You are on page 1of 17

This article was downloaded by: [Pennsylvania State University]

On: 12 August 2013, At: 13:21


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Philosophical Magazine
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tphm20

Determination of fracture toughness


of bulk materials and thin films by
nanoindentation: comparison of
different models
a
Kirsten Ingolf Schiffmann
a
Fraunhofer Institut für Schicht- und Oberflächentechnik,
Bienroder Weg 54E, Braunschweig, Germany
Published online: 28 Jun 2010.

To cite this article: Kirsten Ingolf Schiffmann (2011) Determination of fracture toughness of
bulk materials and thin films by nanoindentation: comparison of different models, Philosophical
Magazine, 91:7-9, 1163-1178, DOI: 10.1080/14786435.2010.487984

To link to this article: http://dx.doi.org/10.1080/14786435.2010.487984

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Philosophical Magazine
Vol. 91, Nos. 7–9, 1–21 March 2011, 1163–1178

Determination of fracture toughness of bulk materials and thin films


by nanoindentation: comparison of different models
Kirsten Ingolf Schiffmann*

Fraunhofer Institut für Schicht- und Oberflächentechnik, Bienroder Weg 54E,


Braunschweig, Germany
(Received 9 October 2009; final version received 19 May 2010)
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

Fracture toughness is an important technical material parameter, which,


for bulk materials, can be determined by macroscopic crack extension
experiments. For thin coatings, this classical method is not applicable.
However, during the last 25 years, techniques have been developed which
allow fracture toughness determination from indentation experiments by
measuring crack lengths generated during controlled indentation. For its
evaluation, a number of different models have been proposed which partly
yield significant different results. In the present work, cube corner
nanoindentation was performed using a Hysitron Triboscope on a
number of different bulk (single crystal silicon, fused silica, single crystal
sapphire) and thin film materials (Si-DLC, aC, SnO2, ZnO:Al, ITO). Crack
lengths were measured by AFM-based methods and six different models,
including a crack-energy-based and a thin film model, were used for
evaluation. The models are validated by comparing mean deviations of
experimental and literature values. Best agreement over all was found for
the two models of Niihara and co-workers for halfpenny [J. Mater. Sci.
Lett. 1 (1982) p.13] and Palmquist cracks [J. Mater. Sci. Lett. 2 (1983)
p.221]. The energy-based approach of Li et al. [Acta Mater. 45 (1997)
p.4453] also yields good results, which is remarkable since it does not
contain any empirical constants. The crack-energy-based model, however,
could not be applied to all materials since it is based on pop-ins in the
indentation curve, which were not observed for all materials. A discussion
of possible sources of error is given.
Keywords: fracture toughness; nanoindentation; cube corner indenter; bulk
materials; thin films

1. Introduction
Fracture toughness KIC is a technical material parameter which describes the
resistance of a material against growth of pre-existing cracks. It is of importance for
all technical applications of materials where strong mechanical loading occurs. For
bulk materials, it can be determined by macroscopic crack extension experiments.
These methods are already standardised by ISO [1,2].

*Email: kirsten.schiffmann@ist.fraunhofer.de

ISSN 1478–6435 print/ISSN 1478–6443 online


ß 2011 Taylor & Francis
DOI: 10.1080/14786435.2010.487984
http://www.informaworld.com
1164 K.I. Schiffmann

A new field of applications is thin coatings, e.g. protective coatings in automotive


or hard disk industry, low-k dielectric materials in microelectronic industry or solar
control coatings on architectural glass. For thin coatings, the classical methods of
fracture toughness determination are not applicable. However, during the last
decades, techniques have been developed which allow fracture toughness determi-
nation from indentation experiments by measuring crack lengths generated during
controlled indentation. Between 1974 and 1989, nearly 20 different equations were
developed to calculate KIC from Vickers indentation experiments. In 1989, Ponton
and Rawling [3,4] extensively reviewed the (at that time) state-of-the art and made a
validation of 19 different expressions for KIC. However, those first Vickers-based
methods were applied only to bulk materials – typically, loads were in the range
50–150N and crack lengths were evaluated by optical methods.
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

Since that time, some of these equations have become more accepted than others,
some have been improved and, in addition, some new approaches have been
published, e.g. a thin film approach and a crack-energy-based model. Nowadays,
nanoindentation generally uses three-sided pyramidal indenters, e.g. Berkovich or
cube-corner, loads are in the range 1–500 mN and crack lengths of 0.5–10 mm are
measured by SEM or AFM-based methods. Therefore, a reappraisal of actually used
theoretical expressions in conjunction with the actual experimental techniques and
applications now seems appropriate. The aim of this work is to apply some
frequently used models to a number of different bulk and thin film materials, and
validate the results by comparison with literature material data, where available.

