You are on page 1of 141

Macro- and Micromechanical and Tribological Properties 229

6
Macro- and
Micromechanical and
Tribological Properties
Bharat Bhushan and Bal K. Gupta
1.0

INTRODUCTION

Mechanical properties of the solid surfaces and thin films are


of interest as the mechanical properties affect friction and wear
performance of interfaces. Among the mechanical properties of
interest, one or more of which can be obtained using commercial and
specialized hardness testers are: elastic-plastic deformation behavior, hardness, Youngs modulus of elasticity, scratch resistance,
film-substrate adhesion, residual stresses, time-dependent creep and
relaxation properties, fracture toughness, and fatigue. For macro
devices, bulk properties are important. However, physical contacts at
sliding interfaces in microelectromechanical systems (MEMS), such
as microsensors, microactuators, micromotors, microgear trains,
microvalves, and magnetic recording heads typically occur at very
low loads, thus, friction and wear of sliding interfaces performance is
primarily controlled by the physical and chemical properties of a few
surface atomic layers.[38][48][56][114][145][275] To protect these devices
against wear and corrosion and to achieve a low coefficient of
friction, ultrathin coatings are sometimes deposited on the moving
components. For understanding and/or estimating the functional
behaviors of mechanical devices of small sizes ranging from a
229

230 Handbook of Hard Coatings

couple of tens of microns to a few millimeters, and involving very


low loads on the orders of a few tens of nanonewtons, to a few
millinewtons, measurements of mechanical and tribological properties of thin coatings or the near-surface region of a bulk material on
micro- and nanoscales are of crucial importance. In the last decade,
a variety of apparatuses have been developed to measure the mechanical and tribological properties on micro- to nanoscales. Some
of these common apparatuses are: depth-sensing mechanical properties microprobe (Nanoindenter), atomic force microscope (AFM),
and friction force microscope (FFM). These apparatuses operate at
very low loads and yield properties of the near surface region.
In this chapter, we will present an overview of techniques, and
apparatuses used for measurement of mechanical and tribological
properties on the macro- to nanoscale. The diamond and amorphous
carbon coatings used extensively for controlling friction, wear, and
corrosion are chosen to exemplify the mechanical and tribological
measurements made on the micro- and nanoscale.

2.0

MEASUREMENT OF MECHANICAL PROPERTIES

Hardness implies the resistance to local deformation. Most


commonly used hardness measurements are: scratch hardness and
static indentation hardness.[272] Scratch hardness depends on the
ability of one material to scratch another, or to be scratched by
another solid. The solid and thin film surfaces are scratched by a
sharp stylus made of hard material typically diamond, and either the
loads required to scratch or fracture the surface, or delaminate the
film or the normal/tangential load-scratch size relationships are used
as a measure of scratch hardness and/or interfacial adhesion.[2][30][37][56][76][131][133][139][147][199][223][224][261][272][273] [287][302]
The methods most widely used in determining the hardness of
materials are (quasi) static indentation methods. Indentation hardness is essentially a measure of their plastic deformation properties
and only to a secondary extent with their elastic properties. In the

Macro- and Micromechanical and Tribological Properties 231

indentation methods, a spherical, conical, or pyramidal indenter is


forced into the surface of the material which forms a permanent
(plastic) indentation in the surface of the material to be examined. The hardness number (GPa or kg/mm2) equivalent to the
average pressure under the indenter, is calculated as the applied
normal load divided by either the curved (surface) area (Brinell,
Rockwell, and Vickers hardness numbers), or the projected
area (Knoop and Berkovich hardness numbers) of the contact between the indenter and the material being tested, under
load.[15][33][39][46][60][151][204][219][272][273][298]
In a conventional indentation hardness test, the contact area is
determined by measuring the indentation size by a microscope after the sample is unloaded. At least, for metals, there is a
little change in the size of the indentation on unloading, so that the
conventional hardness test is essentially a test of hardness under
load, although it is subjected to some error due to varying elastic
contraction of the indentation. [262] More recently, in depthsensing indentation hardness tests, the contact area is determined
by measuring the indentation depth during the loading/unloading
cycle.[38][39][60][65][101][207][211][218][228][230][282][304] Depth measurements have, however, a major weakness arising from piling-up,
and sinking-in of material around the indentation. The measured
indentation depth needs to be corrected for the depression (or the
hump) of the sample around the indentation, before it can be used for
calculation of the hardness.[99][100][111][210][218][230][304] Youngs modulus of elasticity is the slope of the stress-strain curve in the elastic
regime. It can be obtained from the slope of the unloading
curve.[210][218][230] Hardness data can be obtained from a depth sensing instrument without imaging the indentations with high reproducibility. This is particularly useful for small indents required for
hardness measurements of extremely thin films.
In addition to measurements of hardness and Youngs modulus
of elasticity, static indentation tests have been used for measurements of a wide variety of material properties such as elastic-plastic
deformation behavior,[100][111][127][130][218][228][263] flow stress,[272]
scratch resistance and film-substrate adhesion, [2][30][37][39][41][50]

232 Handbook of Hard Coatings

[56][57][76][131][133][139][147][199][223][224][261][263][272][287][301][302][305][306] re-

sidual stresses,[168][183][270] creep,[19][84][144][178][206][290][297] stress relaxation,[85][135][136][166][167][191][207][234][235][302] fracture toughness and


brittleness,[79][172][174][194][220][232] and fatigue.[177][307]
The extended load range of static indentation hardness testing is
shown schematically in Fig. 1. We note that only the lower micro- and
ultra-microhardness, or nanohardness load range can be employed
successfully for measurements of extremely thin (submicron thick)
films. The intrinsic hardness of surface layers or thin films becomes
meaningful only if the influence of the substrate material can be
eliminated. It is therefore generally accepted that the depth of indentation should never exceed 30% of the film thickness.[15] The minimum load for most commercial microindentation testers available is
about 10 mN. Loads on the order of 50 N to 1 mN are desirable if
the indentation depths are to remain few tens of a nanometer. In this
case, the indentation size sometimes reaches the resolution limit of a
light microscope and it is almost impossible to find such a small
imprint if the measurement is made with a microscope after the
indentation load has been removed. Hence, either the indentation
apparatuses are placed in situ a scanning electron microscope (SEM), or
in situ indentation depth measurements are made. The latter measurements, in addition, would offer the advantages to observe the
penetration process itself. In viscoelastic/viscoplastic materials, since
indentation size changes with time, in situ measurements of the
indentation size is particularly useful, which can, in addition, provide
more complete creep and relaxation data of the materials.

Figure 1. Extended load range of static indentation hardness testing.

Macro- and Micromechanical and Tribological Properties 233

In this section, we review various hardness test apparatuses for


measurements of mechanical properties of surface layers of bulk
materials and extremely thin films (submicron in thickness) on
macro- to nanoscale.
2.1

Apparatuses for Hardness Measurement on Macroscale

Macrohardness measurements are generally used to determine


bulk hardness. These include Brinell and Rockwell hardness
testers.[46][272]
Brinell hardness is determined by forcing a hardened sphere
under a known load into the surface of a material, and measuring the
diameter of the indentation left after the test. The load is maintained
for about 30 seconds, and the diameter of indentation at the surface is
measured with an optical microscope (10) after the ball has been
removed. The Brinell hardness number (BHN) or simply the Brinell
number, is defined in kilograms per square millimeter, as the ratio of
the load used to the actual surface area of the indentation, which is, in
turn, given in terms of the imposed load W, ball diameter D, and the
indentation diameter d or depth of indentation h:

Eq. (1)

BHN =

2W

D( D D 2 d 2 )

W
Dh

Hardened steel bearing balls may be used up to 450 BHN, but


beyond this hardness, especially treated steel or tungsten carbide
balls should be used to avoid flattening of the indenter. The standard
size ball is 10 mm, and the standard loads are 3000, 1500, and 500
kg, with 250, 125, and 100 kg sometimes used for softer materials,
(for more details see Method of Brinell Hardness Testing, ASTM
E1061T). The load and the ball diameter should be adjusted to keep
the ratio d/D within the range of 0.3 to 0.5. The nearest edge of the
specimen should be no closer than 2.5 impression diameters, and the
thickness should be more than one diameter. Since d/D is normally
less than 0.5, this means that for a 10-mm ball, the uninterrupted

234 Handbook of Hard Coatings

width and depth of the specimen may have to be as great as 25 and 10


mm, respectively, to avoid spurious side and bottom effects.
Use of the curved area in the Brinell test was originally introduced to try to compensate for the effects of work-hardening. It is
cumbersome to use the surface area of indentation, though, so to
overcome this disadvantage, Meyer hardness is defined in kilograms
per square millimeter, as the load used divided by the projected area of
indentation:

Eq. (2)

Meyer hardness =

4W
d2

In the Rockwell Hardness test, the indenter may be either a


steel ball of some specified diameter or a spherical-tipped conical
diamond of 120 angle with a 0.2-mm tip radius, called a Brale. A
minor load (or perload) of 10 kg is first applied, and this causes an
initial penetration that holds the indenter in place. At this point, the
dial is set to zero, and the major load is applied. The standard loads
are 150, 100, and 60 kg. Upon removal of the major loads, the
reading is taken while the minor load is still on. The hardness number
may then be read directly from the scale that measures penetration.
The Rockwell hardness number is defined by an arbitrary equation of
the following form:
Eq. (3)

R = C1 C 2 t

where C1 and C2 are constants for a given indenter size, shape, and
hardness scale, and t is the penetration depth in millimeters between the major and minor loads. Although Rockwell hardness
increases with Brinell hardness, the two are not proportional and the
dimensions of the Rockwell hardness are not force per unit area. In
fact, the Rockwell hardness number cannot be assigned any dimensions, since it is defined in an arbitrary Eq. (3).
A variety of combinations of indenters and major loads are
possible; the most commonly used are HRB or R B, which uses

Macro- and Micromechanical and Tribological Properties 235

1.59-mm diameter (1/16 inch) ball as the indenter and a major load
of 100 kg; HRC, or RC, which uses a cone as the indenter and a major
load of 150 kg; and HRA or RA, which uses a cone as the indenter
and a major load of 60 kg. Rockwell B is used for soft metals and
Rockwell C and A are used for hard metals.
2.2

Apparatuses for Hardness Measurement on Microscale

Microhardness measurements allow the indenter to be shallow


and of small volume so as to measure the hardness of brittle materials
or thin materials or coatings. The apparatuses include Vickers and
Knoop hardness testers.[272] Both testers use highly polished diamond pyramidal indenters. The Vickers indenter is a diamond in the
form of a square pyramid with face angles of 136 (corresponding to
edge angles of 148.1) (Fig. 2a), and relatively low loads varying
between 1 and 120 kg are used. The Knoop indenter is a rhombicbased pyramidal diamond with longitudinal edge angles of 172.5
and 130 (Fig. 2b). In general, the loads used in Knoop testers vary
from about 0.2 to 4 kg. Smaller loads (as low as 1 to 25 g) also may
be used in Vickers and Knoop hardness measurements, the lengths of
the diagonals of the indentation are measured using a medium
powered microscope after the load is removed. If d is the mean value
of a diagonal in millimeters and W is the imposed load in kilogram,
Vicker hardness number (V or HV), sometimes called diamondpyramid hardness (DPH), is given by the load divided by the actual
surface area, that is,

Eq. (4)

HV =

1.8544W
d2

If l is the long diagonal in millimeters, Knoop hardness number


(K or HK) is given by the load divided by the projected area of the
indentation, that is,

Eq. (5)

HK =

14.229W
l2

236 Handbook of Hard Coatings

(b)

Figure 2. Geometry and indentation with (a) a Vickers indenter and (b) a Knoop
indenter.

The depth to the diagonal for the Vickers indenter is about 7


and depth to the long diagonal (used for the Knoop indenter) is about
30.5. The advantage of the Knoop indenter over the Vickers indenter
in microhardness testing, lie in the fact that a longer diagonal is
obtained for a given depth of indentation or a given volume of
material deformed. The Knoop indenter is thus advantageous when
shallow specimen or thin, hard layers must be tested. The Knoop
indenter is also desirable for brittle materials (such as glass or diamond), in which the tendency for fracture is related to the area of
stressed material.

Macro- and Micromechanical and Tribological Properties 237

2.3

Apparatuses for Hardness Measurement on Nanoscale

In this section, we review nanoindentation hardness apparatuses in which the indent is imaged after the load has been removed,
as well as the depth-sensing indentation apparatuses in which the
load-indentation depth are continuously monitored during the loading and unloading processes. Earlier work by Alekhin, et al.,[6]
Ternovskii, et al.,[279] and Bulychev, et al.,[72] led to the development
of depth-sensing apparatuses. Prototype depth-sensing hardness apparatuses developed by several research groups are reviewed by
Bhushan.[38][39] A commercial depth-sensing nanoindentation hardness test apparatus manufactured by Nano Instruments Inc., is extensively used and is described in detail.
Nanoindentation Hardness Apparatuses With Imaging of
Indents After Unloading. For completeness, we first describe a
commercially available microindentation hardness apparatus (Model
No. Micro-Duromet 4000), that uses a built-in light optical microscope for imaging of indents after the sample is unloaded. It is
manufactured by C. Reichert Optische Werke AG, A-1171, Vienna,
Box 95, Austria.[233] The indenters case is of the size of a microscope objective mounted on the objective revolver. The load range for
this design is from 0.5 mN to 2 N; therefore it is used for thicker films.
A commercial nanoindentation hardness apparatus for use inside a scanning electron microscope (SEM), (Model No. UHMT-3)
for imaging the indents after the sample is unloaded, is manufactured
by Anton Paar K.G., A-8054, Graz, Austria. The apparatus is mounted
on the goniometer stage of the SEM. In this setup, the indenter is
mounted on a double-leaf spring cantilever, and is moved against the
sample by an electromagnetic system to attain the required indentation load, which is measured by strain gages mounted on the leaf
springs.[22][23] After the required load has been reached and the dwell
time has elapsed, the sample is unloaded, and the indentation diagonal is measured by an SEM.
Depth-Sensing Nanoindentation Hardness Apparatus and
Its Modifications. General Description and Principle of Operation.
The most commonly used commercial depth-sensing nanoindentation

238 Handbook of Hard Coatings

hardness apparatus is manufactured by Nano Instruments Inc., 1001


Larson Drive, Oak Ridge, Tennessee 37830. Ongoing development
of this apparatus have been described by Pethica, et al.,[228] Oliver, et
al.,[216] Oliver and Pethica,[217] Oliver and Pharr,[218] and Pharr and
Oliver.[230] This instrument is called Nanoindenter. The most recent
model is Nanoindenter II.[16] The apparatus continuously monitors
the load, and the position of the indenter relative to the surface of the
specimen (depth of an indent), during the indentation process. The
area of the indent is then calculated from a knowledge of the geometry of the tip of the diamond indenter. The load resolution is about
75 nN, and position of the indenter can be determined to 0.1 nm.
Mechanical properties measurements can be made at a minimum
penetration depth of about 20 nm (or a plastic depth of about 15
nm).[216] Description of the instrument which follows, is based on
Ref. 16.
The Nanoindenter consists of three major components: (1) the
indenter head, (2) an optical microscope, and (3) a X-Y-Z motorized
precision table for positioning and translating the sample between
the optical microscope and indenter, Fig. 3a. The loading system
used to apply the load to the indenter consists of a magnet and coil in
the indenter head, and a high precision current source, Fig. 3b. The
displacement sensing system consists of a special three-plate capacitive displacement sensor, used to measure the position of the indenter, Fig. 3b. The indenter column is attached to the moving plate.
This plate-and-indenter assembly is supported by two leaf springs
cut in such a fashion as to have very low stiffness. At the bottom of
the indenter rod, a three-sided pyramidal diamond tip (Berkovich
indenter, discussed next) is generally attached. The indenter head
assembly is rigidly attached to the U beam, and sample rides on the
X-Y-Z table, Fig. 3a. The optical microscope is also attached to the U
beam. The position of an indent on a specimen is selected using the
microscope (max. magnification of 1500). The spatial resolution of
the position of the table in the X-Y plane is 400 nm.

Macro- and Micromechanical and Tribological Properties 239

(a)

(b)
Figure 3. Schematics of the Nanoindenter II (a) showing the major componentsthe indenter head, an optical microscope and a X-Y-Z motorized precision table
and (b) showing the details of indenter head and controls (microscope which is
directly behind indenter and massive U bar are not shown for clarity).[16]

240 Handbook of Hard Coatings

The Nanoindenter also comes with a continuous stiffness measurement device.[217][229] This device makes possible the continuous
measurement of the stiffness of a sample, which allows the elastic
modulus to be calculated, as a continuous function of time (or
indentation depth). Useful data can be obtained from indents with
depths as small as 20 nm. Because of the relatively small time
constant of the measurements, the device is particularly useful in
studies of time-dependent properties of materials.
Weihs, et al.,[295] used acoustic emission (AE) sensors to detect
cracking during indentation tests using Nanoindenter. Acoustic emission (AE), measurement is a very sensitive technique to monitor
cracking of the surfaces and subsurfaces. The nucleation and growth
of cracks result in a sudden release of energy within a solid, then
some of the energy is dissipated in the form of elastic waves. These
waves are generated by sudden changes in stress, and in displacement that accompany the deformation. If the release of energy is
sufficiently large and rapid, then elastic waves in the ultrasonic
frequency regime (acoustic emission) will be generated, and these
can be detected using piezoelectric transducers (PZTs), via expansion and compression of the PZT crystals.[38][249][312] The energy
dissipated during crack growth can be estimated by the rise time of
the AE signal. Weihs, et al.,[295] mounted commercial transducer
with W impregnated epoxy backing for damping underneath the
sample. The transducer converts the AE signal into voltage, that is
amplified by an oscilloscope, and used for continuous display of AE
signal. Any correlation between the AE signal, and the load-displacement curves can be observed (also see Refs. 302 and 306).
The Berkovich Indenter. The main requirements for the indenter are high elastic modulus, no plastic deformation, low friction,
smooth surface, and a well defined geometry that is capable of
making a well defined indentation impression. The first four requirements are satisfied by choosing the diamond material for the tip. A
well defined perfect tip shape is difficult to achieve. Berkovich is a
three-sided pyramid, and provides a sharply pointed tip compared to
the Vickers or Knoop indenters, which are four-sided pyramids and

Macro- and Micromechanical and Tribological Properties 241


have a slight offset (0.51 m).[38][39][273] Because any three nonparallel planes intersect at a single point, it is relatively easy to grind a
sharp tip on an indenter if Berkovich geometry is used. However, an
indenter with a sharp tip suffers from a finite, but an exceptionally
difficult to measure tip bluntness. Experimental procedures have
been developed to correct for the tip shape, to be described later.
Berkovich indenter is a three-sided (triangular-based), pyramidal diamond, with a nominal angle of 65.3 between the (side) face,
and the normal to the base at apex, an angle of 76.9 between edge
and normal, and with a radius of the tip less than 0.1 m (Fig. 4a).[33]
Typical indenter is shaped to be used for indentation (penetration)
depths of 1020 m. The indents appear as equilateral triangles (Fig.
4b), and the height of triangular indent l is related to the depth h as
1
h 2
= cot 76.9 =
6.44
l 3

Eq. (6a)

(a)

(b)

Figure 4. (a) Schematic of a Berkovich tip and (b) impression of a indentation


made by Berkovich tip.

242 Handbook of Hard Coatings

The relationship h(l ) is dependent on the shape of the indenter. The


height of the triangular indent l is related to the length of one side of
the triangle a as
Eq. (6b)

l = 0.866a

and
Eq. (6c)

h
1
=
a 7.44

The projected contact area (A) for the assumed geometry is given as
Eq. (7)

A = 0.433a2 = 23.97h2

The exact shape of the indenter tip needs to be measured for


determination of hardness, and Youngs modulus of elasticity. Since
the indenter is quite blunt, direct imaging of indentations of small
size in the scanning electron microscope is difficult. Determination
of tip area function will be discussed later.
Analysis of Indentation Data. An indentation curve is the
relationship between load W and displacement (or indentation depth or
penetration depth) h, which is continuously monitored and recorded
during indentation. Stress-strain curves, typical indentation curves,
the deformed surfaces after tip removal, and residual impressions of
indentation for ideal elastic, rigid-perfectly plastic, and elastic-perfectly plastic, and real elastic-plastic solids are shown in Fig. 5. For
an elastic solid, the sample deforms elastically according to Youngs
modulus, and the deformation is recovered during unloading. As a
result, there is no impression of the indentation after unloading. For
a rigid-perfectly plastic solid, no deformation occurs until yield
stress is reached, when plastic flow takes place. There is no recovery
during unloading, and the impression remains unchanged. In the case
of elastic-plastic solid, it deforms elastically according to Youngs
modulus, and then it deforms plastically. The elastic deformation is
recovered during unloading. In the case of elastic-perfectly plastic
solid, there is no work hardening.

Macro- and Micromechanical and Tribological Properties 243

Figure 5. Schematics of stress-strain curves, typical indentation curves, deformed surfaces after tip removal, and residual impressions of indentation, for
ideal elastic, rigid-perfectly plastic, elastic-perfectly plastic (ideal) and real
elastic-plastic solids.

All engineering surfaces follow real elastic-plastic deformation behavior with work hardening.[151] The deformation pattern of a
real elastic-plastic sample during and after indentation is shown
schematically in Fig. 6. In this figure, we have defined the contact
depth (hc) as the depth of indenter in contact with the sample under
load. The depth measured during the indentation (h) includes the
depression of the sample around the indentation in addition to the
contact depth. The depression of the sample around the indentation
(hs = h-hc) is caused by elastic displacements and must be subtracted
from the data to obtain the actual depth of indentation or actual
hardness. At peak load, the load and displacement are Wmax and hmax,
respectively, and the radius of the contact circle is a. Upon unloading, the elastic displacements in the contact region are recovered and
when the indenter is fully withdrawn, the final depth of the residual
hardness impression is h f.

244 Handbook of Hard Coatings

Figure 6. Schematic representation of the indenting process illustrating the


depression of the sample around the indentation and the decrease in indentation
depth upon unloading.[218]

Schematic of a load displacement curve is shown in Fig. 7.


Based on the work of Sneddon[255] to predict the deflection of the
surface at the contact perimeter for a conical indenter and a paraboloid
of revolution, Oliver and Pharr[218] developed an expression for hc at
the maximum load (required for hardness calculation) from hmax,
Eq. (8a)

hc = hmax W max / S max

where =0.72 for the conical indenter, =0.75 for the paraboloid of
revolution, and =1 for the flat punch, and Smax is the stiffness (=1/
compliance), equal to the slope of unloading curve (dW/dh) at the
maximum load. Oliver and Pharr[218] assumed that behavior of
Berkovich indenter is similar to that of conical indenter, since crosssectional areas of both types of indenters varies as the square of the
contact depth, and their geometries are singular at the tip. Therefore,
for Berkovich indenter, ~ 0.72. Thus hc is slightly larger than
plastic indentation depth (hp) which is given by
Eq. (8b)

h p = hmax Wmax / S max

Macro- and Micromechanical and Tribological Properties 245


We note that Doerner and Nix[100] had underestimated hc by assuming that hc = hp. Based on the finite element analysis of the indentation process, Laursen and Simo[171] showed that hc cannot be assumed equal to hp for indenters which do not have flat punch
geometry.

Displacement, h
Figure 7. Schematic of load-displacement curve.

For a Vickers indenter with ideal pyramidal geometry (ideally


sharp tip), projected contact area to depth relationship is given
as[38][100]
Eq. (9a)

A = 24 .5hc2

Since the area to depth relationship is equivalent for both typical


Berkovich and Vickers pyramids, Eq. (9) holds for Berkovich indenter as well. Though we have derived a slightly different expression for A(h) presented in Eq. (7) for the assumed Berkovich indenter
geometry, we use Eq. (9) for A(h) in this chapter, as this relationship

246 Handbook of Hard Coatings

is most commonly used in the analysis of the indentation hardness


data.
As shown in Fig. 6, the actual indentation depth, hc, produces
a larger contact area than would be expected for an indenter with an
ideal shape. For the real indenter used in the measurements, the
nominal shape is characterized by an area function F(hc) which
relates projected contact area of the indenter to the contact depth
(Eq. 9a),
Eq. (9b)

1/ 2

= F ( hc )

The functional form must be established experimentally prior to the


analysis.
Determination of Load Frame Compliance and Indenter
Area Function. As stated earlier, measured displacements are the
sum of the indentation depths in the specimen, and the displacements
of suspending springs, and the displacements associated with the
measuring instruments, referred to as load frame compliance. Therefore, to accurately determine the specimen depth, load frame compliance must be known. The exact shape of the diamond indenter tip
needs to be measured because hardness and Youngs modulus of
elasticity depend on the contact areas derived from measured depths.
The tip gets blunt and its shape significantly affects the prediction of
mechanical properties (Figs. 810).
The method used in the past for determination of area function
has been to make a series of indentations at various depths in
materials in which the indenter displacement is predominantly plastic, and measure the size of the indentations by direct imaging.[100][228]
Optical imaging cannot be used to accurately measure submicron
size impressions. Because of the shallowness of the indent impressions, SEM images result in poor contrast.

