You are on page 1of 20

CRC_2308_Ch013.

qxd 11/30/2005 3:48 PM Page 385

13 Nanolayered or Kinking
Nonlinear Elastic Solids
Michel W. Barsoum
Department of Materials Science and Engineering,
Drexel University, Philadelphia, Pennsylvania

CONTENTS

Abstract
13.1 Introduction
13.2 Kinking in Crystalline Solids
13.3 Kinking Nonlinear Elastic Solids
13.4 Theoretical Aspects of Spherical Nanoindentations
13.5 Nanoindentation Results on KNE Solids
13.5.1 Sapphire
13.5.2 The MAX Phases
13.5.3 Graphite
13.5.4 Mica
13.5.5 Hexagonal Boron Nitride
13.6 Bulk vs. Nanoindentation Results
13.7 Summary and Conclusions
Acknowledgments
References

ABSTRACT
Recently we postulated that most solids with a high c/a ratio belong to the same class, which we
designated as kinking nonlinear elastic (KNE).1 The high c/a ratio renders slip, other than basal and
non-basal slip and twinning, prohibitively expensive. Thus only basal slip, that leads to kinking is
activated during deformation. The signature of KNE solids is the formation of fully reversible, hys-
teretic, stress–strain loops, on repeat loadings. This full reversibility, in turn, is due to the formation
of incipient kink bands (IKB), comprised of two nearly parallel dislocation walls of opposite polar-
ity that are attracted to each other; when the load is removed that they annihilate. The energy dissi-
pated per cycle is usually substantial and increases with increasing stress. Recently, we have also
shown that repeated spherical nanoindentations, on the same location, is a technique that is ideally
suited to identify and characterize KNE solids. Examples on Ti3SiC2, mica, hexagonal boron
nitride, graphite, and sapphire are presented. KNE solids tend to be nanolayered, but as shown
herein, sapphire, which is not, also behaves like a KNE solid.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:48 PM Page 386

386 Nanomaterials Handbook

13.1 INTRODUCTION
Graphite, ice, sapphire single crystals, layered semiconductors, high-Tc superconductors, hexagonal
boron nitride (h-BN), Mg, Bi, Zn, the ternary carbides and nitrides (the so-called MAX phases, see
below) mica, and other layered silicates, among others, are not thought to be related. In this chap-
ter we not only show them to be related, but we also demonstrate why they are all members of a
huge new class of solids, best described as kinking nonlinear elastic (KNE) solids.1 In doing so, we
demonstrate how repeated loadings onto the same location with a hemispherical nanoindenter tip is
a powerful technique for characterizing these solids.1–4 The considerable information gleaned from
this technique is outlined herein.
Kinking is a ubiquitous and well-known phenomenon in nature that can be seen easily at many
scales. At the largest scale, kinking and kink bands (KBs) are quite common in geologic formations.
This mechanism has been invoked to explain deformation in single-crystal hexagonal metals, such
as Cd and Zn,5,6 highly constrained rocks, organic crystals, card decks, rubber laminates, oriented
polymer fibers, wood, graphite fibers, laminated C–C and C-epoxy composites, among others.7
Before proceeding further it is important to emphasize that this chapter deals with kinking in
crystalline solids exclusively; kinking in amorphous solids or polymers is not discussed.
Orowan was the first to report on kinking in a metal.5 Later, Hess, and Barrett6 suggested that
the formation of KBs entailed the formation of pairs of dislocations of opposite signs that move in
opposite directions in response to a shear stress. It is the coalescence of these mobile dislocation
walls (MDWs) into a narrow region that gives rise to kink boundaries.6,7 A few years later, Frank
and Stroh (F&S)8 derived an analytic expression, based on energy arguments, that predicted the
remote stress at which a KB nucleus became unstable. Recently, we adapted the F&S theory and
introduced the concept of IKBs discussed in more detail below.9,10
In our previous work we suggested that a sufficient, but not necessary, condition for a solid to
be a KNE solid is a high c/a ratio.1 This condition ensures that only basal plane dislocations are
energetically favorable and eliminates twinning as a possible deformation mechanism. Under those
constraints a solid can either deform by kinking or fracture.
This chapter is structured as follows: the next section describes kinking and its characteristics at
all levels — atomic, microscopic and macroscopic. Section 13.3 deals with the kinking in bulk sam-
ples. The theoretical aspects of kinking — with special emphasis on the F&S model — and spheri-
cal nanoindentations are discussed in Section 13.4. Section 13.5 deals with specific examples,
including mica, sapphire, graphite, Ti3SiC2, and Ti2AlC, as representatives of the MAX phases. The
latter are a novel class of layered, machinable, hexagonal ternary carbides and nitrides that number
over 50, with the chemical formula Mn+1AXn, where M is an early transition metal, A is an A-group
element (mostly IIIA and IVA), and X is C or N.11–13 The last section links the results obtained from
nanoindentations and bulk compression specimens and provides an overview of KNE solids.

