You are on page 1of 9

Surface and Coatings Technology 125 (2000) 322330

www.elsevier.nl/locate/surfcoat

Hard and superhard nanocomposite coatings


J. Musil *
Department of Physics, University of West Bohemia, P.O. Box 314, 30614 Plzen, Czech Republic

Abstract
This article reviews the development of hard coatings from a titanium nitride film through superlattice coatings to nanocomposite
coatings. Significant attention is devoted to hard and superhard single layer nanocomposite coatings. A strong correlation between
the hardness and structure of nanocomposite coatings is discussed in detail. Trends in development of hard nanocomposite
coatings are also outlined. 2000 Elsevier Science S.A. All rights reserved.
Keywords: Hardness; Magnetron sputtering; Mechanical properties; Nanocomposite coatings; Structure

1. Introduction
Hard coatings have been successfully used for protection of materials and particularly to enhance the life of
cutting tools since the 1970s. Both the technological
process of their production and their properties, i.e.
hardness, wear and oxidation resistance, however, are
continuously being improved. Important milestones in
the development of hard coatings are briefly summarized
in Table 1. This table shows a clear effort (i) to decrease
the temperature T at which hard coatings are formed
and (ii) to improve the properties of hard coatings,
particularly to increase the hardness and oxidation
resistance. The oxidation resistance should be increased
up to approximately 1000C because during high-speed
machining the temperature of the tool tip can reach
1000C and the coating should be stable at such high
temperatures.
As to the hardness, the coatings are usually divided
into two groups: (1) hard coatings having a hardness
<40 GPa, and (2) superhard coatings having a hardness
>40 GPa. Compared to a large number of hard materials, there are only a few superhard materials, i.e. cubic
boron nitride (c-BN ), amorphous diamond-like carbon
(DLC ), amorphous carbon nitride (a-CN ) and polyx
crystalline diamond. Moreover, these superhard materials are thermodynamically unstable. This is a serious
disadvantage which strongly limits their utilization in
some applications. For instance, the high chemical
* Tel.: +420-19-279072; fax: +420-19-279071.
E-mail address: musil@kfy.zcu.cz (J. Musil )

affinity of carbon to iron limits the applicability of


diamond coated cutting tools to machining of aluminum,
their alloys and wood only. Similar problems can be
expected when the c-BN coating is used in cutting of
steels due to the chemical dissolution of boron in iron.
These problems stimulated intensive research in this
field, and recently new superhard materials based on
superlattices and nanocomposites were developed.

2. Hard superlattice coatings


Superlattice coatings are nanometre-scale multilayers
composed of two different alternating layers with a
superlattice period, i.e. the bilayer thickness of two
materials, ranging from 5 to 10 nm. The bilayers of
these superlattices can be metal layers, nitrides, carbides
or oxides of different materials or a combination of one
layer made of nitride, carbide or oxide of one metal and
the second layer made of another metal. According to
the composition of the bilayer, superlattice coatings can
be divided into five groups: (1) metal superlattices, (2)
nitride superlattices, (3) carbide superlattices, (4) oxide
superlattices and (5) nitride, carbides or oxides/metal
superlattices.
Experiments show that metal superlattices exhibit a
relatively low hardness. On the contrary, single-crystal
nitride superlattice coatings are superhard materials with
a hardness ranging from 45 to 55 GPa, e.g. TiN/VN,
56 GPa [15]; TiN/( V Nb )N, 41 GPa [16 ];
0.6 0.4
TiN/NbN, 51 GPa [17]; TiN/Nb, 52 GPa [18,19];
TiN/CN , 4555 GPa [20]; ZrN/CN , 4045 GPa [21];
x
x

0257-8972/00/$ - see front matter 2000 Elsevier Science S.A. All rights reserved.
PII: S0 2 5 7- 8 9 7 2 ( 9 9 ) 0 0 58 6 - 1

J. Musil / Surface and Coatings Technology 125 (2000) 322330

323

Table 1
Important steps in development of hard coatings
Coating

Material

H (GPa)

Main characteristics

Single layer
Single layer
Multilayer
Single layer
Single layer
Single layer
Single layer
Single layer
Superlattices
Single layer
Single layer
Single layer

TiN, TiC, Al O
2 3
TiN, TiC
TiC/TiB
2
c-BN
diamond
TiAlN
DLC
CN
x
TiN/VN, TiN/NbN, etc.
nc-MeN/a-nitride
nc-MeN/metal
Ti Al N
0.4 0.6

21, 28, 21
21, 28

CVD at T around 1000C on cemented carbides


PVD at T550C on steel substrates
About 103 phase boundaries TiC/TiB [2]
2
High chemical affinity of C to iron
Chemical dissolution of B in iron [5]
Oxidation resistance up to 800C [6 ]
Amorphous phase [7]
Substoichiometric (x=0.20.35) turbostratic structure [8,9]
Superlattice period 510 nm [10,11]
Superlattice period 510 nm [10,11]
Nanocomposite [13]
Nanocomposite, oxidation resistance up to 950C [14]