Macro- and Micromechanical and Tribological Properties 247

(a)

(b)
Figure 8. (a) Predicted projected contact area as a function of indentation depth
curves for various tip radii and measured data, (b) predicted load as a function of
indentation depth curves for various tip radii and measured data.[254]

248 Handbook of Hard Coatings

Oliver and Pharr [218] proposed an easier method for determining area functions that requires no imaging. Their method is based
only on one assumption, that Youngs modulus is independent of
indentation depth. They also proposed a method to determine loadframe compliance. We first describe the methods for determining of
load frame compliance followed by the method for area function.
They modeled the load frame and the specimen as two springs in
series, thus
Eq. (10)

C = Cs + C f

where C, Cs, and Cf are the total measured compliance, specimen


compliance and load frame compliance, respectively. The measured
compliance C is given by
Eq. (11a)

C = dh/dW

The relationship for the sample compliance Cs (inverse of


stiffness S) for an (Vickers, Knoop, and Berkovich) indenter is given as

Eq. (11b)

Cs =

1
2 Er

1/ 2

where
2
2
1 1 vs 1 vi
=
+
Er
Es
Ei

and dW/dh is the slope of the unloading curve at the maximum load
(Fig. 7), Er, Es, and Ei are the reduced modulus and elastic moduli of
the specimen and the indenter, and s and i are the Poissons ratios
of the specimen and indenter. C (or S) is the experimentally measured compliance (or stiffness) at the maximum load during unloading and A is the projected contact area at the maximum load.
From Eqs. (10) and (11), we get
1/ 2

Eq. (12)

C = Cf +

1

2 Er A

Macro- and Micromechanical and Tribological Properties 249

From Eq. (12), we note that if the modulus of elasticity is constant, a


plot of C as a function of A1/2 is linear and the vertical intercept gives
Cf. It is obvious that most accurate values of Cf are obtained when
the specimen compliance is small, i.e., for large indentations.
Using the measured Cf value, they calculated contact areas for
indentations made at shallow depths on the aluminum with measured
Er, and/or on a harder fused silica surface with published values of
Er, by rewriting Eq. (12) as
Eq. (13)

A=

1
1
2
4 E r (C C f ) 2

from which an initial guess at the area function was made by fitting A
as a function hc data to an eighth order polynomial
Eq. (14)

A = 24.5hc2 + C1 hc + C 2 hc1 / 2 + C 3 h 1c / 4 + + C 8 hc1 / 128

where C1 through C8 are constants. The first term describes the


perfect shape of the indenter; the others describe deviations from the
Berkovich geometry due to blunting of the tip.
2.4

Mechanical Properties by Nanoindentation

Hardness And Modulus Of Elasticity. Berkovich hardness


HB (or HB ) is defined as the load divided by the projected area of the
indentation. It is the mean pressure which a material will support
under load. From the indentation curve, we can obtain hardness at the
maximum load as,
Eq. (15)

HB = Wmax / A

where Wmax is the maximum indentation load and A is the projected


contact area at the peak load. The contact area at the peak load is
determined by the geometry of the indenter and the corresponding
contact depth hc using Eq. (3a) and (4b). Plot of hardness as a
function of indentation depth for polished single-crystal silicon (111),

250 Handbook of Hard Coatings

with and without tip shape calibration, is shown in Fig. 9. We note


that, for this example, tip shape calibration is necessary and the
hardness is independent of corrected depth.

uncorrected shape

corrected shape

Figure 9. Hardness as a function of indentation depth for polished single-crystal


silicon (111) calculated from the area function with and without tip shape
calibration.[100]

Figure 10. Compliance as a function of inverse of indentation depth for tungsten


with and without tip shape calibration. A constant modulus with 1/depth would
be indicated by the straight line. The slope of the corrected curve is 480 GPa,
which compares reasonably well to the known modulus of tungsten (420 GPa).
The small y-intercept of about 0.3 nm/mN is attributed to load frame compliance,
not removed.[100]

Macro- and Micromechanical and Tribological Properties 251

It should be pointed out that hardness measured using this


definition may be different from that obtained from the more conventional definition, in which the area is determined by direct measurement of the size of the residual hardness impression. The reason
for the difference is that, in some materials, a small portion of the
contact area under load is not plastically deformed, and as a result,
the contact area measured by observation of the residual hardness
impression may be less than that at peak load. However, for most
materials, measurements using two techniques give similar results.
Even though during loading, a sample undergoes elastic-plastic deformation, the initial unloading is an elastic event. Therefore,
the Youngs modulus of elasticity or simply called elastic modulus
of the specimen can be inferred from the initial slope of the unloading curve (dW/dh) called stiffness (1/compliance) (at the maximum
load) (Fig. 7). The modulus of elasticity is calculated from Eq. (12).
It should be noted that the contact stiffness is measured only at the
maximum load and no restrictions are placed on the unloading data
being linear during any portion of the unloading.
Hardness And Modulus of Elasticity of Thin Films From
The Composite Response of Film and Substrate. As mentioned
previously, for a thin film on a substrate, if the indentation exceeds
about 30% of the film thickness, then measured hardness is affected
by the substrate properties. A number of researchers have attempted
to derive expressions that relate thin-film hardness to substrate
hardness, composite hardness (measured on the coated substrate),
and film thickness, and so allow the calculations of these quantities
give the remaining three.[34][35][69][70][74][152][244] Here we discuss
two models based on volume law of mixtures, (volume fraction
model) and finite element simulation.
Sargent[244] suggested that the hardness of a film/substrate
composite is determined by a weighted average of the volume of
plastically deformed material in the film (Vf), and that in the substrate (Vs),
Eq. (16)

H =Hf

Vf
V

+ Hs

Vs
V

252 Handbook of Hard Coatings

where V = Vf + Vs. The deformed volumes of film and substrate can


be calculated using expanding spherical cavity model.[151] Burnett
and Rickerby[73][74] found it necessary to incorporate a further weighing factor to deforming volume, to obtain a reasonable fit to experimental data. This factor accounts for the differences in relative sizes
of the plastic zones in the film and substrate. Equation (16) is
modified as for a soft film on the hard substrate,
Eq. (17a)

H =Hf X3

Vf
V

+ Hs

Vs
V

and for the hard film on the soft substrate,

Eq. (17b)

H = Hs

Vf
V

+ Hs X 3

Vs
V

where X is the ratio of plastic zone volumes given as,

Eq. (17c)

Ef Hs
X =
H f Es

They found that n, determined empirically, ranged from 1/2 to 1/3.


Bhattacharya and Nix[34] modeled the indentation process using finite element method to study the elastic-plastic response of
materials. Bhattacharya and Nix[35] calculated elastic and plastic
deformation associated with submicron indentation by a conical
indenter of thin films on substrates, using the finite-element method.
The effects of the elastic and plastic properties of both the film and
substrate on the hardness of the film/substrate composite were studied by determining the average pressure under the indenter as a
function of the indentation depth. They developed empirical equations for film/substrate combinations for which the substrate is either

Macro- and Micromechanical and Tribological Properties 253

harder or softer than the film. For the case of a soft film on a harder
substrate, the effect of substrate on film hardness can be described as

Eq. (18a)

s hc
H
= 1 + f 1 exp
t
f
Hs
Ef
Hs

E s

where Ef and Es are the Youngs moduli, f and s are the yield
strengths and Hf and Hs are the hardnesses of the film and substrate,
respectively. H is the hardness of the composite, hc is the contact
indentation depth, and tf is the film thickness. Similarly for the case
of a hard film on a softer substrate, the hardness can be expressed as

Hf

H
s
H
hc
f

= 1 +
1 exp
1
2

Eq. (18b) H s
tf
E

Hs

f f
Es
s

Composite hardness results were found to depend only very


weakly on Poissons ratio (v), and for this reason, this factor was not
considered in the analysis. In Fig. 11, they show the composite
hardness results as a function of (hc/tf), for cases in which the film
and substrate have different yield strengths. We note that hardness is
independent of the substrate for indentation depths less than about
0.3 of the film thickness, after which the hardness slowly increases/
decreases because of the presence of the substrate. In Fig. 12, they
show the composite hardness results for cases in which the film and
substrate have different Youngs moduli. It is observed that the
variation of hardness with depth of indentation in these cases is
qualitatively similar to cases in which the film and substrate have
different yield strengths, although the hardness changes more gradually than in the previous cases. Burnett and Rickerby[75] and Fabes et
al.[111] have applied Eqs. (19)(21) to calculate the film hardness
from the measured data for various films and substrates.

254 Handbook of Hard Coatings

(a)

(b)
Figure 11. Effect of relative yield strengths of the film and the substrate on the
composite hardness for (a) a soft film on a hard substrate and (b) a hard film on a
soft substrate.[35]

Macro- and Micromechanical and Tribological Properties 255

(a)

(b)
Figure 12. Effect of relative Youngs moduli of the film and the substrate on the
composite hardness for (a) a soft film on a hard substrate and (b) a hard film on a
soft substrate.[35]

256 Handbook of Hard Coatings

Doerner and Nix[100] empirically modeled the influence of the


substrate on the elastic measurement of very thin film in an indentation test using the following expression for the compliance
Eq. (19)
1 v 2f

t f

1
exp

1
A
dh 1 2 E f

=
C=
+b
2
2
dW 2 A 1 v f
t f 1 v i
+
+ E exp

E i
A
s

where the subscripts f, s, and i refer to the film, substrate, and


indenter, respectively. The term A is equal to (24.5)1/2hc for the
Vickers or Berkovich indenter. The film thickness is tf, and b is the y
intercept for the compliance versus 1/depth plot, obtained for the
bulk substrate, which can be neglected in most cases. The weighing
factors [1-exp(-tf / A )] and exp (-tf / A )] have been added to
account for the changing contributions of the substrate and film to
the compliance. The factor can be determined empirically.
Kings analysis[157] verified that Eq. (19) is an excellent functional form for describing the influence of the substrate, and theoretically determined the values of for various indenter shapes. The
value of was found to depend on the indenter shape and size and
film thickness and was found to be independent of Ei/Es. The values
of as a function of A /tf for Berkovich (triangular) indenters are
shown in Fig. 13. The values of are found to be similar for square
and triangular indenters. Bhattacharya and Nix[35] analyzed the deformations of a layered medium in contact with a conical indenter
using the finite-element method. Their analysis also verified the
relationship given in Eq. (19).

Macro- and Micromechanical and Tribological Properties 257

Figure 13. Parameter as a function of normalized indenter size for Berkovich


indenter indenting a layered solid surface.[157]

Viscoelastic/Plastic Properties. Most materials including ceramics, and even diamond are found to creep at temperatures well
below half their melting points, even at room temperature. Indentation creep and indentation load relaxation (ILR) tests are used for
measurement of the time-dependent flow of materials. These offer an
advantage of being able to probe the deformation properties of a thin
film as a function of indentation depth and location.
In the indentation creep test, the hardness indenter maintains its
load over a period of time under well controlled conditions, and
changes in indentation size are monitored.[19][144][178][206][290][297]
Nanoindenters are also used for indentation creep studies.[178][182][234][235] Indentation creep is influenced by a large number of variables, such as the materials plastic deformation properties, diffusion constants, normal load of indenter, duration of the
indentation, and the test temperature.
In a typical ILR test, the indenter is first pushed into the sample
at a fixed displacement rate until a predetermined load or displacement is achieved, and the position of the indenter is then fixed. The
material below the indenter is elastically supported, and will continue to deform in an nonelastic manner, thereby tending to push the
indenter farther into the sample. Load relaxation is achieved by
conversion of elastic strain in the sample into inelastic strain in the
sample. During the test, the load and position of the indenter, and the
specimen are continuously monitored. Normally the indenter motion

258 Handbook of Hard Coatings

is held constant, and the changes in the load are monitored as a


function of time. It is possible to obtain the plastic indentation rate
from the indentation load and total depth information during the
relaxation run.[85][136][207] The resulting load relaxation data are
reported in the form of log (indentation pressure) as a function of log
(plastic indentation strain rate).[135][166][167][302]
The indentation pressure is calculated by dividing the load by
the projected area of the indenter. Once the plastic indentation depth
is known as a function of time, the projected area is determined
experimentally as described earlier. The plastic indentation strain
rate [(1/h)(dh/dt), where h is the current indentation depth] is calculated in the manner similar to that for bulk relaxation data.
The strain rate sensitivity of materials is measured in terms of
stress exponent, n which is defined by the equation,
Eq. (20)

Plastic indentation rate = A (indentation pressure) n

where A and n are the constants. The stress exponent is found as a


slope of a log-log plot of plastic indentation rate (or strain rate) and
indentation pressure. In the ILR test, the continuous change in the
contact area results in continuous changes in both plastic indentation
rate and pressure. Thus, data from a single indentation test, which may
span several orders of magnitude in both strain rate and pressure, are
sufficient to determine the stress exponent. Stress exponent can be used
to define the superplasticity of a material. The variations in stress
exponent reflect the changes that may take place when the substructure
generated at high strain rate approaches equilibrium condition.[190]
Nanofracture Toughness. Fracture toughness, KIc of a material is a measure of its resistance to the propagation of cracks, and the
ratio H/KIc is an index of brittleness, where H is the hardness.
Indentation fracture toughness is a simple technique for determination of fracture toughness.[18][79][81][82][88][134][172][174][220][232] Indentation cracking method is especially useful for measurement of
fracture toughness of thin films, or small volumes. This method is
quite different from conventional methods, in that no special

Macro- and Micromechanical and Tribological Properties 259

specimen geometry is required. Rather, the method relies on the fact


that when indented with a sharp indenter, most brittle materials form
radial cracks, and the lengths of the surface traces of the radial cracks
(for definition of crack length, see Fig. 14) have been found to
correlate reasonably well with fracture toughness. Using simple
empirical equations, fracture toughness can then be determined from
simple measurement of crack length.

Figure 14. Schematic of Vickers indentation with radial cracks.

In microindentation, cracks at relatively high indentation loads


of several hundred grams are on the order of 100 m in length, and
can be measured optically. However, to measure toughness of very
thin films or small volumes, much smaller indentations are required.
However, a problem exists in extending the method to nanoindentation
regime in that there are well defined loads, called cracking thresholds, below which indentation cracking does not occur in most brittle
materials.[169] For a Vickers indenter, cracking thresholds in most
ceramics are about 25 g. Pharr, et al.[232] have found that Berkovich
indenter, (a three-sided pyramid) with the same depth-to-area ratio as
a Vickers indenter, (a four-sided pyramid) has a cracking of the
thresholds very similar to that of the Vickers indenter. They showed
that cracking thresholds can be substantially reduced by using sharp
indenters, i.e., indenters with smaller included tip angles, such as a
three-sided indenter with the geometry of the corner of a cube.
Studies using a three-sided indenter with the geometry of a corner of

260 Handbook of Hard Coatings

a cube have revealed that cracking thresholds can be reduced to loads


as small as 0.5 g, for which indentations and crack lengths in most
materials are submicron in dimension.
Based on fracture mechanics analysis, Lawn, et al.[174] developed a mathematical relationship between fracture toughness and
indentation crack length given as,

Eq. (21)

K IC

E
= B
H

1/ 2

W
3/2
c

where W is the applied load and B is an empirical constant depending


upon the geometry of the indenter, (also see Refs. 172 and 232).
Antis, et al.[18] conducted a study on a number of brittle materials
chosen to span a wide range of toughness. They indented with a
Vickers indenter at several loads, and measured crack length optically. They found a value of B = 0.016 to give good correlation
between the toughness values measured from the crack length, and
the ones obtained using more conventional methods. Mehrotra and
Quinto[196] used a Vickers indenter to measure fracture toughness of
the coatings. Pharr, et al.[232] tested several bulk ceramics listed in
Table 4 using Vickers, Berkovich, and cube corner indenters. They
found that the fracture toughness equation can be applied for the data
obtained with all three indenters, provided a different empirical
constant was used for a cube corner indenter. The constant B for
Vickers and Berkovich indenter was found to be about 0.016 and for
cube corner it was about 0.032. Pharr, et al.[232] further reported that
predominant cracks formed with Vickers or Berkovich indenters are
cone cracks, and with cube corner indenter, predominant cracks were
radial cracks, Fig. 15.

Macro- and Micromechanical and Tribological Properties 261

Figure 15. Indentations in fused quartz made with the cube corner indenter
showing radial cracking at indentation loads of (a) 12 g and (b) 0.45 g.[232]

Chantikul et al.[79] developed a relationship between fracture


toughness and the indentation fracture strength and the applied load

Eq. (22)

K Ic

E
= c
H

1
4
8
f W 3

where is the fracture strength after indentation at a given load and


c is an empirical constant (0.59). Advantage of this analysis is that
the measurement of crack length is not required. Mecholsky et al.[194]
used this analysis to calculate fracture toughness of diamond films
on silicon. They indented the films at various indentation loads of 3
to 9 kg, and then fractured in four point flexure to measure fracture
strength. The data was then used to get fracture toughness. Equation
(22) was found to hold for the measurements. They reported a
fracture toughness of 6 and 12 m thick diamond films on silicon on
the order of 2 MPa m .
Chaing et al.[81] developed a relationship for fracture toughness
of the coating/substrate interface, which can be used as a measure of
indentation adhesion of the coatings. This analysis is presented later
on in the section on adhesion measurements.

262 Handbook of Hard Coatings

Nanofatigue. Delayed fracture resulting from extended service


is called fatigue. Fatigue fracturing progresses through a material via
changes within the material at the tip of a crack, where there is a high
stress intensity. There are several situations: cyclic fatigue, stress
corrosion, and static fatigue. Cyclic fatigue results from cyclic loading of machine components, e.g., the stresses cycle from tension, and
compression occurs in a loaded rotating shaft. Fatigue also can occur
with fluctuating stresses of the same sign, as occurs in a leaf spring,
in a dividing board. In a low flying slider in a head-disk interface,
isolated asperity contacts occur during use, and the fatigue failure
occurs in the multilayered thin-film structure of the magnetic disk.[38]
Asperity contacts can be simulated using a sharp diamond tip in an
oscillating contact with the thin-film disk.
Li and Chu[177] developed a indentation fatigue test, called
impression fatigue. In this test, a cylindrical indenter with a flat end
was pressed onto the surface of the test material with a cyclic load
and the rate of plastic zone propagation was measured.
Wu et al.[307] developed a nanoindentation fatigue test by modifying their Nanoindenter. The cyclic indentation was implemented
by servo controlling the PZT stack to drive the indenter so that the
loadcell output followed a 0.1 Hz sinusoidal loading pattern, and the
latter was specified by the cyclic frequency, and the lower and upper
limits for the desired load range. The lower limit for all the tests was
set at about 0.2 mN to perform a full load cycle indentation fatigue. A
nonzero limit was required in order to activate the load cell servo
control mode. Several maximum loads, namely 4, 16, and 24 mN
were used. Following results can be obtained: (i) endurance limit,
i.e., the maximum load below which there is no coating failure for a
preset number of cycles, (ii) number of cycles at which the coating
failure occurs, and (iii) changes in contact stiffness measured by
using the unloading slope of each cycle which can be used to monitor
the propagation of the interfacial cracks during cyclic fatigue process. They used a conical diamond indenter with a nominal 1 m tip.
Applied load and penetration depth were simultaneously monitored
during the entire test.

Macro- and Micromechanical and Tribological Properties 263

Typical nanoindentation fatigue results for a 0.11-m DC


planar magnetron sputtered amorphous carbon films on (100) silicon
substrate deposited at argon pressure of 30 m torr is shown in Fig.
16. [307] The test was run at a maximum cyclic load of 4 mN and 0.1
Hz for a total of 105 cycles with fracture in the 93rd cycle. SEM
micrograph of the damaged zone (Fig. 16c) shows that the plastic
deformation attributed primarily to the silicon substrate, occurred in
the central indent area, and moreover, the carbon coating spalled
around the indent, and resulted in several isolated carbon flakes as
well as cantilevered flakes. Wu et al. reported that in a single
indentation test, the critical indentation load required to crack the
carbon coating was about 6 mN. The critical indentation load was
extracted by using the criterion of the applied load at which a load
drop appeared along a loading curve. Evidently, the endurance limit
can be significantly lower than the critical load of a single indentation. This scenario is analogous to the fatigue strength versus tensile
strength in macrotests. Wu et al. further reported that the carbon
films deposited at a argon pressure of 6 m torr exhibited endurance
limit of about 24 mN, about a factor of six higher than the film
deposited at 30 m torr. Scratch resistance of the film at 30 m torr was
also poorer as compared to that deposited at 6 m torr.[302][305][306]
Adhesion Strength Measurements. For measurement of adhesion strength of the coating-substrate interface using indentation
method, the coating sample is indented at various loads. At low
loads, the coating deforms with the substrate. However, if the load is
sufficiently high, a lateral crack is initiated and propagated along the
coating-substrate interface. The lateral crack length increases with the
indention load. The minimum load at which the coating fracture is
observed is called the critical load, and is employed as the measure of
coating adhesion (Fig. 17). For relatively thick films, the indentation
is generally made using Brinell hardness tester with a diamond sphere
of 20 m radius,[278] Rockwell hardness tester with a Rockwell C
120o cone with a tip radius of 200 m,[196] or a Vickers pyramidal
indenter.[5][81][180] However, for extremely thin films, a Berkovich
indenter,[263] or a conical diamond indenter with a tip radius of 5 mm
and 30 of included angle,[285] is used in a nanoindenter.

264 Handbook of Hard Coatings

(a)

(b)

(c)

Figure 16. Typical microindentation fatigue results from 0.11-m thick dc


sputtered amorphous carbon on (100) Si, (a) the direct outputs from a strip chart
recorder of applied load (LC) and indenter position (IND) (Maximum load = 4
mN, frequency = 0.1 Hz), (b) a plot of the indentation fatigue loading curve
versus total penetration depth (the plot includes only the load cycles from 91 to
100; note the abrupt depth increase started from the 93rd cycle) and (c) SEM
micrograph of the fatigue indent.[307]

Figure 17. Schematic illustration of the indentation method for adhesion


measurement.

Macro- and Micromechanical and Tribological Properties 265

It should be noted that the measured critical load Wcr is a


function of hardness and fracture toughness in addition to the adhesion of coatings. Chiang et al.[81] have related the measured crack
length during indentation, the applied load and critical load (at which
coating fracture is observed) to the fracture toughness of the substrate-coating interface. A semianalytical relationship derived between the measured crack length c and the applied load W:
1/ 2

Eq. (23)

W
c = 1 cr W 1 / 4
W

where

2 =

1t c3 / 2 H 1 / 2
(K Ic )interface

1 is a numerical constant, tf is the coating thickness, H is the mean


hardness and (KIc )interface is the fracture toughness of the substratecoating interface. Mehrotra and Quinto[196] used this analysis to
calculate fracture toughness of the interface.
Marshall and Oliver[184] estimated adhesion of composites by
measuring the magnitude of shear (friction) stresses at fiber/matrix
interfaces in composites. They used a Berkovich indenter to push on
the end of an individual fiber and measured the resulting displacement of the surface of the fiber below the matrix surface (due to
sliding). The shear stress was calculated from the force-displacement
relation obtained by analysis of the frictional sliding. The force and
displacement measurements were obtained only at the peak of the
load cycle, and the sliding analysis was based on sliding at constant
shear resistance at the interface. These experiments provided measurements of average shear stresses at individual fibers.

266 Handbook of Hard Coatings

2.5

Scratch Resistance/Adhesion Measurements

Macro-, micro-, and nanoscratch techniques are used to measure scratch resistance of surfaces from bulk materials to a few
nanometer thick films. Adhesion describes the sticking together of
two materials. Adhesion strength, in practical sense, is the stress
required to remove a coating from a substrate. Indentation, described
earlier, and scratch on micro-and nanoscales are the two commonly
used techniques to measure adhesion of thin hard films with good
adhesion to the substrate (>70 MPa).[37][46][60][76][199]
Scratching the surface with a fingernail, or a knife, is probably
one of the oldest methods for determining the adhesion of paints and
other coatings. Scratch tests to measure adhesion of films was first
introduced by Heavens.[139] A smoothly round chrome-steel stylus
with a tungsten carbide, or Rockwell C diamond tip (in the form of
120 cone with a hemispherical tip of 200 mm radius),[196][223]
[224][261][287] or Vickers pyramidal indenter,[70][73] for macro-and
microscratching a conical diamond indenter (with a 1 m or 5 m of
tip radius and 60 of included angle), for nanoscratching[56][131]
[132][302][305][306] is drawn across the coating surface. A vertical load is
applied to the scratch tip, and is gradually increased during scratching until the coating is completely removed. The minimum or critical
load at which the coating is detached or completely removed is used
as a measure of adhesion.[2][30][70][73][76][80][119][147][149][153][170][196]
[199][223][225][253][261][287][288][299][301][302][305][306] It is a most commonly used technique to measure adhesion of hard coatings with
strong interfacial adhesion (>70 MPa).
For a scratch geometry shown in Fig. 18, surface hardness H is
given by
Eq. (24)

H=

Wcr
a2

and adhesion strength t is given by [30]

Macro- and Micromechanical and Tribological Properties 267

Eq. (25a)

t = H tan
=

Wcr
a2

a
2
R a 2

1/ 2

or
Eq. (25b)

Wcr
if R >> a
aR

where Wcr is the critical normal load, a is the contact radius and R is
the stylus radius.

Figure 18. Geometry of the scratch.

Burnett and Rickerby,[73] and Bull and Rickerby [70] analyzed


scratch test of a coated sample in terms of three contributions: (i) a
ploughing contribution, which will depend on the indentation stress
field, and the effective flow stress in the surface region, (ii) an
adhesive friction contribution due to interactions at the indentersample interface, and (iii) an internal stress contribution since any
internal stress will oppose the passage of the indenter through the
surface, thereby effectively modifying the surface flow stress. They
derived a relationship between the critical normal load Wcr and the
work of adhesion W ad

268 Handbook of Hard Coatings

Eq. (26)

a 2 2 EWad 2
Wc =

2 t

where E is the Youngs modulus of elasticity and t is the coating


thickness. Plotting of Wc as a function of a2/t1/2 should give a straight
line of the slope (2EWad /t)1/2/2 from which Wad can be calculated.
Bull and Rickerby suggested that either the line slope (interface
toughness) or Wad could be used as a measure of adhesion.
An accurate determination of critical load Wcr sometimes is
difficult. Several techniques, such as (i) microscopic observation
(optical or scanning electron microscope) during the test, (ii) chemical analysis of the bottom of the scratch channel, (with electron
microprobes) and (iii) acoustic emission, have been used to obtain
the critical load.[149][223][224][253][261][287][302][306] The acoustic-emission technique has been reported to be very sensitive in determining
critical load. Acoustic emission starts to increase as soon as cracks
begin to form perpendicular to the direction of the moving stylus. In
some instruments, tangential (or friction), force is measured during
scratching.[16][55][57][131][133][147][287][301][302][306] An instant increase
in the friction force during scratch has also been used as an indicator
of a damage event.
Apparatuses For Scratch Measurements On Macroscale.
Several macroscratch testers are commercially available, such as the
Taber shear/scratch tester model 502 with a no. 13958 diamond
cutting tool (manufactured by Teledyne Taber, North Tonawanda,
NY) for thick films, Revetest automatic scratch tester (manufactured by Centre Suisse dElectronique et de Microtechnique S.A.,
CH-2007, Neuchatel, Switzerland), and scratch tester with friction
force attachment (manufactured by VTT, Technical Research Center
of Finland, Espoo, Finland) for thin films.[223][224][253][261][288] Of all
the macroscratch testers available commercially, Revetest is used
most extensively by numerous research groups to study the adhesion
characteristics of hard coatings.

