Nanoindentation Studies of Materials: Materials Physics. With High-Resolution Load-Displacement Data, Discrete

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 11

Nanoindentation

studies of materials
Nanoindentation has become a commonplace tool for the measurement
of mechanical properties at small scales, but may have even greater
importance as a technique for experimental studies of fundamental
materials physics. With high-resolution load-displacement data, discrete
events including dislocation source activation, shear instability initiation,
and phase transformations can be detected during a nanoindentation
test. Recently-developed capabilities in, for example, high-temperature
nanoindentation testing and in situ imaging of the indented volume, offer
new quantitative details about these phenomena, and present many
opportunities for future scientific inquiry.
Christopher A. Schuh
Department of Materials Science and Engineering, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Room 8-211, Cambridge,
MA 02139 USA
E-mail: schuh@mit.edu

For more than a century, researchers in the mechanical sciences 16-19


stresses . Such measurements have broad application across the
have recognized that surface contacts between materials are highly physical sciences, and there are several recent reviews on the use of
1,2 20-22
dependent on their mechanical properties . Many different nanoindentation for property extraction . However, nanoindentation
indentation and impression tests have been developed in an effort to also lends itself to more fundamental inquiries in materials science. New
measure such mechanical properties from a contact of known capabilities in in situ and ex situ imaging, acoustic emission detection,
geometry. In the past two decades, however, a veritable revolution and high-temperature testing are now being used to probe nanoscale
has occurred in indentation testing, owing to the development of phenomena such as defect nucleation and dynamics, mechanical
new sensors and actuators that allow instrumented indentations to instabilities or strain localization, and phase transformations. This article
be routinely performed on submicron scales. The resulting presents an overview of nanoindentation’s current role and future
technique, termed nanoindentation, has now become ubiquitous for promise in the study of some of these physical phenomena, with special
mechanical property measurements at surfaces. emphasis on experimental approaches.
The most common use of nanoindentation is for the measurement of

hardness and elastic modulus


3-5
, and there has been considerable progress
Nanoindentation
in the measurement of other mechanical parameters as well, including The principal components in a nanoindentation experiment are the test
6-10 11-15 material, the sensors and actuators used to apply and measure the
hardening exponents , creep parameters , and residual

32 MAY 2006 | VOLUME 9 | NUMBER 5 ISSN:1369 7021 © Elsevier Ltd 2006 Open access under CC BY-NC-ND license.
Nanoindentation studies of materials REVIEW FEATURE

(a) (b)

(c)

Fig. 1 Views of the Berkovich diamond geometry commonly used in nanoindentation testing. (a) Profile view as observed in a scanning electron
microscope, showing the pyramidal diamond tip embedded in a braze. (b) Top-down atomic force microscopy image of the tip, illustrating the
three-fold pyramidal symmetry. (c) Line scan height profile of the indenter apex geometry along the line indicated in (b) illustrating the blunting of
the apex into a roughly spherical shape with effective diameter ~1000 nm. [Part (a) is courtesy of C. E. Packard and J. R. Trelewicz,
38
Massachusetts Institute of Technology; parts (b) and (c) are reprinted with permission from . © 2002 Elsevier Ltd.]

mechanical load and indenter displacement, and the indenter tip. The experiments; these are characteristic of energy-absorbing or energy-
latter component is conventionally made of diamond, formed into a sharp, releasing events occurring beneath the indenter tip. The three
symmetric shape such as the three-sided Berkovich pyramid pictured in
examples shown correspond to three different physical phenomena,
Fig. 1. The pyramidal shape is chosen at least in part for its nominal
geometric self-similarity, which makes for relatively simpler analysis using
the methods of continuum mechanics. However, because of the very fine
scale of nanoindentation testing, imperfections in the pyramidal tip shape
are of paramount importance in such analysis, and much effort has been
focused upon methods of characterizing and cataloging tip shapes for
more exact quantitative measurements
3,23-26.

Of particular relevance in this regard is the nature of the tip apex,


which is never atomically sharp and exhibits significant blunting,
as shown in the atomic force microscope scan of Fig. 1c.
During a typical nanoindentation test, force and displacement
are recorded as the indenter tip is pressed into the test material’s
surface with a prescribed loading and unloading profile. The
response of interest is the load-displacement curve (often called
the P-h curve), such as depicted in Fig. 2 for an indentation on
single-crystal Pt(100). The global shape of the P-h curve differs
from one material to the next, and these variations usually reflect
different mechanical properties. Of more interest to the present
article are local details in the P-h curve, which may signal the
operation of discrete physical events beneath the indenter tip.
Fig. 3 shows three examples of P-h curves that exhibit interesting Fig. 2 Example of a typical load-displacement curve (P-h curve)
obtained during nanoindentation of an elastic-plastic material, in this
local perturbations or discontinuities measured in load-controlled case single-crystal Pt(100).