13.2 KINKING IN CRYSTALLINE SOLIDS


As noted above, at the macroscopic level, kinking is manifested in many systems ranging from geo-
logic formations to card decks (Figure 13.1a). In general, kinking will occur whenever a plastic,
anisotropic system is loaded parallel to the deformation or slip plane.
At the microscopic level kinking is observed as bucking in individual grains.14 For the most
part, because kinking is essentially a buckling phenomenon, it is associated with delaminations and
lack of constraints. This is why, for example, when samples with highly oriented grains are loaded
parallel to the easy or basal slip planes they kink at the unconstrained corners rather than in the bulk7
(Figure 13.1b). It is also the reason that, as discussed below, kinking under an indenter is much
more difficult than in simple compression experiments.
Atomically, the so-called stove-pipe configuration, where two kink boundaries are separated from
each other by a dislocation free region is identified as a KB (Figure 13.1c).15 Kink boundaries are
essentially thin regions into which numerous MDWs have collapsed. In other words, MDWs — which
Copyright 2006 by Taylor & Francis Group, LLC
CRC_2308_Ch013.qxd 11/30/2005 3:48 PM Page 387

Nanolayered or Kinking Nonlinear Elastic Solids 387

Kink band

Shear band

(a) (b)

1 µm

(c)

FIGURE 13.1 Kink bands in (a) a deck of cards loaded parallel to the cards,11 (b) Ti3SiC2,7 and (c) TEM of
a KB comprised of two kink boundaries.15 Note similarity to twinning.

are but low-angle grain boundaries — are the precursors of kink boundaries. It is important not to con-
fuse kink boundaries with twins. The latter are crystallographically determined, while the former, also
symmetrical across the kink boundary, can take any angle. The kink boundary angle depends on the
number of MDWs that ultimately end up in it.

13.3 KINKING NONLINEAR ELASTIC SOLIDS


Before reviewing the nanoindentation results of typical KNE solids — most of which have been sin-
gle crystals — it is instructive to review briefly the response of their polycrystalline counterparts
Copyright 2006 by Taylor & Francis Group, LLC
CRC_2308_Ch013.qxd 11/30/2005 3:48 PM Page 388

388 Nanomaterials Handbook

under cyclic loading. There are two typical responses. Type I, where all the loops are closed, includ-
ing the first loop (Figure 13.2a) and type II, where the first loop is open, but all subsequent loops
are closed or fully reversible (Figure 13.2b). The response of some of the MAX phases tested to
date, such as Ti3SiC2 and Ti2AlC, is of type I;9 and that of graphite, sapphire, and h-BN are of type
II.2 The reasons for the differences are not entirely clear but could reflect differences in grain bound-
ary strengths and ease of delaminations; the latter favored by a low c44. What is occurring, however,
is better understood; in type I, only fully reversible IKBs are nucleated. In type II, during the first
cycle, MDWs result in KBs, which in turn lead to a reduction in domain size resulting in harden-
ing. In all subsequent cycles, only fully reversible IKBs form.
In recent chapters, we developed atomistic10 and constitutive models16 for describing KNE
solids. Only the atomistic model, based on the F&S paper,8 is described here. F&S considered the

1000
(a) FG Ti3SiC2

800
Stress (MPa)

600

400

200

0
0 0.001 0.002 0.003 0.004 0.005 0.006
Strain

100 (b) h-BN

80
Stress (MPa)

60 First cycle

40

20 Cycles 2 to 5

0
0 0.0005 0.001 0.0015 0.002 0.0025 0.003 0.0035
Strain

FIGURE 13.2 Compressive stress–strain curves for (a) Ti3SiC2,9 which exhibits type I behavior, and (b)
BN,34 which exhibits type II behavior. Note fully reversible nature of loops.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:48 PM Page 389

Nanolayered or Kinking Nonlinear Elastic Solids 389

problem of the growth of a thin elliptical subcritical kink — with dimensions α and β, such that
α ⬎⬎ β — comprised of opposite sign dislocations (Figure 13.3a). Initially, the thin elliptical kink
is subcritical because it increases the free energy of the system. Using an energy approach, remi-
niscent of Griffith’s, F&S showed that the remote shear stress τ needed to render a subcritical KB
unstable depends on α and is given by8
G2bγ c
τ ⬎ τc ⫽ 冪莦


(13.1)

where b is the Burger’s vector, G the shear modulus (for single crystals G is replaced by c44), and
γc the critical kinking angle given by
b 3兹苶3(1 – v)τloc
γ c ⫽ ᎏᎏ 艐 ᎏᎏ (13.2)
D 2G
with ν being Poisson’s ratio and D the distance between dislocations in the walls (Figure 13.3a).
F&S modeled a two-dimensional single crystal and thus assumed, correctly, that once the
inequality in Equation 13.1 was satisfied, a subcritical KB would rapidly and autocatalytically grow
to the edge of the sample and dissociate into two parallel MDWs (Figure 13.3b). Continued load-
ing would then result in the formation of other IKBs and MDWs within the first one. It is the rep-
etition of this process that ultimately leads to the formation of the KBs.6,7
However, in order to explain our recent results, many of which are discussed in this chapter, we
had to invoke the idea of an IKB.9,10,17 An IKB is one in which the ends remain attached and is
therefore fully reversible (Figure 13.3a).
In a polycrystal, 2α is determined by the thickness of the grains.10 The situation under a nanoin-
denter indented into a single crystal is different. To model this situation we make the plausible
assumption that4

a⫽α (13.3)

where a is the contact radius of the indenter (see below). Once α is known, β, can be determined
from8,17
2α(1 – v)
2β ⫽ ᎏ (σ – σt) (13.4)
2G

where σt is a threshold stress, below which the response is purely elastic.