Macro- and Micromechanical and Tribological Properties 269

Revetest automatic scratch tester is shown schematically in


Fig. 19. This tester has a Rockwell C diamond tip with a cone angle
of 120, and a tip radius of about 200 m. Normal load is applied
through a spring loaded arm in steps of 1 N up to a maximum of 200
N. This tester was further modified by replacing-loaded arm with an
electromagnetic coil to operate at lower load of 10 mN to 30 N with
a detection sensitivity of 10 mN. This tester is equipped with an
acoustic emission detector, and a device which enables the tangential
force in the direction of displacement to be measured. Damage at the
coating-substrate interface usually results in an increase in friction
force, and a generation of an acoustic emission signal.

Figure 19. Schematic illustration of the scratch method for adhesion measurement.

An example of a scratch test, optical micrographs of typical


failure mode of sputtered TiN coating on a steel substrate are presented in Fig. 20a. The acoustic emission curves showing the maximum and minimum signals are plotted as a function of stylus load in
Fig. 20b. As the stylus load was increased to 7.8 N, the coating
started to crack perpendicular to the steel surface at the edge of the
channel. The onset of coating damage was accompanied by a sudden
increase in acoustic emission at loads greater than 7.8 N (region b in
Fig. 20b). The very high normal load can cause the detached coating
particles either to the completely pressed back into the substrate or be
partially removed, thus leaving a smooth surface within the channel.

270 Handbook of Hard Coatings

(a)

(b)
Figure 20. (a) Optical micrographs of channels produced during scribing from
right to left under various stylus loads in Crofer 1700 coated with sputtered
TiN about 1.5 m thick. (b) Acoustic emission signal maxima and minima
taken from the curves recorded within the scratch channel length between 0.5
and 2.5 mm.[149]

Macro- and Micromechanical and Tribological Properties 271

Apparatuses For Scratch Measurement On Microscale. In


principle, the microscratch testers are quite similar to those used for
macroscratch but only with a difference in the normal load and size
and radius of the diamond tip. In the microscratch testers, usually the
normal load ranges from a few hundred N to a few tens of mN and
tips radii varies from about 100 nm to a couple of microns. A
prototype microscratch tester was developed by Wu et al. [307] at IBM
Almaden Research Center by modifying a nanoindenter. With this
instrument, following measurements can be made simultaneously
during a scratch test: applied load and tangential load along the
scratch length (coefficient of friction), critical load, i.e., applied
normal load corresponding to an event of coating failure during a
scratch process, total and plastic depth along the scratch length, the
accumulated acoustic emission (AE) counts versus the scratch length.
The commercially available Nanoindenter, described earlier, has
also been modified for making microscratch measurements.
We describe the microscratch and tangential force attachments
of Nanoindenter which allows making of the scratches of various
lengths at programmable loads. Tangential (friction) forces can also
be measured simultaneously.[16][56] The additional hardware for the
tangential force option includes a set of proximity (capacitance)
probes for measurement of lateral displacement or force in the two
lateral directions along x and y, and a special scratch collar which
mounts around the indenter shaft with hardness indenter, Fig. 21.
The scratch collar consists of an aluminum block, mounted around
the indenter shaft, with four prongs descending from its base. Two of
these prongs hold the proximity probes, and the set screws set them
in place, while the other two prongs hold position screws (and
corresponding set screws). The position screws serve a dual purpose;
they are used to limit the physical deflection of the indenter shaft,
and they are used to lock the indenter shaft in place during tip change
operations. A scratch block is mounted on the end of the indenter
shaft, in line with the proximity probes and the positioning screws.
Finally, the scratch tip itself is mounted on the end of the indenter
shaft, covering the scratch tip. The scratch tip is attached to the

272 Handbook of Hard Coatings

scratch block with two Allen head screws. The scratch tip can be a
Berkovich indenter, or a conventional conical diamond tip with a tip
radius of about 1 to 20 m, and an included angle of 60 to 90
(typically 1 m or 5 m of tip radius with 60 of included
angle.)[56][131][133][302][306] The tip radius does not have to be very
small as it will get blunt readily.

Figure 21. Schematic of the tangential force option hardware (not to scale and
the front and rear prongs not shown).[16]

During scratching, a load is applied up to a specified indentation load or up to a specified indentation depth, and the lateral motion
of the sample is measured. In addition, of course, load and indentation depth are monitored. Scratches can be made either at the constant load, or at ramp up load. Measurement of lateral force allows
the calculations of coefficient of friction during scratching. The
resolution of the capacitance proves to measure tangential load is
about 50 N, therefore, a minimum load of about 0.5 mN can be
measured, or a minimum normal load of about 5 mN should be used
for a sample with coefficient of friction of about 0.1. Microscopy of
the scratch produced at ramp up load allows the measurement of
critical load required to break up of the film (if any) and scratch
width, and general observations of scratch morphology. Additional

Macro- and Micromechanical and Tribological Properties 273

parameters which are used to control the scratch are scratch length
(m), draw acceleration (m/s2), and draw velocity (m/s). The
latter parameters control the speed with which the scratch is performed. The default values of 10 m/s2 and 10 m/s provide safe
rates for performing the scratch. Draw velocity is limited by the
maximum rate of data acquisition (during a scratch the maximum
rate is approximately 2/s) and the length of the desired scratch. Thus,
a scratch with a desired 20 points over 1 mm must have a draw
velocity no greater than 100 m/s.
Wu[302] has used the scratch technique to study the adhesion of
diamondlike carbon and zirconia films deposited on Si(100) substrates. Figure 22 shows the scratch morphology at increasing normal loads, and typical scratch data (normal load, tangential load and
acoustic emission, as well as calculated apparent coefficient of
friction). We note that all three monitored outputs (LC, TG, and AE),
detected the first spallation event of the carbon coating by showing
sudden changes in their output signals. Correlation between the
delamination pattern and the sudden change in the scratch loading is
clearly observed.

Figure 22. Scratch morphology and scratch loading curves of 0.11-m thick
d-c sputtered diamondlike carbon film on a Si substrate.[302]

274 Handbook of Hard Coatings

Gupta and Bhushan[132] measured scratch depths during


scratching and residual depth after scratching. The surface profile of
the coated surface was first obtained by translating the sample at a
low load of about 1.5 mN, which is insufficient to damage the sample
surface. The actual scratches were made by translating the sample
while ramping the loads on the conical tip over different loads, for
instance ranging from 1.5 mN to 45 mN. The actual depth during
scratching was obtained by subtracting the initial profile from the
scratch depth measured during scratching. In order to measure permanent depth, the scratched surface was profiled at a low load of 1.5
mN and was subtracted from the surface profile before scratching. A
typical scratch experiment consisted of seven subsequent steps:
1. Approaching the surface.
2. Indent into sample surface by loading the tip to
1.5 mN.
3. Translating the sample at a constant load of 1.5 mN
at a speed of 5 microns per second.
4. Translating the sample in the opposite direction at
ramping load to a load ranging from 1.5 to 45 mN at
a speed of 5 micron per second.
5. Unloading of the tip to 1.5 mN.
6. Translating the sample at constant load of 1.5 mN
at a speed of 5 micron per second.
7. Final unloading of the tip.
The 500-micron long scratch at ramping normal load was made
during the fourth step, and surface profiles before and after scratch
were obtained during the third and sixth steps, respectively. The
scratch depth profiles obtained during and after the scratch of 20 nm
thick carbon coatings deposited on silicon by cathodic arc, and
sputtering are plotted with respect to initial profile after the
cylindrical curvature is removed, Fig. 23. Reduction in scratch
depth after scratching is observed in Fig. 23. The reduction in
scratch depth after scratching is attributed to an elastic recovery

Macro- and Micromechanical and Tribological Properties 275

after the removal of the normal load. It appears that the scratch depth
after scratching indicates the final depth which reflects the extent of
permanent damage and ploughing of the tip into the sample surface.
We believe that the scratch depths after scratching are probably more
relevant for visualizing the damage that can occur in real applications. The abrupt increase in the coefficient of friction and scratch
depth is associated with damage to the coating. The cathodic arc
coating exhibits an almost constant low coefficient of friction of
about 0.10.15 during the initial stages and an abrupt increase in
friction when normal load exceeds the critical load, the load sufficient to damage the coating. Sputtered carbon coating exhibits a
gradual increase in the coefficient of friction with increasing normal
load from the beginning of the scratch.

Figure 23. Coefficient of friction profile during scratching and scratch depth
during and after scratching as a function of normal load for scratches made on
20 nm thick carbon coatings deposited on silicon by cathodic arc and sputtering
deposition techniques.[132]

276 Handbook of Hard Coatings

Bulk Si exhibits a very low coefficient of friction of 0.05 at the


beginning of a scratch at 2 mN, Fig. 24. The coefficient of friction
remains constant up to 4 mN. The coefficient of friction increases
abruptly from 0.05 to 0.15, and then gradually increases to 0.25, as
the normal load increased from 4 to 18 mN, and to 0.9 when the load
exceeded 18 mN. Particulate debris of submicron size was observed
when the normal load exceeded 18 mN. First abrupt increase in
friction corresponds to an initiation of ploughing of the silicon
surface by the tip, whereas the second abrupt increase in friction
corresponds to a catastrophic damage of the surface.

Figure 24. SEM images of various regions and coefficient of friction profiles as
a function of normal load for 500 m long scratches made on single-crystal
silicon (111) at 2 to 20 mN.[56]

Macro- and Micromechanical and Tribological Properties 277

The dependence of coefficient of friction on increasing normal


load and shape and size of debris generated during scratching can be
used to obtain important information regarding the adhesion of the
coating with the substrate, and how the coating or sample surface is
damaged during scratching.
Apparatuses For Scratch Measurement On Nanoscale.
Nanoscratch measurements were made by Bhushan et al.[53] and
Bhushan[40] using a modified commercial AFM/ FFM (Nanoscope
III from Digital Instruments, Inc. Santa Barbara, Calif.). The modified AFM/ FFM will be described later on in the next section, on the
measurements of friction and wear on nanoscale. The scratches were
made by a three-sided pyramidal diamond tip with a tip radius of
about 100 nm at normal loads ranging from 10 to 100 N. Sample
surfaces were scanned before and after the scratch to obtain the
initial and the final surface topography at a load of about 0.05 N,
over an area larger than the scratches region. AFM images of scratches
made on (111) silicon, PECVD-oxide coated silicon, dry-oxidized
silicon, and C+ implanted (1 10-17 ions cm-2 at 100 keV) silicon are
shown in Fig. 25.[51] As expected, the scratch depth increases with an
increase in normal load. The depths of scratches at 40 N on
PECVD-oxide coated silicon and (111) silicon are about 5 and 20
nm, respectively. PECVD-oxide coated silicon has the largest scratch
resistance followed by dry-oxidized silicon and ion-implanted silicon. Ion implantation showed no improvements on scratch resistance because the depth of the scratches is lower than the depth of
implanted zone beneath the surface.
This study demonstrates that scratches with a depth of a few
nanometers can be made at very low loads using AFM/ FFM. This
help characterizing the scratch resistance/ adhesion of top few top
tens of monolayers of bulk materials or coating of the order of a few
nanometers thickness.

278 Handbook of Hard Coatings

Figure 25. Surface profiles for scratched (a) (111) single-crystal silicon, (b)
PECVD-oxide coated Si, (c) dry-oxidized Si and (d) C+-implanted Si. The loads
used for various scratches at ten cycles are indicated in the plot.[51]

Macro- and Micromechanical and Tribological Properties 279


3.0

MEASUREMENT OF FRICTION AND WEAR

Friction and wear between two moving surfaces depends on


mechanical properties of the mating materials such as: hardness, elastic
modulus, fracture toughness; other physical and chemical properties
such as: thermal conductivity, surface energy, adsorption characteristics, chemical reactivity; surface conditions such as: roughness and
apparent area of contact; and operating conditions such as: load, speed,
interface temperature, and environment. Various friction mechanisms such as adhesion and ploughing, that contribute to friction and
wear, are strongly affected by the properties of mating surfaces and
operating conditions. Surface roughness of mating surfaces and area
of contact between these surfaces play a key role in deciding friction.
For instance, rigid thin-film magnetic disk are deliberately roughened to reduce the area of contact to obtain lower friction between a
head slider and the magnetic disk. The friction and wear tests should
be performed under the close-to-ideal conditions. The conditions and
test geometry for friction and wear tests should be selected on the
basis of the actual operating conditions of the components.
With the advent of new surface imaging tools like atomic force
microscope (AFM), friction force microscope (FFM), and point
contact microscope (PCM), it is possible to perform friction tests at
ultra small loads of a couple of N against a sharp tip of about 10 to
100 nm tip-radius which simulates a single-asperity contact conditions.[39][40][58] Sliding tests with a single-asperity contact are more
close-to-ideal conditions of micromechanical devices, and help understand the failure mechanisms.
In this section, we present the apparatuses used for friction and
wear measurements on macro- to nanoscales.
3.1

Friction and Wear Measurements on Macroscale

Accelerated friction and wear tests on macroscale are conducted to rank the friction and wear resistance of coatings, or bulk
materials to optimize their selection or development for specific
applications. After these coatings or bulk materials have been ranked
by accelerated friction and wear tests, the most promising candidates

280 Handbook of Hard Coatings

(typically from one to three), should be tested in the actual machine


under the actual operating conditions (functional tests). Accelerated
friction and wear tests should accurately simulate the operating
conditions to which the material pair will be subjected. If these tests
are properly simulated, an acceleration factor between the simulated
test, and the functional tests can be empirically determined so that
the subsequent functional tests can be minimized, saving considerable test time. Standardization, repeatability, short testing time, and
simple measuring and ranking techniques are desirable in these
accelerated tests. The coefficient of friction of a material, or a
coating depends not only on the counterface materials, but on the
operating conditions such as speed, load, lubrication, and environment. Shown in Table 1 are a couple of examples of coefficients of
friction and wear rate for various bulk material and coating combinations. We note that with a proper selection of materials and operating
conditions, one can achieve as low coefficient of friction as 0.02
between two solid materials without using any lubricant. In this
section, we will present the design methodology and typical test
geometries for friction and wear tests.
Design Methodology. Design methodology of a friction and
wear test consists of four basic elements: simulation, acceleration,
specimen preparation, and friction and wear measurements. Simulation is most critical, but no other elements should be overlooked.
Proper simulation ensures that the behavior experienced in the
test is identical to that of the actual system. A successful simulation
requires the similarity between the functions of actual system, and
those of the test system, i.e., similarity of inputs and outputs, and of
the functional input-output relations. To obtain this similarity, first
the mating materials, the lubricant, and the operating conditions of
the test should be the same as the actual system requirements.
Selection of the contact geometry depends on the geometry of the
function to be simulated. Other factors besides contact type that
significantly influence the success of a simulation include type of
motion, load, speed, and operating environment (contamination,
lubrication, temperature, and humidity). Specimen preparation plays
a key role in obtaining repeatable/reproducible results.

Macro- and Micromechanical and Tribological Properties 281

Table 1. Typical Values of Coefficients of Friction for Selected

Materials and Coatings in the Ambient Environments (~ 22C, 50%


RH) Unless Otherwise Specified[46][64][142][227]
Materials pair
Bulk
material/Coating
and Substrate

Counterface material

Coefficient of friction

Bulk materials
Aluminum

Aluminum

1.0

Copper

Copper

0.8

Silver

Silver

0.9

Brass

Brass

0.4

Cast iron

Cast iron

0.6

Mild steel
Brass
Bronze
Copper

0.4
0.25
0.25
0.3

Mild steel

0.8

Aluminum
Nickel
Silver
Copper
Brass
Bronze
Babbitt
Graphite
PTFE

0.5
0.7
0.5
0.8
0.3
0.3
0.3
0.15
0.1

Tool steel

Tool steel

0.4

Nickel

Brass
Polyethylene
PTFE
Nickel

0.25
0.25
0.1
1.1

Chromium

Chromium

0.4

Silicon nitride

Silicon nitride

0.3

Mild steel

(Contd.)

282 Handbook of Hard Coatings

Table 1. (Contd.)
Materials pair
Bulk
material/Coating
and Substrate
Tungsten carbide

Counterface material

Coefficient of friction

Tungsten carbide

0.35

Natural diamond

Steel
Natural diamond

0.4
0.05

PTFE

PTFE

0.05

Coatings/Surface
treatments
MoS2 (Sputtered)

Steel

0.050.1 (Ambient)

WC-Co

0.02 (Vacuum)

Steel

0.10.2 (Ambient)

PTFE (Air sprayed)

Steel

0.40.6 (Dry)
0.030.1

Silver/Gold (Sputtered)

Steel

0.10.25

Al2O3 (CVD)

Steel

0.20.5

TiN (Sputtered)

Steel

0.150.5

TiC(Sputtered)

Steel

0.20.5

Diamond(HFCVD)/Si

Steel

0.10.2

a-C:H (Sputtered)

Steel

0.150.3

a-C:H (PEPVD)

Steel

0.150.3

Graphite (air sprayed)

The coefficient of friction is generally measured during a wear


test. It is calculated from the ratio of friction force to applied normal
force. The stationary member of the material pair is mounted on a
flexible member, and the frictional force is measured using the strain
gauges or displacement gauges.[36][38] Common wear measurements
are weight loss, volume loss, or displacement scar width or depth, or
other geometric measures, and indirect measurements such as, time

Macro- and Micromechanical and Tribological Properties 283

required to wear through a coating or load required to cause severe


wear, or a change in the surface finish, size of indentation marks, or
width and depth of scratches. Scanning electron microscopy of worn
surfaces is commonly used to measure microscopic wear. Other less
commonly used techniques include radioactive decay, scanning tunneling microscopy (STM), and atomic force microscopy (AFM).
The resolution of various macroscopic wear measurement techniques are compared in Table 2.

Table 2. Resolution of Several Macroscopic Wear Measurement


Techniques

Typical Test Geometries. The choice of a test geometry


depends on the wear situation to be simulated. In this section, we will
present a few most commonly used test geometries to rank coatings
and materials in terms of their resistance to sliding wear, and rolling
contact fatigue wear. However, depending on the requirements, tests
can be performed with replicas and facsimiles of the actual devices.
Many accelerated test apparatuses are commercially available that
allow control of such factors as sample geometry, applied load,
sliding velocity, ambient temperature, and humidity. Bayer,[24][27]

284 Handbook of Hard Coatings

Benzing et al.[31] Bhushan,[37][38] Bhushan and Gupta,[46] Clauss,[87]


Nicoll,[209] and Yust and Bayer[313] have reviewed the various friction and wear testers that have been used in various tribological
applications.
Sliding Friction And Wear Tests. In the sliding wear test
configurations, ball, pin, cylinder, or ring of one material slides over
the disk, block, or cylinder of another material in the presence or
absence of a lubricant. The most commonly used interface geometries used to rank coatings and materials in terms of their resistance
to sliding wear are schematically shown in Fig. 26, and are compared
in Table 3. With these test geometries, static, or dynamic loading can
be applied, and the tests can be performed either in the presence or
absence of a lubricant. In the pin-on-disk test apparatus (Fig. 26a),
the pin is held stationary, and the disk rotates or oscillates. The pin
can be a non rotating ball, a hemispherically tipped rider, a flat-ended
cylinder, or even a rectangular parallelepiped. In the pin-on-flat test
apparatus (Fig. 26b), a flat moves relative to a stationary pin in
reciprocating motion, such as in a Bowden Leben apparatus. The
pin-on-cylinder test apparatus (Fig. 26c), is similar to the pin-on-disk
apparatus, except that loading of the pin is perpendicular to the axis
of rotation or oscillation. In the thrust-washer test apparatus (Fig.
26d), the flat surface of a washer (disk or cylinder) rotates or
oscillates on the flat surface of a stationary washer, such as in the
Alpha model LFW-3. In the pin-into-bushing test apparatus (Fig. 26e),
the axial force necessary to press an oversized pin into a bushing is
measured, such as in the Alpha model LFW-4. The normal (axial)
force acts in the radial direction, and tends to expand the bushing; the
radial force can be calculated from the material properties, the
interference, and the change in the bushings outer diameter. Dividing the axial force by the radial force gives the coefficient of friction.
In the rectangular flat, on a rotating cylinder test apparatus (Fig. 26f),
two rectangular flats are loaded perpendicular to the axis of rotation
or oscillation of the disk. The crossed cylinders test apparatus (Fig.
26g), consists of a hollow (water cooled) or solid cylinder as the
stationary wear member and a solid cylinder as the rotating or

Macro- and Micromechanical and Tribological Properties 285


oscillating wear member that operated at 90o to the stationary member, such as in Reichert wear tester. The four ball test apparatus (Fig.
26h), also called the Shell four-ball tester, consists of four ball in the
configuration of an equilateral tetrahedron. The upper ball rotates
and slides against the lower three balls, which are held in a fixed
position. This test configuration is extensively used to study liquid
lubricants.

Figure 26. Schematic illustration of typical interface geometries used for sliding
friction and wear tests: (a) pin-on-disk, (b) pin-on-flat, (c) pin-on-cylinder, (d)
thrust washer, (e) pin-into-bushing, (f) rectangular flats on rotating cylinder, (g)
crossed cylinders and (h) four ball.[46]

286 Handbook of Hard Coatings

Table 3. Details of Typical Test Geometries for Sliding Friction and


Wear Testing

Rolling-Contact Fatigue Wear Tests. A number of rollingcontact fatigue (RCF) tests are used for testing materials and lubricants for rolling-contact applications such as rolling element bearings, and gears. In RCF test apparatus, basically a pair of driven
rollers are pressed against one another, and the surface damage on
roller surfaces is monitored with the number of cycles. The surface
damage could be the appearance of cracks, change in surface texture
or roughness, or spalling of material. In general, rolling-contact
fatigue (RCF) wear is compared in terms of number of cycles
sufficient to result a specific damage on the rollers. The most commonly used interface geometries used to rank coatings and materials
in terms of their resistance to rolling-contact fatigue wear are schematically shown in Fig. 27. The disk-on-disk test apparatus (Fig.
27a) uses two disks or a ball-on-disk rotating against each other
on their outer surfaces (edge loaded). The rotating four ball test

Macro- and Micromechanical and Tribological Properties 287

apparatus (Fig. 27b) consists of four balls in the configuration of an


equilateral tetrahedron. The rotating upper ball is deadweight loaded
against the three supporting balls (positioned 120 apart), which
orbit the upper ball in rotating contact. The rolling-element-on-flat
test apparatus consists of three balls or rollers equispaced by a retainer
that are loaded between a stationary flat washer and a rotating
grooved washer (Fig. 27c). The rotating washer produces ball motion, and serves to transmit load to the ball and the flat washer.[42]

Figure 27. Schematic illustration of typical interface geometries used for


rolling-contact fatigue wear tests: (a) disk-on-disk, (b) rotating four ball, (c)
balls-on-flat.[46]

3.2

Friction and Wear Measurements on Micro- and Nanoscale

Friction and wear measurements on a micro-to nanoscale can


be performed using a nanoindenter in scratch mode which was
described earlier, or an atomic force microscope, or friction force

288 Handbook of Hard Coatings

microscope.[41][58][185] The coefficient of friction can be measured


using a nanoindenter in the scratch mode by monitoring the friction
force during scratching at low loads on the order of a few hundred
microNewtons. The selection of the normal load should be such that
it is insufficient to cause damage to the sample surface, or to result in
ploughing of the tip into the sample surface. In these experiments the
diamond tip can be replaced with tips of different materials to study
the friction of a coating against a variety of materials.
The atomic force microscope (AFM), was developed by Binning, et al.,[59] to measure ultrasmall forces on the order of 1 nN or
less present between the AFM tip surface, and a sample surface.
Binning and co-workers measured these small forces by measuring
the motion of a cantilever beam having an ultrasmall mass. For
imaging surfaces, AFM can be thought of as a nanometer scale
profiler. A sharp tip at the end of a cantilever is brought into contact
with a sample surface by moving the sample with piezoelectric
transducers. During initial contact, the atoms at the end of the tip
experience a very weak repulsive force due to electronic orbital
overlap with the atoms in the sample surface. The force acting on
the tip causes a lever deflection which is measured by tunneling, [59] capacitive detectors,[192] or optical detectors such as laser
interferometry.[109][185] The optical techniques are believed to be
reliable and easily implementable detection methods. Mate et al.[185]
were the first to modify an AFM in order to measure both normal and
friction forces; the modified instrument is generally called the
friction force microscope (FFM), or lateral force microscope
(LFM). This group measured the atomic-scale friction of a tungsten tip sliding on a basal plane of a single grain of highly oriented
pyrolytic graphite (HOPG). Since then, several research groups have
developed FFMs of various design. Ruan and Bhushan[242] Bhushan
et al.[53][58] and Bhushan[40][41] have modified a commercial AFM/
FFM (Nanoscope III from Digital Instruments, Inc., Santa Barbara,
Calif.) to conduct studies of friction, scratching, wear, indentation,
lubrication, and nanofabrication.

Macro- and Micromechanical and Tribological Properties 289

A modified AFM/FFM is shown in Fig. 28. Simultaneous


measurements of friction force, and surface roughness can be made
using this instrument. The sample is mounted on a piezoelectric tube
(PZT) scanner which can precisely scan the sample in the horizontal
(xy) plane, and can move the sample in the vertical (z) direction. A
sharp tip at the free end of a cantilever is brought in contact with the
sample. A laser beam from a laser diode is focused on to the back of
a cantilever near its free end. The cantilever is tilted downwards at
about 10o with respect to the horizontal plane. The beam is reflected
from the cantilever, and is directed through a mirror on a split
photodetector with four quadrants. Two quadrants (top and bottom)
of the detector are used for topography measurements. As the sample
is scanned under the tip, topographic features of the sample cause the
tip to deflect in the vertical direction. This tip deflection will change
the direction of reflected laser beam, changing the intensity difference between the top and bottom photodetectors (AFM signal). For
topographic imaging, applied normal force is kept constant, a feedback current is used to modulate the voltage applied to the PZT
scanner to adjust height of the PZT. The PZT height variation is a
direct measure of surface roughness of the sample. For measurement
of friction force being applied at the tip surface during sliding, the
other two (left and right) quadrants of the photodetector (arranged
horizontally) are used. The sample is scanned back and forth in a
direction orthogonal to the long axis of the cantilever beam. Friction
force between the sample and the tip will produce a twisting of the
cantilever. As a result, the laser beam will be reflected out of the
plane defined by the incident beam, and the beam reflected vertically
from an untwisted cantilever. This produces an intensity difference
between the left and right detectors (FFM signal), and is directly
related to the degree of twisting, and hence, to the magnitude of
friction force. One problem associated with this method, is that any
misalignment between the laser beam, and the photodetector axis
would introduce error in the measurement. However, by following
the procedures developed by Ruan and Bhushan[242] in which the
average FFM signal as the sample is scanned in two opposite direction,

290 Handbook of Hard Coatings

and is subtracted from the friction profiles of each scan, the misalignment effect can be eliminated. By following procedures developed
by Ruan and Bhushan[242] voltage corresponding to normal and
friction forces can be converted to force units. By making measurements at various normal loads, the average value of the coefficient of
friction is obtained which can then be used to convert the friction
profile to the coefficient of friction profile. Thus, any directionality
and local variation of friction can be measured. Surface topography data can be measured simultaneously with the friction data, and
a local relationship between the two profiles can be established.