2. Theoretical
The fracture toughness KIC is defined by the following equation:
pffiffiffiffiffiffiffiffiffi
KIC ¼ 2E , ð1Þ
where KIC ¼ critical stress intensity factor during mode I type loading (¼pure
tension), E ¼ Young’s modulus,  ¼ specific fracture energy, i.e. work for creating a
crack of unit area.
Evans et al. [5] and Lawn et al. [6] developed an equation which allows the
determination of KC from indentation experiments1 by measuring the length of
cracks emanating from the edges of the indent impression. It has the form:
 1=2
E P
KC ¼ A , ð2Þ
H c3=2
where: E ¼ Young’s modulus, H ¼ hardness, P ¼ load, c ¼ crack length measured
from the centre of the indent to the crack tip. A is an empirical constant which
depends on the indenter and the crack geometry.
Cook and Pharr [8] have classified different crack geometries: halfpenny cracks,
radial or Palmquist cracks, median cracks, conical cracks and lateral cracks, where
the first four are vertical cracks, while the latter are horizontal cracks. From a top
view of the impression, the halfpenny, Palmquist and median cracks cannot
clearly be distinguished and, therefore, assumptions are often made as to what
Philosophical Magazine 1165

type of crack exists. In most cases, it is believed that cracks created by indentation
are halfpenny or Palmquist cracks.
From the work of Evans and Lawn [5,6], a number of modifications of
Equation (2) have been developed. The selection of models compared in this work is
summarised in Table 1. Jang and Pharr [9] developed an analytical expression which
correlates constant A with the indenter opening angle (Equation (3)). Niihara and
co-workers [10,11] modified Equation (2) by introducing a different exponent of
(E/H), namely 2/5 instead of ½. They gave two expressions, one for halfpenny cracks
with c/a  2.5 (Equation (4)) and one for Palmquist cracks with a low crack-to-indent
size ratio 0.25  l/a  2.5 (Equation (5)). Equation (6) by Laugier and Ouchterlony
[12,13] gives a combined dependence on a, l and c, and a different power of (E/H)2/3.
Thurn and Cook [14] modified the basic expression for thin coatings. They assumed
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

through-thickness cracks, i.e. vertical extension of the cracks, are limited by the film
thickness. Additionally, residual stress of the coating, which may increase or decrease
fracture toughness, is considered (Equation (7)). Finally, Li et al. [15] made a new
approach starting directly from Equation (1) and calculating the fracture toughness
by measuring the fracture energy from pop-in events of the indentation curve and by
estimating the area F of the crack (Equation (8)). Fracture energy can be calculated
from the area U in Figure 1a. Other authors also used area U0 [16]. The crack area
depends on crack geometry. In our cases, generally, very short halfpenny cracks are
assumed where the crack area is estimated as shown in Figure 1c as the difference
between the quarter circle area c2/4 and the triangle area a2/2.

3. Experimental
All experiments were performed with a Hysitron Triboscope nanoindenter attached
to a DI 3100 AFM platform. The maximum possible load of this indenter is 32 mN.
This has the effect that, for some materials, the crack threshold is not reached or the
cracks are very short and difficult to detect. To reduce fracture threshold, a cube
corner diamond indenter was used, with a tip radius of 250 nm. For some special
measurements, a tip with a 30 nm tip radius was also used. Tip radii were determined
by SEM imaging. The indentation experiments are load-controlled and consist of 5 s
loading to the maximum load of 32 mN, 10 s hold time and 5 s unloading. Each test
was repeated at least five times, which allowed 15 crack lengths to be evaluated.
The triboscope allowed in situ imaging of the created impressions and cracks by
scanning the indenter tip over the surface with a minimum load of some micro
Newtons. The parameters c, l and a were determined from these AFM-like images.
The sensitivity of detecting cracks could be strongly enhanced by image processing
using filter techniques, such as contrast enhancement, edge enhancement or high-
pass filtering. For comparison, a selection of indentations was evaluated by
conventional AFM using a sharp silicon tip with 10 nm tip radius and by scanning
electron microscopy (SEM).
As samples, four bulk materials were tested: single crystal silicon (100), fused
silica, single crystal sapphire (0001) and highly polished polycrystalline steel. In
addition, a number of different thin film materials were measured: TiN, Si-DLC, aC,
SnO2, ZnO:Al, ITO. Details of these samples are summarised in Table 2.
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

1166

Table 1. Models for fracture toughness determination from indentation crack measurements. E ¼ Young‘s modulus, H ¼ hardness, P ¼ load,
 ¼ Poisson ratio,  ¼ half opening angle of indenter ¼ 35.3 (cube corner), 65.3 (Berkovich), c ¼ crack length from indent centre to crack tip,
l ¼ crack length from indent corner to crack tip, a ¼ indent size from centre to corner (a þ l ¼ c), tf ¼ film thickness,  R ¼ residual stress, U ¼ crack
energy from pop-in (Figures 1a and b), F ¼ crack area ¼ c2/4  a2/2 (Figure 1c).

Eq. Model Constants Indenter/crack geometry Source


2=3
 E 1=2 P
3 KC ¼ A H c3=2 A ¼ 0:0352
1 ðcot Þ Three-sided pyramids & cone/Halfpenny cracks [9]
 E 2=5 P
4 KC ¼A H A ¼ 0.033 Vickers/Short halfpenny cracks [10,11]
P
 E 2=5 c3=2
5 KC ¼A H pffi A ¼ 0.009 Vickers/Short Palmquist cracks [10,11]
a1=2 a E l2=3 P
6 KC ¼ x l H c3=2
x ¼ 0.015(Vickers) Vickers & Berkovich/Palmquist cracks [12,13]
K.I. Schiffmann

x ¼ 0.016(Berkovich)
 E
 1=3 P
7 KC ¼ tf0 H c1=2
þ 0R t1=2
f 0 ¼ constant Cube corner &Vickers/Through coating cracks [14]
0 ¼ 0.0061 (cube)
qffiffiffiffiffiffiffiffiffiffiffi 0 ¼ 0.00053 (Vickers)
E U
8 KC ¼ 12 F
Cube corner/Any crack shape [15]
Philosophical Magazine 1167

(a)

Load
U„

Displacement
(b)
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

Load

Displacement

(c)

c
a
a

Fracture area

Figure 1. Schematic load–displacement curves with pop-in due to fracture. (a) Area U (or U0 )
is used by different authors to calculate fracture energy. (b) In the case of ‘‘sliding’’ pop-in, the
fracture area may be computed by interpolation. (c) Crack area F is computed from the
quarter circle c2/4 minus triangle a2/2.

Table 2. Materials subjected to nanoindentation fracture tests. tf ¼ film thickness,


E ¼ Young’s modulus, H ¼ hardness, hmax ¼ maximum indentation depth at 32 mN.

No. Material tf [mm] Substrate H [GPa] E [GPa] hmax/tf

1 Single crystal silicon (100) – – 12 170 –


2 Fused silica (amorphous SiO2) – – 9 72 –
3 Single crystal sapphire Al2O3 (0001) – – 31 420 –
4 ASP 23 tool steel (polished) – – 11 310 –
5 TiN (O, Ar, C 5 1 at%) 2.5 Steel 26 550 –
6 Si-DLC (diamond-like carbon, 4.7 Steel 15.5 150 0.2
Si ¼ 9 at%, H ¼ 18 at%)
7 aC (amorphous, H-free carbon) 0.45 Si 35.6 350 2
8 SnO2 (tempered & not tempered) 1.2 Glass 14 131 1
9 ZnO:Al (Al ¼ 1.5 at%) 0.8 Glass 10 110 1.9
10 In2O3:Sn (ITO) þ 40 nm SiOxNy 1.2 Glass 12 133 1
toplayer
1168 K.I. Schiffmann

4. Results
Initially, indentations using two different cube corner indenters, one with a 250 nm
tip radius and one with a 30 nm tip radius, were evaluated. It is well known that the
opening angle of the indenter has a significant effect on the fracture threshold [9,17],
such that the smaller the opening angle, the lower the crack threshold. One may
expect that the same is true for tip radius, but the opposite was found. The tests were
performed on sapphire, Si-DLC and silicon. It is found that, using the sharp
indenter, no radial cracks could be generated on sapphire or Si-DLC, while, with the
blunt tip, short cracks could be created at the same load. In the case of sapphire, the
sharp indenter produced slip planes running away from the edges of the imprint
along preferential crystallographic directions (Figure 2a). Even if the indenter was
oriented with the corners showing into the direction of slip planes, no cracks could be
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

enforced. In the case of Si-DLC coating, so-called picture frame cracks occurred
instead of radial cracks (Figure 2b). Both the sharp and blunt indenter only created
cracks on silicon (Figure 2c). The reason for this effect is unclear. A larger tip radius
will reduce the contact pressure at a given load which will increase the threshold
force for the onset of plasticity, but also the driving force for crack formation should
be reduced. Actual models for indentation fracture always use ideal conical or
pyramidal tip geometry and the effect of tip rounding has not been investigated to
date. This is a topic which needs further systematic investigation.
A full evaluation of nine indents on silicon, with the sharp and blunt cube corner
indenter, respectively, showed that mean crack length c and mean indent size a were
identical for both tips with high precision. In other words, the tip radius seems to
influence the fracture threshold but not the resulting fracture toughness. Therefore,
all following experiments were performed using the blunt cube corner indenter with a
250 nm tip radius.
In the next step, evaluations of crack length and indent size using the
nanoindenter tip for imaging (NI) were validated by comparing the results of
conventional atomic force microscopy imaging with a standard sharp silicon tip
(AFM) and by scanning electron microscopy (SEM). For this comparison,
impressions in silicon and fused silica were used, since cracks are large and easy to
recognise in SEM. It was found that, for individual crack lengths determined by the
three methods, a significant scattering of results was found. However, the mean
values for crack length c, entered into the formulae of Table 1, showed only minor
deviations, smaller than the scattering within each method (Figure 3). For indent size
a, there was a small trend where SEM yielded slightly larger values than NI and
AFM, especially for fused silica. Finally, it cannot be decided which is the true value.
SEM- and AFM-based methods are simply based on different contrast mechanisms.
In SEM, edges are well detected, while AFM is sensitive to height differences. Both
criteria can be used for detecting the corner of an indent and, finally, it depends on
the operator to decide where the corner is.
In a third step, literature was evaluated to find reference values for the materials
under investigation. For bulk materials, many reliable values were found, partly
determined by classical crack extension experiments, partly determined by indenta-
tion (Table 3). Also for aC and DLC-coatings several values were found, but they
show large scattering (2 . . . 11 MPa m1/2). This is probably a result of real material
Philosophical Magazine 1169
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

Figure 2. Indentations created by the sharp cube corner indenter (30 nm tip radius) into (a)
sapphire, showing slip planes, (b) Si-DLC, exhibiting picture frame cracks and (c) silicon with
normal radials cracks. For comparison, indentations created with a blunt cube corner tip
(250 nm tip radius) in the same materials (d, e, f).
1170 K.I. Schiffmann

SEM NI AFM
5000

Crack length c [nm]


4000

3000

2000

1000

0
Fused silica Si, 0° Si,45°
Sample
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

SEM NI AFM
2000
Indent size a [nm]

1500

1000

500

0
Fused silica Si, 0° Si,45°
Sample

Figure 3. Comparison of crack length c and indent size a determined by scanning electron
microscopy (SEM), scanning nanoindenter tip (NI) and conventional atomic force microscopy
(AFM). As samples, fused silica and silicon were used. On silicon, impressions were done
under substrate orientation 0 (Si, 0 ) and rotated by 45 (Si, 45 ).

differences due to different preparation conditions. It is well known that aC and


DLC coatings may have large variations in material properties (e.g.
hardness ¼ 10 . . . 50 GPa, depending on the hydrogen content and sp3 bonding
fraction). Therefore, these values are only of limited value for validation of the
present measurements. For the optical coatings, SnO2, ZnO and ITO, only one or
two literature values were found and all were determined by indentation techniques.
To compare the six models, the root-mean-square deviation2 of experimental and
literature KC values was calculated for each equation. This can be done in different
ways: (a) all samples are considered in computing the rms deviation; (b) all but the
aC and DLC values are considered, because these latter values are not well defined
and extremely scattered; (c) only the results of the three bulk samples are considered,
since for these the most reliable reference values are available. In the following
discussion, option (b) is used for validation of the models.
For steel and TiN samples, no fracture occurred up to a 32 mN load. Therefore,
these two samples have not been considered further. A summary of all results for the
other samples is given in Table 3 and Figure 4. Figure 5 shows the rms deviations
for all models.
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

Table 3. Measured fracture toughness KC [MPa m1/2] for all materials evaluated by Equations (3)–(8) and mean literature value KIC (Lit) including
minimum and maximum values of KIC(Lit).

KC KC KC KC KC KC KIC
Sample (Equation (3)) (Equation (4)) (Equation (5)) (Equation (6)) (Equation (7)) (Equation (8)) (Lit) Min Max Sources

Silicon 0.96  0.07 0.43  0.03 0.42  0.02 0.31  0.04 Bulk 0.45  0.22 0.95 0.70 1.25 [9,17,18,19]
Fused silica 0.68  0.08 0.34  0.04 0.27  0.02 0.27  0.05 Bulk 0.34  0.02 0.77 0.60 1.20 [17,20,21,22,23]
Sapphire 8.94  0.50 3.90  0.22 3.04  0.19 7.39  1.06 Bulk 2.97  1.41 3.25 2.10 4.50 [5,17,22,24,25,26]
DLC/steel 4.13  0.38 1.91  0.18 1.54  0.28 3.28  0.90 0.07  0.002 1.23  0.07 3.65 1.30 10.1 [16,27,28,29]
aC/Si 4.80  0.29 2.22  0.13 1.76  0.13 3.51  0.59 0.87  0.02 3.34  0.44 4.84 1.90 10.9 [14,15,16,30,31]
ZnO:Al/glass 2.37  0.06 1.15  0.03 1.10  0.13 3.14  0.47 0.41  0.003 No pop-ins 0.90 0.85 0.95 [33,34]
SnO2/glass 2.57  0.28 1.15  0.13 0.95  0.11 2.35  0.37 0.25  0.04 No pop-ins 1.90 Only one value [32]
no-temp
Philosophical Magazine

SnO2/glass 2.32  0.23 1.04  0.10 0.81  0.17 1.78  0.77 0.21  0.06 1.62  0.32 1.90 Only one value [32]
temp
ITO/glass 3.54  0.62 1.49  0.25 1.26  0.39 2.91  1.33 0.28  0.04 No pop-ins 2.10 Only one value [32]
1171
1172 K.I. Schiffmann

Literature Eq(3) Eq(4) Eq(5) Eq(6) Eq(7) Eq(8)


10
9
8
KIc [MPa*m1/2]

7
6
5
4
3
2
1
0
Fused silica Sapphire Silicon a-C/Si DLC/steel ZnO/glass SnO2/glass SnO2/glass ITO/glass
no temp temp
Sample
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

Literature Eq(3) Eq(4) Eq(5) Eq(6) Eq(7) Eq(8)


10.0
KIc [MPa*m1/2]

1.0

0.1
Fused silica Sapphire Silicon a-C/Si DLC/steel ZnO/glass SnO2/glass SnO2/glass ITO/glass
no temp temp
Sample

Figure 4. Fracture toughness for all materials and all equations compared with literature
mean values in a linear and logarithmic scale.

1.0
ALL samples
ALL without aC&DLC
0.8 BULK samples only
RMS (Klit-Kexp)

0.6

0.4

0.2

0.0
Eq(3) Eq(4) Eq(5) Eq(6) Eq(7) Eq(8)
Equation

Figure 5. Root-mean-square deviation between literature and experimental values for all
equations and different subsets of materials: (a) includes all materials, (b) includes all materials
without aC and DLC and (c) includes only the three bulk materials.
Philosophical Magazine 1173