Attractive force
Free
surface Delamination
2 plane

D 


Mobile
2 dislocation
walls
(a) (b)

FIGURE 13.3 Schematic of (a) an IKB with length 2α and width 2β. As long as the ends remain attached, the
force is attractive. (b) IKB under an indenter showing MDWs bracketing an IKB; upon removal of the load only
the latter disappears. Note the delamination plane, which must accompany the IKB to MDW transformation.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:48 PM Page 390

390 Nanomaterials Handbook

Recently, we have shown that the number of IKBs per unit volume, Nk, is given by10,17
2G2γ c
Nk 艐 ᎏᎏ ε (13.5)
πα (1 – v)2[σ – σt]2 1KB
3

where εIKB is the nonlinear strain under the indenter due to the formation of the IKBs. In other
words, εIKB = εtot − εel, where εtot is the total strain and εel the elastic strain. In this model and in the
following discussion, the formation of dislocation pileups is ignored, because their contributions to
the inelastic strain is small compared to those from kinking.10,17
Once Nk, α and β are determined, a near complete picture of the dislocation configuration under
the indenter comes into focus. The results are shown schematically in Figure 13.4 and can be sum-
marized as follows: below σt the response is linear elastic; at σ ⬎ σt a large number of short IKBs are
introduced in the material (Figure 13.4a). With increasing load the number of IKBs actually decrease
at the same time that their lengths increase (Figure 13.4b). This somewhat counterintuitive result
comes about because with increasing stress, the IKBs get longer and wider, merge and thus decrease
Nk. Without a merging mechanism, the IKBs could not penetrate deeper into the crystal while main-
taining the shape dictated by Equation 13.4. How that is accomplished is unclear at this time, but most
probably occurs by the annihilation of adjacent IKB walls that have opposite polarities.
As the loads are increased further, the IKBs penetrate deeper into the sample and when the ten-
sile stresses that develop at their tip exceed the delamination stresses of the material, delaminations
occur below the indenter (Figure 13.4c). With delaminations, the walls making up the IKBs imme-
diately become mobile and are swept away from under the indenter and end up as pileups or
mounds on either side of the indenter (Figure 13.4c).
As clearly demonstrated in this chapter, the formation of fully reversible, reproducible hysteretic
stress–strain loops upon load cycling is one of the signatures of IKBs (e.g., Figure 13.2a). The area
within each loop, Wd, represents the energy dissipated per unit volume per cycle as a result of the to
and fro motion of the dislocations comprising the IKBs. Wd is related to the energy dissipated by a
dislocation line of unit length sweeping a unit area, Ω, by10,17
Ω G2γ c
ᎏ ⫽ ᎏᎏ W (13.6)
b πNkα3(σ – σt)2 d
In deriving Equation 13.6, Wd is assumed to be solely due to the movement of IKB related disloca-
tions; the motion of pileup dislocations is again neglected. With that caveat, we believe Ω/b is a
material property that is proportional to, if not equal to, the critical resolved shear stress (CRSS) of
basal plane dislocations involved in the IKBs. This comment notwithstanding, more work is needed
to establish the relationship between Ω/b and the CRSS.

2a

Pileup
RR

IKBs
Delaminations
(a) (b) (c)

FIGURE 13.4 Schematic of microstructure under a spherical nanoindenter with (a) low loads; (b) higher
load, but before IKB to KB transformation. Note that size and density of IKBs is less at higher loads. (c) Loads
corresponding to stresses greater than delamination stresses. Note plastic deformation and pileups around the
indentation marks.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:48 PM Page 391

Nanolayered or Kinking Nonlinear Elastic Solids 391

13.4 THEORETICAL ASPECTS OF SPHERICAL NANOINDENTATIONS


According to Hertz, applying a load of P using a spherical indenter of radius R (Figure 13.5a)
results in an elastic penetration depth, he, relative to the original surface given by

冢 冣冢 冣
2/3 2/3
3P 1
he ⫽ ᎏᎏ ᎏᎏ (13.7)
4E* R
E* is a reduced modulus given by
1 1 –νs2 1 –νi2
ᎏᎏ ⫽ ᎏᎏ ⫹ ᎏᎏ (13.8)
E* Es Ei
The subscripts s and i refer to the specimen and the indenter, respectively.
Geometrically, as long as hc << R, the actual contact diameter 2a is related to hc by

a ⫽ 兹2苶R
苶h苶苶c (13.9)

where hc is the depth at which the indenter and surface are no longer conformal (see Figure 13.5a).3
If one further assumes that18

hc ⫽ htot ⫺ he/2 (13.10)

a can be readily calculated from Equation 13.9 once he is calculated from Equation 13.7. This
assumption, shown schematically in Figure 13.5b, is a very good approximation as long as the sys-
tem remains in the elastic regime.
Furthermore, in the purely elastic regime,18,19

hc 艐 htot /2 (13.11)

combining Equations 13.7 to 13.12 it can be shown that

冢冣
P 4E* a
ᎏᎏ2 ⫽ ᎏᎏ ᎏᎏ (13.12)
πa 3π R
The left-hand side is the indentation stress and a/R can be considered to be an indentation strain.
Hence a plot of P/π a2 vs. a/R should, in the elastic regime, yield straight lines with slopes that are
proportional to the E*.