MAY 2006 | VOLUME 9 | NUMBER 5


REVIEW FEATURE Nanoindentation studies of materials

(a) (b) (c)

27 28
Fig. 3 Three P-h curves that exhibit interesting discontinuities from indentations on (a) pure Pt(100) , (b) Pd-Ni-Cu-P metallic glass ,
29
and (c) single-crystal Si(100) . Some of the relevant discontinuities are indicated by arrows in each case.

in three different materials spanning various states of Because of the very small volume of material sampled in a
bonding and structural order: nanoindentation, these events are detected in a discrete fashion in
• Dislocation activity is detected during a shallow indentation real time, and can be isolated and studied in detail. In this article, we
27 will focus our attention on the three specific phenomena listed
into single-crystal Pt ,
above, although other discrete events (such as fracture or twinning)
• Shear localization into ‘shear bands’ is measured in a Pd-
28 may be detected during nanoindentation as well.
based amorphous alloy , and
• A phase transformation with a significant volume increase is Incipient plasticity
29
detected during unloading of an indentation on Si . The largest body of literature using nanoindentation for fundamental
materials science is that concerned with the detection and
30-48
understanding of plastic yield on the nanoscale , such as for the
case of the indentation on pure Pt shown in Fig. 3a. These studies of
‘incipient plasticity’ often focus upon the very earliest stages of the
mechanical contact, where the transition from elastic to plastic
deformation can be observed. In this context, the inevitable blunting
of the Berkovich tip turns out to be something of a benefit rather than
an experimental difficulty; in the earliest stages of contact at which
plastic yield first occurs, the geometry of the nanoindenter can often
be approximated as spherical. With this fortuitous geometry, it
becomes possible to predict the expected elastic response using the
Hertzian law for mechanical contacts, based on isotropic continuum
elasticity. This law predicts a simple power-law form for the elastic
3/2
portion of the load-displacement curve, P ∝ h , with a
proportionality constant that is fully specified by the radius of the
blunted indenter tip and the elastic properties of the two contacting
2
materials . This expression has been found accurate when
compared with the earliest stages of experimental P-h curves from a
40-47
variety of materials . An example from indentations on (0001)-
oriented single crystals of 4H-SiC is shown in Fig. 4.
With theoretical expectations for the elastic response given by
Fig. 4 A load-displacement (P-h) curve for 4H-SiC indented by a the Hertzian theory, the onset of plastic deformation during
46
roughly spherical tip of ~300 nm radius . The view here is nanoindentation can nominally be identified by the first point at
magnified to better illustrate the initial deviation from elastic
contact, as predicted using the Hertzian contact law. The yield which the experimental data deviate from the elastic curve. In Fig. 4,
point corresponds to a discontinuity, or ‘pop-in’ event.