5. Discussion
The smallest deviation from literature values was found for Equations (4) and (5).
These models both use an exponent of 2/5 for (E/H), compared to a power of 1/2 or
2/3 used in the other models. There was no major difference between Equations (4)
and (5), although Equation (4) assumes halfpenny cracks while Equation (5) assumes
Palmquist cracks. The reason why Palmquist cracks also seem to work is that not
p
only l but a l is used in the expression, which makes it more appropriate for very
short cracks. These results are consistent with those of Ponton and Rawling [3,4],
who found that Equations (4) and (5) of Niihara et al. [10,11] show the second- and
third-best correlation between Kexp and KLit of the 19 equations evaluated.3 They
also concluded from their extensive analysis that, generally, halfpenny and Palmquist
equations give equally valid data regardless of the crack profile developed in the
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

material.
Equations (3) and (6) yielded less good agreement with literature, especially when
only bulk samples are considered. This is mainly due to sapphire for which much too
high values are computed; significant deviations also exist for other materials (e.g. ZnO
and ITO). It seems that for very short cracks, as in the case of sapphire, the exponent 2/5
of (E/H ) might give a better description than ½ or 2/3 of Equations (3) and (6).
Equation (7), which was developed especially for thin coatings, always yielded
very small values. Since nothing was known about the stress state of the coatings, the
model has been used without the stress term. Actually, coatings often exhibit
compressive stress which tends to close cracks and would more likely lead to higher
toughness. Further assumptions of model 7 are that the crack length is large
compared to film thickness and the crack depth is equal to film thickness, i.e. the
cracks is constrained within the coating. For metallic substrates, the latter will be the
case but, for actual samples, silicon and glass were used as substrates, both of which
have low fracture toughness. Additionally, the indentation depth during the
experiments, except for the thick DLC coating, were in the range of the coating
thickness or even larger. Therefore, it is very probable that the cracks will have
propagated into the substrate or may even have nucleated in the substrate.
Evaluation using crack energy (Equation (8)) was not possible for all samples
since pop-ins were not always visible or analysable, but, in most cases, it worked. In
these cases, this model yielded rather good results, which is remarkable since it does
not use any empirical fitted constants and the approximation of fracture area is
rather raw.
There are a number of possible sources of error which should be considered when
valuating the method of indentation-based fracture toughness determination.
Cracking is an inherently statistical process which will lead to a large intrinsic
scattering of results. An error of up to 40% is often given as a typical error for
fracture toughness.
A certain systematic error may occur due to the fact that it is difficult to decide
where the impression corner is and where the crack tip ends. A comparison of SEM,
AFM and NI scanned cracks (Figure 3) showed that all three methods yield similar
results but, in all probability the real crack tip cannot be resolved by all methods.
This will lead to a certain underestimation of crack length and, hence, overestimation
of fracture toughness.
1174 K.I. Schiffmann

Probably, the most severe source of error is the unknown crack geometry.
All models make assumptions about crack geometry, either of halfpenny type4 or
Palmquist type, which has major consequences for the crack area considered. If the
true crack geometry is different, the model may yield completely wrong results. To
demonstrate deviating crack geometries, focussed ion beam (FIB) has been used to
create a sequence of cross-sectional slices through the cracked region of the silicon
sample. Figure 6 gives a three-dimensional reconstruction of the crack pattern below
the surface. For three impressions investigated, it was found that vertical crack
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

Figure 6. SEM images: (a) top view and (b) 3D reconstruction of nanoindenter impression
into silicon showing a vertical crack which splits into two horizontal cracks.
Philosophical Magazine 1175

always splits into two horizontal cracks below the surface. Additionally, there were
not only cracks from the impression corners but also from the edges. In other words,
the crack geometry differed significantly from ideal half- or quarterpenny type. The
reason for this deviation is the single crystal character of the sample, which prefers
certain crack directions. Nevertheless, if the total crack area is comparable to that of
an ideal halfpenny crack with the same visible surface crack length, the models may
yield more or less acceptable results. On the other hand, FIB investigations on fused
silica showed classical three-fold radial cracks emanating from the centre of the
impression (Figure 7), which has also been recorded by Tandon [35] on soda-lime
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