50 [3]
90 [4]
65
5060
~50
~50
~50
~32

TiN /CN, 2050 GPa [22,23]. The overall hardness of


x
the superlattice coating is therefore greater than that of
the materials of the individual components of the bilayer,
e.g. H
#52 GPa, H =21 GPa, H =14 GPa.
TiN/NbN
TiN
NbN
This hardness enhancement is a very complex phenomenon. In spite of this, several models which explain
multilayers strengthening have already been developed
[2427]. The model of Shinn et al. [26 ] shows that (i)
a difference in elastic modulus between the two layer
materials is required to increase the hardness of the
superlattice film, and (ii) the coherency strain at the
interface between the two layers has only a minor effect.
The model of Chu and Barnett [27] is based on restricted
dislocation movement within and between layers in the
superlattice coating. It predicts a peak in hardness when
there is a difference in shear modulus between two layer
materials and sharp interfaces between layers.
So far, less attention has been devoted to carbide
multilayer coatings. These coatings can also be
superhard, up to 55 GPa, e.g. TiC/VC, 52 GPa [28];
TiC/NbC, 4555 GPa [29]. Considerable attention has
been devoted to nitride/metal superlattice coatings. This
combination of bilayer materials, i.e. the hard nitride
with a more ductile metal, makes it possible to improve
the toughness of the coating while retaining its relatively
high hardness (30 GPa), e.g. TiN/Ti, 36.8 GPa [30];
WN/W, 34 GPa [30]; HfN/Hf, 50 GPa [30]; TiN/Ni,
35 GPa [31]; TiN/Ni Cr , 32 GPa [31]; NbN/Mo,
0.4 0.1
33 GPa [32,33]; NbN/W, 30 GPa [32,33]. An improvement in the toughness of the coating increases its adhesion to the substrate, which is of fundamental
importance for coating applications.
At present, practically no data are available on oxide
superlattice coatings. This is probably because the deposition rate of oxide films using both d.c. and r.f. reactive
magnetron sputtering is, compared with that of metals,
too low. Recently, this situation strongly changed.
Considerable progress has been made in magnetron
deposition of oxide films, such as Al O , ZrO , TiO ,
2 3
2
2

etc. using pulsed d.c. sputtering [34,35]. For instance,


the deposition rate for clear Al O films formed using
2 3
this technique can reach 78% of the metal deposition
rate, which is about 25 times the rate obtainable with
r.f. power [1]. Therefore, oxide films can be produced
at economical rates and so we can expect that research
will be intensified for oxide superlattice coatings. For
instance, Sproul already reported on the deposition of
multilayer nanometre-scale oxide films composed of
alternating layers of Al O and ZrO at a high rate onto
2 3
2
glass, silicon, and high-speed steel substrates [1]. Results
are very encouraging because the energy delivered to
the growing film during the pulse operation can stimulate
film crystallization at low deposition temperatures.
In conclusion it is worthwhile to note that superhard
coatings in the form of superlattices represent a very
important milestone in the development of superhard
materials and the understanding of the origin of the
superhardness. However, the maximum hardness of the
superlattice coating is very strongly dependent on the
superlattice period l, see for instance Fig. 3 presented
in Ref. [15]. The strong dependence of H on l may
cause large variations in the coating hardness, H, when
deposited in industrial machines because it is difficult to
ensure the same thickness of all superlattice layers on
all coated objects, particularly when they have a complex
shape. Similar variations in H can also be caused by the
interdiffusion of elements in neighbouring layers at high
service temperatures. These problems can be avoided if
the superlattice coating is replaced with a single-layer
nanocomposite coating.

3. Nanocomposite materials
Materials are composed of grains separated by grain
boundaries. The size of grains in currently produced
materials, which can be called the conventional materials, varies over a wide range from about 100 nm to

324

J. Musil / Surface and Coatings Technology 125 (2000) 322330

several hundred millimetres, corresponding to


monocrystals. This means that the number of atoms in
grains is always considerably greater than that
in boundary regions. The behaviour of such materials
is determined mainly by the bulk of grains in which
dislocations play a decisive role. Properties of these
materials are continuously improved by optimizing their
composition, structure and the technological processes
used for their formation. No fundamental qualitative
changes in properties of the conventional materials can,
however, be expected.
Completely new properties are exhibited by nanocrystalline materials with a grain size of about 10 nm or
less. The behaviour of these materials is determined
mainly by processes in boundary regions because the
number of atoms in the grains is comparable to or
smaller than that in the boundary regions. Under these
conditions dislocations do not exist [36 ], because grain
boundaries prevent their formation, and the boundary
regions play a decisive role in the material deformation.
A new deformation mechanism, called grain boundary
sliding, replaces the dislocation activity which is the
dominant deformation process in conventional materials
[37]. All these facts result in new unique properties of
nanocrystalline materials. In the case where the size of
grains decreases below 5 nm, the participation of atomic
forces in material formation has to be considered and
the formation of nanocrystalline subatomic structures
can be expected [38].
The new unique properties of nanocrystalline materials are the main driving force stimulating their development. These materials can be prepared only by a method
which simultaneously ensures a high rate of nucleation
and a low growth rate of grains. This can be achieved
relatively easily when the nanocrystalline materials are
produced in the form of films. The most suitable method
for the production of nanocrystalline films is magnetron
sputtering.