Figure 28. Schematic of AFM/FFM instrument.[53]

Bhushan et al.[53][58] have measured microfriction data for


magnetic tapes and magnetic disks by FFM, by following the aforementioned procedure of microfriction measurements. Figure 29 are
the surface roughness profile, the slope of surface roughness profile
taken along the sample sliding direction, and friction profile of an
unlubricated, textured, rigid disk, respectively. No direct correlation
between the surface profiles, and the corresponding friction profiles
is observed in these figures, e.g., high and low points on the friction

Macro- and Micromechanical and Tribological Properties 291

profile do not correspond to high and low points on the roughness


profile, respectively. However, there is a correlation between the
slope of the roughness profiles, and the corresponding friction profiles. Bhushan and Ruan[52] have presented evidence of similar
correlation on a microscale of 500 nm 500 nm for the tape and disk
samples. A ratchet mechanism explains the correlation of friction to
the local surface slope, with the ascending edge of an asperity having
a larger friction force than that at the descending edge. They also
reported that the macroscale coefficient of friction is about a factor of
five larger than that of the average microscale coefficient of friction
of the corresponding samples.[52] Larger macro coefficient of friction may be the result of plowing effect associated with the macroscale
measurement, since visible wear scars were observed on the samples
after the measurement.[58] Bhushan and Ruan[52] had also observed
the directionality in the local variation of microscale friction data as
the samples were scanned in either direction, resulting from the
scanning direction, and the anisotropy in the surface topography.
Bhushan et al.[53] have also made wear measurement on
microscale by scanning the sample (in 2D) using a FFM. Surface
profiles of the wear scars generated in a single cycle on an aspolished, unlubricated disk at different loads are shown in Fig. 30.
Wear depths as a function of normal load for the as-polished disks
(lubricated and unlubricated) are shown in Fig. 31. The normal force
for the imaging was about 0.5 N, and the loads used for the wear
were 20, 50, 80, and 100 N. They reported that wear takes place
relatively uniformly across the disk surface, and appears to be essentially independent of the lubrication for the disks studied. For both
lubricated and unlubricated disks, the wear depth increases slowly
with load at low loads with almost the same wear rate. As the load is
increased to about 60 N, wear increases rapidly with load. The wear
depth at 50 N is about 14 nm, slightly less than the thickness of the
carbon film. The rapid increase of wear with load at loads larger than
60 N is an indication of the breakdown of the carbon coating on the
disk surface. Wear depths as a function of number of cycles for the
as-polished disks (lubricated and unlubricated) are shown in Fig. 32.

292 Handbook of Hard Coatings

Again, for both unlubricated and lubricated disks, wear initially


takes place slowly with a sudden increase between 40 and 50 cycles at
10 N. The sudden increase occurred after 10 cycles at 20 N. This
rapid increase is associated with the breakdown of the carbon coating.
The preceding microfriction and microwear results illustrate
that friction force microscopy can be used for various tribological
studies on microscale. By this technique it is possible to identify
normal load and number of cycles sufficient to initiate damage to the
20-nm thick hard carbon overcoat on a thin-film magnetic disk.

Figure 29. A 200 nm 200 nm scan of an unlubricated, textured disk. (a)


Surface roughness profile ( = 2.9 nm), (b) slope of the roughness profile (mean
= -0.03, = 0.22), and (c) friction profile (mean = 5.7 nN, = 1.6 nN) for a
normal load of 140 nN.[53]

Macro- and Micromechanical and Tribological Properties 293

Figure 30. Surface profile of an as-polished unlubricated disk showing the worn
region (center 2 m 2 m) after one cycle of wear. The normal load and the
number of test cycles are indicated in the figure.[53]

Figure 31. Wear depth as a function of load for both lubricated and unlubricated
as-polished disks after one cycle.[53]

294 Handbook of Hard Coatings

Figure 32. Wear depth as a function of number of cycles for as-polished,


lubricated and unlubricated disks at 10 N and for an as-polished, unlubricated
disk at 20 N.[53]

3.3

Friction and Wear Measurements in MEMS

Reduction of friction and wear are critical to successful performance of micromechanical elements requiring relative motion. Since
electrostatic forces, van der Waals forces, and viscous drag forces in
liquids change in proportion to the surface area of the element, these
become dominant over electromagnetic forces, and inertial forces,
which are in proportion to the volume of the element. In MEMS, the
electrostatic and van der Waals forces are the major sources of the
normal force which directly affect the friction force. Wear of sliding
components also need to be minimized. One of the MEMS devices in
which friction and wear issues are critical, is a micromotor. Friction
and wear studies have been conducted on actual or simulated MEMS
components, as well as on the coupons in conventional accelerated
tribological tests.
To in-situ measure static friction of rotor-bearing interface in a
micromotor, Tai and Muller[274] measured the starting torque (voltage)
and pausing position for different starting positions under a constant
bias voltage. A friction-torque model was used to obtain the coefficient

Macro- and Micromechanical and Tribological Properties 295

of static friction. To in-situ measure kinetic friction of the turbine


and gear structures, Gabriel et al.[114] used a laser-based measurement system to monitor the steady-state spins and decelerations. Lim
et al.[179] designed and fabricated a polysilicon microstructure to insitu measure static friction of various films. The microstructure
consisted of shuttle suspended above the underlying electrode by a
folded beam suspension. A known normal force was applied, and
lateral force was measured to obtain coefficient of static friction.
Gabriel et al.[114] reported wear of turbine after spinning at 2500 rps
for 5 minutes. Mehregany et al.[195] developed a quantitative method
for in-situ wear measurements in micromotors. They used a wobble
micromotor under electric excitation for quantitative wear measurement. Since the gear ratio of the wobble micromotor depends on the
bearing clearance, changes in the gear ratio can be a direct measure
of wear in the motor bearing.
Conventional accelerated friction and wear test apparatuses
have been used on a coupon level to measure friction and wear of
various thin films. To measure static friction, Deng et al.[93][95] used
a measurement technique shown in Fig. 33, with two millimetersized flat coupons in controlled environment. A polyvinylidene
difluoride (PVDF) bimorph cantilever was used to generate a repeatable tangential force from 0 to 0.8 mN. The normal force was applied
by electrostatic attraction through a bias voltage, and it was in the
range of 1 mN. A capacitance-voltage (C-V) meter that supplied the
clamping voltage to the samples was also used to detect the displacement of the samples. When the samples slid over each other, the
charge equilibrium at the interface was disturbed, and the air gap
between the two surfaces was changed. The change in the charge
corresponded to the displacement of the sample. During an experiment, the electrostatic clamping force was applied using the bias
voltage of a high frequency CV meter, and the capacitance was
measured simultaneously. Next, an adjustable DC voltage was applied to the PVDF bimorph cantilever to generate the tangential force
that pushed the sample forward until a displacement was observed as
a sudden change of capacitance indicated by the CV meter. The

296 Handbook of Hard Coatings

coefficient of static friction was then obtained from a plot of tangential force as a function of normal force. Figure 34 shows the tangential force as a function of normal force for the sputtered SiO2 coating
from a fused silica target against itself in different environments. The
tangential force versus normal force plots for the SiO2/ SiO2 pair in
air before baking is 0.54, in air after baking is 0.21, and in high
vacuum of 10-5 torr is 0.36.[95]

Figure 33. Schematic of the setup used to measure static friction between two
flats samples at very low loads.[95]

Figure 34. Tangential force versus normal force plots for SiO2/ SiO 2, pair in
ambient air as well as in high vacuum (10-5 torr). The coefficient of friction is
0.54 before baking the sample and 0.21 after. The coefficient of friction is 0.36 at
10-5 torr after the sample was baked.[95]

Deng et al.[93][95] developed another setup to measure static


friction between a flat coupon, and a three-millimeter radius aluminum bullet coated with various materials in controlled environment,
Fig. 35. The normal force on the bullet was applied by placing a 10 N

Macro- and Micromechanical and Tribological Properties 297

weight rotating on a threaded rod. The arm holding the bullet was
translated horizontally by a translation table using a piezoelectric
positioning device. The push of the piezoelectric positioning device
was controlled by a high voltage power supply, supervised by a
programmable controller. The tangential force on a bullet was measured by a charge mode piezoelectric force sensor. The sliding test
was performed by lowering the bullet onto the substrate, adjusting
the normal force on the bullet, and pushing the bullet forward using
the translational device until slippage was observed on the forcedisplacement curve. The bullet was then raised. Deng and Ko[95]
used this setup to measure friction between various material/ coating
pairs in ambient air, in UHV (8 10-10 torr), argon, and oxygen
environments. The coefficients of friction obtained by using this
setup are compared in Table 4. The values of coefficients of friction
obtained by using this setup were found to be very close to those
values obtained from in situ friction measurements in micromotors.[95]

Figure 35. Schematic of the setup used to measure static friction between a
coated bullet and a flat in controlled environment.[95]

298 Handbook of Hard Coatings

Table 4. Measurement Results from the Setup Shown in Fig. 35.


SiNx: PECVD Silicon Nitride[95]

Noguchi et al.[212] measured static and kinetic friction between


a millimeter-sized flat, and coated coupons using an experimental
setup shown in Fig. 36. The millimeter sized flat mounted on a
mover was driven electrostatically on the surface of a fixed flat to
measure static friction. The gap between the mover and the fixed
plate was 0.5 mm. The driving electrode was attached to the backside
of the fixed plate. The AC voltage was applied between the mover and
the driving electrode. The static friction was calculated from the minimum voltage required to start the flat moving. The kinetic friction was
measured from the transient measurement of the flat traveling across
the gap. When a voltage was applied between the flat and the driving
electrode, the flat moved toward the insulating glass pulled by the
electrostatic force. The transit time during which the mover crosses the
gap was measured using a timer, a video tape recorder, and a monitor.

Macro- and Micromechanical and Tribological Properties 299

(a)

(b)
Figure 36. Schematic of the setup used to measure static and kinetic friction
between two flat samples.[212]

For friction and wear studies, Suzuki et al.[268] used a magnetic


disk drive with a cylindrical rider sliding against a disk surface (see
Ref. 38). They mounted the rider on a strain-gauged I beam to
measure coefficient of static and kinetic friction. Beerschwinger et
al.[28] conducted wear studies by sliding a micromachined sample
against a flat disk in a conventional pin-on-disk arrangement. To
better simulate MEMS interactions, they used specially fabricated
microstructures sliding against a disk surface. Microstructures with

300 Handbook of Hard Coatings

different apparent area of contact and heights were micromachined


on single-crystal silicon and polycrystalline silicon by using photolithography. The SEM images of micromachined surfaces, sample
holder with circularly arranged holes for the placement of triangles,
and pin-on-disk type of experimental setup for sliding tests are
shown in Fig. 37.[28] Different geometry, and size of micromachined
surface gives different contact pressures ranging from 50 to 1700 Pa.
The wear rate was calculated from the changes in the dimensions of
a microstructure on a sample measured from the surface profiles
before and after the wear tests. The wear rate of LPCVD diamondlike
carbon (DLC), PECVD silicon nitride (SiN), silicon dioxide (SiO)
by wet hydrogen at 1050C; and p-type (100) single-crystalline
silicon (SCS) microstructures sliding against a DLC coated silicon
substrate as a function of sliding distance at a contact pressure of 170
Pa, and a normal load of 30 N, are compared in Fig. 38. The DLC
coated microstructure exhibits the lowest wear rates.

4.0

MECHANICAL AND TRIBOLOGICAL PROPERTIES


OF DIAMOND AND AMORPHOUS CARBON
COATINGS

Mechanical, friction and wear properties of hard coatings vary


over a wide range with their physical and metallurgical properties,
which are strongly influenced by the deposition conditions and the
interfacial chemistry. Among the most widely used hard coatings are
polycrystalline diamond, amorphous carbon nitrides, and carbides of
various metals. Lately, more attention has been paid to boron nitride,
and carbon-nitrogen compound coatings because of their outstanding hardness (> 50 GPa). Coatings of amorphous carbon with and
without hydrogen have been produced by numerous researchers to
protect surfaces of mechanical components such as magnetic thinfilm rigid disk and recording heads, and MEMS devices against
wear. In this section, we present the mechanical and tribological

Macro- and Micromechanical and Tribological Properties 301

properties of polycrystalline diamond, and amorphous carbon


coatings in detail because of their potential applications in various
devices. For completeness, a brief overview of various CVD, and
PVD techniques used to deposit polycrystalline diamond and amorphous carbon coatings is also given.

Drilled holes in
circular arrangements

Specimen holder

Frequency
counter
Power supply
Test structure

X-rotation/
translation
Y-translation
Y-rotation

Substrate
Adjustable
base plate
DC motor

Figure 37. (a) SEM images of triangles micromachined on silicon by photolithography; total area of contact of triangle A and B is 1.88 10-1 mm-2 and 2.36
10 -2 mm -2, respectively. (b) Sample holder with circular arranged holes for the
placement of test triangles. (c) Schematic of the set-up used for sliding triangles
on a rotating disk. [28]

302 Handbook of Hard Coatings

Figure 38. Comparison of wear rates of various coatings pairs at a contact


pressure of 170 Pa and a normal force of 30 N obtained by using the set-up
shown in Fig. 37.[28]

4.1 Deposition Processes


Polycrystalline diamond, and amorphous carbon coatings have
been deposited by various deposition conditions, and have been
extensively studied by numerous researchers for their applications in
tribology, optics, and electronics. Depending on the deposition conditions, one can synthesize coatings of almost pure diamond with
very high hardness, as well as very soft coatings of carbon polymers.[12][14][46][86][176][189][238][239][241][256][257][284][286] The sp3-bonding in amorphous carbon results in many of the outstanding properties of diamond such as high hardness, elastic modulus, low friction,
optical transparency, and chemical inertness. Atomic hydrogen is
believed to be responsible for formation of sp3-bonding.[12][14]
A wide variety of CVD techniques have been used to synthesize pure diamond films at low pressures on a variety of nondiamond substrates since the breakthrough of diamond deposition on
nondiamond substrates.[96][154][259] These techniques have been reviewed from time to time by many researchers. [10][11][13][20]
[77][97][98][148][159][208][256] [260][310][311] Not much success has been
achieved in synthesizing diamond coatings by PVD. In general,
CVD polycrystalline diamond coatings are deposited from a carbon-

Macro- and Micromechanical and Tribological Properties 303

containing precursor gas, and formation of nondiamond carbon is


prevented by the presence of a selective enchant, such as atomic
hydrogen. To obtain high deposition rates, energetic and efficient
decomposition of precursor gas and diamond deposition at lower
substrate temperatures; CVD techniques have been assisted by a
wide variety of thermal-hot filament,[175][187] electrons,[247] laser,[158]
plasma-DC/RF discharge,[266] DC arc jet plasma,[4] DC plasma
jet, [113][165][214] RF thermal plasma, [162][186][188] microwave
plasma,[154] electron cyclotron resonance,[103][150][155][267] microwave
plasma torch,[198] and combustion flame processes.[141] In plasmaassisted CVD techniques, precursor gases are activated by plasma
generated by electrical discharges at various frequencies or induction heating. The role of plasma is to produce atomic hydrogen and
proper carbon precursor for the growth of diamond film. In plasma
processes, the atomic hydrogen is produced by electron impact
dissociation of molecular hydrogen, and its kinetic energy depends
on the impact energy of electrons. The kinetic energy of atomic
hydrogen produced by plasma is much higher than that produced by
thermal cracking.
These assisted CVD techniques, although different in their
deposition reactions and deposition configurations, have a few common features: the presence of a gas-phase nonequilibrium in the
immediate vicinity of the substrate generated by some sort of gasphase activation, and growth of diamond on elevated substrate temperatures of 500 to 1200C. The deposition rates of diamond coatings deposited by various CVD techniques have been observed to vary
from 0.1 mh-1 to approximately 1 mm/h. Usually higher growth
rates are associated with higher pressure processes such as, combustion and thermal plasma, which run at pressures up to 1 atm. It has
been reported that DC plasma jet CVD process resulted in the highest
deposition rate of 930 mh-1, whereas the microwave-assisted CVD
resulted in high quality diamond films on non-diamond substrates such
as Si, Mo, and WC over an area of 40 cm2. However, deposition rates
in the microwave-assisted CVD were low (0.1 mh-1).
The nucleation density, growth mechanisms, and quality of
diamond films (characterized by sp3-bonding fraction), grain size,
and presence of nondiamond component are strongly influenced by

304 Handbook of Hard Coatings

numerous deposition parameters in a very complex manner. It is


difficult to set up any trend towards the dependence of deposition
parameters on the properties of diamond films. The parameters of
particular importance in CVD depositions are: appropriate combination of carbon, hydrogen and/or oxygen in precursor gas mixture, substrate temperature, total pressure of precursor gas mixture,
plasma density, and substrate biasing. The substrate temperature
from 500 to 1200C is essential for diamond deposition. Films
deposited above 1200C are found to be graphitic, whereas films
deposited below 500C are amorphous. The substrate temperature is
also found to affect the deposition rates,[258][260] and crystallite orientation of diamond grains. The total pressure of hydrocarbon-hydrogen gas mixture determines the recombination length, the life time,
and the drift distance of atomic hydrogen.[17] The precursor gas
temperature affects the deposition rates, higher gas temperatures
yield higher deposition rates. In the low-pressure plasma-assisted
CVD, the deposition rates were found to increase linearly with the
plasma density.[292] The substrate biasing which affects the energy of
electron and/or ion bombardment on the growing film controls the
kinetics of growth, nucleation density, which in turn affects diamond
quality, and physical properties of the films.[205][247][264]
Diamond films synthesized by different CVD techniques can
range from rough well-faceted films with a crystallite size of a couple of
microns, to optically smooth films with a crystallite size of a couple of
nanometers. Thermochemical, chemomechanical, and plasma/ion beam/
laser techniques have been explored to polish diamond films.[1][48][54][61]
[63][110][126][140][250][280][281][286][314] A summary of these polishing techniques is given in Table 5.[54] Thermochemical and chemomechanical
techniques are contact techniques, so these can be used for surfaces that
are planar. Plasma, ion beam, and laser techniques do not require bulk
sample heating, and are noncontact techniques that can be used on
nonplanar surfaces, although these are line-of-sight processes. The
material removal rates using these techniques have thus far been
small. High-energy ion beams have the potential for high material
removal rates. However, heat generated during sputtering results in
graphitization of the diamond film, so the laser technique emerges as
the most desirable among the noncontact techniques.

Macro- and Micromechanical and Tribological Properties 305

Table 5. Summary of Various Polishing Techniques[54]

306 Handbook of Hard Coatings

Like diamond deposition, amorphous carbon coatings have


also been deposited CVD techniques by using a hydrocarbon gas
mixed with hydrogen as a starting materials, and the gas mixture is
activated by similar thermal or plasma activation processes. RF glow
discharge-assisted CVD followed by microwave discharge-assisted
CVD are the most widely used for deposition of dense hydrogenated
carbon coatings.[32][68][78][160][186][200][215] Depending on the deposition technique, a-C or a-C:H films can have carbon phases of different chemical composition (%H), atomic arrangement (especially the
coordination of the nearest neighbors and the degree and scale of order),
and electronic structure (binding type). It is believed that deposition
of sp3-bonded carbon requires the depositing species to have kinetic
energies on the order of 100 eV or higher, well above those obtained
in thermal processes like evaporation (~ 0.1 eV). The species must
then be quenched into the metastable configuration via rapid energy
removal. Excess energy, such as that provided by substrate heating,
is detrimental to the achievement of a high sp3-fraction.
In PVD techniques, amorphous carbon coatings are deposited
by thermal evaporation, DC/RF glow discharge sputtering, or single/
dual ion beams sputtering,[3][309] ion plating,[296] direct ion beam
(DIB),[106][107][271][293] and vaporizing by arc discharge, [7][9][67]
[91][181] or a laser source (such as CO laser),[112] in argon or argon
2
and hydrogen gas mixture, and the growing carbon film is bombarded with energetic ions of 20500 eV energy, which is believed to
be necessary to promote sp3-bonding. To achieve good adhesion of
a-C and a-C:H films on nondiamond substrates, and materials that do
not readily form carbides, thin silicon interlayers of a couple of
nanometers are commonly applied.[121][122][125]
The common feature of various PVD techniques is that the
carbon species arrive with an energy significantly greater than that
represented by the substrate temperature. Generically, high kinetic
energy of carbon species, which depends on the deposition configuration and substrate biasing, results in higher degree of sp3-bonding,
and improved adhesion and mechanical properties of the coatings.
The extent to which the properties of a film are affected by ion

Macro- and Micromechanical and Tribological Properties 307

bombardment is dependent on both the kinetic energy, and the flux


density of bombarding ions, as well as the ions to be deposited atom
ratio to the mass ratio of the ions relative to the surface atoms. Ionsurface interactions during deposition include several processes:
increased adatom migration, desorption of adsorbed surface impurities, displacement of atoms in the near surface region, recoil implantation, sputtering, re-sputtering of deposited species, working gas
entrapment. The severity of these interactions depends on the ion
energy and dose which in turn affect the film density, stoichiometry,
growth kinetics (rates), residual stresses, hardness, and other mechanical properties.[237][276]
The structure and properties of a-C or a-C:H films are strongly
influenced by the deposition conditions, deposition configurations,
and the constituents of precursor gas in a very complex manner. The
energy of carbon species which depends on the kinetic energy
involved in a particular PVD technique, and substrate biasing in the
CVD techniques, is one of the most sensitive deposition parameters.
The films deposited at low kinetic energy of a few tens of electronvolts
by ion plating technique using benzene were found to be polymer
like; their physical properties and IR spectra resemble those of
polystyrene.[296] The formation of amorphous carbon occurred over
an energy range of several hundred electronvolts. The features of
carbon coatings deposited at different kinetic energies by ion plating
using benzene are shown in Fig. 39.[296] The concentration of hydrogen was found to be decreasing exponentially up to an ion energy of
1 keV. The films deposited at 150 eV energy did not exhibit any C-O
bonds, even when the film was exposed to air. Furthermore, some
graphite traces were found in the films deposited at energies over 1
keV. Weissmantel et al.[296] believe that secondary radiation damage
due to impinging energetic ions is responsible for the formation of
graphitic carbon. The films were nonuniform in thickness when
kinetic energy exceeded 3 keV. This study suggests that the properties of amorphous carbon coatings are very sensitive to the kinetic
energy of the carbon species which can be manipulated through
substrate biasing, ion beam energy, and laser irradiation.

308 Handbook of Hard Coatings

Figure 39. General properties of amorphous carbon coatings deposited at


various acceleration voltages (or kinetic energies) by ion plating from benzene.[296]

The sp3-bonded carbon fraction in a-C:H films deposited by


RF capacitively coupled plasma-assisted CVD, using methane-based
precursor gas as a function of substrate biasing is shown in Fig. 40.[277]
The sp2-bonded carbon fraction predominates at the highest biasing
where the structure is highly ordered graphitic.

Figure 40. Effect of substrate biasing on the sp3-bonded carbon fraction of


diamond films deposited by RF capacitively coupled plasma-assisted CVD using
methane based precursor gas.[277]

Macro- and Micromechanical and Tribological Properties 309

The influence of composition of the precursor gas (Ar+CH4


and H2+CH4) on the film density and hydrogen content of a-C:H
films deposited by MW/RF PACVD is shown in Fig. 41. We note
that the density profile as a function of Ar concentration passes
through a peak (1.57 g cm-3 ) at about 2040 vol.% Ar. The
hydrogen content increases with H2 concentration in MW/RF
PACVD a-C:H films, whereas it decreases with H2 concentration in
RF PCVD a-C:H films.[236]

Figure 41. Effect of composition of precursor gas on the film density of diamond
films deposited by MW/RF plasma-assisted CVD using methane mixed with
argon and hydrogen.[236]

310 Handbook of Hard Coatings

The dependence of ion energy on the film density of a-C films


deposited by ion beam deposition is illustrated in Fig. 42.[146][193][240]
McKenzie et al.[193] observed that the density profile of a-C films as
a function of ion energy passes through a peak (highest density
~3.15 g cm-3) at 30 eV. Ishikawa et al.[146] also found a peak in
density profile and other properties with ion energy but at much
higher ion energy of 150300 eV, Fig. 42.

Figure 42. Effect of ion energy on the film density of amorphous carbon
films deposited by ion beam deposition with e/m selection.[146][193][240]

The dependence of substrate biasing on the film density and


hydrogen content of a-C:H films deposited by RF capacitively coupled
plasma-assisted CVD using benzene-based precursor gas,[163] methane-based precursor gas,[277] and hydrocarbon vapor, [300] and RF
inductively coupled plasma-assisted CVD using methane-based precursor gas[222] is shown in Fig. 43. We note that in most of the cases,
film density increases and hydrogen content decreases linearly with
the substrate biasing. However, the density profile as a function of
substrate biasing of a-C:H films deposited by methane-based precursor gas[277] passes through a peak at 200400 V, similar to the
variation observed for ion beam deposited a-C coatings, Fig. 43.