Figure 7. SEM images: (a) top view and (b) 3D FIB reconstruction of nanoindenter
impression into fused silica showing three-fold vertical crack.
1176 K.I. Schiffmann

silica glass. Here, it is not a pure quarterpenny but a more elliptical shape, because
the depth of crack below the impression is smaller than outside the impression.
Finally, in the case of thin films, the coating thickness, substrate material and
potential intrinsic stress will influence the fracture toughness measurement.
Equations (3)–(7) and (9) do not even consider these aspects. Only Equation (8)
tries to account for thin films. Nevertheless, this model is only valid for the special
case of substrate materials with high toughness, which lead to an enclosure of cracks
within the coating and large crack lengths in comparison to film thickness, which was
not the case in the present investigation.
In addition to a discussion of experimentally caused sources of error, it should be
mentioned that there are also principle doubts on the approach of the equations.
Quinn and Bradt [7], in a critical discussion, claimed that indentation-based fracture
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

measurement is not a valid method to determine the fracture toughness KIC because
(a) the deformation of the material during indentation is a complex three-
dimensional strain and the material is left in a complex residual stress state, which
does not meet the conditions of the fracture toughness definition; (b) none of the
equations are based on an accurate stress intensity factor solution, but formulae are
the result of dimensional analysis matched by empirical factors to fit a certain set of
samples; (c) as shown by Ponton and Rawling [3,4] none of the equations are able to
generate accurate results for all types of material or even make a unambiguous
ranking of their material toughness.
Thus, the indentation-based KC should not necessarily be identified with the well-
defined fracture toughness KIC from a standardised crack extension experiment but,
perhaps, should be taken as an empirical value characterising the fracture behaviour
under the condition of a sharp body pressed into the surface.

6. Conclusion
Indentation fracture toughness KC was determined by nanoindentation experiments
on different bulk and coating materials using a cube corner indenter. The following
results were found:
(a) Decreasing the tip radius of curvature of a cube corner indenter will increase
the fracture threshold or change the fracture characteristic but had no
influence on crack length (in the case of silicon).
(b) Evaluation of crack length by scanning the indenter tip (NI), by conventional
AFM and by SEM, yielded similar results. The differences were smaller than
the scatter within each method.
(c) Equations (4) and (5) yielded the best agreement with literature values, which
is consistent with the work of Ponton and Rawling [3,4]. Both equations used
a (E/H)2/5 dependency for KC, which may be more adequate for short cracks
than the common (E/H)1/2 or (E/H)2/3 dependence of Equations (3), (6) and
(7). Nevertheless, the mean deviations of Equations (4) and (5) were still in
the range of 40%, which gave the right magnitude of the result but not a
precise value.
(d) Equation (8), using fracture energy, yielded surprisingly good results. It has
the advantage that no empirical fitting constant is used and the hardness of
Philosophical Magazine 1177

the material does not enter the evaluation. However, the method was only
applicable when cracks and pop-ins were present.
(e) Better models considering the substrate effect would be desirable for
application to thin films. A recent approach by Morris et al. [36] especially
considered the wedging effect of acute cube corner indenters and the different
elastic moduli of the film and substrate.

Acknowledgements
The author acknowledges M. Reichelt and A. Heiss from GfE RWTH Aachen for preparation
and imaging of the FIB cross-sections.
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

Notes
1. In the following, KC is used instead of KIC if fracture toughness is determined by an
indentation method, which is not a pure mode I-type loading and will not meet the precise
definition of fracture toughness (see discussion in [7]).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 PN
2. rmsðKIC Þ ¼ NðN1Þ ðKLit  Kexp Þ2 :
sample¼1

3. The equation with the best correlation in the paper of Ponton and Rawling [3,4] also uses
the exponent 2/5 for (E/H), but it gave slightly worse results when applied to our data set.
4. In case of cube corner symmetry, one would expect three-fold quarterpenny geometry of
the crack.

References

[1] ISO 12737:2005, Metallic materials. Determination of plane-strain fracture toughness,