4. Methods to control the size and orientation of grains in


sputtered films
The main task in the development of nanocomposite
materials is to master control of their growth mechanism. Therefore, further basic processes used for controlling the size and crystallographic orientation of
grains in growing films are given. There are two fundamental processes: (1) low-energy ion bombardment and
(2) mixing process [39,40].
4.1. Low-energy ion bombardment
The process of low-energy ion bombardment controls
the growth mechanism of the film by the energy delivered
to the growing film by bombarding ions. The ion

bombardment is a strongly non-equilibrium process


which heats the growing film at an atomic level.
Therefore, it is called atomic scale heating (ASH ). The
ion bombardment significantly differs from conventional
heating because the kinetic energy of bombarding ions
is transferred into very small areas of atomic dimensions
and then very quickly conveyed into their close vicinity,
i.e. the ASH is accompanied by extremely fast cooling
rates of about 1014 K/s [41]. ASH can replace conventional heating and so produce dense films corresponding
to zone T in the Thornton structural zone model when
sputtering is carried out at low pressures of about 0.1 Pa
and lower [42].
Ion bombardment of the growing film can restrict
the grain growth and permit the formation of nanocrystalline films. The size and crystallographic orientation
of grains can be controlled by the energy and flux of
bombarding ions. This control of the film structure is,
however, accompanied by film heating and is not convenient for all applications.
4.2. Mixing process
The mixing process is based on the addition of one
or several elements to a base, one element material. As
at least two elements are present in the film, alloy films
are formed by this process. The mixing process is an
efficient method convenient for production of nanocrystalline films. Compared with ion bombardment, no substrate bias and heating are necessary to form the films
with nanocrystalline structure. Also, metastable hightemperature phases can be formed on unheated substrates [43]. This is connected with efficient ASH, caused
by condensing sputtered atoms, and subsequent
extremely fast cooling at an atomic level.

5. Nanocrystalline alloy films


The structure of alloy films depends on the amount
and type of elements added to a base, one element
material. Recent experiments carried out in our laboratory show that there are two groups of binary metal
alloy films. The films of the first group are characterized
by
relatively
narrow
X-ray
reflection
lines
( FWHM1). The films of the second group are very
fine grained (nanocrystalline) or X-ray amorphous films
characterized by very broad low-intensity reflections
( FWHM>1). Several examples are given in Table 2.
Here, together with the grain size, d, two characteristic parameters of the alloyed materials, i.e (i) difference
in atomic radii of the alloy elements and (ii) enthalpy
of the alloy formation DH , are also given. These quantif
ties are often used to predict the creation of an amorphous state [5153]. This state is expected to be formed
in the case when (i) the atomic size difference is greater

325

J. Musil / Surface and Coatings Technology 125 (2000) 322330

Table 2
Typical features of X-ray reflection line from selected alloy films sputtered on unheated substrates (T=RT ) at I =1 A, U =U and p =0.50.7 Pa
d
s
fl
Ar
Alloy

2H ()

I (cps)
hkl

FWHM ()

d (nm)

DH (kJ/mol )
f

r (nm)
A

r (nm)
B

Dr/ ra

Reference

Narrow reflections
NiCr (80/20 wt.%)
CrNi (60/40 wt.%)
ZrY (70/30 at.%)
TiSi (90/10 at.%)
CuCr (60/40 at.%)

51.74
46.50
46.56
52.21
51.57

22642
7627
89231
756
1500

0.2995
0.2788
0.2910
0.3770
0.9360

51.97
59.94
49.58
36.16
12.38

4
9
14
65
20

0.1246

0.125

0.003

[44]

0.160
0.144
0.117

0.181
0.1176
0.125

0.123
0.202
0.066

[45]
[46 ]
[47]

Broad reflections
TiCu (50/50 at.%)
TiAl (60/40 at.%)
ZrCu (70/30 at.%)

46.70
45.80
43.55

15
45
135

4.7020
5.1018
9.8495

1.79
1.49
0.93

9.6
57
12

0.144
0.144
0.160

0.117
0.143
0.117

0.207
0.007
0.255

[48]
[49]
[50]

a Dr/ r is the difference in atomic radii, r=(r +r ).


A B

than 0.15 [52] and (ii) the enthalpy of alloy formation


is negative and large. As can be seen from Table 2, these
two materials parameters, Dr/ r and DH , are not,
f
however, sufficient to predict the structure of binary
metal alloy films formed by sputtering. At present, it is
not known which combinations of these two elements
will form films with narrow and broad X-ray reflection
lines, respectively. Despite these problems there exists a
solution which allows the preparation of nanocrystalline
alloy films.
Alloy films with narrow X-ray reflection lines can be
converted into films with broad X-ray reflection lines if
nitrogen is added, i.e. when the nitride of the alloy film
is formed, see e.g. Ref. [44]. This means that either
binary metal alloys or their nitrides form nanocrystalline
films. The structure of the nanocrystalline film can be
controlled by the substrate bias, substrate temperature
and by the amount of nitrogen incorporated into the
film, i.e. by the energy delivered to the growing film.