Macro- and Micromechanical and Tribological Properties 311

Figure 43. (a) Effect of substrate biasing on the film density of amorphous
carbon films deposited by rf capacitively coupled plasma-assisted CVD using
benzene-based precursor gas,[163] methane based precursor gas,[207] and hydrocarbon vapor,[300] and (b) effect of substrate biasing on hydrogen content of
amorphous carbon films deposited by rf capacitively coupled plasma-assisted
CVD using benzene-based precursor gas,[163] methane based precursor gas,[207]
and hydrocarbon vapor;[300] and by rf inductively coupled plasma-assisted CVD
using methane based precursor gas.[222]

4.2

Mechanical Properties

The mechanical properties of carbon coatings depend on their


physical properties, growth conditions, sp3- sp2-bonded carbon ratio,
the amount of hydrogen in the coating, and adhesion of the coating
to the substrate, which are influenced by the precursor material,
kinetic energy of the carbon species prior to deposition, deposition
rate, substrate temperature, and substrate biasing.[11][47][83][89]
[91][221][236][238][243][245][246][248][265][283][284][289] In this section, we

312 Handbook of Hard Coatings

present some representative results on the mechanical and tribological properties on the micro- and nanoscale, of diamond, a-C, and aC:H coatings reported by various authors.
Hardness And Youngs Modulus of Elasticity. High cohesive energy, short bond length and covalent bonding in diamond
result in the highest hardness and Youngs modulus of elasticity. In
polycrystalline diamond films, even in the presence of high strength
of interatomic forces, the mechanical properties significantly vary
with the operative deformation mechanisms, which vary with the
microstructure and other structural features like voids, impurities,
defects, and texture.[265] The micromechanical characterization studies of a-C and a-C:H films deposited by various PVD and CVD
techniques have been done by various research groups. [56][83]
[116][132][133][246] Since these coatings are much smoother than polycrystalline diamond films, it is comparatively easy to obtain reproducible hardness, and elastic modulus values by nanoindentation. In
this section, we focus on the micromechanical properties of diamond
and a-C and a-C:H coatings. The information on the hardness and
elastic modulus measured by conventional methods can be found in
various reviews.[236][283][284]
Very few studies have been conducted on the micromechanical
properties measurements of diamond films. Beetz et al.[29] and OHern
and McHargue[213] have made nanoindentation measurements on
polycrystalline diamond films deposited on silicon by HFCVD using
methane and hydrogen mixture precursor gas. Beetz et al.[29] deposited films at two methane concentrations (0.11 and 0.99% CH4 in
H2), 1 kPa partial pressure, and at about 800900C substrate temperature. The films deposited at lower methane concentration exhibit
larger size grain size of about 58 m; and the films deposited at
higher methane concentration exhibit smaller size grain size of
about 0.5 m. The load-displacement plots of nanoindentation, and
SEM images of two polycrystalline diamond films deposited by
Beetz et al. [29] are shown in Fig. 44.[29] We note that the coarsegrained films exhibit almost an elastic deformation, whereas
the fine-grained diamond films exhibit a little hysteresis in the

Macro- and Micromechanical and Tribological Properties 313

unloading curve, indicating some amount of plastic deformation, the


consequence of which is a lower hardness. Beetz et al.[29] reported
that the higher methane concentration in the precursor gas led to a
decrease in grain size, and incorporation of sp2-bonded carbon,
indicated by Raman Studies, which results into a decrease in hardness of the diamond films. OHern and McHargue[213] deposited
films at a methane concentration of 1% CH4 in H2 and a substrate
temperature of about 1100C. The load-displacement plots of a
HFCVD polycrystalline diamond film and type II a natural diamond
are compared in Fig. 45.[213] We note that loading and unloading
curves of natural diamond overlap each other indicating a perfect
elastic deformation during nanoindentation. However, the unloading
curve of the diamond film shows a little plastic deformation.

Figure 44. Load-displacement plots of nanoindentation and SEM images of


HFCVD diamond films deposited by HFCVD using (a) 0.11% CH4 in H 2 gas and
(b) 0.99% CH4 in H 2 gas.[29]

314 Handbook of Hard Coatings

Figure 45. Comparison of load-displacement plots of nanoindentations made on


a type II a natural diamond and a HFCVD diamond film.[213]

The load-displacement plots of three nanoindentations and the


SEM image of a 3 m thick polycrystalline diamond film having
average grain size of about 23 m, deposited on (100) silicon by
MPCVD using 1% methane in a H2 mixture, are shown in Fig.
46.[246] The loading and unloading curves of these indentations show
a very little hystereses during unloading consistent with the observations of Beetz et al.[29] and OHern and McHargue.[213] Savvides and
Bell[246] made indentations at two different indentation loads of 67
and 100 mN (corresponding to 0.06 and 1.07 mN contact force,
respectively), and obtained hardnesses and elastic moduli of 80 and
533 GPa and 56102 and 1050 GPa, respectively. The hardness and
elastic modulus values of polycrystalline diamond films are quite
close to those of natural diamond which varies from 80 GPa for

Macro- and Micromechanical and Tribological Properties 315


polycrystalline diamond to 104 GPa for single-crystal diamond.[66]
It is important to mention that since in nanoindentation the depths of
indentation are very shallow, rough surface morphology with large
grain size result in a wide scatter in values; thus for accurate and
reproducible measurements, rough polycrystalline films need to be
polished.

Figure 46. Load-displacement plots of nanoindentations and SEM images of


HFCVD diamond films deposited by MPCVD using 1% CH4 in H2 gas.[246]

316 Handbook of Hard Coatings

Seino et al.[251] and Seino and Nagai[252] deposited polycrystalline diamond films by MPCVD with 0.51 micron grain size at 4 kPa
partial pressure, 0.1 to 5% methane in a H2 mixture, and 500 W
microwave power, and measured the hardness by Vickers indentation at 210 mN, and elastic modulus of a free standing film by a
vibrating reed apparatus. The effect of methane concentration in
precursor gas on the hardness and elastic modulus is shown in Fig.
47.[251] We note that films deposited at lower methane concentration
of 0.1% exhibit the highest hardness of about 94 GPa and elastic
modulus of about 960 GPa and hardness and elastic modulus decrease with increasing methane concentration resulting in an increase in sp2-bonded carbon fraction in diamond films. The high
hardness, and elastic modulus values of MPCVD diamond films
deposited at lower methane concentration are very close to those of
natural diamond ranging from 80 to 100 GPa.[66]

Figure 47. Effect of methane concentration on the hardness and elastic modulus
of diamond films deposited by MPCVD using 0.1 to 5% CH4 in H2 gas. [251]

The data reported by various authors indicate that the polycrystalline diamond films exhibit hardness and elastic modulus values
very close to those of natural diamond. More systematic study on
micromechanical characterization is needed to better understand the
dependence of deposition parameters, effects of diamond crystallite orientation, surface morphology, grain size, and role of atomic
hydrogen in the growth process on the mechanical behavior of
diamond films.

Macro- and Micromechanical and Tribological Properties 317

The load-displacement plots of three indentations made on a 5-m


thick a-C film deposited on silicon by ion-assisted magnetron sputtering are shown in Fig. 48.[246] Compared to polycrystalline diamond films (Fig. 46), these films exhibit much higher indentation
depths because of their lower hardness. The dependence of ion
energy of carbon species in magnetron sputtering, which consequently affect the sp3-to-sp2-bonded carbon ratio on the hardness is
shown in Fig. 49. We note that the hardness decreases with increasing ion energy and sp3-to-sp 2-bonded carbon ratio.

Figure 48. Load-displacement plots of three nanoindentations made on 5-m


thick amorphous carbon films deposited by low-energy ion-beam assisted
magnetron sputtering.[246]

Anders et al.[9] deposited hydrogen free amorphous carbon


films by cathodic arc deposition technique by changing substrate
pulsed bias voltage, and duty bias cycle %. The variation in hardness
with pulsed bias voltage, and pulsed bias duty cycle is shown in Fig.
50. We note that the hardness profile as a function of pulsed bias
voltage attains a maxima of 35 GPa at about -100 V and as a function
bias duty cycle % attains a maxima of 45 GPa at 55%. The authors
believe that the deposition conditions resulting in higher hardness
probably formed increased fraction of sp3-bonded carbon.

318 Handbook of Hard Coatings

Figure 49. The dependence of hardness of amorphous carbon films on ion


energy per depositing carbon atom and the sp3/sp2 bonding ratio. The films were
deposited by low-energy ion-beam assisted magnetron sputtering.[246]

Cho et al.[83] deposited a-C:H films with varying hydrogen


content by magnetron sputtering of a graphite target at a pressure of
10 mtorr by changing the power density of hydrogen-argon plasma
from 0.1 to 10 watt cm-2. The variation in the hardness, measured by
Nanoindenter, with hydrogen content in the film is shown in Fig. 51.
We note that the hardness decreases from 14 to 3 GPa as hydrogen
content increases from about 0.4 to 6%. Cho et al.[83] found from
Raman and electron energy loss spectrometry (EELS) measurements
that hydrogen incorporation in the a-C:H films enhances the crystallinity of the sp2-bonded carbon clusters, as well as increases the sp2bonded carbon fraction. The lower hardness at higher hydrogen
content is attributed to the higher sp2-bonded carbon fraction in the
a-C:H films.

Macro- and Micromechanical and Tribological Properties 319

Figure 50. The dependence of hardness of amorphous carbon films deposited by


cathodic arc deposition (a) on substrate pulsed bias voltage for 10 and 33% and
pulsed duty cycles, and (b) on pulsed bias duty cycle for pulsed and dc bias
voltages of -100.[9]

Figure 51. Effect of hydrogen content of the hardness of a-C:H films deposited
by magnetron sputtering using a graphite target and by varying hydrogen concentration from 0.5 to 15 % in Ar + H2 gas mixture.[83]

320 Handbook of Hard Coatings

The variations in hardnesses with substrate biasing of a-C:H


films deposited by RF capacitively coupled plasma-assisted CVD
under different deposition conditions: (i) methane-based precursor
gas biasing from -100 to -1250 V,[116] (ii) 75% CH4 and 25% Ar,
biasing from -50 to -600 V, substrate temperature 200C,[92] and (iii)
benzene-based precursor gas at 3 Pa partial pressure[163] are compared in Fig. 52. We note that films deposited from methane-based
precursor gas exhibit a peak maxima in the hardness profile at about
-200 V, whereas the films deposited from benzene-based precursor
exhibit a continuing increase in hardness with substrate biasing
voltage. Furthermore, low biasing voltage of about -100 V resulted
in soft polymeric films and high biasing voltage of about -600 V
resulted in the graphitic films.[116]

Figure 52. The effect of substrate bias voltage on the hardness of amorphous
carbon films deposited by rf capacitively coupled plasma-assisted CVD using
methane based precursor gas and biasing from -100 to -1250 V,[116] and 75% CH4
and 25% Ar precursor gas and biasing from -50 to -600 V,[92] and benzene-based
precursor gas at 3 kPa partial pressure.[163]

Pappas and Hopwood[222] studied the effect of various deposition parameters on the hardness, measured by Knoop indentation, of
a-C:H films deposited on silicon by RF induction plasma-assisted
CVD. Shown in Fig. 53 are the effects of substrate biasing, RF
induction power, Ar concentration in methane-based precursor gas,
and precursor gas flow on the hardness of a-C:H films. Like RF

Macro- and Micromechanical and Tribological Properties 321

capacitively coupled plasma-assisted CVD deposited a-C:H films,


the hardness profile of these films as a function of substrate biasing
also passes through a maxima of about 30 GPa peak at -100 V. The
hardness decreases with RF induction power and methane gas flow
and increases with argon concentration in methane-based precursor.
The variation in hardness with deposition conditions can be related
to the variations in sp3-to-sp2-bonded carbon ratio resulting from
changes in the deposition conditions.

(a)

(b)

(c)

(d)

Figure 53. The dependence of hardness of amorphous carbon films on various


deposition parameters (a) substrate biasing, (b) rf induction power, (c) argon
percentage in methane, and (d) methane gas flow rate. These films were deposited by rf inductively coupled plasma-assisted CVD technique.[222]

The hardness and elastic modulus of metal-containing amorphous carbon films (Me-C:H) were measured using Nanoindenter by

322 Handbook of Hard Coatings

Wang et al.[291] Figure 54 shows the dependence of metal concentrations on the hardness, and elastic modulus of Ti-C:H films deposited
on silicon by DC magnetron sputtering and Ta-C:H films deposited
by combining RF capacitively coupled plasma-assisted CVD at three
substrate temperatures, i.e., 20, 270, and 20270C in the presence
of methane-based plasma. Both the hardness and elastic modulus
profiles as a function of metal concentration in films passes through
a peak at about 50 to 60% metal. A model for explaining the increase
in hardness and elastic modulus of Me-C:H with metal concentration
need to be developed.

Figure 54. The dependence of hardness and elastic modulus Me-C:H films
deposited by rf plasma CVD combined with reactive rf glow discharge sputtering
using argon+acetylene precursor gas at different percentages of acetylene and
substrate temperatures. Films shown by S1 were deposited at 270C, S2 at 20C,
and S3 at temperatures between 20 to 270C.[291]

Macro- and Micromechanical and Tribological Properties 323

To study the effect of deposition technique on the mechanical


properties of amorphous carbon coating, Gupta and Bhushan[132][133]
deposited 20- to 400-nm thick amorphous carbon coatings on silicon
by four deposition processes: (i) cathodic arc, (ii) ion beam deposition, (iii) RF-plasma enhanced chemical vapor deposition (RFPECVD), and (iv) RF sputtering. Hardness and elastic modulus
profiles, measured by Nanoindenter, as a function of depth of carbon
coatings deposited by various techniques are compared in Fig. 55.
Cathodic arc carbon coatings exhibit the highest hardness of 38 GPa,
and elastic modulus of 350 GPa of all coatings examined in this
study. The high hardness and elastic modulus of cathodic arc carbon
coatings are followed by ion beam carbon and PECVD/ sputtered
carbon coatings. The high hardness and elastic modulus of cathodic
arc carbon are attributed to the high kinetic energy of carbon species
involved in the cathodic arc deposition. The difference in hardness
and elastic modulus of carbon coatings deposited by various deposition techniques is attributed to their varying sp3- to sp2-bonding
ratio, and the amount of hydrogen. The reduction in hardness, and
elastic modulus of cathodic arc carbon coatings with the indentation
depth is attributed to the increased contributions of the silicon
substrate (having a lower hardness of 11 GPa) at larger depths.
The hardness and elastic modulus data on a-C and a-C:H films
reported by several research groups suggest that mechanical behavior of these films is very sensitive to sp3- to sp2-bonding ratio and
hydrogen content which can be optimized by changing the deposition.
Adhesion. Usage of the diamond, and amorphous carbon
coatings to protect against wear and corrosion, and to serve as a solid
lubricant coating for reduced friction, requires high adhesion of these
coatings. The adhesion forces between the coating and the substrate
must be sufficiently high so that the forces due to residual stresses
can be neutralized. Diamond coatings are found to adhere well with
materials that have a capability of carbide and silicide formation,
such as, Si, SiC, Si3N4, quartz, and iron alloys.[11][105][296] On the
contrary, almost no adhesion of PECVD deposited amorphous carbon coatings was observed on some materials like Co, Cr, Ni, Al2O3,

324 Handbook of Hard Coatings

and ZrO2 and their alloys.[105][121][143] In general, high adhesion of a


coating can be achieved by increasing the nucleation density, forming a compositionally graded junction between the coating and
substrate, and applying a thin interlayer.

Figure 55. The hardness and elastic modulus, measured by nanoindentation, of


400-nm thick amorphous carbon films deposited on silicon by cathodic arc,
direct ion beam, RF plasma-enhanced CVD, and dc magnetron sputtering.[132]

A variety of surface pretreatments to achieve a high nucleation


density which in turn enhances adhesion of diamond coatings for
various substrates have been discussed in an earlier section. In the
case of thin amorphous carbon coatings (~ 20 nm thick), a post
deposition treatment by ion bombardment of inert gas species resulting into ion beam mixing at the interface can also be used to enhance
the adhesion. Grill et al.[121][122][125] studied the effect of silicon and
metallic interlayers on the adhesion of a-C:H coatings deposited by
RF plasma CVD, and concluded that a 2 to 4 nm thick film of silicide
forming metal like Si is very effective in improving the adhesion of a
carbon coating on nondiamond substrate.

Macro- and Micromechanical and Tribological Properties 325

To study the effect of deposition techniques on the adhesion of


amorphous carbon coating, Gupta and Bhushan[132] deposited 20-nm
thick amorphous carbon coatings on silicon by four deposition processes: (i) cathodic arc, (ii) ion beam deposition, (iii) RF-plasma
enhanced chemical vapor deposition (RF-PECVD), and (iv) RF
sputtering. The adhesion of these coatings was compared by scratching at ramping loads ranging from 1 to 15 mN using Nanoindenter,
and monitoring the friction and scratch depth while scratching.[132]
The coefficient of friction and scratch depth profiles as a function of
increasing normal load and SEM images of two regions over scratches:
at the beginning of the scratch (indicated by A on friction profile),
and at the point of initiation of damage at which the coefficient of
friction increases abruptly to a very high value (indicated by B on
friction profile), made on various 20-nm thick carbon and SiC
coatings and uncoated substrate are shown in Fig. 56. The abrupt
increase in the coefficient of friction and scratch depth is associated
with damage to the coating. Most of the coatings exhibit a steady low
coefficient of friction of about 0.10.15 during the initial stages and
an abrupt increase in friction when normal load exceeds the critical
load, the load sufficient to damage the coating. Only sputtered
carbon coating exhibits a gradual increase in the coefficient of
friction with increasing normal load from the beginning of the
scratch. In a separate study, Bhushan and Gupta[55] have shown that a
steady low coefficient of friction during the initial stages of a scratch
is associated with the sliding of the tip without significantly ploughing into the sample, and abrupt increase in friction is associated with
catastrophic failure, as well as significant ploughing of the tip into
the sample. The gradual increase in friction with normal load during
scratching is associated with the ploughing of the tip into the sample;
and the depth of ploughing increases with increasing normal load.[55]
The friction profiles shown in Fig. 56 suggests that most of the
coatings except sputtered carbon coating did not get damaged before
the critical load, and got damaged catastrophically as the normal load
exceeded the critical load limit. In contrast, the sputtered carbon
coatings were damaged right from the beginning of the scratch.

326 Handbook of Hard Coatings

Figure 56. The adhesion, measured by microscratching, of 20-nm thick amorphous carbon films deposited on silicon by cathodic arc, direct ion beam, rf
plasma-enhanced CVD, and DC magnetron sputtering.[132]

The scratch depth profiles obtained during and after the scratches
on all samples are plotted with respect to initial profile after cylindrical curvature removed, Fig. 56. Reduction in scratch depth after
scratching is observed. The reduction in scratch depth after scratching is attributed to an elastic recovery after the removal of the normal
load. It appears that the scratch depth after scratching indicates the
final depths which reflect the extent of permanent damage and

Macro- and Micromechanical and Tribological Properties 327

ploughing of the tip into the sample surface. We believe that the
scratch depths after scratching are probably more relevant for visualizing the damage that can occur in real applications.
The higher scratch resistance/adhesion of cathodic arc carbon
coatings is attributed to an atomic intermixing at the coating-substrate interface because of high kinetic energy (2 keV) plasma
formed during the cathodic arc deposition process.[9] The atomic
intermixing at the interface provides a graded compositional transition between the coating and the substrate materials. In all other
coatings used in this study, the kinetic energy of the plasma was
insufficient for atomic intermixing.
4.3

Tribological Properties

In this section, we present the friction and wear characteristics


of a few selected diamond and amorphous carbon coatings. The
friction and wear properties of coating are strongly influenced by
their physical properties and the chemistry in the interface of the
mating surfaces. In many cases environment plays a crucial role in
affecting the friction and wear characteristics of a material. Friction
and wear properties of polycrystalline diamond films are found to be
strongly influenced by their surface morphology, grain size of crystallite size, and amount of nondiamond carbon phase in the diamond
matrix.[21][115][117][118][129][161][164][226] Surface roughness plays an
important role in friction and wear properties of CVD diamond
films.[49][54][71][129][137][138][202][203] In this section, we will show how
surface morphology and surface roughness affect the friction of
diamond films, and the wear of the counterface materials.
Miyoshi et al.[203] studied friction and wear characteristics of
diamond films of different surface morphologies deposited on silicon, and a-SiC by high-pressure MPCVD at varying partial pressures ranging from 5 to 40 torr, substrate temperatures ranging from
8501000C, microwave power ranging from 500 to 1000 watts, and
deposition times ranging from 10 to 21 h. Shown in Fig. 57 are the

328 Handbook of Hard Coatings

SEM images of these films indicating that the grain size of diamond
crystallites varies from 20100 nm to 3.3 mm. The fine-grained films
(grain size ~ 20100 nm) are smooth, and contain a significant
amount of nondiamond carbon, whereas the coarse-grained films
(grain size ~ 1.13.3 mm) are rough and contain a little amount of
nondiamond carbon. The coefficients of friction of these films against
bulk diamond pins at 1N in humid air and dry nitrogen are shown in
Fig. 58. As expected, smooth films exhibit low coefficients of
friction of 0.030.04 in humid air, as well as in dry nitrogen. Rough
films exhibit slightly higher coefficients of friction of 0.050.07.
Measurements in high vacuum of 10-7 torr show very high coefficient of friction approaching 1.51.8.[203] The authors believe that in
humid air and nitrogen, friction arises from abrasion from a diamond
pin sliding on the diamond film, and high vacuum adhesion between
the sliding surfaces. They also observed that the friction in air and
nitrogen increases with the surface roughness, whereas in vacuum it
is independent of surface roughness.

(a)

(b)

(c)

Figure 57. SEM images of diamond films deposited by MPCVD. (a) Fine-grain
(20-100 nm) diamond film on (100) silicon; surface roughness rms = 15 nm. (b)
Medium-grain (1100 nm) diamond film on (100) silicon; surface roughness rms
= 63 nm. (c) Coarse-grain (3300 nm) diamond film on -SiC silicon; surface
roughness rms = 160 nm.[203]

Macro- and Micromechanical and Tribological Properties 329

Figure 58. Coefficient of friction as a function of sliding distance of diamond


films whose surface morphologies are shown in Fig. 57, slid against a bulk
diamond pin (a) in humid air with 40% RH and (b) in dry nitrogen at 1 N normal
load, 86 mm min-1 sliding speed, and 25C ambient temperature.[203]

Gangopadhyay and Tamor[115] studied the friction and wear


properties of diamond films deposited on silicon by MPCVD with
three different surface morphologies, Fig. 59 (i.e., faceted,
cauliflowered, and smooth) by changing the composition of CO+H2
precursor gas. The grain size and surface roughness (cla value) of
these films were measured as 0.21.5 mm and 29 nm; 13 mm and
21 nm; 0.10.3 mm and 11 nm, respectively. Shown in Fig. 60a are
the coefficients of friction of these diamond films sliding in reciprocating mode against themselves, silicon nitride, zirconia, and steel at
4.2 N at ambient temperature and relative humidity of 55% to 85%.
The average wear rates of counterface materials are compared in Fig.
60b. All material combinations except diamond film sliding against
another diamond film, exhibit high coefficients of friction ranging
from 0.38 to 0.90. Furthermore, in all cases faceted films exhibit
higher friction and wear than those of smooth films. The smooth
films also exhibit formation of a wear track during sliding tests. The
wear of the diamond film sliding against steel occurred by the
fracture of the tips of the grains, although evidence of delamination
was also observed. Similar wear mechanisms were observed for a
diamond film sliding against another diamond film.

330 Handbook of Hard Coatings

(a)

(b)

(c)

Figure 59. SEM images of diamond films deposited by MPCVD at 900oC on


silicon using different compositions of precursor gas; (a) rough and faceted
films (grain size ~ 0.21.5 m, rms roughness ~ 29 nm) using 20% CO+80% H2,
(b) cauliflower-type film (grain size ~ 1-3 m, rms roughness ~ 21 nm) using
0.5% CH4+N2+97% H2, (c) smooth film (grain size ~ 0.1-0.3 m, rms roughness
~ 11 nm) using 2% CH4+ 98% H2.[115]

Figure 60. (a) Coefficient of friction of diamond films whose surface morphologies
are shown in Fig. 59, slid against a diamond coated silicon nitride, silicon nitride,
alumina, zirconia, and steel at 4.2N load, 35 mm s-1 sliding speed, and 5585%
relative humidity and (b) average wear rate of slider materials slid for 1 h against
the three diamond films.[115]

Macro- and Micromechanical and Tribological Properties 331


Gupta et al.[129] deposited 20 m thick diamond films on
silicon using HFCVD technique from a mixture of methane (typically 1%) diluted in hydrogen at 800C substrate temperature. Prior
to deposition, the silicon substrates were abraded with 4 to 8 m and
submicron (0.5 m and less) diamond powder in order to obtain
coarse-grain (grain size ~ 10 to 15 m), and fine-grain (grain size
~ 3 to 5 m) films, respectively. SEM and AFM images of these asdeposited coarse-grain film, chemomechanically- and laser-polished
coarse-grain film, and as deposited fine grain film are shown in Fig.
61. The friction profiles of as deposited coarse-grain (rms ~650 nm)
film, as-deposited fine-grain (rms ~180 nm) film, laser polished (rms
~ 97 nm) coarse-grain film, and natural diamond (rms ~ 1.72 nm),
slid against an alumina ball are compared in Fig. 62. The polishing of
diamond films resulted in a significant reduction of the coefficient of friction from about 0.4 to 0.09. Diamond films with finegrains (~ 35 m) exhibit considerably lower coefficients of friction
(0.20) as compared to the films having larger grains (~ 1015 m).
Wear of the mating alumina ball sliding against diamond surfaces
was considerably reduced after polishing. The wear tracks on the
mating alumina ball surface, and diamond films were examined by
optical microscopy, and the degree of wear was estimated in terms of
the diameter of the round wear scar on the ball, and the amount of the
wear debris collected on the film. Figure 63 shows the optical
micrographs of as-deposited, and chemomechanically-polished
coarse-grain diamond films, as-deposited fine-grain diamond films
and of natural diamond after a wear test. We note that the wear scar
on the alumina ball slid against the polished diamond film is comparable to that slid against natural diamond after a wear test. Further,
the amount of alumina transferred on to the polished diamond film
during sliding is considerably lower than that for the as-deposited
coarse-grain diamond film (Fig. 63). The debris seen on the natural
diamond surface comes from the fracture of the asperities on the
alumina ball surface. It was observed that most of the debris in
natural diamond is generated during the early stage of sliding. The
role of surface roughness on the friction of diamond coatings is

332 Handbook of Hard Coatings

shown in Fig. 64. The coefficient of friction decreased from 0.4 to


0.09 as the rms roughnesses of polished films were reduced from 650
to 170 nm, suggesting the important role of surface roughness. This
study demonstrated that the friction and wear characteristics of CVD
diamond films can be controlled by their grain size and surface
roughness; and the friction and wear characteristics of polished CVD
diamond films are comparable to natural diamond.

Figure 61. SEM and AFM images of (a) as-deposited coarse-grained films,
(b) as-deposited fine-grained films polished diamond films, (c) chemomechanically-polished coarse-grained films, and (d) laser-polished coarse-grained
films, deposited by HFCVD.[129]

Macro- and Micromechanical and Tribological Properties 333

Figure 62. The coefficient of friction profile as a function of sliding distance of


as-deposited and polished HFCVD diamond films slid against an alumina ball at
1N in ambient conditions in reciprocating mode.[129]

Since a-C and a-C:H coatings can be deposited at lower temperatures (< 200300C) on substrates ranging from soft polymers to
hard ceramic substrates without significantly affecting their surface
topography, numerous studies have been performed to correlate the
friction and wear properties of carbon films to their properties, and
subsequently to their deposition conditions.[9][106][107][116][120][122]
[125][131][133][143][293][294] Like hardness and elastic modulus, friction
and wear properties of a-C and a-C:H coatings vary dramatically
with the deposition conditions. The environment plays a crucial role
in the friction and wear properties of a-C:H coatings because of their
lower thermal stabilities, the a-C:H coatings start to graphitize by
hydrogen loss above 400C.[102][197][202] In this section, we present
some representative data to illustrate the effect of deposition conditions, and the environment on the friction and wear properties of a-C
and a-C:H coatings. For more detailed information reader can refer
to a review by Grill.[120]

334 Handbook of Hard Coatings


Diamond surface

Ball surface

Figure 63. The optical images of the diamond surface and the alumina ball after
the sliding tests on (a) as-deposited coarse-grained films, (b) chemomechanicallypolished coarse-grained films, (c) as-deposited fine-grained films polished films,
and (d) natural diamond.[129]

Macro- and Micromechanical and Tribological Properties 335

Figure 64. Effect of surface roughness on the coefficient of friction of HFCVD


diamond films slid against an alumina ball at 1N normal load in ambient
conditions in reciprocating mode.[129]

The effect of ion beam energy and substrate temperature on the


friction and wear properties of a-C coatings deposited by direct ion
beam on aluminum substrate using a ring-cusp ion source operating
on methane gas was studied by Wei et al.[293] The variation in the
coefficients of friction as a function of number of reciprocating
cycles, of a-C coatings deposited at different ion beam energies, and
substrate temperatures sliding against an alumina ball at 0.2 N
normal load under 50 10% relative humidity and 21 2 oC temperature are shown in Fig. 65. The coatings deposited at 450 to 550 eV
ion beam energy and 50100oC substrate temperatures exhibit the
lowest coefficient of friction of about 0.080.12 at 0.2 N normal load
and the highest number of cycles required to initiate damage at 5N
normal load.
The effect of pulsed bias voltage on the coefficient of friction
of hydrogen free amorphous carbon films deposited by cathodic arc
deposition technique, sliding against a diamond sphere of 2 mm
radius at 0.3 N normal load in ambient environment is shown in Fig.
66.[9] We note that the friction profile as a function of pulsed bias
voltage attains a minima of 0.09 at about -100 V. The authors believe
that the deposition conditions at -100 V biasing resulting in higher
hardness and mass density probably formed increased fraction of

336 Handbook of Hard Coatings

sp3-bonded carbon which is responsible for lower friction. Anders et


al.[9] also observed that for a given bias voltage, increasing the
pulsed bias duty cycle results in lower coefficient of friction.

Figure 65. Coefficient of friction of a-C coatings deposited at different ion beam
energies and substrate temperatures sliding against an alumina ball at 0.2 N
normal load under 50 10% relative humidity and 21 2C temperature.[293]

Figure 66. The coefficient of friction of a-C coatings deposited by cathodic arc
deposition on silicon slid against diamond at 0.3 N as a function of pulsed bias
voltage for pulse duty cycles of 10 and 33%.[9]

Macro- and Micromechanical and Tribological Properties 337

Friction Coefficient

Gangopadhyay et al.[116] deposited a-C:H films by RF capacitively coupled plasma-assisted CVD using methane-based precursor
gas at substrate biasing ranging from -100 to -1250 V corresponding
to ion energies from 30 to 400 eV. The effect of substrate biasing on
the coefficient of friction of a-C:H films slid against steel ball at 4.2
N in ambient humidity of 70% RH are compared in Fig. 67. We note
that films deposited at the lowest (-100 V) and the highest (-1250 V)
biasing exhibit high coefficient of friction in the range of 0.2 to 0.3,
while films deposited at the biasing from -200 to -300 V, exhibit the
lowest coefficient of friction of about 0.08. The lowest coefficient of
friction in these films can related to their high hardness of about 16
GPa, Fig. 52. Gangopadhyay et al. [116] have also studied the effect of
humidity on the friction and wear properties of the films deposited at
a substrate biasing of -350 V, and found that both friction and wear
increases with the relative humidity, Fig. 68. They observed a more
uniform transfer film on a steel ball at lower humidity (4%) than the
film at higher humidity (67%). Enke et al.[104] reported a low coefficient of friction of about 0.010.02 for RF PECVD deposited a-C:H
films using acetylene and argon mixture precursor gas at a relative
humidity of 1% and an increase in the coefficient of friction to 0.19
as relative humidity was increased to 100%.

The film was removed in about a minute. The


values are at the end of tests.

Bias Voltage (-V)


Figure 67. The coefficient of friction of a-C:H films slid against steel at 4.2 N in
ambient conditions. These films were deposited voltage by rf capacitively coupled
plasma-assisted CVD at varying substrate biasing from -100 to -1250 V.[116]

338 Handbook of Hard Coatings

Figure 68. Effect of relative humidity on (a) the coefficient of friction and (b)
wear volume of a-C:H films deposited at -350 V bias voltage by rf capacitively
coupled plasma-assisted CVD properties of a-C:H films.[116]

Memming et al.[197] reported an increase in the coefficient of


friction from about 0.04 to 0.25 for RF-PECVD a-C:H films, deposited from acetylene based precursor gas, with an increase in relative
humidity from 1 to 100%. They measured an extremely low coefficient of friction of 0.02 in 10-8 torr vacuum environment; this is
contrary to the high coefficient of friction of about 1.6 of polycrystalline diamond films. Furthermore, Memming et al.[197] observed an
increase in the coefficient of friction from about 0.02 to 0.60 when
dry nitrogen was replaced with dry oxygen. Based on AES analyses
of worn coatings, the authors concluded that in case of films that
exhibit a low coefficient of friction, a carbonaceous material was
transferred from the carbon coating to the steel ball, while in case of

Macro- and Micromechanical and Tribological Properties 339

films that exhibit a high coefficient of friction (in dry oxygen) iron
was transferred to the coating. Miyoshi et al.[202] deposited 0.25-m
thick for 30 kHz AC-PECVD a-C:H films, using methane and butane
precursor gases. They found low coefficients of friction of 0.02 and
0.06 between these films and diamond in dry nitrogen and ambient
air, respectively. The a-C:H coatings exhibit coefficients of friction
of 0.15 and 0.25 when slid against silicon nitride in dry nitrogen and
humid air, respectively. This increase in friction is due to the silicon
oxide films produced on the surface of silicon nitride pins in humid air.
Erdemir et al.[108] studied the nature of transfer layers formed
on M50 steel balls during sliding against 1.5 m thick ion-beam
deposited carbon coatings on M50 steel disks. These coatings exhibited a low coefficient of friction of about 0.12 at initial stage of
sliding against M50 steel ball in nitrogen at 5N normal load under
ambient temperature. The coefficient of friction decreased with
continuing sliding, and reached to a value of about 0.07 toward the
end of 1 km sliding test, and to about 0.03 toward the end of a long
test. Ultra low friction of these carbon films was attributed to the
formation of transfer films that were rich in carbon and had a
disordered graphitic structure.
Memming et al.[197] annealed a-C:H films at 550C to study the
effect of hydrogen depletion during annealing on the friction and
wear properties. They found the annealing resulted in a high coefficient of friction of 0.68 in vacuum, as well as in dry nitrogen.
Miyoshi et al.[202] also observed a similar increase in the coefficient
of friction on annealing; Miyoshi and co-workers suggested the
formation of a graphite layer through a two stage process: a carburization stage resulting in hydrogen depletion and a polymerization
stage resulting in the formation of graphite crystallites.
Grill et al.[123] reported the effects of substrate temperature and
substrate biasing on the friction and wear properties of RF PECVD
a-C:H coatings deposited on silicon using acetylene. The coefficient
of friction of a-C:H films (hydrogen content-40%) sliding against
a steel ball in 4070%, relative humidity was decreased from
0.35 0.04 to 0.20 0.04 as the substrate temperature increased from

340 Handbook of Hard Coatings

100oC to 250C. The substrate biasing did not show any significant
variation in friction. The insensitivity of the coefficient of friction to
the annealing temperature was similar to what was reported by
Memming et al.[197] and Miyoshi[201] in humid air.
Kim et al.[156] reported the effects of oxygen and humidity on
friction and wear of RF PECVD a-C coatings deposited on silicon
from methane and hydrogen mixture. The coefficient of friction of
these films against silicon nitride ball varied from 0.2 in 50% humid
argon and 100% humid air, to 0.06 in dry argon, Fig. 69. The
variations in friction were attributed to a transfer layer, produced by
friction, that covered the contact surface of the ball in all cases.[156]
Low friction in dry argon was resulted from the material (unoxidized
DLC debris) transferred from the carbon coating to the silicon nitride
ball. The oxidized DLC debris (carbonyl compounds) formed on the
a-C coating, as well as on the silicon nitride ball in air, and humid
argon increases the friction.[156] The lowest wear rates were observed
in dry environments. The humidity increases the wear rates of a-C
coatings in 50 and 100% argon, and humidity saturated air by five times.
Wear of these coatings was dominated by the tribochemistry in air
and humid argon, but in dry argon it is controlled by adhesive wear.

Figure 69. Coefficient of friction and wear rate of a-C coatings sliding against
silicon nitride ball in air and argon as a function of relative humidity. The test
conditions were as follows: silicon nitride ball diameter-7.94 mm, sliding speed18.7 0.8 mm s-1, load- 9.8 N, environment-ambient temperature.[156]

Macro- and Micromechanical and Tribological Properties 341


Donnet et al.[102] studied the tribochemistry of RF PECVD
deposited carbon coatings on silicon sliding against 52,100 steel
balls in a vacuum range from 10-7 to 50 Pa, dry nitrogen, and ambient
air. In all cases, a transfer layer was observed on the steel ball during
the first 100 cycles, associated with relatively a high coefficient of
friction of about 0.20.3. Friction decreased to 0.0060.008 beyond
100 cycles in a vacuum below 10-1 Pa, and to 0.010.07 in a vacuum
of 1050 Pa and in dry nitrogen. A high vacuum was found to be
associated with the ultralow friction and low wear. On the contrary,
a poor vacuum, and inert atmosphere was found to be associated with
the low friction, and moderate wear. In vacuum and dry nitrogen
atmospheres, the wear particles were consisted of amorphous hydrogenated carbon. Ambient air was found to be associated with the
relatively high friction and severe wear, coupled with the formation
of roll-shaped debris of amorphous carbon containing iron oxide
precipitates.[102]
Based on the friction and wear data just presented, it is apparent that the friction and wear properties of amorphous carbon coatings are very sensitive to the sp3-bonded carbon fraction and the
amount of hydrogen present in the coatings as well as the operating
environment.

5.0

CLOSURE

The micromechanical and microtribological properties measurements can be made on ultrathin coatings, and on the skin of bulk
materials by simulating single asperity contact scratching and sliding situations by using a nanoindenter, an atomic force microscope,
and a friction force microscope. Depth profiles of various mechanical properties such as elastic-plastic deformation behavior, hardness,
Youngs modulus of elasticity, fracture toughness, fatigue, and coating-substrate adhesion can be obtained by making nanoindentations
at different loads by using a nanoindenter. The evolution of wear on

342 Handbook of Hard Coatings

a typical surface can be observed by using an atomic force microscope or a friction force microscope.
Diamond films deposited by plasma enhanced CVD exhibit
hardness, elastic modulus, friction, and wear properties comparable
to those of natural diamond. High friction (0.4) of rough CVD
diamond films having grain size of diamond crystallites in the range
of 1020 m and rms roughness of about 700 nm can be reduced to
0.08 by polishing by using chemomechanical, and laser polishing
techniques. Properties of amorphous and hydrogenated amorphous
carbon coatings can be tailored by optimizing the deposition parameters. Amorphous carbon coatings with a hardness of more than 40
GPa can be deposited by cathodic arc deposition. The friction and
wear properties and wear mechanisms of amorphous and hydrogenated amorphous carbon coatings are found be very sensitive to the
test environment.

REFERENCES
1. Ageev, V. P., Armeyev, V., Yu, Chapliev, N. I., Kuzmichov, A. V.,
Pimenov, S. M., and Ralchenko, V. G., Laser Processing of Diamond
and Diamond-Like Films, Materials and Manufacturing Processes,
8:18 (1993)
2. Ahn, J., Mittal, K. L., and Macqueen, R. H., Hardness and Adhesion
of Filmed Structures as Determined by the Scratch Technique, in:
Adhesion Measurement of Thin Films, Thick Films, and Bulk
Coatings, (K. L. Mittal, ed.), STP 640, pp. 134157, ASTM,
Philadelphia (1978)
3. Aisenberg, S., Properties and Applications of Diamond Like Carbon
Films, J. Vac. Sci. Technol. A, 2:369371 (1984)
4. Akatsuka, F., Hirose, Y., and Komaki, K., Rapid Growth of Diamond
Films by Arc Discharge Plasma CVD, Jpn. J. Appl. Phys.,
27:L1600L1602 (1988)
5. Alba, S., Loubet, J. L., and Vovelle, L., Evaluation of Mechanical
Properties and Adhesion of Polymer Coatings by Continuous
Hardness Measurements, J. Adhesion Sci. Technol., 7:131140
(1993)

Macro- and Micromechanical and Tribological Properties 343


6. Alekhin, V. P., Berlin, G. S., Isaev, A. V., Kalei, G. N., Merkulov,
V. A., Skvortsov, V. N., Ternovskii, A. P., Krushchov, M. M.,
Shnyrev, G. D., and Shorshorov, M. K., Micromechanical Testing
by Micromechanical Testing of Materials by Microcompression,
Zavod. Lab., 38:619621 (1972)
7. Anders, A., Anders, S., and Brown, I. G., Macroparticle-Free Thin
Films Produced by an Efficient Vacuum Arc Deposition Technique,
J. Appl. Phys., 74:42394241 (1993)
8. Anders, S., Anders, A., Brown, I. G., Wei, B., Komvopoulos, K.,
Ager, J. W., III, and Yu, K. M., Effect of Vacuum Arc Deposition
Parameters on the Properties of Amorphous Carbon Thin Films,
Surface Coat. Technol., 68/69:388393 (1994)
9. Anders, A., Anders, S., Brown, I. G., Dickinson, M. R., and MacGill,
R. A., Metal Plasma Immersion Ion Implantation and Deposition
Using Arc Plasma Sources, J. Vac. Sci. Technol. B, 12:815820
(1994)
10. Angus, J. C., Diamond and Diamond-Like Films, Thin Solid
Films 216:126133 (1992)
11. Angus, J. C., Koidl, P., and Domitz, S., Carbon Thin Films, in
Plasma Deposited Thin Films, (J. Mort and F. Jensen, eds.), pp. 89
127, CRC Press, Boca Raton, FL (1986)
12. Angus, J. C., and Hayman, C. C., Low Pressure Metastable Growth
of Diamond and Diamondlike Phase, Science, 241:913921 (1988)
13. Angus, J. C., Wang, Y., and Sunkara, M., Metastable Growth of
Diamond and Diamond-Like Phases, Annu. Rev. Mater. Sci., 21:221
248 (1991)
14. Angus, J. C., Argoitia, A., Gat, R., Li, Z., Sunkara, M., Wang, L.,
and Wang, Y., Chemical Vapor Deposition of Diamond, Philos.
Trans. R. Soc. London, Ser. A, 342:195208 (1993)
15. Anonymous, Standard Test Method for Microhardness of Materials,
ASME Designation: E38473:359379 (1979)
16. Anonymous, NanoIndenterTMII Operating Instructions, Nano
Instruments, Inc., 1001 Larson Drive, Oak Ridge, TN 37830 (1991)
17. Anthony, T. R., Methods of Diamond Making, in Diamond and
Diamond-Like Films, (R. E. Clausing, L. L. Horton, J. C. Angus, and
P. Koidl, eds.), pp. 555575, Plenum Press, New York, NY (1991)

344 Handbook of Hard Coatings

18. Antis, G. R., Chantikul, P., Lawn, B. R., and Marshall, D. B., A
Critical Evaluation of Indentation Techniques for Measuring Fracture
Toughness: I Direct Crack Measurements, J. Am. Ceram. Soc.,
64:533538 (1981)
19. Atkins, A. G., Silverio, A., and Tabor, D., Indentation Hardness
and the Creep of Solids, J. Inst. Metals, 94:369378 (1966)
20. Bachmann, P. K., and Lydtin, H., High Rate Versus Low Rate
Diamond CVD Methods, in: Diamond and Diamond-Like Films
and Coatings, (R. E. Clausing, L. L. Horton, J. C. Angus, and P.
Koidl eds.), pp. 829853, Plenum, New York (1991)
21. Bachmann, P. K., Lade, H., Leers, D., and Wiechert, D. U., Wear
Testing of CVD Diamond Films, Diamond and Related Materials,
3:799804 (1994)
22. Bangert, H., Wagendristel, A., and Aschinger, H.,
Ultramicrohardness Tester for Use in a Scanning Electron
Microscope, Colliod and Polymer Sci., 259:238240 (1981)
23. Bangert, H., and Wagendristel, A., Ultralow Low Hardness Testing
of Coatings in a Scanning Electron Microscope, J. Vac. Sci. Technol.
A, 4:29562958 (1986)
24. Bayer, R. G. (ed.), Selection and Use of Wear tests for Metals, STP
615, ASTM, Philadelphia (1976)
25. Bayer, R. G. (ed.), Wear tests for Plastics: Selection and Use, STP
701, ASTM, Philadelphia (1979)
26. Bayer, R. G. (ed.), Selection and Use of Wear tests for Coatings,
STP 769, ASTM, Philadelphia (1982)
27. Bayer, R. G., Mechanical Wear Prediction and Prevention, Marcel
Dekker Inc., New York, NY (1994)
28. Beerschwinger, U., Albrecht, T., Mathieson, D., Reuben, R. L.,
Yang, S. J., Taghizadeh, M., Wear at Microscopic Scales and Light
Loads for MEMS Applications, Wear, 181183:426435 (1995)
29. Beetz, C. P. Jr., Cooper, C. V., and Perry, T. A., Ultralow-Load
Indentation Hardness and Modulus of Diamond Films Deposited by
Hot-Filament-Assisted CVD, J. Mater. Res., 5:25552561 (1990)
30. Benjamin, P., and Weaver, C., Measurement of Adhesion of Thin
Films, Proc. R. Soc. London, Ser. A, 254:163176 (1960)

Macro- and Micromechanical and Tribological Properties 345


31. Benzing, R. J., Goldblatt, I., Hopkins, V., Jamison, W., Mecklenburg,
K., and Peterson, M. B., Friction and Wear Devices, 2d Ed., ASLE,
Park Ridge, IL (1976)
32. Berg, S., and Anderson, L. P., Diamond-Like Carbon Films
Produced in Butane Plasma, Thin Solid Films, 58:117120 (1979)
33. Berkovich, E. S., Three-Faceted Diamond Pyramid for MicroHardness Testing, Indus. Diamond Rev., 11:129132 (1951)
34. Bhattacharya, A. K., and Nix, W. E., Finite Element Simulation of
Indentation Experiments, Int. J. Solids Struct., 24:881891(1988)
35. Bhattacharya, A. K., and Nix, W. D., Analysis of Elastic and
Plastic Deformation Associated with Indentation Testing of the
Thin Films on Substrates, Int. J. Solids Struct., 24:12871298
(1988)
36. Bhushan, B., Stick-Slip Induced Noise Generation in WearLubricated Complaint Rubber Bearing, ASME Journal of Tribology,
102:201212 (1980)
37. Bhushan, B., Overview of Coating Materials, Surface Treatments,
and Screening Techniques for Tribological Applications-Part 2:
Screening Techniques, Testing of Metallic and Inorganic Coatings,
(W. B. Harding and G. A. DiBari, eds.), STP 947, pp. 310319,
ASTM, Philadelphia (1987)
38. Bhushan, B., Tribology and Mechanics of Magnetic Storage Devices,
Second Edition, Springer-Verlag, New York (1996)
39. Bhushan, B., Nanomechanical Properties of Solid Surfaces and
Thin Films, in Handbook of Micro/Nanotribology, (B. Bhushan,
ed.), Second Edition, CRC Press, Boca Raton, FL (1999)
40. Bhushan, B., Micro/Nanotribology and Micro/Nanomechanics of
MEMS, Handbook of Micro/Nanotribology, (B. Bhushan, ed.),
Second Edition, CRC Press, Boca Raton, FL (1995)
41. Bhushan, B., (ed.) Handbook of Micro/Nanotribology, Second
Edition, CRC Press, Boca Raton, FL (1999)
42. Bhushan, B., and Sibley, L. B., Silicon Nitride Rolling Bearing for
Extreme Operating Conditions, ASLE Trans., 25: 417428 (1982)

346 Handbook of Hard Coatings

43. Bhushan, B., Williams, V. S., and Shack, R. V., In-Situ


Nanoindentation Hardness Apparatus for Mechanical
Characterization of Extremely Thin Films, ASME J. Tribology,
110:563571(1988)
44. Bhushan, B., and Doerner, M. F., Role of Mechanical Properties
and Surface Texture in the Real Area of Contact of Magnetic Rigid
Disks, ASME J. Tribology, 111:452458 (1989)
45. Bhushan, B., and Blackman, G. S., Atomic Force Microscopy of
Magnetic Rigid Disks and Sliders and Its Applications to Tribology,
ASME J. Tribology, 113:453458 (1991)
46. Bhushan, B., and Gupta, B. K., Handbook of Tribology: Materials,
Coatings and Surface Treatments, McGraw Hill, New York, NY
(1991)
47. Bhushan, B., Kellock, A. J., Cho, N. H., and Ager, J. W.,
Characterization of Chemical Bonding and Physical Characteristics
of Diamond-like Amorphous Carbon and Diamond Films, J. Mater.
Res., 7:404410 (1992)
48. Bhushan, B., and Venkatesan, S., Mechanical and Tribological
Properties of Silicon for Micromechanical Applications: A Review,
Adv. Info. Storage Syst., 5:211239 (1993)
49. Bhushan, B., Subramaniam, V. V., Malshe, A., Gupta, B. K., and
Ruan, J., Tribological Properties of Polished Diamond Films, J.
Appl. Phys., 74:41744180 (1993)
50. Bhushan, B., and Koinkar, V. N., Nanoindentation Hardness
Measurements Using Atomic Force Microscopy, Appl. Phys. Lett.,
64:16531655 (1994)
51. Bhushan, B., and Koinkar, V. N., Tribological Studies of Silicon
for Magnetic Recording Applications, J. Appl. Phys.,
75:57415746 (1994)
52. Bhushan, B., and Ruan, J., Atomic-Scale Friction Measurements
Using Friction Force Microscopy Part II-Application to Magnetic
Media, ASME J. Tribology, 116:389396 (1994)

Macro- and Micromechanical and Tribological Properties 347


53. Bhushan, B., Koinkar, V. N., and Ruan, J. A., Microtribology of
Magnetic Media, I MechE: Proc. Instn. Mech. Engrs., 208:1729
(1994)
54. Bhushan, B., Subramaniam, V. V., and Gupta, B. K., Polishing of
Diamond Films, Diamond Films and Technology, 4:7197 (1994)
55. Bhushan, B., and Gupta, B. K., Micromechanical Characterization
of Ni-P Coated Aluminum-Magnesium, Glass, and Glass-Ceramic
Substrates, and Finished Magnetic Thin-Film Rigid Disks, Adv.
Info. Storage Syst., 6:208216 (1995)
56. Bhushan, B., Gupta, B. K., and Azarian, M. H., Nanoindentation,
Microscratch, Friction and Wear Studies of Coatings for Contact
recording Applications, Wear, 181183:743758 (1995)
57. Bhushan, B., Gupta, B. K., Sundaram, R., Dey, S., Anders, S.,
Anders, A., Brown, I. G., and Reader, P. D., Development of Hard
Carbon Coatings for Thin-Film Tape Heads, IEEE Trans. Mag.,
31:29762978 (1995)
58. Bhushan, B., Israelachvili, J. N., Landman, U., Nanotribology:
Friction, Wear, and Lubrication on the Atomic Scale, Nature,
374:607616 (1995)
59. Binnig, G., Quate, C. F., and Gerber, C., Atomic Force Microscope,
Phys. Rev. Lett., 56:930933 (1986)
60. Blau, P. J., and Lawn, B. R. (eds.), Microindentation Techniques in
Materials Science and Engineering, STP 889, ASTM, Philadelphia
(1986)
61. Bogli, U., Blatter, A., Pimenov, S. M., Smolin, A. A., and Konov, V.
I., Smoothening of Diamond Films with an ArF Laser, Diamond
and Related Materials, 1:782788 (1992)
62. Bouilov, L. L., Chapliev, N. I., Konov, V. I., Pimenov, S. M.,
Smolin, A. A., and Spitsyn, B. V., Excimer Laser Etching and
Polishing of Diamond Films, J. Electrochem. Soc., 91:357 (1991)
63. Bovard, B. G., Zhao, T., and Macleod, H. A., Oxygen-ion Beam
Polishing of a 5-cm-diameter Diamond Film, Appl. Opt.,
31:23662369 (1992)

348 Handbook of Hard Coatings

64. Bowden, F. P., and Tabor, D., The Friction and Lubrication of
Solids, pp. 172175, Clarendon Press, Oxford Part I (1950), and
Part II (1964)
65. Bravman, J. C., Nix, W. D., Barnett, D. M., and Smith, D. A. (eds.),
Thin Films: Stresses and Mechanical Properties, Symp. Proc.,
Mat. Res. Soc., Vol. 130, Pittsburgh, PA (1989)
66. Brookes, C. A., and Brookes, E. J., Diamond in Perspective: a
Review of Mechanical Properties of Natural Diamond, Diamond
and Related Materials, 1:1317 (1991)
67. Brown, I. G., Anders, A., Anders, S., Dickinson, M. R., Ivanov, I.
C., McGill, R. A., Yao, X. Y., Yu, K. M., Plasma Synthesis of
Metallic and Composite Thin Films with Atomically Mixed Substrate
Bonding, Nucl. Instrum. Method Phys. Res., B 80/81:12811287
(1993)
68. Bubenzer, A., Dischler, B., Brandt, G., and Koidl, P., RF Plasma
Deposited Amorphous Hydrogenated Hard Carbon Thin Films:
Preparation, Properties, and Applications, J. Appl. Phys.,
54:45904595 (1983)
69. Buckle, H., The Science of Hardness Testing and its Research
Applications, (J. W. Westbrook and H. Conrad, eds.), pp. 453491,
Am. Soc. Metals, Metals Park, OH (1973)
70. Bull, S. J., and Rickerby, D. S., New Developments in the Modeling
of the Hardness and Scratch Adhesion of Thin Films, Surf. Coatings
Technol., 42:149164 (1990)
71. Bull, S. J., Chalker, P. R., Johnston, C., and Moore, V., The Effect
of Roughness on the Friction and Wear of Diamond Thin Films,
Surf. Coat. Technol., 68/69:603610 (1994)
72. Bulychev, S. I., Alekhin, V. P., and Shorshorov, M. K., Studies of
Physico-Mechanical Properties in Surface Layers and Microvolumes
of Materials by the Method of Continuous Application of an
Indenter, Fizika Khim. Obrab. Materialov., No. 5 (1979)
73. Burnett, P. J., and Rickerby, D. S., The Relationship Between
Hardness and Scratch Adhesion, Thin Solid Films, 154:403416
(1987)

Macro- and Micromechanical and Tribological Properties 349


74. Burnett, P. J., and Rickerby, D. S., The Mechanical Properties of
Wear Resistant Coatings I: Modelling of Hardness Behavior, Thin
Solid Films, 148:4150 (1987)
75. Burnett, P. J., and Rickerby, D. S., The Mechanical Properties of
Wear-Resistant Coatings II: Experimental Studies and Interpretation
of Hardness, Thin Solid Films, 148:5165 (1987)
76. Campbell, D. S., Mechanical Properties of Thin Films, Handbook
of Thin Film Technology, (L. I. Maissel and R. Glang, eds.), Ch. 12,
McGraw-Hill, NY (1970)
77. Catherine, Y., Preparation Techniques for Diamond-Like Carbon,
Diamond and Diamond-Like Films and Coatings, (R. E. Clausing,
L. L. Horton, J. C. Angus, and P. Koidl, eds. ), pp. 193227, Plenum
Press, New York (1991)
78. Catherine, Y., and Couderc, P., Electrical Characteristics and
Growth Kinetics in Discharges used for Plasma Deposition of
Amorphous Carbon, Thin Solid Films, 144:265280 (1986)
79. Chantikul, P., Anstis, G. R., Lawn, B. R., and Marshall, D. B., A
Critical Evaluation of Indentation Techniques for Measuring Fracture
Toughness: II, Strength Method, J. Am. Ceram. Soc., 64:539543
(1981)
80. Cheng, W., Ling, E., and Finnie, I., Median Cracking of Brittle
Solids Due to Scribing with Sharp Indenters, J. Am. Ceram. Soc.,
73:580586 (1990)
81. Chiang, S. S., Marshall, D. B., and Evans, A. G., Simple Method
for Adhesion Measurement, Surfaces and Interfaces in Ceramics
and Ceramic-Metal Systems, (J. Pask and A. G. Evans, eds.), pp.
603612, Plenum, New York (1981)
82. Chiang, S. S., Marshall, D. B., and Evans, A. G., The Response of
Solids to Elastic/Plastic Indentation: I. Stresses and Residual
Stresses, J. Appl. Phys., 53:298311 (1982)
83. Cho, N. H., Krishnan, K. M., Veirs, D. K., Rubin, M. D., Hopper, C.
B., Bhushan, B., and Bogy, D. B., Chemical Structure and Physical
Properties of Diamond-like Amorphous Carbon Films Prepared by
Magnetron Sputtering, J. Mater. Res., 5:25432554 (1990)

350 Handbook of Hard Coatings

84. Chu, S. N. G., and Li, J. C. M., Impression Creep: A New Creep
Test, J. Mater. Sci., 12:22002208 (1977)
85. Chu, S. N. G., and Li, J. C. M., Localized Stress Relaxation by
Impression Testing, Mater. Sci. Eng., 45:167171 (1980)
86. Clausing, R. E., Horton, L. L., Angus, J. C., Koidl, P., Diamond and
Diamond-Like Films and Coatings, NATO ASI Series, Vol. 266,
Plenum, New York, NY (1991)
87. Clauss, F. J., Solid Lubrication and Self-Lubricated Solids, Academic,
New York, NY (1972)
88. Cook, R. F., and Pharr, G. M., Direct Observation and Analysis of
Indentation Cracking in Glasses and Ceramics, J. Am. Ceram. Soc.,
73:787817 (1990)
89. Cuomo, J. J., Pappas, D. L., Bruley, J., Doyle, J. P., Saenger, K. L.,
Vapor Deposition Processes for Amorphous Carbon Films With
sp 3 Fractions Approaching Diamond, J. Appl. Phys.,
70:17061711 (1991)
90. Cuomo, J. J., Doyle, J. P., Bruley, J., and Liu, J. C., Ion Beam
Sputtered Diamond-Like Carbon with Densities of 2.9 g/cc, J. Vac.
Sci. Technol. A, 9:22102215 (1991)
91. Cuomo, J. J., Pappas, D. L., Lossy, R., Doyle, J. P., Bruley, J.,
DiBello, G. W., and Krakow, W., Energetic Carbon Deposition at
Oblique Angles, J. Vac. Sci. Technol. A, 10:34143418 (1992)
92. Dekempeneer, E. H. A., Jacobs, R., Smeets, J., Meneve, J., Eersels,
L., Blanpain, B., Roos, J., Oostra, D. J., RF Plasma-Assisted
Chemical Vapor Deposition of Diamond-Like Carbon: Physical and
Mechanical Properties, Thin Solid Films, 217:5661 (1992)
93. Deng, K., Ko, W. H., and Michal, G. M., A Preliminary Study on
Friction Measurements in MEMS, 6th Intl. Conf. Solid-State Sensors
and Actuators (Transducers 91), San Francisco, CA, pp. 213216,
IEEE, New York, NY (1991)
94. Deng, K., and Ko, W. H., Static Friction of Diamond-Like Carbon
Film in MEMS, Sensors and Actuators, A35:4550 (1992)
95. Deng, K., and Ko, W. H., A Study of Static Friction Between
Silicon and Silicon Compounds, J. Micromech. Microeng.,
2:1420 (1992)

Macro- and Micromechanical and Tribological Properties 351


96. Deryaguin, B. V., and Fedoseev, D. V., Growth of Diamond and
Graphite from the Gas phase, Nauka, Moscow (1977)
97. Deshpandey, C. V., and Bunshah, R. F., Diamond and Diamondlike
Films: Deposition Processes and Properties, J. Vac. Sci. Technol.
A, 7:22942302 (1989)
98. DeVries, R. C., Synthesis of Diamond Under Metastable
Conditions, Annu. Rev. Mater. Sci., 17:161187 (1987)
99. Doerner, M. F., Gardner, D. S., and Nix, W. D., Plastic Properties
of Thin Films on Substrates as Measured by Submicron Indentation
Hardness and Substrate Curvature Techniques, J. Mater. Res.,
1:845851 (1986)
100. Doerner, M. F. and Nix, W. D., A Method for Interpreting Data
from Depth-Sensing Indentation Instruments, J. Mater. Res.,
1:601609 (1986)
101. Doerner, M. F., Oliver, W. C., Pharr, G. M., and Brotzen, F. R.,
(eds.), Thin Films: Stresses and Mechanical Properties II, MRS
Proceedings, Vol. 188, Mat. Res. Soc., Pittsburgh, PA (1990)
102. Donnet, C., Belin, M., Auge, J. C., Martin, J. M., Grill, A., Patel, V.
V., Tribochemistry of Diamond-Like Carbon Coatings in Various
Environments, Surf. Coat. Technol., 68/69:626631 (1994)
103. Eddy, C. R., Youchison, D. L., Sartwell, B. D., and Grabowski, K.
S., Deposition of Diamond onto Aluminum by Electron Cyclotron
Resonance Microwave Plasma-Assisted CVD, J. Mater. Res.,
7:32553259 (1992)
104. Enke, K., Dimigen, H., and Hubsch, H., Frictional Properties of
Diamondlike Carbon Layers, Appl. Phys. Lett., 36:291292 (1980)
105. Erck, R. A., Nichols, F. A., Dierks, J. F., Pull-Test Adhesion
Measurements of Diamondlike Films on Silicon carbide, Silicon
Nitride, Aluminum Oxide, and Zirconium Oxide, J. Vac. Sci.
Technol. A, 12:15831586 (1994)
106. Erdemir, A., Switala, M., Wei, R., and Wilbur, P., A Tribological
Investigation of the Graphite-to-Diamond-Like Behavior of
Amorphous Carbon Films Ion Beam Deposited on Ceramic
Substrates, Surface Coat. Technol., 50:1723 (1991)

352 Handbook of Hard Coatings

107. Erdemir, A., Nichols, F. A., Pan, X. Z., Wei, R., and Wilbur, P. J.,
Friction and Wear Performance of Ion-Beam-Deposited DiamondLike Carbon Films on Steel Substrates, Diamond and Related
Materials, 3:119125 (1993)
108. Erdemir, A., Bindal, C., Pagan, J., Wilbur, P., Characterization of
Transfer Layers on Steel Surfaces Sliding Against Diamondlike
Carbon in Dry Nitrogen, Thin Solid Films, (1995)
109. Erlandsson, R., McClelland, G. M., Mate, C. M., and Chiang, S.,
Atomic Force Microscopy Using Optical Interferometry, J. Vac.
Sci. Technol. A, 6:266270 (1988)
110. Evans, T., and James, P. F., A Study of the Transformation of
Diamond to Graphite, Proc. R.. Soc. London, Ser. A, 277:260269
(1964)
111. Fabes, B. D., Oliver, W. C., McKee, R. A., and Walker, F. J., The
Determination of Film Hardness From the Composite Response of
Film and Substrate to Nanometer Scale Indentation, J. Mater. Res.,
7:30563064 (1992)
112. Fujimori, S., Kasai, T., and Inamura, T., Carbon Film Formation
by Laser Evaporation and Ion Beam Sputtering, Thin Solid Films,
92:7180 (1982)
113. Furakawa, R., Uyama, H., and Matsumoto, O., Diamond Deposition
With Plasma Jet at Reduced Pressure, IEEE Trans. Plasma Sci.,
18:930933 (1990)
114. Gabriel, K. J., Behi, F., Mahadevan, R., Mehregany, M., In Situ
Friction and Wear Measurement in Integrated Polysilicon
Mechanisms, Sensors and Actuators, A21A23:184188 (1990)
115. Gangopadhyay, A. K., and Tamor, M. A., Friction and Wear
Behavior of Diamond Films Against Steel and Ceramics, Wear,
169:221229 (1993)
116. Gangopadhyay, A. K., Vassell, W. C., Tamor, M. A., Willermet, P.
A., Tribological Behavior of Amorphous Hydrogenated Carbon
Films on Silicon, ASME J. Tribology, 116:454462 (1994)
117. Gardos, M. N., Tribology and Wear Behavior of Diamond, in
Synthetic Diamond: Emerging CVD Science and Technology, (K. E.
Spear and J. P. Dismukes, eds.), pp. 419504, John Wiley, New
York, NY (1994)

Macro- and Micromechanical and Tribological Properties 353


118. Gardos, M. N., and Soriano, B. L., The Effect of Environment on
The Tribological Properties on Polycrystalline Diamond Films, J.
Mater. Res., 5:25992609 (1990)
119. Greene, J. E., Woodhouse, J., and Pestes, M., A Technique for
Detecting Critical Loads in the Scratch Test for Thin-Film Adhesion,
Rev. Sci. Instrum., 45:747749 (1974)
120. Grill, A., Review of the Tribology of Diamond-Like Carbon,
Wear, 168:143153 (1993)
121. Grill, A., Meyerson, B. S., and Patel, V. V., Interface Modifications
for Improving the Adhesion of a-C:H Films to Metals, J. Mater.
Res., 3:214217 (1988)
122. Grill, A., Meyerson, B. S., and Patel, V. V., Diamond-like Carbon
Films by RF Plasma-Assisted Chemical Vapor Deposition from
Acetylene, IBM J. Res. Develop., 34:849857 (1990)
123. Grill, A., Patel, V. V., and Meyerson, B. S., Optical and Tribological
Properties of Heat-Treated Diamond-like Carbon, J. Mater. Res.,
5:25312537 (1990)
124. Grill, A., Horng, C. T., Meyerson, B. S., Patel, V. V., and Russak,
M. A., Magnetic Head Slider Having a Protective Coating Thereon,
U S Patent No. 5,159,508, (Oct. 27, 1992)
125. Grill, A., and Meyerson, B. S., Development and Status of
Diamondlike Carbon, Synthetic Diamond: Emerging CVD Science
and Technology, (K. E. Spear and J. P. Dismukes, eds.), pp. 91141,
John Wiley, New York, NY (1994)
126. Grogan, D. F., Zhao, T., Bovard, B. G., and Macleod, H. A.,
Planarizing Technique for Ion-Beam Polishing of Diamond Films,
Appl. Opt., 31:14831487 (1992)
127. Gupta, B. K., Chevallier, J., and Bhushan, B., Tribology of Ion
Bombarded Silicon for Micromechanical Applications, ASME J.
Tribology, 115:392399 (1993)
128. Gupta, B. K., Bhushan, B., and Chevallier, J., Modification of
Tribological Properties of Silicon by Boron Ion Implantation,
Tribol. Trans., 37:601607 (1994)
129. Gupta, B. K., Malshe, A., Bhushan, B., Subramaniam, V. V., Friction
and Wear Properties of Chemomechanically Polished Diamond
Films, ASME J. Tribology, 116:445453 (1994)

354 Handbook of Hard Coatings

130. Gupta, B. K., and Bhushan, B., The Nanoindentation Studies of Ion
Implanted Silicon, Surface Coat. Technol., 68/69:564570 (1994)
131. Gupta, B. K., and Bhushan, B., Wear Resistant Coatings for Contact
Recording Applications, IEEE Mag. Trans., 31:30123014 (1995)
132. Gupta, B. K., and Bhushan, B., Micromechanical Properties of
Amorphous Carbon Coatings Deposited by Different Deposition
Techniques, Thin Solid Films, 270:391398 (1995)
133. Gupta, B. K., and Bhushan, B., Mechanical and Tribological
Properties of Hard Carbon Coatings for Magnetic Recording Heads,
Wear, 190:110122 (1995)
134. Harding, D. S., Oliver, W. C., and Pharr, G. M., Cracking During
Nanoindentation and its use in the Measurement of Fracture
Toughness, Vol. 356, Mat. Res. Soc., Pittsburgh, PA (1985)
135. Hannula, S. P., Stone, D., and Li, C. Y., Determination of TimeDependent Plastic Properties of Metals by Indentation Load
Relaxation Techniques, Electronic Packaging Materials Science,
(E. A. Giess, K. N. Tu, and D. R. Uhlmann, eds.), 40:217224, Mat.
Res. Soc., Pittsburgh, PA (1985)
136. Hart, E. W., and Solomon, H. D., Load Relaxation Studies of
Polycrystalline High Purity Aluminium, Acta Metall., 21:295307
(1973)
137. Hayward, I. P., Friction and Wear Properties of Diamond and
Diamond Coatings, Surface Coat. Technol., 49:554559 (1991)
138. Hayward, I. P., Singer, I. L., and Seitzman, L. E., Effect of
Roughness on the Friction of Diamond on CVD Diamond Coatings,
Wear, 157:215227 (1992)
139. Heavens, O. S., Some Factors Influencing the Adhesion of Films
Produced by Vacuum Evaporation, J. Phys. Rad., 11:355360
(1950)
140. Hirata, A., Tokura, H., and Yoshikawa, M., Smoothing of Diamond
Films by Ion Beam Irradiation, in: Application of Diamond Films
and Related Materials, (Y. Tzeng, M. Yoshikawa, M. Murakawa,
and A. Feldman, eds.), pp. 227232, Elsevier Sci. Pub., Amsterdam,
Netherlands (1991)

Macro- and Micromechanical and Tribological Properties 355


141. Hirose, Y., Ananuma, S., Okada, N., Komaki, K., Proc. 1st Intl.
Sypm. on Diamond and Diamond-Like Films, The Electrochemical
Society, 89(12):80, Pennington, NJ (1989)
142. Holmberg, K., and Matthews, A., Coating Tribology: Properties,
Techniques and Applications in Surface Engineering, Elsevier,
London, UK (1994)
143. Holmberg, K., Koskinen, J., Ronkainen, H., Vihersalo, J., Hirvonen,
J. P., and Likonen, J., Tribological Characteristics of Hydrogenated
and Hydrogen-Free Diamond-like Carbon Coatings, Diamond Films
and Technology, 4:113129 (1994)
144. Hooper, R. M., and Brookes, C. A., Incubation Periods and
Indentation Creep in Lead, J. Mater. Sci., 19:40574060 (1984)
145. Howe, R. T., Surface Micromachining for Microsensors and
Microactuators, J. Vac. Sci. Technol. B, 6:18091813 (1988)
146. Ishikawa, J., Takeiri, Y., Ogawa, K., and Takagi, T., Transparent
Carbon Film Prepared by Mass-Separated Negative-Carbon-IonBeam Deposition, J. Appl. Phys., 61:25092515 (1987)
147. Jacobson, S., Jonsson, B., and Sundquist, B., The Use of Fast
Heavy Ions to Improve Thin Film Adhesion, Thin Solid Films,
107:8998 (1983)
148. Jansen, F., Machonkin, M., and Kuhman, D. E., The Deposition of
Diamond Films by Filament Techniques, J. Vac. Sci. Technol. A,
8:37853790 (1990)
149. Je, J. H., Gyarmati, E., and Naoumidis, A., Scratch Adhesion Test
of Reactively Sputtered TiN Coatings on a Soft Substrate, Thin
Solid Films, 136:5767 (1986)
150. Jin, S., and Moustakas, T. D., Growth of Diamond Films by ECR
Plasma-Assisted at Low Pressures and Temperatures, Diamond
and Related Materials, 2:13551359 (1993)
151. Johnson, K. L., Contact Mechanics, Cambridge University Press,
Cambridge, UK (1985)
152. Jonsson, B., and Hogmark, S., Hardness Measurements of Thin
Films, Thin Solid Films, 114:257269 (1984)

356 Handbook of Hard Coatings

153. Julia-Schmutz, C., and Hintermann, H. E., Microscratch Testing to


Characterize the Adhesion of Thin Layers, Surface Coat. Technol.,
48:1 (1991)
154. Kamo, M., Sato, Y., Matsumoto, S., and Setaka, N., Diamond
Synthesis From Gas Phase in Microwave Plasma, J. Cryst. Growth,
62:642644 (1983)
155. Kawarada, M., Mar, K., and Kiraki, A., Large Area Chemical
Vapor Deposition of Diamond Particles and Films using MagnetoMicrowave Plasma, Jpn. J. Appl. Phys., 26:L1032L1034 (1987)
156. Kim, D. S., Fischer, T. E., and Gallois, B., The Effects of Oxygen
and Humidity on Friction and Wear of Diamond-Like Carbon Films,
Surf. Coat. Technol., 49:537542 (1991)
157. King, R. B., Elastic Analysis of Some Punch Problems for Layered
Medium, Int. J. Solids Struc., 23:16571664 (1987)
158. Kitahama, K., Hirata, K., Nakamatsu, H., Kawai, S., Fujimori, N.,
Imai, T., Yoshino, H., and Doi, A., Synthesis of Diamond by
Laser-Induced Chemical Vapor Deposition, Appl. Phys. Lett.,
49:634635 (1986)
159. Klages, C. P., Chemical Vapor Deposition of Diamond, Appl.
Phys. A, 56:513526 (1993)
160. Kobayashi, K., Mutsukura, N, and Machi, Y., Deposition of Hard
Carbon Films by the RF Glow Discharge Method, Thin Solid
Films, 158:233238 (1988)
161. Kohzaki, M., Higuchi, K., Noda, S., and Uchida, K., Tribological
Characteristics of Polycrystalline Diamond Films Produced by
Chemical Vapor Deposition, J. Mater. Res., 7:17691777 (1992)
162. Kohzaki, M., Uchida, K., Higuchi, K., and Noda, S., Large-Area
High Speed Diamond Deposition by RF Induction Thermal Plasma
Chemical Vapor Deposition Method, Jpn. J. Appl. Phys.,
32:L438L440 (1993)
163. Koidl, P., Wild, C., Dischler, B., Wagner, J., and Ramsteiner, M.,
Plasma Deposition, Properties and Structure of Amorphous
Hydrogenated Carbon Films, Materials Science Forum,
52&53:4170 (1989)

Macro- and Micromechanical and Tribological Properties 357


164. Kuo, C. T., Yen, T. Y., Huang, T. H., and Hsu, S. E., Adhesion and
Tribological Properties of Diamond Films on Various Substrates,
J. Mater. Res., 5:25152523 (1990)
165. Kurihara, K., Sasaki, K., Kawarada, M., and Koshino, N., High
Rate Synthesis of Diamond of DC Plasma Jet Chemical Vapor
Deposition, Appl. Phys. Lett., 52:437438 (1988)
166. LaFontaine, W. R., Yost, B., Black, R. D., and Li, C., Indentation
Load Relaxation Experiments on Al/Si Metallizations, Symp. Proc.,
188:165170, Mat. Res. Soc., Pittsburgh, PA (1990)
167. LaFontaine, W. R., Yost, B., Black, R. D., and Li, C. Y., Indentation
Load Relaxation Experiments with Indentation Depth in the
Submicron Range, J. Mater. Res., 5:21002106 (1990)
168. LaFontaine, W. R., Paszkiet, C. A., Korhonen, M. A., and Li, C. Y.,
Residual Stress Measurements of Thin Aluminum Metallizations
by Continuous Indentation and X-ray Stress Measurement
Techniques, J. Mater. Res., 6:20842090 (1991)
169. Lankford, J., Threshold-Microfracture During Elastic/Plastic
Indentation of Ceramics, J. Mater. Sci., 16:11771182 (1981)
170. Laugier, M., The Development of Scratch Test Technique for the
Determination of the Adhesion of Coating, Thin Solid Films,
76:289294 (1981)
171. Laursen, T. A., and Simo, J. C., A Study of The Mechanics of
Microindentation Using Finite Elements, J. Mater. Res.,
7:618626 (1992)
172. Lawn, B., Fracture of Brittle Solids, 2nd Ed., Cambridge Univ.
Press (1993)
173. Lawn, B., and Wilshaw, R., Review Indentation Fracture: Principles
and Applications, J. Mater. Sci., 10:10491081 (1975)
174. Lawn, B. R., Evans, A. G., and Marshall, D. B., Elastic/Plastic
Indentation Damage in Ceramics: The Median/Radial Crack System,
J. Am. Ceram. Soc., 63:574581 (1980)
175. Lee, Y. H., Richard, P. D., Bachmann, K. J., and Glass, J. T., BiasControlled Chemical Vapor Deposition of Diamond Thin Films,
Appl. Phys. Lett., 56:620622 (1990)

358 Handbook of Hard Coatings

176. Lettington, A. H., Applications of Diamond-like Carbon Thin


Films, Philos. Trans. R.. Soc. London, Ser. A, 342:287296 (1993)
177. Li, J. C. M., and Chu, S. N. G., Impression Fatigue, Scr. Metall.,
13:10211026 (1979)
178. Li, W. B., Henshall, J. L., Hooper, R. M., and Easterling, K. E., The
Mechanism of Indentation Creep, Acta Metall. Mater.,
39:30993110 (1991)
179. Lim, M. G., Chang, J. C., Schultz, D. P., Howe, R. T., and White, R.
M., Polysilicon Microstructures to Characterize Static Friction,
IEEE Proc. MEMS90, (J. E. Wood and R. T. Howe, eds.), Catalogue
# 90CH 28324, IEEE, pp. 8288, New York (1990)
180. Lin, M. R., Ritter, J. E., Rosenfeld, L., and Lardner, T. J., Measuring
the Interfacial Shear Strength of Thin Polymer Coatings on Glass,
J. Mater. Res., 5:11101117 (1990)
181. Lossy, R., Pappas, D. L., Roy, R. A., and Cuomo, J. J., Filtered Arc
Deposition of Amorphous Diamond, Appl. Phys. Lett., 61:171173
(1992)
182. Lucas, B. N., and Oliver, W. C., The Elastic, Plastic and Time
Dependent Properties of Thin Films as Determined by Ultra Low
Load Indentation, Thin Films: Stresses and Mechanical Properties
III, (W. E. Nix, J. C. Bravman, E. Arzt, and L. B. Freund, eds.),
239:337341, Mat. Res. Soc., Pittsburgh, PA (1992)
183. Marshall, D. B., and Lawn, B. R., Residual Stress Effects in Sharp
Contact Cracking Part 1 Indentation Fracture Mechanics, J. Mat.
Sci., 14:20012012 (1979)
184. Marshall, D. B., and Oliver, W. C., Measurement of Interfacial
Mechanical Properties in Fiber-Reinforced Ceramic Composites,
J. Am. Ceramic Soc., 70:542548 (1987)
185. Mate, C. M., McClelland, G. M., Erlandsson, R., and Chiang, S.,
Atomic-Scale Friction of a Tungsten Tip on a Graphite Surface,
Phys. Rev. Lett., 59:19421945 (1987)
186. Matsumoto, O., Toshima, H., and Kanzaki, Y., Effect of Dilution
Gases in Methane on the Deposition of Diamond-Like Carbon in a
Microwave Discharge, Thin Solid Films, 128:341351 (1985)