P Pmax

Elastic/plastic
R loading

Elastic
2a loading
htot Elastic
hc unloading
≈a

h
hr he
hc he /2

(a) (b) htotal

FIGURE 13.5 Schematic of (a) indentation at maximum load, and (b) of corresponding load–displacement-
nanoindentation curve. In order to convert these results to stress–strain the approximation that hc 艐 ht⫺he/2 is made.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:48 PM Page 392

392 Nanomaterials Handbook

As discussed below the response under the repeated loadings of a spherical nanoindenter
depends on applied stress; at stresses below the stress needed to initiate the IKBs, the response is
linear-elastic. At stresses high enough to initiate IKBs but not sunder them, i.e., not to cause delam-
inations, the response is of type I. At even higher stresses the response is of type II. It is crucial to
note that the IKB to KB transition must be accompanied by delaminations that, as discussed below,
more often than not, result in massive pop-ins.

13.5 NANOINDENTATION RESULTS ON KNE SOLIDS


In the following sections, our recent nanoindentation results on sapphire, graphite, Ti3SiC2, h-BN,
and mica are reviewed.

13.5.1 SAPPHIRE
Typical load–displacement results of a sapphire single crystal loaded normal to its basal planes by
with a spherical diamond indenter of 1 µm radius are shown in Figure 13.6a.20 Unless otherwise
noted, the indenter is typically pushed into the same location five consecutive times. This cycling
is crucial because it is only when the third and subsequent cycles fully reproduce the second cycle
that a solid can be unambiguously characterized as KNE (see below). The results shown in Figure
13.6a clearly show that at around 30 mN, a pop-in of the order of 100 nm is observed. When the
load–displacement curves were converted into stress–strain curves (not shown) the stresses at which
the pop-ins occurred were 艐 40 GPa. Figure 13.6b plots the load–displacement results for a region
indented 24 times, zooming on the middle section of the loops. The reproducibility of these hys-
teretic loops is noteworthy, especially that the maximum normal stresses in every cycle were of the
order 40 GPa.20 This reproducibility is consistent with the formation of fully reversible IKBs.
In metals, the theoretical shear stress needed to create a dislocation in a perfect crystal is
assumed to be 艐 G/2 π .21 Recent nanoindentation results and theoretical models have verified this
assumption; in some cases the measured strengths were actually higher than G/2 π.22–24 If one
assumes the maximum shear stress, τm, under the indenter to be 艐 σn/2 (20 GPa in this case) and
noting that c44 for sapphire is 148 GPa,25 it follows that τm for sapphire is at least equal to c44/7.5,
which, gratifyingly, is 艐 G/2 π . The very high stresses needed to initiate the IKBs are compelling
evidence that the small volume probed by the indenter must have been perfect, or near perfect.
In general, and despite pop-ins of the order 2–3 µm, it was not easy to find the locations of the
indentations in the SEM or with an AFM. In a few cases, we were able to trace the indentation and
confirmed that material pileup was seen around the indenter.20 In agreement with previous work on
the nanoindentation of sapphire single crystals,26–29 there was no evidence for surface linear features
associated with slip lines or twins, when the loading was parallel to the c-axis. When the loading was
parallel to the a-axis, linear features were observed, again in agreement with the previous work.26,28,29
In order to find the indentations, a square array was introduced into the surface. When the dis-
tance between the indentations was 艐 6 µm, in some areas, the surface was damaged exposing the
locations of the nanoindentations marks, and as important it provided direct evidence for the
formation of kink bands.20
The implications of the results shown herein are not inconsequential. They unambiguously
show that dislocations, most probably in the form of IKBs, are mobile even at room temperature.
Furthermore, if the factors by which IKBs in sapphire nucleate can be understood, and lowered sig-
nificantly, it may be possible to impart some limited ductility to an otherwise very brittle solid.

13.5.2 THE MAX PHASES


Like other KNE solids, the response of the MAX phases to indentations, depends on the applied
load. For example, for a fine-grained (3–5 µm) Ti3SiC2 polycrystalline samples in the low load
regime (Figure 13.7a), the response, while slightly noisy, was fully reversible.30 When these results
are converted into stress–strain (Figure 13.7b), the response on initial loading and initial unloading
Copyright 2006 by Taylor & Francis Group, LLC
CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 393

Nanolayered or Kinking Nonlinear Elastic Solids 393

Sapphire C
100

80

Load (mN)

60

1
40 2
3
Pop-in 4
5
20 6
7
8

0
0 100 200 300 400 500 600 700 800
(a) Displacement (nm)

50

45
C-plane

40

35
Load (mN)

30

25 Cycle 8

Cycle 18
20
Cycle 22
15 Cycle 24

10
650 700 750 800
(b) Displacement (nm)

FIGURE 13.6 Typical load–displacement curves obtained when sapphire single crystals repeatedly loaded —
parallel to the c-axis — with a 1 µm diameter hemisphere,20 (a) first to 50 mN and then to 100 mN. Note pop-
in at 艐 10 mN. Also note lack of hysteretic loops when loaded to 50 mN. (b) Same as (a), but focused on the
middle of the hysteretic loops showing reproducibility despite maximum stresses of 艐 40 GPa. This indentation
was repeated 24 times in the same location. Reproducibility of loops is noteworthy.20

is linear-elastic. Note that the loops, even the first, are fully reversible indicating that the response
is of type I.
At higher loads the response is of type II (Figure 13.8a). When these results, obtained on coarse-
grained (100 µm diameter and 艐 15 µm thick) Ti2AlC polycrystalline samples, are converted into
Copyright 2006 by Taylor & Francis Group, LLC
CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 394

394 Nanomaterials Handbook

stress–strain curves (Figure 13.8b), the response is one where an elastic region
is followed by a plastic region at 艐 3 GPa.17 The response is clearly of type II. The slope of the lines
upon initial loading and unloading correspond to the Young’s moduli values of Ti2AlC.30
The results shown in Figure 13.7 and Figure 13.8 were obtained on polycrystalline samples and the
response on the first loading in the different locations was quite stochastic.31 However, the response of
the second and subsequent cycles, was much more reproducible from location to location and harder
than during the first cycle. As noted above, reproducibility and hardening after the first indent is a char-
acteristic of KNE solids, and results when the indenter essentially creates its own “equilibrium”
microstructure.3 The hardening is due to the formation of smaller domains under the indenter as a result
of the deformation. Despite being layered solids, the MAX phases in general, and Ti3SiC2 and Ti2AlC
in particular, are reasonably tough and we were thus not able directly to observe this domain reduction.

10
Cycle 1
Cycle 2
8 Cycle 3
Cycle 4
Cycle 5
Load (mN)

0
0 10 20 30 40 50 60
(a) Displacement (nm)

Cycle 1
3 Cycle 2
Cycle 3
Cycle 4
Cycle 5
Stress (GPa)

Elastic
1 response
Elastic
response

0
0 0.02 0.04 0.06 0.08 0.1
(b) Strain (a /R )

FIGURE 13.7 (a) Typical load–displacement curves of fine-grained Ti3SiC2 loaded to 10 mN.31 (b) Stress–
strain curves of results shown in (a). The solid inclined lines shown in (b) represent the elastic response, assum-
ing a c33 of 艐 325 GPa.35 Note response is of type I.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 395

Nanolayered or Kinking Nonlinear Elastic Solids 395

50
Cycle 1
Cycle 2
40 Cycle 3
Cycle 4
Cycle 5
30

Load (mN) 20

10

0
0 50 100 150 200 250 300
(a) Displacement (nm)

2.5

2
Stress (GPa)

Cycle 1
Cycle 2
1.5 Cycle 3
Cycle 4
Cycle 5
1

0.5

0
0 0.05 0.1 0.15 0.2
(b) Strain (a /R )

FIGURE 13.8 (a) Typical load–displacement curves of a coarse-grained Ti2AlC2 polycrystalline sample
loaded to 50 mN.17 (b) Stress–strain curves of results shown in (a). Note yield point at 艐 3 GPa. Solid inclined
lines represent the elastic response, assuming a modulus of 艐 278 GPa.30

Graphite and mica are more brittle and direct evidence for domain reduction is shown below (see Figure
13.11). As important, direct TEM evidence for the formation of KBs and delaminations under a
Berkovich indenter in epitaxial thin films of Ti3SiC2 — loaded parallel to the c-axis — exists.32
It is important to note that despite loads corresponding to stresses of the order of 6–8 GPa, SEM
observations of areas indented showed little trace of the indentations (Figure 13.9a). At the highest
stresses, however, a clear crater is observed (Figure 13.9b). The slip lines along the basal planes of
some grains are clearly visible.
Every MAX phase tested to date — Cr2AlC, Ti3GeC2, Ti3AlC2, and Zr2InC, among others —
has been shown to be a KNE solid. These results will be published in the near future.

13.5.3 GRAPHITE
Similar to sapphire and the MAX phases, the response of graphite single crystals loaded normal to
their basal planes with a 13.5 µm radius hemispherical diamond indenter depends on load.2 At low

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 396

396 Nanomaterials Handbook

(a)

Magn Det WD 10 µm
2500× SE 6.2

(b)

Magn Det WD 10 µm
2500× SE 10.1

FIGURE 13.9 Typical SEM micrographs of a fine-grained polycrystalline Ti3SiC2 sample loaded to (a)
200 mN, and (b) 500 mN.31 Deformation in (b) is significant. In contradistinction, and in full accord with
the notion that IKBs are fully reversible, the crater formed at 200 mN is much more shallow and almost
imperceptible.

loads (not shown) the response is of type I.2 At 艐 10 mN or 艐 0.5 GPa, the response changes from
type I to type II. At higher loads, the response is clearly of type II (Figure 13.10a). In Figure 13.10b,
the slope of the solid inclined line corresponds to an elastic modulus of 36.5 GPa, which is the value
of c33 reported in the literature for graphite.33 Note clear evidence for a yield point around 0.4 GPa.
Postindentation SEM micrographs of most surfaces loaded to 200 mN showed no trace of
indentation. In a few instances, a faint circumferential crack pattern was observed (see inset in
Figure 13.10a).
In one experiment a massive, 60 µm, pop-in was registered.2 When the specimen surface was
examined in the SEM a graphitic “rose” with sixfold symmetry was clearly visible2 (Figure 13.11a).
Higher magnification SEM micrographs of the center of the crater provide unambiguous evidence

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 397

Nanolayered or Kinking Nonlinear Elastic Solids 397

200

150
5 µm

Load (mN)
Multiple
indents
100
First indent

Cycle 1
Cycle 2
50
Cycle 3
Cycle 4
Cycle 5
0
0 1000 2000
(a) Displacement (nm)

1.5

1.25
First indent

1
Stress (GPa)

Elastic
0.75
Multiple
indents
0.5
Cycle 1
Cycle 2
0.25 Cycle 3
Cycle 4
Cycle 5
0
0 0.2 0.4 0.5
(b) Strain (a/R )

FIGURE 13.10 (a) Typical load–displacement curves for a single-crystal graphite sample loaded parallel to
its c-axis with a 13.5 µm indenter. Inset is an SEM micrograph of sample after loading to 200 mN, where only
a faint network of circumferential cracks is visible. (b) Stress–strain curves for results shown in (a). Note yield
at 艐 0.4 GPa. Response is of type 1.

for the break up of the single crystal into smaller domains alluded to above.2 The nanolaminate
nature of graphite and direct evidence for the formation of KBs is shown in Figure 13.11b. That
graphite forms KBs had been known for a long time.33 The fundamental role IKBs play in the defor-
mation process was not appreciated until quite recently. The deformation of polycrystalline graphite
was finally understood only by understanding that role.2

13.5.4 MICA
The response of mica single crystals, loaded normal to their basal planes, with a 13.5µm spherical dia-
mond indenter was found to be a strong function of the quality of mica sheets used.1 The response of

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 398

398 Nanomaterials Handbook

FIGURE 13.11 SEM micrograph of a (a) graphite single crystal that experienced a 60 µm pop-in.2 (b) Same
as (a), but at a higher magnification and with emphasis on the center of the crater. (c) Mica single crystal, after
a 艐 2 µm pop-in.1

the best quality mica (Grade A) was linear-elastic up to, in some cases, 8GPa (cycles 1 and 2 in Figure
13.12a). In other cases, the linear elastic regime was interrupted by massive pop-ins4 (cycle 3 in Figure
13.12a). Surprisingly, and despite these massive pop-ins, in some cases, the response upon reloading
was again linear-elastic (cycles 4 and 5 in Figure 13.12a). The pop-ins, however, create a crater that
reduces the stress under the indenter (Figure 13.12b). The fact that the response after a pop-in of
艐1µm, is elastic is intriguing and not well understood. Clearly dislocations must have nucleated and

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 399

Nanolayered or Kinking Nonlinear Elastic Solids 399

3.0
500
Cycle 3
Cycle 1
Cycle 2 2.5
400 Cycle 3
Cycle 4
Cycle 5 2.0 First

Stress (GPa)
pop-in
Load (mN)

300
1.5

200
1.0 Cycle 1
Cycles Cycle 2
Cycle 5
100 1 and 2 Cycles
4 and 5 0.50

0 0.0
0 500 1000 1500 2000 2500 0.0 0.050 0.10 0.15 0.20 0.25
(a) Displacement (nm) (c) Strain (a/R )

10
0.6

8 0.5 20 GPa;
Elastic response
Contact stress (GPa)
61 GPa elastic
modulus
Stress (GPa)

Cycles 0.4
6 1 and 2 Cycle 3
(red) 50
0.3
4
Cycle 1
0.2 Cycle 1
Cycle 2 Cycle 2
Cycle 3
Cycle 3
2 Cycle 4
Cycle 5 0.1
Cycles
4 and 5
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.05 0.1 0.15 0.2 0.25
(b) Strain (a /R ) (d) Strain (a /R )

FIGURE 13.12 Load–displacement curves obtained when single-crystal mica samples — loaded parallel to
the c-axis with a 13.5 µm radius spherical nanoindenter — were indented. (a) Grade A. (b) Stress–strain curves
of results shown in (a). (c) Grade B. Here the response is fully reversible, but the difference between the first
and subsequent cycles is small. Small gap on unloading is an artifact of the experiment,1 (d) Stress–strain
curves for Grade C mica. Here the note slope on initial loading is roughly half than that in (c).4

moved, which in turn implies delaminations. The pop-in also most probably resulted in a penny-shaped
delaminated volume under the indenter. Why no evidence for IKB formation — i.e., fully reversible
hysteretic loops — is observed is not clear. One possibility is that the domains formed after the mas-
sive pop-ins are too small for an IKB to nucleate in them. For the latter to occur, a higher stress is prob-
ably needed; a situation not unlike the one shown in Figure 13.6a. These comments notwithstanding,
more work is needed, some of which is ongoing, to understand better this intriguing behavior.
When a lesser grade mica (Grade B) was used, the response was of type II1 (Figure 13.12c). It
is important to note that the small openings at the end of the repeat cycles shown in this figure are
artifacts of the experiment; the actual load–displacement response (not shown) is fully reversible.
This was confirmed several times by sometimes indenting the same location repeatedly — in one
case up to 100 cycles — and then examining the indentation location in the SEM. In all cases, no
trace of the indentation was found, which would not have been the case had the response not been
fully reversible. Another important clue is the area encompassed within the load–displacement
curves: for a KNE solid this areas asymptote to a constant value, after a few cycles.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 400

400 Nanomaterials Handbook

When an even more defective mica (Grade C) was indented, the response was unique (Figure
13.12d).4 The modulus upon loading was significantly lower than the one measured in Figure
13.12c, i.e., < c33. The reason for this state of affairs is not entirely clear at this time, but suggests
that the nucleation of IKBs in this mica is significantly easier than in either Grade A or B.
As noted above, in some cases, massive pop-ins — of the order of 2 µm (Figure 13.12a) —
were observed. SEM micrographs of the indented regions (Figure 13.11c) showed large craters

h-BN
10

Cycle 1
Cycle 2
Load (mN)

6 Cycle 3
Cycle 4
Cycle 5

0
0 500 1000 1500 2000 2500 3000 3500
(a) Displacement (nm)

h-BN
40

30
Stress (MPa)

20

Cycle 1
10 Cycle 2
Cycle 3
Cycle 4
Cycle 5

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
(b) Strain (h /a)

FIGURE 13.13 (a) Load–displacement curves of polycrystalline h-BN samples loaded to 10mN with a 13.5µm
radius spherical nanoindenter.34 (b) Stress–strain curves of results shown in (a). Note that here the penetration depth
is so large as to make Equation 13.12 invalid, as evidenced by the negative slopes upon initial unloading.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 401

Nanolayered or Kinking Nonlinear Elastic Solids 401

reminiscent of those formed in graphite (Figure 13.11a). Here again, the nanolaminate nature of
mica, direct evidence for kinks and, as important, the break up of the single crystal into smaller ones
is unambiguous.

13.5.5 HEXAGONAL BORON NITRIDE


Typical stress–strain curves derived from nanoindentation load–displacement (Figure 13.13a)
curves in polycrystalline h-BN are shown in Figure 13.13b. The response here is unlike any
other material explored to date. The response is initially elastic (solid line) is followed by
yield and considerable hardening. Upon reloading, fully reversible loops are observed. In
this case, however, the penetration is so deep that Equation 13.12 breaks down. This is manifested
by slopes that, upon initial unloading, are too steep to be physically meaningful. Note the same
problem is encountered in Figure 13.12b and may explain the positive curvature of the curves.

1000
Nanoindentation

Mica
100
Irreversible work, Wd (MJ/m3)

10

1
Uniaxial compression

Graphite
Fine-grained Ti3SiC2
0.1

Coarse-grained Ti3SiC2
0.01

0.001
10 100 1000 10000 100000
(a) Maximum stress (MPa)

Nanoindentation
10−3 results

10−4

10−5
Wd /E

Graphite
10−6

−7 Ti3SiC2
10

10−8

0.0001 0.001 0.01 0.1


(b) /E

FIGURE 13.14 Log–log plot of (a) Wd vs. σmax for select KNE solids. (b) Same plot as in (a) but with axes
normalized by appropriate Young’s modulus (bulk) or c33 for single-crystal nanoindentation work.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 402

402 Nanomaterials Handbook

13.6 BULK VS. NANOINDENTATION RESULTS


Figure 13.14 is a log–log plot of Wd vs. σ for a number of solids obtained from nanoindentation
experiments and compression of polycrystalline bulk samples such as those shown in Figure 13.2.
The agreement between the two sets of results for Ti3SiC2 and graphite over five orders of magni-
tude is gratifying and confirms that our procedure to converting load–displacement curves to
stress–strain is reasonably valid. When the results shown in Figure 13.14a are normalized by the
appropriate elastic constants the different curves collapse on one universal curve (Figure 13.14b)
strongly suggesting that the same underlying mechanism — kinking — is operative in all cases.1

13.7 SUMMARY AND CONCLUSIONS


Kinking nonlinear elastic solids are an important newly identified and fairly ubiquitous class of
solids. Repeated spherical nanoindentations, into the same location, on single and polycrystalline
samples is a powerful technique to identify KNE solids and obtain important information about
what is occurring under the indenter. The technique also clearly delineates the important and cru-
cial role that IKBs and KBs play in the deformation of these solids.

ACKNOWLEDGMENTS
This work was funded by the NSF (DMR-0503711) and ARO (DAAD19-03-1-0213). The help of
my current (S. Basu, A. Zhou) and former (A. Murugaiah) graduate students in preparing this man-
uscript is greatly appreciated. Many fruitful discussions with my colleague Prof. S. Kalidindi were
also critical is shaping this manuscript.

REFERENCES
1. Barsoum, M.W., Murugaiah, A., Kalidindi, S.R., and Zhen, T., Phys. Rev. Lett., 92, 255508, 2004.
2. Barsoum, M.W., Murugaiah, A., Kalidindi, S.R., and Gogotsi, Y., Carbon, 42, 1435–1445, 2004.
3. Murugaiah, A., Barsoum, M.W., Kalidindi, S.R., and Zhen, T., J. Mater. Res., 19, 1139–1148, 2004.
4. Murugaiah, A., Basu, S., Kalidindi, S.R., and Barsoum, M.W., submitted, 2005, unpublished results.
5. Orowan, E., Nature, 149, 463–464, 1942.
6. Hess J.B., and Barrett, C.S., Trans. AIME, 185, 599–606, 1949.
7. Barsoum M.W. and El-Raghy, T., Metall. Mater. Trans., 30A, 363–369, 1999.
8. Frank, F.C. and Stroh, A.N., Proc. Phys. Soc., 65, 811–821, 1952.
9. Barsoum, M.W., Zhen, T., Kalidindi, S.R., Radovic, M., and Murugahiah, A., Nature Mater., 2,
107–111, 2003.
10. Barsoum, M.W., Zhen, T., Zhou, A., Basu, S., and Kalidindi, S.R., Phys. Rev. B., 71, 134101, 2005.
11. Barsoum, M.W. and El-Raghy, T., Am. Sci., 89, 336–345, 2000.
12. Barsoum, M.W. and Radovic, M., in Encyclopedia of Materials Science and Technology, Buschow,
R.W.C.K.H.J., Flemings, M.C., Kramer, E.J., Mahajan, S., and Veyssiere, P., Eds., Elsevier,
Amsterdam, 2004.
13. Barsoum, M.W., Prog. Solid State Chem, 28, 201–281, 2000.
14. El-Raghy, T., Zavaliangos, A., Barsoum, M.W., and Kalidindi, S.R., J. Amer. Cer. Soc., 80, 513–516,
1997.
15. Barsoum, M.W., Farber, L., El-Raghy, T., and Levin, I., Met. Mater. Trans., 30A, 1727–1738, 1999.
16. Kalidindi, S.R., Zhen, T., and Barsoum, M.W., submitted.
17. Zhou, A.G., Barsoum, M.W., Basu, S., Kalidindi, S.R., and El-Raghy, T., Acta Mater., in print.
18. Iwashita, N., Swain, M.V., Field, J.S., Ohta, N., and Bitoh, S., Carbon, 39, 1525–1532, 2001.
19. Johnson, K.L., Indentation Contact Mechanics, Cambridge, Cambridge University Press, 1985.
20. Basu, S., Barsoum, M.W., and Kalidindi, S.R., J. Appl. Phys., in print.
21. Courtney, T.H., Mechanical Behavior of Materials, McGraw-Hill, New York, 1990.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 403

Nanolayered or Kinking Nonlinear Elastic Solids 403

22. Li, J., Van Vliet, K.J., Zhu, T., Yip, S., and Suresh, S., Nature, 418, 307, 2002.
23. Van Vliet, K.J., Li, J., Zhu, T., Yip, S., and Suresh, S., Phys. Rev. B, 67, 104105, 2003.
24. Kelchner, C.L., Plimpton, S.J., and Hamilton, J.C., Phys. Rev. B, 58, 11085–11088, 1998.
25. Wachtman, J.B.J., Tefft, W.E., Lam, D.C.J., and Stenchfield, R.P., J. Res. NBS, 64A, 213, 1960.
26. Nowak R., and Sakai, M., Acta Mater., 47, 2879–2891, 1994.
27. Nowak, R., Sekino, T., Maruno, S., and Niihara, K., Appl. Phys. Lett., 68, 1063–1065, 1996.
28. Nowak, R., Manninen, T., Heiskanen, K., Sekino, T., Hikasa, A., Niihara, K., and Takagi, T., Appl.
Phys. Lett., 83, 5214–5216, 2003.
29. Tymiak, N.I., Daugela, A., Wyrobek, T.J., and Warren, O.L., Acta Mater., 52, 553–563, 2004.
30. Radovic, M., Ganguly, A., Barsoum, M.W., Zhen, T., Finkel, P., Kalidindi, S.R., and Lara-Curzio, E.,
submitted.
31. Murugaiah, A., Ph.D. thesis, Drexel University, 2004, p. 56.
32. Molina-Aldareguia, J.M., Emmerlich, J., Palmquist, J., Jansson, U., and Hultman, L., Scripta Mater.,
49, 2003.
33. Kelly, B.T., Physics of Graphite, Applied Science Publishers, London, 1981.
34. Zhen, T., Ph.D. thesis, Drexel University, 2004.
35. Holm, B., Ahuja, R., and Johansson, B., Appl. Phys. Lett., 79, 1450, 2001.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch013.qxd 11/30/2005 3:49 PM Page 404

Copyright 2006 by Taylor & Francis Group, LLC

You might also like