6. Nanocomposite films based on nitrides of binary metal


alloys
Nanocomposite films consist of at least two phases.
Experiments show that the incorporation of N into the
growing film is not sufficient to produce the nanocomposite film with fully separated phases. To achieve this,
the substrate has to be heated, see for instance the
formation of ZrCuN [50], NiCrN [44] and TiNiN
[54] nanocomposite films.
The formation of nanocomposite structures is connected with a segregation of the one-phase to grain
boundaries of the second phase, and this effect is responsible for stopping of the grain growth. There are,
however, two open questions. (1) What is a minimum
temperature T necessary to start the segregation proseg
cess. and are there some factors which could be used to
decrease T ? (2) What is a thermal stability of nanoseg
composite films? Recent experiments indicate that rare-

earth elements, such as yttrium, drastically reduce the


grain size in metals, see for instance Ref. [55].

7. Hard and superhard nanocomposite coatings


At present, significant effort is devoted to master the
formation of nanocomposite coatings using magnetron
sputtering because this technology can easily be scaled
up for industrial use. The nanocomposite coatings are
produced using so-called selective reactive magnetron
sputtering [56 ]. In this deposition process, one element
of the alloy is converted into nitride and the second
element of the alloy is transported into the growing film
unreacted.
Recently, new hard (<40 GPa) supertough material
of the type nc-TiC/a-C [57] with a remarkable plasticity
(40% during nanoindentation deformation) and new
superhard (40 GPa) materials of the type
nc-MeN/a-Si N [12] with a high elastic recovery (up to
3 4
80%) in the form of nanocomposite coatings were developed; here nc- and a- denote the nanocrystalline and
amorphous phases, respectively, and Me=Ti, W, V, Zr,
etc. are transition metals. Even though both types of
material are composed of nanocrystalline grains embedded in an amorphous matrix, they exhibit completely
different physical properties. This is due to the different
structures of the supertough and superhard nanocomposite coatings.
A systematic investigation carried out on systems
ZrCuN [13] and TiAlN [49] showed that superhard
nanocomposite coatings can be composed not only of
two hard phases as proposed by Veprek et al. [58] but
also in the case when only one phase is hard, e.g. a
nc-ZrN/Cu nanocomposite [13]. This means that there
are two groups of superhard nanocomposite coatings:
1. nc-MeN/nitride (e.g. a-Si N , a-TiB , etc.);
3 4
2
2. nc-MeN/metal (e.g. Cu, Ni, Y, Ag, Co, etc.).
The hardness H of films of both groups can be continuously varied from low values of about 10 GPa to very

326

J. Musil / Surface and Coatings Technology 125 (2000) 322330

high values, achieving up to 5070 GPa. Moreover, it is


worthwhile to note that the hardness of nanocomposite
films is strongly correlated with their structure.
7.1. Structure of superhard nanocomposite coatings
Superhard films with H40 GPa are formed only
when their structure is close to X-ray amorphous. This
structure corresponds to a transition from the crystalline
structure to an amorphous one. The main factors which
govern the formation of films with an X-ray amorphous
structure are: the energy delivered to the growing film,
the substrate temperature T , the type of elements forms
ing the film [their (i) mutual solubility or immiscibility,
(ii) ability to form intermetallic compounds, (iii) chemical affinity, and (iv) binding energy], gap of elements
immiscibility, enthalpy of the alloy formation and the
content of individual phases in the nanocomposite film,
e.g. the content of soft phase in a composite of the type
nc-Me/metal.
The structure of hard (<40 GPa) and superhard
(40 GPa) nanocomposite coatings is significantly
different. A systematic investigation of the correlation
between the hardness and structure in the ZrCuN and
TiAlN nanocomposite films showed that (a) the hard
films are characterized by many reflections from polyoriented grains of both phases, (b) the superhard films
are two-phase nanocomposites one phase of which has
a nanocrystalline structure and the second is X-ray
amorphous, and (c) the maximum hardness is achieved
only when all grains are oriented in the same direction
and the size of grains has an optimum value of approximately several tens of nanometres.
7.2. Classification of hard nanocomposite coatings
Till now, various different types of hard nanocomposite coating have been prepared:
1. nc-MeN/a-nitride, e.g. nc-MeN/a-Si N (Me=Ti, W,
3 4
V ) [58,61], nc-TiN/a-Si N [62];
3 4
2. nc-MeN/nc-nitride, e.g. nc-TiN/nc-BN [58];
3. nc-MeC/a-C, e.g. nc-TiC/DLC [7];
4. nc-MeN/ metal, e.g. nc-ZrN/Cu [13], nc-( Ti,Al )/AlN
[49,60], nc-CrN/Cu [47];
5. nc-MeN or MeC/a-boron compounds, e.g.
nc-Ti(B,O)/quasi-a-(TiB , TiB and B O ) [63], Ti
2
2 3
BC [64];
6. nc-WC+nc-WS /DLC [65];
2
7. nc-MeC/a-C+a-nitride,
e.g.
nc-Mo C/a-C+a2
Mo N [59].
2
This survey shows that all hard nanocomposite coatings
contain one or two hard crystalline phases. The second
phase is more complicated. It is either amorphous (e.g.
a-Si N ) or crystalline (e.g. nc-BN [58]). Sometimes, the
3 4
content of the second phase in the nanocomposite
coating is very low at approximately 12 wt.% (e.g. Cu