Macro- and Micromechanical and Tribological Properties 359


187. Matsumoto, S., Sato, Y., Kamo, M., and Setaka, N., Vapor
Deposition of Diamond Particles From Methane, Jpn. J. Appl.
Phys., 21:L183L185 (1982)
188. Matsumoto, S., Hino, M., and Kobayashi, T., Synthesis of Diamond
Films in a RF Induction Thermal Plasma, Appl. Phys. Lett.,
51:737739 (1987)
189. Matthews, A., and Bachmann, P. K., Diamond and Diamond-Like
Carbon Coatings, Elsevier, Laussane, Switzerland (1990)
190. Mayo, M. J., and Nix, W. D., A Micro-Indentation Study of
Superplasticity in Pb, Sn, and Sn-38Wt% Pb, Acta Metall.,
36:21382192 (1988)
191. Mayo, M. J., Siegel, R. W., Narayanasamy, A., and Nix, W. D.,
Mechanical Properties of Nanophase TiO 2 as Determined by
Nanoindentation, J. Mater. Res., 5:10731082 (1990)
192. McClelland, G. M., Erlandsson, R., and Chiang, S., Atomic Force
Microscopy: General Principles and a New Implementation, In:
Review of Progress in Quantitative Nondestructive Evaluation, (D.
O. Thompson and D. E. Chimenti, eds.), 6B:13071314 Plenum
Press, New York (1987)
193. McKenzie, D. R., Muller, D., and Pailthorpe, B. A., CompressiveStress-Induced Formation of Thin-Film Tetrahedral Amorphous
Carbon, Phys. Rev. Lett., 67:773776 (1991)
194. Mecholsky, J. J., Tsai, Y. L., and Drawl, W. R., Fracture Studies of
Diamond on Silicon, J. Appl. Phys., 71:48754881 (1992)
195. Mehregany, M., Senturia, S. D., and Lang, G. H., Measurement of
Wear in Polysilicon Micromotors, IEEE Trans. Electron Devices,
39:11361143 (1992)
196. Mehrotra, P. K., and Qunito, D. T., Techniques for Evaluating
Mechanical Properties of Hard Coatings, J. Vac. Sci. Technol.,
A3:24012405 (1985)
197. Memming, R., Tolle, H. J., and Wierenga, P. E., Properties of
Polymeric Layers of Hydrogenated Amorphous Carbon Produced
by a Plasma-Activated Chemical Vapor Deposition Process II:
Tribological and Mechanical Properties, Thin Solid Films,
143:3141 (1986)

360 Handbook of Hard Coatings

198. Mitsuda, Y., Yoshida T., and Akashi K., Development of a New
Microwave Plasma Torch and Its Application to Diamond Synthesis,
Rev. Sci. Instrum., 60:249252 (1989)
199. Mittal, K. L., (ed.), Adhesion Measurements on Thin Coatings,
Thick Coatings and Bulk Coatings, STP 640, ASTM, Philadelphia
(1978)
200. Miyasato, T., Kawakami, Y., Kawano, T., Hiraki, A., Preparation
of sp3-Rich Amorphous Carbon Film by Hydrogen Gas Reactive
RF-Sputtering of Graphite, and Its Properties, Jpn. J. Appl. Phys.,
23:L234L237 (1984)
201. Miyoshi, K., Studies of Mechanochemical Interactions in the
Tribological Behavior of Materials, Surf. Coat. Technol.,
43/44:799812 (1990)
202. Miyoshi, K., Wu, R. L. C., and Garscadden, A., Friction and Wear
of Diamond and Diamondlike Carbon Coatings, Surf. Coat.
Technol., 54/55:428434 (1992)
203. Miyoshi, K., Wu, R. L. C., Garscadden, A., Barnes, P. N., and
Jackson, H. E., Friction and Wear of Plasma-Deposited Diamond
Films, J. Appl. Phys., 74:44464454 (1993)
204. Mott, B. W., Microindentation Hardness Testing, Butterworths,
London (1957)
205. Moustakas, T. D., Growth of Diamond CVD Method and Effect of
Process Parameters, in: Synthetic Diamond: Emerging CVD
Science and Technology, (K. E. Spear and J. P. Dismukes, eds.),
pp. 145192, John Wiley, New York, (1994)
206. Mulhearn, T. O., and Tabor, D., Creep and Hardness of Metals: A
Physical Study, J. Inst. of Metals, 89:712 (196061)
207. Nastasi, M., Parkin, D. M., and Gleiter, H., (eds.), Mechanical
Properties and Deformation Behavior of Materials Having UltraFine Microstructures, Kluwer Academic Pub., Dordrecht,
Netherlands (1993)
208. Nemanich, R. J., Growth and Characterization of Diamond Thin
Films, Annu. Rev. Mater. Sci., 21:535558 (1991)

Macro- and Micromechanical and Tribological Properties 361


209. Nicoll, A. R., A Survey of Methods Used for the Performance
Evaluation of High Temperature Coatings, in: Coatings for High
Temperature Applications, (E. Lang, ed.), pp. 269339, Applied
Science Publishers, London (1983)
210. Nix, W. D., Mechanical Properties of Thin Films, Metall. Trans.,
20A:22172245 (1989)
211. Nix, W. D., Bravman, J. C., Arzt, E., and Freund, L. B., (eds.), Thin
Films: Stresses and Mechanical Properties III, Symp Proc., Vol
239, Mat. Res. Soc., Pittsburgh, PA (1992)
212. Noguchi, K., Fujita, H., Suzuki, M., Yoshimura, N., The
Measurements of Friction on Micromechatronics Elements, Proc.
IEEE Micro Electro Mechanical Systems, (M. Esashi and H. Fujita,
eds.), pp. 148153, IEEE, New York (1991)
213. OHern, M. E., and McHargue, C. J., Mechanical Properties Testing
of Diamond and Diamond-Like Films by Ultra-Low Load
Indentation, in: Diamond and Diamond-Like Films and Coatings,
(R. E. Clausing, L. L. Horton, J. C. Angus, P. Koidl, eds.), pp.
715721 Plenum Press, New York (1991)
214. Ohtake, N., and Yoshikawa, M., Diamond Film Preparation by Arc
Discharge Plasma Jet Chemical Vapor Deposition in the Methane
Atmosphere, J. Electrochemical Soc., 137:717722 (1990)
215. Ojha, S. M., Norstorm, M., and McCulluch, D., The Growth
Kinetics and Properties of Hard and Insulating Carbonaceous Films
Grown in an RF Discharge, Thin Solid Films, 60:213225 (1980)
216. Oliver, W. C., Hutchings, R., and Pethica, J. B., Measurement of
Hardness at Indentation Depths as Small as 20 nm, Microindentation
Techniques in Materials Science and Engineering, (P. J. Blau and B.
R. Lawn, eds.), STP 889, pp. 90108, ASTM, Philadelphia, PA
(1986)
217. Oliver, W. C., and Pethica, J. B., Methods for Continuous
Determination of the Elastic Stiffness of Contact Between Two
Bodies, U.S. Patent No. 4,848,141 (July 18, 1989)
218. Oliver, W. C., and Pharr, G. M., An Improved Technique for
Determining Hardness and Elastic Modulus Using Load and
Displacement Sensing Indentation Experiments, J. Mater. Res.,
7:15641583 (1992)

362 Handbook of Hard Coatings

219. ONeill, H., Hardness Measurement of Metals and Alloys, Chapman


and Hall, London (1967)
220. Palmquist, S., Jernkontorets Ann., 141:300 (1957)
221. Pappas, D. L., Saenger, K. L., Bruley, J., Krakow, W., and Cuomo,
J. J., Pulsed Laser Deposition of Diamond-Like Carbon Films, J.
Appl. Phys., 71:56755684 (1992)
222. Pappas, D. L., and Hopwood, Deposition of Diamond-Like Carbon
Using a Planar Radio Frequency Induction Plasma, J. Vac. Sci.
Technol. A, 12:15761582 (1994)
223. Perry, A. J., The Adhesion of Chemically Vapor-Deposited Hard
Coatings on Steel-The Scratch Test, Thin Solid Films, 78:7793
(1981)
224. Perry, A. J., Scratch Adhesion Testing of Hard Coating, Thin
Solid Films, 197:167180 (1983)
225. Perry, A. J., Scratch Adhesion Testing: A Critique, Surface
Enginnering, 2:183189 (1986)
226. Perry, S. S., Ager, J. W., III., Somorjai, G. A., Combined Surface
Characterization and Tribological (Friction and Wear) Studies of
CVD Diamond Films, J. Mater. Res., 8:25772586 (1993)
227. Peterson, M. B., and Winer, W. O., (eds.), Wear Control Handbook,
ASME, New York (1980)
228. Pethica, J. B., Hutchings, R., and Oliver, W. C., Hardness
Measurements at Penetration Depths as Small as 20 nm, Phil.
Mag., A48:593606 (1983)
229. Pethica, J. B., and Oliver, W. C., Mechanical Properties of
Nanometer Volumes of Material: Use of the Elastic Response of
Small Area Indentations, Thin Films: Stresses and Mechanical
Properties, (J. C. Bravman, W. D. Nix, D. M. Barnett, and D. A.
Smith, eds.), 130:1323, Mat. Res. Soc., Pittsburgh, PA (1989)
230. Pharr, G. M., and Oliver, W. C., Measurement of Thin Film
Mechanical Properties Using Nanoindentation, MRS Bulletin, pp.
2833 (July 1992)
231. Pharr, G. M., Oliver, W. C., and Brotzen, F. R., On the Generality
of the Relationship Among Contact Stiffness, Contact Area, and
Elastic Modulus During Indentation, J. Mater. Res., 7:613617
(1992)

Macro- and Micromechanical and Tribological Properties 363


232. Pharr, G. M., Harding, D. S., and Oliver, W. C., Measurement of
Fracture Toughness in Thin Films and Small Volumes Using
Nanoindentation Methods, Mechanical Properties and Deformation
Behavior of Materials Having Ultra-Fine Microstructures, (M.
Nastasi, D. M. Parkin, and H. Gleiter, eds.), pp. 449461, Kluwer
Academic Pub., Dordrecht, Netherlands (1993)
233. Pulker, H. K., and Salzmann, K., Micro-/Ultramicro Hardness
Measurements with Insulating Films, SPIE Thin Film Technologies,
652:139144 (1986)
234. Raman, V., and Berriche, R., Creep Behavior of Sputtered TiN
Films Using Indentation Testing, Thin Films: Stresses and
Mechanical Properties II, (M. F. Doerner, W. C. Oliver, G. M.
Pharr, and F. R. Brotzen, eds.), pp. 171176, Mat. Res. Soc.,
Pittsburgh, PA (1990)
235. Raman, V., and Berriche, R., An Investigation of the Creep Processes
in Tin and Aluminum Using a Depth-Sensing Indentation
Technique, J. Mater. Res., 7:627638 (1992)
236. Raveh, A., Martinu, L., Hawthorne, H. M., and Wertheimer, M. R.,
Mechanical and Tribological Properties of Dual-Frequency PlasmaDeposited Diamond-Like Carbon, Surf. Coat. Technol., 58:4555
(1993)
237. Rickerby, D. S., and Matthews, A., Advanced Surface Coatings: A
Handbook of Surface Engineering, Chapman and Hall, New York
(1991)
238. Robertson, J., Amorphous Carbon, Advances in Phys.
35:317374 (1986)
239. Robertson, J., Structure and Electronic Properties of DiamondLike Carbon, in: Diamond and Diamond-Like Carbon Films, (R. E.
Clausing, et al., eds.), pp. 331356, Plenum Press, New York (1991)
240. Robertson, J., Properties of Diamond-Like Carbon, Surf. Coat.
Technol., 50:185203 (1992)
241. Robertson, J., Deposition of Diamond-Like Carbon, Philos. Trans.
R. Soc. London, Ser. A, 342:277286 (1993)
242. Ruan, J., and Bhushan, B., Atomic-Scale Friction Measurements
Using Friction Force Microscopy: Part I-General Principles
and New Measurement Techniques, ASME J. Tribology,
116:378388 (1994)

364 Handbook of Hard Coatings

243. Rubin, M., Hopper, C. B., Cho, N. H., and Bhushan, B., Optical and
Mechanical Properties of DC Sputtered Carbon Films, J. Mater.
Res., 5:2538 2542 (1990)
244. Sargent, P. M., Use of the Indentation Size Effect on Microhardness
of Materials Characterization, Microindentation Techniques in
Materials Science and Engineering, (P. J. Blau, and B. R. Lawn,
eds.), STP 889, pp. 160174, ASTM, Philadelphia (1986)
245. Savvides, N., Diamondlike Thin Films and Their Properties,
Materials Science Forum, 52 /53:407426 (1989)
246. Savvides N., and Bell, T. J., Microhardness and Youngs Modulus
of Diamond and Diamondlike Carbon Films, J. Appl. Phys.,
72:27912796 (1992)
247. Sawabe, A., and Inuzuka, T., Growth of Diamond Thin Films by
Electron Assisted Chemical Vapor Deposition, Appl. Phys. Lett.,
46:146147 (1985)
248. Scheibe, H. J., and Schultrich, B., DLC Film Deposition by LaserArc and Study of Properties, Thin Solid Films, 246:92102 (1994)
249. Scruby, C. B., An Introduction to Acoustic Emission, J. Phys. E:
Sci. Instrum., 20:946953 (1987)
250. Seal, M., The Effect of Surface Orientation on the Graphitization
of Diamond, Phys. Stat. Sol., 3:658664 (1963)
251. Seino, Y., Hida, N., and Nagai S., Mechanical Properties of Diamond
Thin Film Prepared by Chemical Vapor Deposition, J. Mater. Sci.
Lett., 11:515517 (1992)
252. Seino, Y., and Nagai S., Temperature Dependence of Youngs
Modulus of Diamond Thin Film Prepared by Microwave Plasma
Chemical Vapor Deposition, J. Mater. Sci. Lett., 12:324325 (1993)
253. Sekler, J., Steinmann, P. A., and Hintermann, H. E., The Scratch
Test: Different Critical Load Determination Techniques, Surf. Coat.
Technol., 36:519529 (1988)
254. Shih, C. W., Yang, M., and Li, J. C. M., Effect of Tip Radius on
Nanoindentation, J. Mater. Res., 6:26232628 (1991)
255. Sneddon, I. N., The Relation Between Load and Penetration in the
Axisymmetric Boussinesq Problem for a Punch of Arbitrary Profile,
Int. J. Eng. Sci., 3:4757 (1965)

Macro- and Micromechanical and Tribological Properties 365


256. Spear, K. E., Diamond-Ceramic Coating of the Future, J. Am.
Ceram. Soc., 72:171191 (1989)
257. Spear, K. E., and Dismukes, J. P., Synethtic Daimond: Emerging
CVD Science and Technology John Wiley, New York, NY (1994)
258. Spitsyn, B. V., Chemical Crystallization of Diamond from the
Activated Vapor Phase, J. Crystal Growth, 99:11621167 (1990)
259. Spitsyn, B. V., and Deryaguin, B. V., A Technique of Diamond
Growth on a Diamond Face, USSR Patent No. 339,134 (May 5,
1980)
260. Spitsyn, B. V., Bouilov, L. L., and Deryaguin, B. V., Diamond and
Diamond-Like Coatings: Deposition from the Vapor Phase, Structure
and Properties, Prog. Crystal Growth and Character., 17:79170
(1988)
261. Steinmann, P. A., Tardy, Y., and Hintermann, H. E., Adhesion
Testing by the Scratch Test Method: The Influence of Intrinsic and
Extrinsic Parameters on the Critical Load, Thin Solid Films,
154:333349 (1987)
262. Stilwell, N. A., and Tabor, D., Elastic Recovery of Conical
Indentation, Proc. Phys. Soc., 78: 169179 (1961)
263. Stone, D., LaFontaine, W. R., Alexopoulos, P. S., Wu, T. W., and
Li, C. Y. , An Investigation of Hardness and Adhesion of SputterDeposited Aluminum on Silicon by Utilizing a Continuous
Indentation Test, J. Mater. Res., 3:141147 (1988)
264. Stoner, B. R., Ma, G. H. M., Wolter, S. D., and Glass, J. T.,
Characterization of Bias-Enhanced Nucleation of Diamond on
Silicon by in vacuo Surface Analysis and Transmission Electron
Microscopy, Phys. Rev. B, 45:1106711084 (1992)
265. Sundgren, J. E., and Hentzell, H. T. G., A Review of the Present
Status of the Art in Hard Coatings Grown from the Vapor Phase, J.
Vac. Sci. Technol. A, 4:22592279 (1986)
266. Suzuki, K., Sawabe, A., and Yasuda, H., Inuzuka, T., Growth of
Diamond Thin Films by DC Plasma Chemical Vapor Deposition,
Appl. Phys. Lett., 50:728729 (1987)
267. Suzuki, J. I., Kawarada, H., Mar, K. S., Wei, J., Yokota, Y., and
Hiraki, A., The Synthesis of Diamond Films at Lower Pressure and
Lower Temperature Using Magneto-Microwave Plasma CVD, Jpn.
J. Appl. Phys., 28 L281L283 (1989)

366 Handbook of Hard Coatings

268. Suzuki, S., Matsuura, T., Uchizawa, M., Yura, S., and Shibata, H.,
Friction and Wear Studies on Lubricants and Materials Applicable
to MEMS, Proc. IEEE Micro Electro Mechanical Systems, (M.
Esashi, and H. Fujita, eds.), pp. 143147, IEEE, New York (1991)
269. Suzuki, S., Matsuura, T., Karino, I., Sakai, Y., and Shibata, H.,
Quality and Wear Durability of Diamond-Like Carbon Protective
Films for Thin Film Magnetic Media, Tribology and Mechanics of
Magnetic Storage Systems, STLE SP-29, pp. 144150 ,STLE, Park
Ridge, IL (1990)
270. Swain, M. V., Hagan, J. T., Field, J. E., Determination of the
Surface Residual Stresses in Tempered Glasses by Indentation
Fracture Mechanics, J. Mater. Sci., 12:19141917 (1977)
271. Swec, D. M., Mirtich, M. J., and Banks, B. A., Ion Beam and
Plasma Methods of Producing Diamondlike Carbon Films, 32nd
Annual International Technical Symposium on Optical and
Optoelectronic Applied Science and Engineering, Proc. SPIE,
Bellingham, WA (1989)
272. Tabor, D., The Hardness of Metals, Clarendon Press, Oxford, UK
(1951)
273. Tabor, D., The Hardness of Solids, Rev. Phys. Technol.,
1:145179 (1970)
274. Tai, Y. C., and Muller, R. S., Frictional Study of IC Processed
Micromotors, Sensors and Actuators, A21A23:180183 (1990)
275. Tai, Y. C., and Muller, R. S., IC Processed Electrostatic Synchronous
Micromotors, Sensors and Actuators, A20:4955 (1989)
276. Takagi, T., Surface Interactions during Thin Film Deposition, J.
Vac. Sci. Technol. A, 2:382388 (1984)
277. Tamor, M. A., Vassell, W. C., and Carduner, K. R., Atomic
Constraint in Hydrogenated Diamondlike Carbon, Appl. Phys.
Lett., 58:592594 (1991)
278. Tangena, A. G., and Hurkx, G. A. M., The Determination of StressStrain Curves of Thin Layers Using Indentation Tests, ASME J.
Eng. Mat. Technol., 108:230232 (1986)
279. Ternovskii, A. P., Alekhin, V. P., Shorshorov, M. K., Khrushchov,
M. M., and Skvortsov, V. N., Micromechanical Testing of Materials
by Depression, Zavod. Lab., 39:16201624 (1973)

Macro- and Micromechanical and Tribological Properties 367


280. Thronton, A. G., and Wilks, J., Diamond Research 1974, Suppl.
Ind. Diamond Rev., 3942 (1974)
281. Tokura, H., Yang, C. F., and Yoshikawa, M., Study on the Polishing
of Chemically Vapor Deposited Diamond Films, Thin Solid Films,
212:4955 (1992)
282. Townsend, P. H., Weihs, T. P., Sanchez, J. E., and Borgesen, P.
(eds.), Thin Films: Stresses and Mechanical Properties IV, Vol.
308, Mat. Res. Soc., Pittsburgh, PA (1993)
283. Tsai, H., Structure and Physical Properties of Amorphous
Hydrogenated Carbon (a-C:H) Films, Materials Science Forum,
52/53:71102 (1989)
284. Tsai, H., and Bogy D. B., Characterization of Diamond-like Carbon
Films and Their Application as Overcoats on Thin-Film Media for
Magnetic Recording, J. Vac. Sci. Technol. A, 5:32873312 (1987)
285. Tsukamoto, Y., Kuroda, H., Sato, A., and Yamaguchi, H.,
Microindentation Adhesion Tester and its Applications to Thin
Films, Thin Solid Films, 213:220225 (1992)
286. Tzeng, Y., Yoshikawa, M., Murakawa, M., and Feldman, A., (eds.),
Applications of Diamond Films and Related Materials, Elsevier
Science Pub., Amsterdam, The Netherlands (1991)
287. Valli, J., A Review of Adhesion Test Method for Thin Hard
Coatings, J. Vac. Sci. Technol. A, 4:30073014 (1986)
288. Valli, J., Makela, U., and Matthews, A., Assessment of Coating
Adhesion, Surface Engineering, 2:4953 (1986)
289. Vandentop, G. J., Kawasaki, M., Nix, R. M., Brown, I. G., Salmeron,
M., and Somorjai, G. A., Formation of Hydrogenated Amorphous
Carbon Films of Controlled Hardness From a Methane Plasma,
Phys. Rev., B41:32003210 (1990)
290. Walker, W. W., The Science of Hardness Testing and its Research
Applications, (J. H. Westbrook, and H. Conrad, eds.), pp. 258273,
Am. Soc. Metals, Metals Park, OH (1973)
291. Wang, M., Schmidt, K., and Reichelt, K., Characterization of
Metal-Containing Amorphous Hydrogenated Carbon Films, J.
Mater. Res., 7:667676 (1992)

368 Handbook of Hard Coatings

292. Wei, J., Kawarada, H., Suzuki, J. I., and Hiraki, A., LowTemperature Synthesis of Diamond Films using Magneto-Microwave
Plasma CVD, Jpn. J. Appl. Phys., 29:L1483L1485 (1990)
293. Wei, R., Wilbur, P. J., Erdemir, A., and Kustas, F. M., The Effects
of Beam Energy and Substrate Temperature on the Tribological
Properties of Hard-Carbon Films on Aluminum, Surf. Coat.
Technol., 51:139145 (1992)
294. Wei, R., Wilbur, P. J., and Liston, M. J., Effects of Diamond-Like
Hydrocarbon Films on Rolling Contact Fatigue of Bearing Steels,
Diamond and Related Materials, 2:898903 (1993)
295. Weihs, T. P., Lawrence, C. W., Derby, C. B., and Pethica, J. B.,
Acoustic Emissions During Indentation Tests, Thin Films: Stresses
and Mechanical Properties III, Symp. Proc., (W. D. Nix, J. C.
Bravman, E. Arzt, and L. B. Freund, eds.), 239:361370, Mat. Res.
Soc., Pittsburgh, PA (1992)
296. Weissmantel, C., Bewilogua, K., Breuer, K., Dietrich, D., Ebersbach,
U., Erler, H. J., Rau, B., and Reisse, G., Preparation and Properties
of Hard i-C and i-BN Coatings, Thin Solid Films, 96:3144 (1982)
297. Westbrook, J. H., Proc. Am. Soc. Test. Mater., 57:873 (1957)
298. Westbrook, J. H., and Conrad, H., (eds.), The Science of Hardness
and its Research Applications, Am. Soc. Metals, Metals Park, Ohio
(1973)
299. White, R. L., Nelson, J., and Gerberich, W. W. , Residual Stress
Effects in the Scratch Adhesion Testing of Tantalum Thin Films,
Thin Films: Stresses and Mechanical Properties IV, (P. H. Townsend,
T. P. Weihs, J. E. Sanches, and P. Borgesen, eds.), 308:141146,
Mat. Res. Soc., Pittsburgh, PA (1993)
300. Wood, P., Wydeven, T., and Tsuji, O., Effect of Temperature on
the Deposition Rate and Properties of Hydrogenated Amorphous
Carbon Films, Surf. Coat. Technol., 49:399405 (1991)
301. Wu, T. W., Microscratch Test for Ultra-Thin Films, Thin Films:
Stresses and Mechanical Properties II, (M. F. Doerner, W. C.
Oliver, G. M. Pharr, and F. R. Brotzen, eds.), Symp. Proc.,
188:191205, Mat. Res. Soc., Pittsburgh, PA (1990)
302. Wu, T. W., Microscratch and Load Relaxation Tests for Ultra-Thin
Films, J. Mater. Res., 6:407426 (1991)

Macro- and Micromechanical and Tribological Properties 369


303. Wu, T. W., The AC Indentation Technique And Its Applications,
Mater. Chem. Phys., 33:1530 (1993)
304. Wu, T. W., Hwang, C., Lo. J., and Alexopoulos, P., Microhardness
and Microstructure of Ion-Beam-Sputtered, Nitrogen Doped NiFe
Films, Thin Solid Films, 166:299308 (1988)
305. Wu, T. W., Burn, R. A., Chen, M. M., and Alexopoulos, P. S.,
Micro-Indentation and Micro-Scratch Tests on Sub-Micron Carbon
Films, Symp. Proc., 130:117121, Mat. Res. Soc., Pittsburgh, PA
(1989)
306. Wu, T. W., Shull, A. L., and Lin, J., Microscratch Test on Carbon
Films as Thin as 20 nm, Symp. Proc., 188:207212, Mat. Res. Soc.,
Pittsburgh, PA (1990)
307. Wu, T. W., Shull, A. L., and Berriche, R. , Microindentation
Fatigue Tests on Submicron Carbon Films, Surf. Coat. Technol.,
47:696709 (1991)
308. Wu., T. W., and Lee, C. K., The Micro-wear Technique and its
Application to Ultrathin Film Systems, J. Mater. Res., 9:805811
(1994)
309. Wyon, C., Gillet, R., and Lombard, L., Properties of Amorphous
Carbon Films Produced by Magnetron Sputtering, Thin Solid Films,
122:203216 (1984)
310. Yarbrough, W. A., Inspector, A., and Messier, R., The Chemical
Vapor Deposition of Diamond, Materials Science Forum, 52/
53:151174 (1989)
311. Yarbrough, W. A., and Messier, R., Current Issues and Problems in
the Chemical Vapor Deposition of Diamond, Science,
247:688696 (1990)
312. Yeak-Scranton, C. E., Novel Piezoelectric Transducer to Monitor
Head-Disk Interactions, IEEE Trans. Magn., Mag. 22:10111016
(1986)
313. Yust, C., and Bayer, R. G. (eds.), Selection and Use of Wear Tests
for Ceramics, STP 1010, ASTM, Philadelphia (1988)
314. Zhao, T., Grogan, D. F., Bovard, B. G., and Macleod, H. A.,
Diamond Film Polishing With Argon and Oxygen Ion Beams, in:
SPIE Diamond Optics III, 1325:142151(1990)

You might also like