34 MAY 2006 | VOLUME 9 | NUMBER 5


Nanoindentation studies of materials REVIEW FEATURE

remain under active discussion. Details of the atomistic mechanisms


behind the pop-in have largely been provided by the simulation
53
community. For example, Kelchner et al. first demonstrated via
Fig. 5 Atomistic simulation result for indentation of Al with a spherical indenter molecular dynamics simulation the possibility that yield may be
57
tip, illustrating the nucleation of a dislocation beneath the indented surface . associated with the homogeneous nucleation of dislocation loops beneath
Only miscoordinated atoms are visible, revealing the position of the surface a nanoindentation on pure Au – a result confirmed by many later
(colored pink and bowed out under the load of the indenter) as well as the first
crystal defect that formed at the yield point (colored in blue). (Reprinted 37,54-57
simulations conducted on various other model materials . An
courtesy of K. J. Van Vliet, Massachusetts Institute of Technology and with
57 example of the subsurface nucleation of a crystal defect during simulated
permission from . © 2005 American Physical Society.)
nanoindentation of Al is shown in Fig. 5. Some of these simulation efforts
have led to considerable quantitative insight into the process of
the point of departure is denoted by an arrow. In a great variety of 57,58
homogeneous dislocation nucleation . However, the connection
41,43-45 40,42
materials, including metals , alloys , between experiment and simulation remains somewhat tenuous, as
38,49,50 46,51 simulation time scales are generally several orders of magnitude smaller
intermetallics , and ceramics , this yield point has been
than in experiments, while the indenter radius usually simulated (~1-10
found to occur at a discontinuity in the P-h curve, sometimes referred
nm) is much smaller than achieved in experiments (~50-500 nm). A
to as a ‘pop-in’ event. During this event the indenter travels without a
measured increase in applied load (for a load-controlled experiment), 59
multiscale modeling study by Knap and Ortiz demonstrates the critical
or the load is rapidly released at a constant displacement (for a effect of this disparity. A 7 nm indenter radius led to homogeneous
43,52 dislocation nucleation that was easily observed as a discontinuity in the
displacement-controlled experiment) . Indentations performed to
lower, subcritical loads, usually exhibit ideal reversibility and leave no P-h curve, while a larger and more realistic tip radius of 70 nm led to
trace of a residual impression on the specimen surface, confirming considerable dislocation activity without a pop-in event in the P-h curve.
that the first pop-in marks the onset of irreversible flow. Furthermore, The implication is that, in experiments, significant undetectable dislocation
the burst-like character of the first pop-in suggests that strain is activity may precede the first pop-in, which would suggest that the pop-in
accommodated by an abrupt avalanche of atomic activity beneath event is not, in fact, a homogeneous dislocation nucleation event, but
the indenter, such as might be expected for activation of a dislocation rather a heterogeneous process of, for example, dislocation source
source. activation or multiplication. In this case, the nominally ‘elastic’ portion of
The events that control the first pop-in event during the P-h curve, which fits the predictions of Hertzian contact theory,

nanoindentation have been the subject of considerable research and apparently

61
Fig. 6 A series of in situ TEM micrographs from the work of Minor et al. , illustrating the response of an Al grain to nanoindentation. In the first
frame, the indenter has not yet made contact with the grain below it. The second frame illustrates an elastic contact with strain contours. The
third frame is taken just after the first plastic deformation has occurred and lattice dislocations are observed. The fourth frame illustrates the
61
development of a complex dislocation network within the Al grain. (Reproduced with permission from . © 2004 Materials Research Society.)

MAY 2006 | VOLUME 9 | NUMBER 5


REVIEW FEATURE Nanoindentation studies of materials

Fig. 7 demonstrates this effect using a statistical approach, where the


same indentation was repeated numerous times at many different
locations on the metal surface. The cumulative distribution of pop-in
loads measured in this exercise is plotted in Fig. 7, and can be seen to
shift as the test temperature is increased; higher temperatures generally
lower the load required to initiate yield, a result that conforms with
expectations for a thermally activated deformation mechanism.
Statistical data such as those in Fig. 7 are also amenable to deeper
65,66
quantitative analysis , and have been used to assess the activation
energy and activation volume for the yield event in Pt. In the case of a
single-crystal Pt surface prepared by electrochemical polishing, such
analysis has revealed very low values for both the characteristic energy
and volume of the rate-limiting process for yield. This suggests that the
first pop-in may be associated with a heterogeneous event, such as the
activation of a dislocation source from preexisting surface features or
subcritical dislocations.

Development of dislocation networks


Fig. 7 High-temperature nanoindentation data illustrating the cumulative
Beyond the initial yield point in nanoindentation, additional pop-in events
distribution of measured yield loads for many identical tests performed
65 are usually observed at higher loads (see, for example, Fig. 3a for
on the same surface of Pt(110) . As the test temperature increases, the
loads required to cause the first pop-in are shifted to lower values, an indentation of Pt), which are associated with further dislocation motion,
effect which is consistent with a model for thermally-activated, stress- multiplication, and the evolution of a complex defect structure. Such
assisted deformation (green lines).
structural evolution is generally quite complicated and varies significantly
between materials of different structure, chemistry, and crystallographic
would contain some amount of superimposed plastic orientation. Because of these complexities, analysis of the P-h curve
displacement too subtle to detect. alone seems unlikely to yield a detailed physical picture, although imaging
Nanoindentation experiments have not provided nearly as much of the indented volume, by a variety of techniques, has provided some
atomic-level insight into incipient plasticity as have simulations, although
76-85
new experimental techniques that have emerged in the past few years very useful insights . For example, Rojo and
27,60-67
are quickly remedying this situation . Minor, Stach, 81-83