in nc-ZrN/Cu films [13]). In such a case, it is very


difficult, without a HRTEM investigation, to determine
if the second phase is crystalline or amorphous because
the X-ray reflections from a small quantity of grains are
below the detection limit. This means that the second
phase can be incorrectly interpreted as amorphous when
determined from X-ray diffraction analysis only. On the
basis of these facts hard nanocomposite coatings can be
divided into two main groups: (1) crystalline/amorphous
nanocomposites
and
(2)
crystalline/crystalline
nanocomposites.
At present, no definite methodology exists on how to
select the combination of elements to produce films with
nanocrystalline and/or X-ray amorphous structure. Such
films can be formed from nitrides of alloys composed
of elements which exhibit a wide miscibility gap and
contain one element forming hard nitride, for instance
nitrides of binary alloys formed of immiscible elements
such as CuCr, ZrY, etc. or transition metal
nitride/boride and nitride/carbide systems [64]. A miscibility gap between intermetallics is also sufficient to
produce hard nanocomposite coating, e.g. ZrCuN film.
There is also a possibility to produce the nanocomposite
coating from nitrides of alloys whose elements form a
solid solution, e.g. Ti Al N film. This is enabled by
1x x
the existence of a gap in the alloy composition x where
the structure of the film is quasi-X-ray amorphous [66 ].
7.3. Mechanical properties of hard nanocomposite
coatings
A short survey of basic mechanical characteristics of
some recently prepared nanocomposite films is given in
Table 3. For comparison, the characteristics of some
selected hard bulk materials and hard amorphous
carbon films are also given in this table.
The hard nanocomposite coating is characterized not
only by its hardness H but also by its Youngs modulus
E and elastic recovery W . The determination of these
e
quantities for thin films is, however, difficult because
they strongly vary with the load L used in their measurement. Despite these problems, Figs. 1 and 2 display H
as a function E1=E/(1n2) and the elastic recovery
W as a function H, because these dependencies were
e
measured under the same conditions in one laboratory
and on different nanocomposite systems. The high hardness of the material is only one parameter which ensures
scratch and abrasion resistance. Protective overcoat
films must be highly resistant also to plastic deformation
during contact events. This requires a low Youngs
modulus E since, according to Johnson analysis, the
load P needed to initiate plastic deformation when a
y
rigid sphere of radius r is pressed into the coating is
proportional to H3/E2 [74]. The ratio H3/E2 is a parameter which controls the resistance of materials to plastic
deformation. The likelihood of plastic deformation is

327

J. Musil / Surface and Coatings Technology 125 (2000) 322330


Table 3
Comparison of hard bulk materials, hard single layer films and selected hard and superhard nanocomposite coatings
Material
Bulk materials
Diamond
Boron
Sapphire
Amorphous films
DLC
a-C (cathodic arc)
Nanocomposite single layer films
nc-TiN/Si N
3 4
nc-TiN/BN
nc-W N/a-Si N
2
3 4
TiBC
TiBN
Zr Cu N
98 2
W Ni N
86.7 8.3 5
W Si N
68 14 18
nc-Mo C/a-(C+Mo N )
2
2
Ti Al N
45 55
Ti Al N
60 40
ZrYN
CrNiN
Ti Si N
75 25
Ti C
( TiC/a-C )
0.32 0.68

H (GPa)

E1=E/(1n2) (GPa)

100
35
30

1050
470
441

65
>59

550
>395

48
69
51
71
54
54
55
45
49
47
40
41
32
29
32

~565
585
560
486
~500
394
510

440
409
650
319
253
256
370

W (%)
e

H3/E12

d (nm)

Reference

0.91
0.19
0.14

[67]
[68]
[69]

8090

0.91
~1.3

[7]
[74]

80.5

81

67
74

77
74
67
60

~0.34
0.96
0.42
1.52
0.63
1.03
0.64

0.61
0.62
0.15
0.66
0.50
0.36
0.239

4.5
9
3.5
~1
~1
35

27
30

1050

[70]
[58]
[58,70,71]
[64]
[64]
[13]
[75]
[61]
[59]
[49]
[14]
[45]
[72]
[46 ]
[73]

a Denotes data not given in the references or not determined.

Fig. 1. H as a function of E/(1n2) for ZrCuN, ZrYN, CrNiN, TiSiN and TiAlN nanocomposite films sputtered at different deposition
conditions, i.e. T , U , i and p .
s s s
N2

reduced in materials with high hardness and low modulus [74]. In general, a low modulus is also desirable as
it allows the given load to be distributed over a wider
area. The ratio H3/E2 of nanocomposite films spreads
over a very wide range from about 0.15 to 1.52, see
Fig. 3 and Table 3. Data given in Table 3 and in Figs. 1
3 show that the elastic recovery W and the resistance
e

of materials to plastic deformation can be controlled by


the film hardness H and its elastic modulus E.
A spread of experimental points around the straight
lines in Figs. 13 is connected with the variation in the
film structure induced particularly by different (i) deposition conditions and (ii) chemical composition of the
film. Therefore, H and E of nanocomposite coatings

328

J. Musil / Surface and Coatings Technology 125 (2000) 322330

Fig. 2. Elastic recovery upon indentation W as a function of H for ZrCuN, ZrYN, CrNiN, TiSiN and TiAlN nanocomposite films
e
sputtered at different deposition conditions, i.e. T , U , i and p .
s s s
N2

Fig. 3. The ratio H3/E2 characterizing the resistance of the material to plastic deformation as a function of H for ZrCuN, ZrYN, CrNiN,
TiSiN and TiAlN nanocomposite films sputtered at different deposition conditions, i.e. T , U , i and p .
s s s
N2

prepared by plasma CVD and magnetron-assisted laser


deposition processes differ from those prepared by magnetron sputtering. A further systematic investigation is
necessary to master the preparation of materials with
prescribed properties.

8. Conclusions
The main results obtained in the development of hard
and superhard nanocomposite coatings can be summarized as follows.
1. Hard (<40 GPa) coatings are characterized by a high

plastic deformation increasing with decreasing H up


to about 70% for H#10 GPa.
2. Superhard (40 GPa) coatings are characterized by
a high elastic recovery increasing with increasing H
up to about 85% for H#70 GPa.
3. There are two types of superhard nanocomposite
film: (1) nc-MeN/a-nitride and (2) nc-MeN/metal.
This means that the superhard film can be composed
either of two hard phases or of one hard phase and
the second soft phase.
4. The hardness of nanocomposite coatings correlates
well with their structure. A decrease of the film
crystallinity, characterized by a decrease of the inten-

J. Musil / Surface and Coatings Technology 125 (2000) 322330

sity of X-ray reflection lines and by an increase of


their broadening, results in an increase of hardness.
5. The structure of superhard films is close to X-ray
amorphous.
The nanocomposite films are fascinating materials since
novel structures and new physical properties are
expected. The development of nanocomposite films is,
however, only at the beginning. Many problems are
unsolved and many questions remain open. Therefore,
further systematic research in this field is required.
Further research is focusing on the following problems:
(i) understanding of the origin of superhardness; (ii)
correlation between the mechanical parameters of materials and process parameters; (iii) dramatic changes in
crystallographic orientation of grains in alloy films, their
hardness and elastic recovery induced by incorporation
of nitrogen; (iv) formation of nanocomposite films with
controlled hardness, elastic modulus and elastic recovery
and new functional properties; and (v) investigation of
materials having very small grains of about 1 nm in size.

Acknowledgements
The author would like to thank his Ph.D. students
H. Hruby, I. Leipner, H. Polakova, F. Regent, Z.
Soukup and P. Zeman for the preparation of many
nanocomposite coatings of different chemical composi erstvy and Dr. M. Kolega for X-ray
tion and Dr. R. C
characterization of these coatings. This work was supported in part by the Grant Agency of the Czech
Republic under Project No. 106/96/K245 and by the
Ministry of Education of the Czech Republic under
Project Nos. VS 96/059 and ME 173/1999.

References
[1] W.D. Sproul, Science 273 (1996) 889.
[2] H. Holleck, Ch. Kuhl, H. Schultz, J. Vac. Sci. Technol. A 3 (6)
(1985) 234.
[3] H. Holleck, J. Vac. Sci. Technol. A 4 (6) (1986) 2661.
[4] C.A. Brookes, in: J.E. Field (Ed.), The Properties of Diamond,
Academic Press, New York, 1979, p. 383.
[5] B.M. Kramer, P.K. Judd, J. Vac. Sci. Technol. A 3 (1985) 2439.
[6 ] W.D. Munz, J. Vac. Sci. Technol. A 4 (6) (1986) 2717.
[7] A.A. Voevodin, J.S. Zabinski, Diamond Relat. Mater. 7 (1998)
463.
[8] H. Sjostrom, S. Stafstrom, M. Boman, J.-E. Sundgren, Phys. Rev.
Lett. 76 (1996) 56.
[9] H. Sjostrom, L. Hultman, J.-E. Sundgren, S.V. Hainsworth, T.F.
Page, G.S.A.M. Teunissen, J. Vac. Sci. Technol. A 14 (1996) 56.
[10] U. Helmersson, S. Todorova, S.A. Barnett, J.-E. Sundgren, L.C.
Markert, J.E. Greene, J. Appl. Phys. 62 (1987) 481.
[11] X. Chu, S.A. Barnett, M.S. Wong, W.D. Sproul, Surf. Coat. Technol. 57 (1993) 13.
[12] S. Veprek, S. Reiprich, S. Li, Appl. Phys. Lett. 66 (20) (1995)
2640.
[13] J. Musil, P. Zeman, H. Hruby, P. Mayrhofer, ZrN/Cu nanocom-

329

posite film novel superhard material, Proc. ICMCTF99, April


1218, San Diego, CA (1999), paper no. B1-2-9.
[14] Y. Min, Y. Makino, M. Nose, K. Nogi, Thin Solid Films 228
(1999) 204.
[15] U. Helmersson, S. Todorova, S.A. Barnett, J.-E. Sundgren, L.C.
Markert, J.E. Greene, J. Appl. Phys. 62 (1987) 481.
[16 ] P.B. Mirakami, L. Hultman, S.A. Barnett, Appl. Phys. Lett. 57
(25) (1990) 2654.
[17] M. Shinn, L. Hultman, S.A. Barnett, J. Mater. Res. 7 (4)
(1992) 901.
[18] X. Chu, M.S. Wong, W.D. Sproul, S.L. Rohde, S.A. Barnett,
J. Vac. Sci. Technol. A 10 (4) (1992) 1604.
[19] X. Chu, S.A. Barnett, M.S. Wong, W.D. Sproul, Surf. Coat. Technol. 57 (1993) 13.
[20] D. Li, X.W. Lin, S.C. Chen, V.P. Dravid, Y.W. Chung, M.S.
Wong, W.D. Sproul, Appl. Phys. Lett. 68 (9) (1996) 1211.
[21] M.L. Wu, W.D. Qian, Y.W. Chung, Y.Y. Wang, M.S. Wong,
W.D. Sproul, Thin Solid Films 308/309 (1997) 113.
[22] H. Jensen, J. Sobota, G. Sorensen, Surf. Coat. Technol. 94/95
(1997) 174.
[23] H. Jensen, J. Sobota, G. Sorensen, J. Vac. Sci. Technol. A 16 (3)
(1998) 1180.
[24] J.E. Kryanowski, Mater. Res. Soc. Symp. Proc. 239 (1992) 509.
[25] S.A. Barnett, in: M. Fracombe, J.A. Vossen ( Eds.), Physics of
Thin Films, Academic Press, New York, 1993, p. 1.
[26 ] M. Shinn, L. Hultman, S.A. Barnett, J. Mater. Res. 7 (1992) 901.
[27] X. Chu, S.A. Barnett, J. Appl. Phys. 77 (1995) 4403.
[28] B.G. Wendler, P. Kula, K. Jakubowski, S. Fauvry, Ph. Kapsa, L.
Vincent, D. Heper, Proc. 36th Tagung der Deutschen Gesellschaft
fur Obrflachen- und Galvanotechnik, October 79, Schwabisch
Gmund, Germany (1998) 611.
[29] B.G. Wendler, M. Molinaro, Proc. 10th Int. Summer School on
Modern Plasma Surface Technology, May 1014, Mielno, Poland
(1998) 423436.
[30] K.K. Shih, D.B. Dove, Appl. Phys. Lett. 61 (6) (1992) 654.
[31] X. Chu, M.S. Wong, W.D. Sproul, A.S. Barnett, Mechanical
properties and microstructures of polycrystalline metal/ceramic
superlattices, TiN/Ni and TiN/Ni Cr , paper presented at
0.9 0.1
ICMCTF-20, San Diego, CA, 1993.
[32] A. Madan, X. Chu, S.A. Barnett, Appl. Phys. Lett. 68 (16)
(1996) 2198.
[33] A. Madan, Y.-Y. Wang, S.A. Barnett, C. Engstrom, H.
Ljungcrantz, L. Hultman, M. Grimsditch, J. Appl. Phys. 84 (2)
(1998) 776.
[34] J.M. Schneider, W.D. Sproul, A.A. Voevodin, A. Matthews,
J. Vac. Sci. Technol. A 15 (3) (1997) 1084.
[35] J.M. Schneider, W.D. Sproul, Reactive pulsed dc magnetron sputtering and control, Handbook of Thin Film Process Technology,
IOP, Bristol, 1998, pp. A5.1:1A5.1:12.
[36 ] S. Veprek, S. Reiprich, Thin Solid Films 268 (1995) 64.
[37] R.W. Siegel, Nanophase materials: Structure, defects and properties, Proc. ATC Int. Symp. (ACTA), October 2930, Tokyo,
Japan (1996) 1.
[38] P.Yu. Butyagin, Colloid J. 59 (1997) 112.
[39] J. Musil, J. Vlcek, Czech. J. Phys. 48 (10) (1998) 1209.
[40] J. Musil, J. Vlcek, Magnetron sputtering of alloy and alloy-based
films, Thin Solid Films 343344 (1999) 47.
[41] K.D. Leedy, J.M. Rigsbee, J. Vac. Sci. Technol. A 14 (1996) 2202.
[42] J. Musil, Vacuum 50 (3/4) (1998) 363.
[43] J. Musil, A.J. Bell, J. Vlcek, T. Hurkmans, J. Vac. Sci. Technol.
A 14 (4) (1996) 2247.
[44] J. Musil, F. Regent, J. Vac. Sci. Technol. A 16 (6) (1998) 3301.
[45] J. Musil, H. Polakova, Hard nanocomposite ZrYN coatings.
Correlation between hardness and structure, 2nd AsianEuropean
Int. Conf. on Plasma Surface Engineering, September 1519, Beijing, China (1999), Paper No. Sat-0A3-2.

330

J. Musil / Surface and Coatings Technology 125 (2000) 322330

[46 ] J. Musil, H. Polakova, V. Cibulka, Czech. J. Phys. 49 (3)


(1999) 359.
[47] J. Musil, I. Leipner, M. Kolega, Surf. Coat. Technol. 115 (1)
(1999) 32.
[48] J. Musil, Z. Soukup, unpublished results.
[49] J. Musil, H. Hruby, Superhard nanocomposite Ti Al N films
1x x
prepared by magnetron sputtering, Proc. 14th Int. Symp. on
Plasma Chemistry, August 26, Praha, Czech Republic, Vol. III
(1999) 16051610.
[50] J. Musil, P. Zeman, Vacuum 52 (1999) 269.
[51] W.L. Johnson, Progr. Mater. Sci. 30 (1986) 81.
[52] G. Weigang, H. Hecht, G. von Minnigerode, Z. Phys. B 96
(1995) 349.
[53] J.R. Ding, D.Y. Che, H.B. Yhang, K. Tao, B.X. Liu, Appl. Phys.
Lett. 60 (8) (1992) 994.
[54] M. Misina, J. Musil, S. Kadlec, Surf. Coat. Technol. 110 (1998)
168.
[55] Z. Liu, R. Singh, K. Poole, R.J. Diefendorf, J. Harris, K. Cannon,
J. Vac. Sci. Technol. B 15 (1997) 1990.
[56 ] L. May, W.R. Allen, A.L. Glower, J.C. Mabon, J. Vac. Sci. Technol. B 13 (1995) 361.
[57] A.A. Voevodin, S.V. Prasad, J.S. Zabinski, J. Appl. Phys. Lett.
82 (2) (1997) 855.
[58] S. Veprek, P. Nesladek, A. Niederhofer, F. Glatz, M. Jlek, M.
Sma, Surf. Coat. Technol. 108/109 (1998) 138.
[59] M. Benda, J. Musil, Vacuum 55 (1999) 171.
[60] Y. Min, Y. Makino, M. Nose, K. Nogi, Thin Solid Films 339
(1/2) (1999) 238.
[61] C. Louro, A. Cavaliero, Hardness versus structure in WSiN
sputtered coatings, 6th Int. Conf. on Plasma Surface Engineering,
PSE-98, September 1418, Garmisch-Partenkirchen, Germany
(1998), paper no. ThWB5.

[62] M. Diserens, J. Patscheider, F. Levy, Surf. Coat. Technol. 108/


109 (1998) 241.
[63] C. Mitterer, P. Losbichler, F. Hofer, P. Warbichler, W. Gissler,
Vacuum 50 (3/4) (1998) 313.
[64] C. Mitterer, P.H. Mayrhofer, M. Beschliesser, P. Losbichler, P.
Warbichler, F. Hofer, P.N. Gibson, W. Gissler, H. Hruby,
J. Musil, J. Vlcek, Proc. ICMCTF99, April 1216, San Diego,
CA (1999), paper no. BP-16.
[65] A.A. Voevodin, J.P. ONeill, J.S. Zabinski, Surf. Coat. Technol.
(1999) submitted.
[66 ] U. Wahlstrom, L. Hultman, J.-E. Sundgren, F. Abidi, I. Petrov,
J.E. Greene, Thin Solid Films 235 (1993) 62.
[67] J.E. Field, The Properties of Diamond, Academic Press,
London, 1979.
[68] S.M. Gorbatkin, R.L. Rhoades, T.Y. Tsui, W.C. Oliver, Appl.
Phys. Lett. 65 (1994) 2672.
[69] W.C. Oliver, G.M. Pharr, J. Mater. Res. 7 (1991) 1564.
[70] S. Veprek, Thin Solid Films 317 (1998) 449.
[71] S. Veprek, Thin Solid Films 297 (1997) 145.
[72] J. Musil, F. Regent, Structure and microhardness of magnetron
sputtered nanocrystalline CrNiN films, Proc. 14th Int. Symp. on
Plasma Chemistry, August 26, Praha, Czech Republic, Vol. III
(1999) 16171622.
[73] A.A. Voevodin, J.S. Zabinski, J. Mater. Sci. 33 (1998) 319.
[74] T.Y. Tsui, G.M. Pharr, W.C. Oliver, C.S. Bhatia, R.L. White, S.
Anders, A. Anders, I.G. Brown, Mater. Res. Soc. Symp. Proc.
383 (1995) 447.
[75] A. Cavaleiro, B. Trindale, M.T. Viera, Deposition and characterization of fine-grained WNiC/N ternary films, 6th Int. Conf.
on Plasma Surface Engineering, PSE-98, Garmisch-Partenkirchen, Germany, Book of Abstracts (1998) 265.

You might also like