth temperature, accompanied by higher
of 25120 8C and 120290 8C would be dominated by the residual stress and higher hardness. When increasing the
growth temperature to Ts=120 8C, the critical load and
fracture toughness are enhanced due to decreasing the local
dislocation pile-ups for larger grain size. The critical load is
maximum when Ts=120 8C, and this decreases with
increasing growth temperature due to larger thermal
stresses introduced in the deposited film.

Acknowledgements

The work is supported by Nation Science Counsel,


Fig. 10. Fracture toughness as a function of growth temperature for the ZrC Taiwan under the project number of NSC91-2116-E006-
films. 060. We are also grateful for the use of the sputter
136 C.-S. Chen et al. / Thin Solid Films 479 (2005) 130136

equipment in the Semiconductor Laboratory, which is [16] M.F. Doerner, W.D. Nix, J. Mater. Res. 1 (1986) 601.
supported and maintained by the Department of Materials [17] B.R. Lawn, A.G. Evans, D.B. Marshall, J. Am. Ceram. Soc. 63 (1980)
574.
Science and Engineering at National Cheng-Kung Univer- [18] P.A. Steinmann, H.E. Hintermann, J. Vac. Sci. Technol., A, Vac. Surf.
sity, Taiwan. Additionally, the authors wish to thank H. L. Films 3 (1985) 2394.
Chen and C. Y. Chen for helpful discussion. [19] P.J. Burnett, D.S. Rickerby, Thin Solid Films 154 (1987) 403.
[20] Y. Xie, H.M. Hawthorne, Surf. Coat. Technol. 155 (2002) 121.
[21] Powder Diffraction File, Joint Committee on Powder Diffraction
Standards, ICDD, USA, 1978, Card 730477.
References [22] B.D. Cullity, Elements of X-ray Diffraction, second ed., Addison-
Wesley Pub. Com., London, 1978, p. 102.
[1] T. Xie, W.A. Mackie, P.R. Davis, J. Vac. Technol., B 14 (1996) 2090. [23] A.D. Gates, J.L. Robins, Thin Solid Films 149 (1987) 113.
[2] S. Mecabih, N. Amrane, Z. Nabi, B. Abbar, H. Aourag, Phys. A 285 [24] I. Petrov, P.B. Barna, L. Hultman, J.E. Greene, J. Vac. Sci. Technol.,
(2000) 392. A, Vac. Surf. Films 21 (2003) 117.
[3] L.E. Toth, Transition Metal Carbides and Nitrides, Academic Press, [25] P.B. Barna, M. Adamik, Thin Solid Films 317 (1998) 27.
New York, 1971. [26] H. Hatherly, W.B. Hutchinson, An Introduction to Textures in Metals,
[4] K. Minato, T. Ogawa, K. Fukuda, H. Sekino, Y. Nozawa, I. Takahashi, Chamcleon Press, London, 1979.
J. Nucl. Mater. 224 (1995) 85. [27] D.S. Rickerby, A.M. Jones, A.J. Perry, Surf. Coat. Technol. 36 (1988)
[5] K. Minato, K. Fukuda, H. Sekino, A. Ishikawa, E. Oeda, J. Nucl. 631.
Mater. 252 (1998) 13. [28] J.L. Loubet, J.M. Georges, O. Marchesini, G. Meille, ASME J. Tribol.
[6] W.A. Mackie, P.R. Davis, IEEE Trans. Electron Devices 36 (1989) 106 (1984) 43.
220. [29] S.E. Hsu, G.R. Edwards, J.C. Shyne, O.D. Sherby, J. Mater. Sci. 12
[7] K. Minato, T. Ogawa, T. Koya, H. Sekino, T. Tomita, J. Nucl. Mater. (1977) 131.
279 (2000) 181. [30] H. Conrad, J. Narayan, Scripta Mater. 42 (2000) 1025.
[8] Q. Zhang, J. He, W. Liu, M. Zhong, Surf. Coat. Technol. 162 (2003) [31] J. Schiotz, F.D.D. Tolla, K.W. Jacobsen, Nature 391 (1998) 561.
140. [32] G.E. Fougere, L. Riester, M. Ferber, J.R. Weertman, R.W. Siegel,
[9] H. Berndt, A.-Q. Zeng, H.-R. Stock, P. Mayer, Surf. Coat. Technol. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 204
74/75 (1995) 369. (1995) 1.
[10] L. DAlessio, R. Teghil, M. Zaccagnino, I. Zaccardo, V. Marotta, D. [33] P.G. Sanders, J.A. Eastman, J.R. Weertman, Acta Mater. 45 (1997)
Ferro, G.D. Maria, A. Santagata, Appl. Surf. Sci. 168 (2000) 284. 4019.
[11] A.A. Volinsky, N.R. Moody, W.W. Gerberich, Acta Mater. 50 (2002) [34] T.D. Shen, C.C. Koch, T.Y. Tsui, G.M. Pharr, J. Mater. Res 10 (1995)
441. 2892.
[12] W.A. Mackie, T. Xie, P.R. Davis, J. Vac. Sci. Technol., B 13 (1995) [35] P. Patsalas, C. Charitidis, S. Logothetidis, Surf. Coat. Technol. 125
2459. (2000) 335.
[13] P.R. Willmott, H. Spillmann, Appl. Surf. Sci. 197/198 (2002) 432. [36] E. Kelesoglu, C. Mitterer, M.K. Kazmanli, M. Urgen, Surf. Coat.
[14] E.O. Hall, Proc. Phys. Soc. Lond., B 64 (1951) 747. Technol. 116 (1999) 133.
[15] N.J. Petch, J. Iron Steel Inst. 174 (1953) 25.

S-ar putea să vă placă și