60-64
coworkers have used scanning tunneling microscopy on indented
surfaces of (100) Au to reveal the mechanisms by which dislocation
Morris, and coworkers have pioneered the development of a
activity relocates matter around the impression site. Fig. 8 shows two
hybrid technique incorporating a nanoindenter into a transmission
images of the same surface, after successive indentations placed at the
election microscope (TEM), providing in situ images of structural
same position. The material ejected during indentation ‘piles up’ in a
evolution beneath an indentation. Fig. 6 shows a series of images
61
from their work on Al , where the initial contact with a
dislocation-free grain can be observed, followed by the initiation
of dislocation activity and the interaction of the dislocation
60
structure with grain boundaries. In a related study, Minor et al.
showed that the first dislocation activity corresponded closely
with a significant discontinuity in the P-h curve.
Further quantitative details on the nature of incipient plasticity have
also recently been revealed through the use of high-temperature
nanoindentation. A number of authors have attempted to extend
nanoindentation methods to nonambient conditions in the past (a) (b)
14,15,34,68-73
decade , although in general these efforts did not yield Fig. 8 Scanning tunneling microscope images of a surface of Au(100) after
nanoindentation illustrating the redistribution of material around the
sufficient resolution or stability to study the first pop-in in great detail.
indentation site. Between frames (a) and (b), the impression was enlarged by
Some recent approaches have achieved the required conditions for
indenting on the same location to a higher load. In the dashed box, a
74,75 stationary screw dislocation can be identified, as well as evidence for the
temperatures up to a few hundred degrees centigrade , and have
motion of another screw dislocation to create a surface ledge. (Reprinted
shown that the first pop-in is thermally activated for at least one metal –
courtesy of E. Carrasco, O. R. de la Fuente, M. A. González, and J. M. Rojo
Pt27,65,66. 83
and with permission from . © 2003 American Physical Society.)

36 MAY 2006 | VOLUME 9 | NUMBER 5


Nanoindentation studies of materials REVIEW FEATURE

(a) (b)

Fig. 9 Atomic force micrographs of nanoindentations in an Al-Fe-Gd metallic glass. Around the three-fold symmetric impression, a series of semicircular
shear band traces can be seen. Image (a) illustrates the surface produced after a slow indentation (1 nm/s), while image (b) shows that resulting from a
92
rapid indentation (100 nm/s). (Reproduced courtesy of W. H. Jiang and M. Atzmon and with permission from . © 2003 Materials Research Society.)

series of geometric terraces reflective of the underlying crystal One of the most quantitatively fruitful approaches to the study of
symmetry, and the termination points of some individual screw shear bands by nanoindentation has been through the examination of
dislocations can be discerned. The geometry of such surface rate and temperature effects on the P-h curve. Fig. 10 shows a set of
traces (and the geometry of the impression itself) can be linked loading curves from the same Pd-Ni-P metallic glass described above,
to the dislocation slip systems of the crystal. The motion of but acquired at different applied loading rates spanning four decades. A
individual dislocations has been captured by comparing frames transition in the pop-in behavior is observed, with discontinuities
83 becoming fewer and less pronounced as the indentation rate rises. This
such as shown in Fig. 8a and 8b .
behavior is typical of many different metallic glasses with widely

Mechanical instabilities
In crystalline metals, the motion and mutual interaction of dislocations
gives rise to work hardening mechanisms that encourage stable plastic
flow. In contrast, in amorphous metals (or metallic glasses) there are
no dislocations per se, and plastic deformation is inherently unstable,
occurring in bursts of highly localized strain referred to as ‘shear
banding’ events. For example, the Berkovich impression shown in
Fig. 9a was made on an amorphous Al alloy, and the presence of shear
bands is readily observed as a series of steps around the periphery of
the indentation.
Nanoindentation testing has recently become an important tool
for fundamental studies of shear banding in metallic glasses, owing
to its ability to resolve individual shear events under well-controlled
86
conditions . The P-h curve for amorphous Pd-Ni-P shown in Fig. 3b
is a prototypical response for a metallic glass indented slowly. The
loading portion of this curve exhibits an impressive series of pop-in
87-89
events and, from the earliest such observations , it has been
believed that each of these events correlates with a single shear
90
banding event beneath the indenter. Moser et al. recently
performed the first in situ observation of shear band formation during
Fig. 10. Typical P-h curves for the loading portion of a nanoindentation
nanoindentation in a scanning electron microscope and directly 91
test on a Pd-Ni-Cu-P metallic glass , offset from one another by 50 nm
correlated pop-in events in the P-h curve with the appearance of
for clarity of presentation. The number and size of pop-in discontinuities
individual surface steps at the periphery of the indenter. in the curves increases significantly as the indentation rate is lowered.
MAY 2006 | VOLUME 9 | NUMBER 5
REVIEW FEATURE Nanoindentation studies of materials

beneath the indenter. The instability that gives rise to discontinuities


in this case results from the interaction of lattice dislocations with
solute atoms, which leads to instantaneous negative strain rate
sensitivity and localization of flow. Chinh et al. have used data such
as these to study the statistics of flow serration, as well as the
influence of solute concentration on the instability.