International Organization for Standardization, Geneva, Switzerland.
[2] ISO 17281:2002, Plastics. Determination of fracture toughness (GIC and KIC) at
moderately high loading rates (1 m/s), International Organization for Standardization,
Geneva, Switzerland.
[3] C.B. Ponton and R.D. Rawlings, Mater. Sci. Tech. 5 (1989) p.865.
[4] C.B. Ponton and R.D. Rawlings, Mater. Sci. Tech. 5 (1989) p.961.
[5] A.G. Evans and E.A. Charles, J. Am. Ceram. Soc. 59 (1976) p.371.
[6] B.R. Lawn, A.G. Evans and D.B. Marshall, J. Am. Ceram. Soc. 63 (1980) p.574.
[7] G.D. Quinn and R.C. Bradt, J. Am. Ceram. Soc. 90 (2007) p.673.
[8] R.F. Cook and G.M. Pharr, J. Am. Ceram. Soc. 73 (1990) p.787.
[9] J.I. Jang and G.M. Pharr, Acta Mater. 56 (2008) p.4458.
[10] K. Niihara, R Morena and D.P.H. Hasselman, J. Mater. Sci. Lett. 1 (1982) p.13.
[11] K. Niihara, J. Mater. Sci. Lett. 2 (1983) p.221.
[12] M.T. Laugier, J. Mater. Sci. Lett. 6 (1987) p.897.
[13] F. Ouchterlony, Eng. Fract. Mech. 8 (1976) p.447.
[14] J. Thurn and R.F. Cook, J. Mater. Sci. 39 (2004) p.4809.
[15] X. Li, D. Diao and B. Bhushan, Acta Mater. 45 (11) (1997) p.4453.
[16] M.D. Michel, L.V. Muhlen, C.A. Achete and C.M. Lepienski, Thin Solid Films 496
(2006) p.481.
[17] G. Pharr, Mater. Sci. Eng. A 253 (1998) p.151.
[18] D. Casellas, J. Caro, S. Molas, J. Prado and I. Valls, Acta Mater. 55 (2007) p.4277.
[19] F. Ericson, S. Johansson and J.A. Schweitz, Mater. Sci. Eng. A 105/106 (1988) p.131.
1178 K.I. Schiffmann

[20] Material Database, AZO Materials, Warriewood, NSW, Australia. Available at:
www.azom.com.
[21] Product datasheet, Insaco Inc., Quakertown, PA, USA. Available at: www.insaco.com.
[22] T. Scholz., G. Schneider, J. Munos-Saldana and M.V. Swain, Appl. Phys. Lett. 84 (16)
(2004) p.3055.
[23] G.M. Pharr, D.S. Harding and W.C. Oliver, in Mechanical Properties and Deformation
Behavior of Materials having Ultra-Fine Microstructures, M. Nastasi, D.M. Parkin and
H. Gleiter, eds., Kluwer, Dordrecht, 1993, pp.449–461, NATO ASI Series E.
[24] T. Guillot, J.L. Henshall and R.M. Hooper, Int. J. Refract. Metals Hard Mater. 16 (1998)
p.323.
[25] M. Iwasa, T. Ueno and R.C. Bradt, J. Soc. Mater. Sci. Jpn. 30 (1981) p.1001.
[26] Product datasheet, Accuratus Corp., Phillipsburg, NJ, USA. Available at:
www.accuratus.com.
[27] S. Zhang, D. Sun, Y. Fu and H. Du, Surf. Coat. Tech. 198 (2005) p.74.
Downloaded by [Pennsylvania State University] at 13:21 12 August 2013

[28] M. Nastasi, P. Kodali, K.C. Walter and J.D. Embury, J. Mater. Res. 14 (5) (1999) p.2173.
[29] Z.-H. Xie, R. Singh, A. Bendavid, P.J. Martin, P.R. Munroe and M. Hoffman, Thin Solid
Films 515 (2007) p.3196.
[30] K. Jonnalagadda, S.W. Cho, I. Chasiotis, T. Friedmann and J. Sullivan, J. Mech. Phys.
Solids 56 (2008) p.388.
[31] H.D. Espinosa, B. Peng, N. Moldovan, T.A. Friedmann, X. Xiao, D.C. Mancini,
O. Auciello, J. Carlisle, C.A. Zorman, Proceedings of the 11th International Conference on
Fracture, Turin, Italy, March 20–25, 2005.
[32] J. Chen and S. Bull, J. Phys. D Appl. Phys. 40 (2007) p.5401.
[33] Y. Hayashi, Y.H. Choa, J.P. Singh, K. Niihara, in Proceedings of the Ceramic Matrix
Composites Symposium, American Ceramic Society Annual Meeting, Cincinnati, OH,
May 3–6, 1998.
[34] A.K. Mukhopadhyay, M.R. Chaudhuri, A. Seal and S.K. Dalui, Bull. Mater. Sci. 24 (2)
(2001) p.125.
[35] R. Tandon, J. Eur. Ceram. Soc. 27 (2007) p.2407.
[36] D.J. Morris and R.F. Cook, J. Mater. Res. 23 (9) (2008) p.2429.

You might also like