Phase transformations
Many materials undergo phase transformations when subjected to
large hydrostatic stresses, and the pressure beneath a nanoindenter
is generally quite high (on the order of several gigapascals). In the
case of some diamond-cubic semiconductors – including Si and Ge –
the mean contact pressure of hardness indentations closely matches
Fig. 11 Load-displacement data acquired during indentation of Al-Mg alloys 105-107
103
the critical pressure to trigger a structural transformation . The
by Chinh et al. illustrating numerous serrations resulting from mechanical
instability. The concentration dependence of this phenomenon can also be interest in indentation-induced phase transformations has been
observed by comparing the different curves. (Reproduced courtesy of steadily increasing for decades, and the wide adoption of
N. Q. Chinh, F. Csikor, Zs. Kovács, and J. Lendvai and with
103 nanoindentation techniques has spurred further study.
permission from . © 2000 Materials Research Society.)
The P-h curve shown in Fig. 3c is a typical example of the indentation
different chemical compositions, and has been documented by 29
response seen in Si for relatively high loads . When sharp indentation
a number of authors working with a variety of different impressions are examined ex situ, thin sheets of material extruded
28,74,89,91-95
nanoindenters . The rate effect around the periphery of the indenter can be observed, as shown in Fig.
is also associated with a shift 29,108,109
12 . This extensive plastic flow has been interpreted as
in the distribution of shear band traces on the surface of the evidence for the transformation from diamond cubic Si to the metallic
material. Figs. 9a and 9b compare indentations performed at (and malleable) β-Sn phase under the pressure of the
relatively lower and higher rates, respectively, from which it is indentation
108,109
. More direct evidence for a semiconducting/metallic
clear that faster indentations produce a higher number density transition under the indenter has also been provided by in situ electrical
of more closely-spaced shear bands. 106,107,110,111
measurements performed by several authors .
Quantitative analysis of these kinds of P-h data has led to significant
The sequence of events during the unloading portion of an
insight into the process of shear localization in metallic glasses. For
indentation are less clearly resolved, although the problem has been
74
example, a collection of high-temperature experiments was used to 29,84,110,112-122
studied extensively . Various ex situ TEM and Raman
identify a ‘critical’ submicron-scale size for shear banding, below which
spectroscopy studies have been conducted, and an impressive number
strain is localizing and the shear band is growing, and above which strain
of polymorphs appear to be involved – eight or more in Si, seven or
has completely localized into the now rapidly propagating shear band. In
more in Ge – depending upon the indenter shape, applied load, and rate
complementary work combining indentation with ex situ TEM
of unloading. These effects have been summarized by Domnich and
observations, other researchers have studied deformation-induced
122
structural evolution and nanocrystallization events in metallic Gogotsi , who extensively reviewed the literature up to 2001.

86,96-98
glasses . These various studies represent initial efforts in the
exploration of a very complex topic, and there is not yet broad agreement
as to the mechanisms of shear localization in metallic glasses, or how the
glass structure and structural evolution impacts the deformation.
Nonetheless, these kinds of studies clearly represent an avenue for
significant scientific advances in the future.

Fig. 12. Scanning electron micrograph of a cube-corner geometry indentation


99-104
on Si(100) illustrating a copious amount of sheet-like material that has been
Shear banding in amorphous metals is just one example of a mechanical instability that can be studied using nanoindentation, and similar instabilities in other materials are amenable to the same kinds
extruded out from beneath the indenter; this is taken as evidence of pressure-
of experimental approaches. A good example of this is the so-called ‘jerky flow’ of solid solution alloys, which has recently been considered by Chinh et al. using depth-sensing indentation.
29
induced transformation to a plastic metallic state during indentation .
29
(Reproduced with permission from . © 2005 Elsevier Ltd.)
Sample data for Al-Mg alloys acquired by this group are shown in Fig.
11, which shows P-h curves with various degrees of flow serration

38 MAY 2006 | VOLUME 9 | NUMBER 5


Nanoindentation studies of materials REVIEW FEATURE

123
Fig. 13 Dark-field TEM image showing a cross-sectional view of indented Si . The regions labeled 1-4 are amorphous and the bands 2-
123
4 are aligned with the expected slip planes in diamond-cubic Si. (Reproduced with permission from . © 2002 Taylor and Francis, Ltd.)

The pop-out phenomenon observed during unloading in Fig. 3c is Summary


attributed to the reversion of the high-pressure β-Sn phase to metastable Nanoindentation presents considerable potential for understanding
121 discrete atomic rearrangements under stress, ranging from phase
rhombohedral or body-centered cubic phases . The transformation
carries an attendant increase in volume that can do mechanical work transformations in small volumes of material to the motion and formation
against the indenter tip, producing the pop-out effect. of individual defects such as lattice dislocations or shear bands. By itself,
The presence of amorphous regions after indentation of Si has been nanoindentation often allows the detection of such events under a
84,114-118,123 reasonably well-defined stress state. When the technique is augmented
widely observed in TEM studies , and represents an
interesting departure from the equilibrium phase diagram. Although by additional capabilities, such as a high-temperature testing stage, in
several authors have suggested that deviatoric stresses or plastic situ electron microscopy, or ex situ measurements by scanning probe
deformation may impact the evolution of metastable phases under the microscopy or Raman spectroscopy, near-atomic level details of
indenter, this remains an area where additional work is required. For deformation physics can be deduced. As these techniques become more
123 refined and their synergy with nanoindentation equipment improves, it is
example, Tachi Suprijadi et al. have used ex situ cross-sectional TEM
to observe the structure of indented Si, and identified amorphous bands expected that the resulting experimental data will play a key role in the
aligned with the expected slip systems for the diamond-cubic structure development of consistent theories of material behavior at the nanoscale.
(Fig. 13). This hints at the involvement of plastic flow in the transformation
sequence beneath the indenter, which is not captured by static
equilibrium arguments alone. Thus, nanoindentation may offer the
possibility of studying nonequilibrium phase transformations (so-called Acknowledgments
‘driven’ transformations) with new levels of consistency and control.
CAS acknowledges the support of the US Office of Naval
Research (contract N00014-04-1-0669).

REFERENCES
1. Timoshenko, S. P., History of Strength of Materials, McGraw-Hill, New York, 7. Chollacoop, N., et al., Acta Mater. (2003) 51, 3713
(1953)
8. Bucaille, J. L., et al., Acta Mater. (2003) 51, 1663
2. Johnson, K. L., Contact Mechanics, Cambridge University Press, UK (1985)
9. Kucharski, S., and Mróz, Z., Mater. Sci. Eng. A (2001) 318, 65
3. Oliver, W. C., and Pharr, G. M., J. Mater. Res. (1992) 7, 1564
10. Ma, D., et al., J. Appl. Phys. (2003) 94, 288
4. Oliver, W. C., and Pharr, G. M., J. Mater. Res. (2004) 19, 3
11. Fischer-Cripps, A. C., Mater. Sci. Eng. A (2004) 385, 74
12. Oyen, M. L., and Cook, R. F., J. Mater. Res. (2003) 18, 139
5. Bhushan, B., In Handbook of Micro/Nano Tribology, Bhushan, B., (ed.) CRC Press,
Boca Raton, Florida (1999), 433
13. Storåkers, B., and Larsson, P.-L., J. Mech. Phys. Solids (1994) 42, 307
6. Dao, M., et al., Acta Mater. (2001) 49, 3899

MAY 2006 | VOLUME 9 | NUMBER 5 39


REVIEW FEATURE Nanoindentation studies of materials

14. Lucas, B. N., and Oliver, W. C., Metall. Mater. Trans. A (1999) 30, 601 69. Syed-Asif, S. A., and Pethica, J. B., Philos. Mag. A (1997) 76, 1105
15. Takagi, H., et al., Philos. Mag. (2003) 83, 3959 70. Smith, J. F., and Zheng, S., Surf. Eng. (2000) 16, 143
16. Suresh, S., and Giannakopoulos, A. E., Acta Mater. (1998) 46, 5755 71. Beake, B. D., and Smith, J. F., Philos. Mag. A (2002) 82, 2179
17. Carlsson, S., and Larsson, P.-L., Acta Mater. (2001) 49, 2179 72. Beake, B. D., et al., Z. Metallkd. (2003) 94, 798
18. Carlsson, S., and Larsson, P.-L., Acta Mater. (2001) 49, 2193 73. Volinsky, A. A., et al., J. Mater. Res. (2004) 19, 2650
19. Swadener, J. G., et al., J. Mater. Res. (2001) 16, 2091 74. Schuh, C. A., et al., Acta Mater. (2004) 52, 5879
20. Cheng, Y.-T., and Cheng, C.-M., Mater. Sci. Eng. R (2004) 44, 91 75. Schuh, C. A., et al., J. Mater. Res. (2006) 21, 725
21. Fischer-Cripps, A. C., Nanoindentation, Springer, New York, USA (2002) 76. Rabe, R., et al., Thin Solid Films (2004) 469-470, 206
22. VanLandingham, M. R., J. Res. Natl. Inst. Stand. Technol. (2003) 108, 249 77. Viswanathan, G. B., et al., Acta Mater. (2005) 53, 5101
23. Thurn, J., and Cook, R. F., J. Mater. Res. (2002) 17, 1143 78. Nibur, K. A., and Bahr, D. F., Scripta Mater. (2003) 49, 1055
24. VanLandingham, M. R., et al., Measurement Sci. Technol. (2005) 16, 2173 79. Gaillard, Y., et al., Philos. Mag. Lett. (2003) 83, 553
25. McElhaney, K. W., et al., J. Mater. Res. (1998) 13, 1300 80. Gaillard, Y., et al., Acta Mater. (2003) 51, 1059
26. Meneve, J. L., et al., Appl. Surf. Sci. (1996) 101, 64 81. de la Fuente, O. R., et al., Phys. Rev. Lett. (2002) 88, 036101
27. Lund, A. C., et al., Appl. Phys. Lett. (2004) 85, 1362 82. de la Fuente, O. R., et al., Philos. Mag. (2003) 83, 485
28. Schuh, C. A., and Nieh, T. G., Acta Mater. (2003) 51, 87 83. Carrasco, E., et al., Phys. Rev. B (2003) 68, 180102
29. Jang, J. I., et al., Acta Mater. (2005) 53, 1759 84. Page, T. F., et al., J. Mater. Res. (1992) 7, 450
30. Gane, N., and Bowden, F. P., J. Appl. Phys. (1968) 39, 1432 85. Stelmashenko, N. A., et al., Acta Metall. Mater. (1993) 41, 2855
31. Gerberich, W. W., et al., Acta Metall. Mater. (1995) 43, 1569 86. Schuh, C. A., and Nieh, T. G., J. Mater. Res. (2004) 19, 46
32. Mann, A. B., and Pethica, J. B., Appl. Phys. Lett. (1996) 69, 907 87. Wang, J. G., et al., J. Mater. Res. (2000) 15, 798
33. Corcoran, S. G., et al., Phys. Rev. B (1997) 55, R16057 88. Golovin, Yu. I., et al., Scripta Mater. (2001) 45, 947
34. Bahr, D. F., et al., J. Mater. Res. (1999) 14, 2269 89. Schuh, C. A., et al., J. Mater. Res. (2002) 17, 1651
35. Mann, A. B., Philos. Mag. A (1999) 79, 577 90. Moser, B., et al., Adv. Eng. Mater. (2005) 7, 388
36. Tymiak, N. I., et al., Acta Mater. (2001) 49, 1021 91. Schuh, C. A., et al., Philos. Mag. (2003) 83, 2585
37. Li, J., et al., Nature (2002) 418, 307 92. Jiang, W. H., and Atzmon, M., J. Mater. Res. (2003) 18, 755
38. Chiu, Y. L. , and Ngan, A. H. W., Acta Mater. (2002) 50, 1599 93. Concustell, A., et al., J. Mater. Res. (2005) 20, 2719
39. Shibutani, Y., and Koyama, A., J. Mater. Res. (2004) 19, 183 94. Nieh, T. G., et al., Intermetallics (2002) 10, 1177
40. Gerberich, W. W., et al., Acta Mater. (1996) 44, 3585 95. Zhang, G. P., et al., Scripta Mater. (2005) 52, 1147
41. Michalske, T. A., and Houston, J. E., Acta Mater. (1998) 46, 391 96. Kim, J.-J., et al., Science (2002) 295, 654
42. Bahr, D. F., et al., Acta Mater. (1998) 46, 3605 97. Jiang, W. H., et al., J. Appl. Phys. (2003) 93, 9287
43. Kiely, J. D., and Houston, J. E., Phys. Rev. B (1998) 57, 12588 98. Kramer, M. J., et al., J. Non-Cryst. Solids (2005) 351, 2159
44. Suresh, S., et al., Scripta Mater. (1999) 41, 951 99. Bérces, G., et al., J. Mater. Res. (1998) 13, 1411
45. Gouldstone, A., et al., Acta Mater. (2000) 48, 2277 100. Chinh, N. Q., et al., Mater. Sci. Eng. A (2002) 324, 219
46. Schuh, C. A., and Lund, A. C., J. Mater. Res. (2004) 19, 2152 101. Bérces, G., et al., Acta Mater. (1998) 46, 2029
47. Bahr, D. F., and Vasquez, G., J. Mater. Res. (2005) 20, 1947 102. Bérces, G., et al., J. Mater. Res. (1998) 13, 1411
48. Pethica, J. B., and Tabor, D., J. Adhesion (1982) 13, 215 103. Chinh, N. Q., et al., J. Mater. Res. (2000) 15, 1037
49. Chiu, Y. L., and Ngan, A. H. W., Acta Mater. (2002) 50, 2677 104. Chinh, N. Q., et al., J. Mater. Res. (2004) 19, 31
50. Wang, W., et al., Acta Mater. (2003) 51, 6169 105. Gerk, A. P., and Tabor, D., Nature (1978) 271, 732
51. Wang, X., and Padture, N. P., J. Mater. Sci. (2004) 39, 1891 106. Clarke, D. R., et al., Phys. Rev. Lett. (1988) 60, 2156
52. Warren, O. L., et al., Z. Metallkd. (2004) 95, 287 107. Gridneva, I. V., et al., Phys. Status Solidi A (1972) 14, 177
53. Kelchner, C. L., et al., Phys. Rev. B (1998) 58, 11085 108. Pharr, G. M., et al., J. Mater. Res. (1991) 6, 1129
54. Lilleodden, E. T., et al., J. Mech. Phys. Solids (2003) 51, 901 109. Kailer, A., et al., J. Appl. Phys. (1997) 81, 3057
55. Zhu, T., et al., J. Mech. Phys. Solids (2004) 52, 691 110. Mann, A. B., et al., J. Mater. Res. (2000) 15, 1754
56. Miller, R. E., et al., Acta Mater. (2004) 52, 271 111. Bradby, J. E., et al., Phys. Rev. B (2003) 67, 085205
57. Van Vliet, K. J., et al., Phys. Rev. B (2003) 67, 104105 112. Zarudi, I., et al., J. Mater. Res. (2003) 18, 758
58. Ronald, R. E.., J. Mech. Phys. Solids (2004) 52, 1507 113. Zarudi, I., et al., Appl. Phys. Lett. (2003) 82, 1027
59. Knap, J., and Ortiz, M., Phys. Rev. Lett. (2003) 90, 226102 114. Zarudi, I., et al., Appl. Phys. Lett. (2003) 82, 874
60. Minor, A. M., et al., Appl. Phys. Lett. (2001) 79, 1625 115. Ge, D. B., et al., J. Appl. Phys. (2003) 93, 2418
61. Minor, A. M., et al., J. Mater. Res. (2004) 19, 176 116. Lloyd, S. J., et al., J. Mater. Res. (2001) 16, 3347
62. Ohmura, T., et al., J. Mater. Res. (2004) 19, 3626 117. Lloyd, S. J., et al., Proc. R. Soc. A (2005) 461, 2521
63. Soer, W. A., et al., Acta Mater. (2004) 52, 5783 118. Bradby, J. E., et al., Appl. Phys. Lett. (2000) 77, 3749
64. Minor, A. M., et al., Philos. Mag. (2005) 85, 323 119. Patriarche, G., et al., J. Appl. Phys. (2004) 96, 1464
65. Schuh, C. A., et al., Nature Mater. (2005) 4, 617 120. Pharr, G. M., et al., Scripta Mater. (1989) 23, 1949
66. Mason, J. K., et al., Phys. Rev. B (2006) 73, 054102 121. Domnich, V., et al., Appl. Phys. Lett. (2000) 76, 2214
67. Tymiak, N. I., et al., Acta Mater. (2004) 52, 553 122. Domnich, V., and Gogotsi, Y., Exp. Methods Phys. Sci. (2001) 38, 355
68. Farber, B. Ya., et al., Philos. Mag. A (1998) 78, 671 123. Tachi Suprijadi, M., et al., Philos. Mag. Lett. (2002) 82, 133

40 MAY 2006 | VOLUME 9 | NUMBER 5

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy