Sunteți pe pagina 1din 6

A Generalized Description of the Elastic

Properties of Nanowires
Andreas Heidelberg,

Lien T. Ngo,

Bin Wu,

Mick A. Phillips,

Shashank Sharma,

Theodore I. Kamins,

John E. Sader,

and John J. Boland*


,
Department of Chemistry and the Center for Research on AdaptiVe Nanostructures and
NanodeVices (CRANN), Trinity College Dublin, Dublin 2, Ireland, Quantum Science
Research, Hewlett-Packard Laboratories, 1501 Page Mill Road, MS 1123, Palo Alto,
California 94304, and Department of Mathematics and Statistics, UniVersity of
Melbourne, Victoria 3010, Australia
Received January 6, 2006; Revised Manuscript Received April 3, 2006
ABSTRACT
We report a model of nanowire (NW) mechanics that describes force vs displacement curves over the entire elastic range for diverse wire
systems. Due to the clamped-wire measurement configuration, the force response in the linear elastic regime can be linear or nonlinear,
depending on the system and the wire displacement. For Au NWs the response is essentially linear since yielding occurs prior to the onset
of the inherent nonlinearity, while for Si NWs the force response is highly nonlinear, followed by brittle fracture. Since the method describes
the entire range of elastic deformation, it unequivocally identifies the yield points in both of these materials.
Knowledge and understanding of the mechanical properties
of nanowires (NWs) are extremely important for emerging
applications in the areas of nanocomposite strengtheners,
nanoscale interconnects, and the active components in
nanoelectromechanical (NEMS) devices. In recent years,
researchers have investigated the mechanical properties of
various nanowire systems using different techniques.
1-18
These include the use of a nanostressing stage within a
scanning electron microscope,
1,2
vibrational studies within
a transmission electron microscope (TEM),
3-5
and atomic
force microscope (AFM) manipulations.
6-18
Each involves
pinning the nanowire at a particular location and measuring
its response to thermal excitation or mechanical deformation.
The clamped-clamped beam configuration has been widely
used to study the mechanical properties of NWs.
6-14
Using
this configuration, we recently reported a new AFM-based
method that enables the full spectrum of NW mechanical
properties to be measured, including the elastic properties,
plastic deformation, and failure.
11
This technique involved
clamping nanowires over trenches on a SiO
2
substrate and
then measuring their response under the lateral load from
an AFM tip (Figure 1). This method has several advantages
over those reported earlier in that it employs well-defined
clamping points and eliminates wire-substrate friction
effects.
Despite the simplicity and widespread use of the clamped-
beam configuration, vastly different force (F) vs displacement
(d) curves have been reported for different systems. For
example, in the case of metallic NWs, we reported F-d
curves that were linear for small displacements after which
there was an abrupt change in slope associated with the yield
point in these materials.
11
Linear F-d curves were also
reported for single-wall carbon nanotube bundles,
7,10,14
mul-
tiwalled carbon nanotubes,
6,9
MoS
2
ropes,
8
and even litho-
graphically defined Si beams.
12,13
Nonlinear F-d curves have
also been reported in measurements involving carbon
* To whom correspondence may be addressed. E-mail: jboland@tcd.ie.

Department of Chemistry and the Center for Research on Adaptive


Nanostructures and Nanodevices, Trinity College Dublin.

Quantum Science Research, Hewlett-Packard Laboratories.

Department of Mathematics and Statistics, University of Melbourne.


Figure 1. Schematic of the experimental setup for lateral nanowire
manipulation: (a) before the manipulation and (b) during the
manipulation. The nanowire under investigation is suspended over
a trench in the substrate and fixed by Pt deposits at the trench edges.
During the manipulation the lateral and normal cantilever deflection
signals are simultaneously recorded.
NANO
LETTERS
2006
Vol. 6, No. 6
1101-1106
10.1021/nl060028u CCC: $33.50 2006 American Chemical Society
Published on Web 05/13/2006
nanotubes
15-17
and for poly(ethylene oxide) and glass nanofi-
bers.
18
These observations have led to the use of models that
treat F-d curves as being exclusively linear or nonlinear.
The limitations of using these exclusively linear or nonlinear
models are yet to be investigated in the literature. Further-
more, a comprehensive model that accounts for the detailed
shape of these F-d curves over the entire elastic region is
yet to be described and validated for nanowire systems.
Here we provide a complete description of the elastic
properties in a double-clamped beam configuration over the
entire elastic regime for diverse wire systems. Within this
model we show that it is possible to perform a comprehensive
analysis of force-displacement (F-d) curves using a single
closed-form analytical description and not only to extract
linear material constants such as the Youngs modulus (E)
but also to describe the entire elastic range and hence to
identify the yield points for dramatically different systems
such as Si and Au nanowires. The E-modulus obtained for
the Si nanowires is diameter independent and comparable
to that of bulk Si. We demonstrate that the linear F-d curves
previously reported for Au NWs are a consequence of
mechanical yielding at small displacements prior to the onset
of a nonlinear geometric effect associated with tensile
stretching. In the case of Si NWs the F-d curves are highly
nonlinear at large deformations and exhibit brittle failure
without measurable plastic deformation.
Si nanowires were grown epitaxially in 111 directions
on (100) oriented Si substrates using chemical vapor deposi-
tion catalyzed by Au nanoparticles.
19
The nanowires were
then sonicated off the supporting substrate and dispersed in
IPA. Au NWs were prepared by electrochemical deposition
into pores of highly ordered anodized aluminum oxide
(AAO) templates
11,20
and were dispersed in toluene after
being liberated from the template. Both types of wires were
then deposited over trenches on a SiO
2
substrate and clamped
at the trench edges by electron beam induced deposition of
Pt. The clamping points are robust and, although they carry
the largest tensile or compressive stresses, do not show failure
or movement after manipulations. After sample preparation,
AFM lateral manipulations were carried out using a Digital
Instruments Nanoman System with closed-loop x-y-z-
scanner. During the manipulation the lateral and normal force
signals were recorded. A more detailed description of the
substrate preparation and manipulation procedures can be
found elsewhere.
11
The upper panel in Figure 2 shows a
typical F-d curve during the bending test for a Si nanowire
with a radius r of 75 nm. On approach of the cantilever to
the nanowire, the lateral force acting on the AFM cantilever
is essentially zero as seen in Figure 2. At the point of contact
between AFM tip and Si nanowire, the force increases
linearly at first, but after a displacement d, of about 75 nm
a nonlinear increase can be observed. (The normal load
applied by the AFM tip to the nanowire was also measured,
but it had no significant influence on the nanowire mechanics
and is omitted for clarity.) Increased loading of the nanowire
ultimately results in failure without plastic deformation,
consistent with a brittle material like Si. The lower panel in
Figure 2 shows the F-d curve obtained for a Au NW (r )
113 nm). The latter is characterized by a linear region
followed by an abrupt change in slope previously attributed
to the yield point of the material, followed by plastic
deformation. In contrast with the Si case, there is little
evidence for nonlinear behavior in the elastic region of the
F-d curve.
To account for these different behaviors, we have to
consider that in linear elastic beam bending theory, the
Youngs modulus is normally obtained from a measurement
of the beam displacement as a function of the applied load.
For a clamped-clamped beam used here, the resulting curve
is described by the following well-known equation
21
where F
center
is the load applied to the center of the beam,
z
center
is the resulting displacement of the beam at the load
point, E is the Youngs modulus of the beam, I ) R
4
/4 is
the moment of inertia (assuming a cylindrical beam), and L
is the beam length. Importantly, this equation predicts a linear
relationship between the applied load and the resulting
displacement and thus requires that measurements also
exhibit this property. However, as the beam is displaced, an
axial tensile force is inherently induced due to the stretching
of the beam. This force affects the total stress experienced
by the beam, leading to an enhancement of its rigidity.
21
Here
we present a simple closed form solution to this problem
that is valid for any beam displacement, enabling the
determination of the Youngs modulus irrespective of the
applied load.
Figure 2. Comparison of typical F-d curves obtained by bending
experiments for a Si and a Au nanowire. (upper panel) F-d curve
recorded during manipulation of a 75 nm radius Si nanowire. After
an initial linear rise upon contact of the AFM tip, the F-d curve
clearly shows the nonlinear elastic behavior of the Si wire, followed
by brittle failure of the structure. (Lower panel) F-d curve taken
during a manipulation of a 113 nm radius Au wire. A pseudolinear
elastic region is visible followed by a yield point after which plastic
deformation occurs. The wire was unloaded prior to failure.
F
center
)
192EI
L
3
z
center
(1)
1102 Nano Lett., Vol. 6, No. 6, 2006
To begin we note that the governing beam equation,
21
which includes the effects of axial forces, is
where w is the deflection of the beam, x is the spatial
coordinate along the beam length, T is the tension along the
beam, and is the Dirac delta function. Note that z
center
)
w(L/2). The tension T is determined from the strain along
the axis of the beam, and is given by
Equations 2 and 3 are solved subject to the usual clamped
boundary conditions at both ends of the beam, i.e.
Solving eqs 2-4 gives the required solution
where
where R is related to the maximum deflection z
center
by the
following transcendental equation
We note that f(R) g1, so that inclusion of the tensile force
increases the effective rigidity of the beam. Also, R is directly
connected to the tension in the rod via the relation
and represents the ratio of axial stresses due to extension
and bending.
Whereas eqs 5-7 present an exact analytical solution,
21
it is complex in nature, requiring the numerical solution of
a transcendental equation. To overcome this limitation, we
note that the solution to eq 7 possesses the following
asymptotic solutions
where
An approximate and explicit expression for R is then
obtained by constructing a Pade approximant using eq 8,
i.e.
which exhibits a maximum error of only 2.1% over the entire
range of . Equations 5, 6, and 10 give the generalized form
we seek and enable the applied force to be determined for
any beam displacement.
Equation 9 indicates that the solution, eq 1, is valid
provided the maximum displacement z
center
is less than the
radius of the beam. As z
center
is increased to a value larger
than the beam radius, the solution becomes progressively
nonlinear with respect to the displacement, leading finally
to a cubic dependence on z
center
. Thus, an approximate
solution can be obtained by superimposing the small and
large deflection limits, namely
a result which exhibits a maximum error of 18%. This
approximate solution (which is not used in any analysis
detailed here) contains both linear and nonlinear terms widely
reported in the literature. The linear dependence accounts
for elastic bending of the beam, whereas the cubic depen-
dence describes the tensile stresses that are induced along
the length of the beam at large displacements.
Figure 3a shows a plot of the rigidity enhancement
function f(R) as a function of (which is directly
proportional to the maximum displacement, see eq 9). In parts
b-d of Figure 3 the effect of this nonlinearity on the
measured force curves for a cylindrical beam is illustrated.
First, we consider the case where the maximum deflection
is one radius (Figure 3b). It can clearly be seen that for
deflections smaller than half of one radius, the standard
theory (eq 1) accurately predicts the behavior of the
nanowire, with observable deviations occurring as the
deflection approaches one radius. Note that the results have
been normalized so that a slope of unity corresponds to the
standard linear theory. Next, we show the force curve for a
slightly larger extension up to twice the radius of the
nanowire (Figure 3c). As can be seen, the curve becomes
increasingly nonlinear as the deflection is increased past one
radius. For larger deflections, the nonlinearity grows, leading
EI
d
4
w
dx
4
- T
d
2
w
dx
2
) F
center
(x - (L/2)) (2)
T )
EA
2L

0
L
(
dw
dx
)
2
dx (3)
w )
dw
dx
) 0 at x ) 0 and L (4)
F
center
)
192EI
L
3
f(R)z
center
(5)
f(R) )
R
48 -
192 tanh(R /4)
R
(6)
R cosh
2
(R /4)
2 + cosh(R /2) - 6
sinh(R /2)
R
(
1 - 4
tanh(R /4)
R
)
2
)
z
center
2
(
A
I
)
(7)
R )
TL
2
EI
R )
{
12
5
-
3
875

2
, f0
2 , f
(8)
) z
center
2
(
A
I
)
(9)
R )
6(140 + )
350 + 3
(10)
F
center
)
192EI
L
3
z
center(
1 +
A
24I
z
center
2
)
(11)
Nano Lett., Vol. 6, No. 6, 2006 1103
to an ultimate cubic-like dependence on the deflection (see
Figure 3d). However, it is important to emphasize that the
gradual change in the slope of the F-d curve makes it
difficult to exclusively identify linear and nonlinear regions
and necessitates the use of the generalized eqs 5, 6, and 10,
which are valid for all displacements in the elastic region.
Whenever the linear or nonlinear term is used by itself, as
is presently the case in the literature, it provides a very poor
description of the mechanical properties. For example, an
exclusively linear treatment fails to recognize that the F-d
data beyond the linear region is still the result of elastic
deformation, except that the dominant stresses are now tensile
and associated with bending the wire beyond one radius.
This model can be directly compared with the force curves
recorded using Si and Au nanowires.
22
Figure 4a shows a
fit (using the generalized eqs 5, 6, and 10) for the F-d curve
for the Si wire shown in Figure 2a, and which gives an
E-modulus of 190 GPa. The only adjustable parameter in
the fit was the E-modulus. It is important to note that the
excellent fit in Figure 4a covers the entire displacement
range, from the first contact point between tip and nanowire
to the sharp force-drop point, which indicates failure without
plastic deformation, hence demonstrating that Si nanowires
are brittle materials. In Figure 5 the E-moduli determined
for nanowires with radii between 50 and 100 nm are shown.
The average modulus E )158 GPa is essentially independent
of the radius over the range investigated and close to the
value reported for bulk Si(111) (169 GPa).
23
The scatter in
the data presented here can be explained by the fact that the
measured E-modulus is dominated by estimates of the
physical dimensions of the wire and AFM cantilever/tip
11
and is not a reflection of the model itself. Indeed the accuracy
of the model is significantly better than that of present
experimental methods, which are limited by physical dimen-
sion measurements, but is expected to become more impor-
tant as better nanomechancial techniques are developed.
The real importance of this method is the ability to
unambiguously identify the elastic deformation regime for
a particular material. For example, in the case of the Au
nanowire system, we previously demonstrated that the force
response is essentially linear and suggested that the change
in slope after this initial linear region was due to yielding.
11
This analysis is borne out by the application of the present
model to this system. The lower curve in Figure 2 shows
the F-d curve for an Au nanowire (r ) 113 nm) that is
essentially linear up to an obvious break point, which occurs
at a wire displacement of about 110 nm. Figure 4b shows
that these data can be fit using eqs 5, 6, and 10 up to the
break point, thus demonstrating the break point is indeed
the yield point of the material.
Given the data for a particular Au NW, and assuming the
modulus is a materials constant for the system under study,
this model can be used to predict the generalized elastic
behavior for any Au nanowire. The fact that the Au NW in
Figure 4b yielded at a wire displacement of 110 nm reflects
the detailed microstructure of the wire. Hardening of the wire,
whether by grain boundary or impurity hardening, will
increase the yield point but the F-d data must still follow
the model (red curve) in Figure 4b, except that the point
where the force data deviates from the model occurs at larger
displacements (shown as open circles in Figure 4b). Using
this approach, it is possible to accurately study the evolution
Figure 3. (A) Plot of the rigidity enhancement function f(R) as a function of . (B) Calculated F-d curve for a circular cylinder showing
the nonlinearity for a maximum deflection of one radius. The force values have been normalized so that a slope of unity corresponds to the
standard linear theory. (C) Simulated and normalized F-d curve for a maximum displacement of two radii. (D) F-d curve for large
deflections showing a growing nonlinearity leading to the ultimate cubic dependence on the deflection.
1104 Nano Lett., Vol. 6, No. 6, 2006
of the mechanical properties of a family of NWs through a
series of work-hardening or annealing treatments.
25
In conclusion we have introduced a model which ac-
curately accounts for the mechanical properties of nanowires
in a clamped-clamped beam configuration over the entire
elastic range and provides a comprehensive methodology for
the analysis of a broad range of nanowire systems. A
nonlinear response is expected for small wires such as single
carbon nanotubes for all significant displacements whereas
an initial linear response followed by an approximate cubic
dependence is expected for larger diameter wires. In general,
however, a detailed description of such F-d curves requires
a solution of the generalized equations developed here.
Whether this nonlinear behavior is observed depends on the
intrinsic mechanical properties of the wire, and for systems
such as metals that are susceptible to mechanical yielding
the nonlinear response is suppressed. Even in this case,
however, for an accurate measurement of elastic properties
it is necessary to account for nonlinear effects that result
from build-up of tensile stresses in the wire. Finally, this
model allows the yield point to be identified in NW systems
and hence facilitates studies of the evolution of NW
mechanical properties during work hardening and annealing.
Acknowledgment. The work at the Department of
Chemistry and the Center for Research on Adaptive Nano-
structures and Nanodevices at Trinity College Dublin is
supported by Science Foundation Ireland under Grant 00/
PI.1/C077A.2. The research by J.E.S. is supported by a
Science Foundation Ireland Walton Fellowship and was
performed while on leave at Trinity College Dublin. The
work at Hewlett-Packard is partially supported by the U.S.
Defense Advanced Research Projects Agency (DARPA).
References
(1) Yu, M.-F.; Lourie, O.; Dyer, M. J.; Moloni, K.; Kelly, T. F.; Ruoff,
R. S. Science 2000, 287, 637-640.
(2) Yu, M.-F.; Files, B. S.; Arepalli, S.; Ruoff, R. S. Phys. ReV. Lett.
2000, 84, 5552-5555.
(3) Krishnan, A.; Dujardin, E.; Ebbesen, T. W.; Yianilos, P. N.; Treacy,
M. M. J. Phys. ReV. B 1998, 58, 14013-14019.
(4) Poncharal, P.; Wang, Z.-L.; Ugarte, D.; de Heer, W. A. Science 1999,
283, 1513-1516.
(5) Chopra, N. G.; Zettl, A. Solid State Commun. 1998, 105, 297-300.
(6) Wong, E.-W.; Sheehan, P. E.; Lieber C. M. Science 1997, 277, 1971-
1975.
(7) Kis, A.; Kasas, S.; Babic, B.; Kulik, A. J.; Benoit, W.; Briggs, G. A.
D.; Schonenberger, C.; Catsicas, S.; Forro, L. Phys. ReV. Lett. 2002,
89, 248101.
(8) Kis, A.; Mihailovic, D.; Remskar, M.; Mrzel, A.; Jesih, A.; Piwonski,
I.; Kulik, A. J.; Benoit, W.; Forro, L. AdV. Mater. 2003, 15, 733-
736.
(9) Salvetat, J.-P.; Kulik, A. J.; Bonard, J. M.; Briggs, G. A. D.; Stockli,
T.; Metenier, K.; Bonnamy, S.; Beguin, F.; Burnham, N. A.; Forro,
L. AdV. Mater. 1999, 11, 161-165.
(10) Salvetat, J.-P.; Bonard, J. M.; Thomson, N. H.; Kulik, A. J.; Forro,
L.; Benoit, W.; Zuppiroli, L. Appl. Phys. A 1999, 69, 255-260.
(11) Wu, B.; Heidelberg, A.; Boland, J. J. Nat. Mater. 2005, 4, 525-
529.
(12) Namazu, T.; Isono, Y.; Tanaka, T. J. Microelectromech. Syst. 2000,
9, 450-459.
(13) Sundararajan, S.; Bushan, B.; Namazu, T.; Isono, Y. Ultramicroscopy
2002, 91, 111-118.
(14) Salvetat, J.-P.; Briggs, G. A. D.; Bonard, J. M.; Bacas, R. R.; Kulik,
A. J.; Stockli, T.; Burnham, N. A.; Forro, L. Phys. ReV. Lett. 1999,
82, 944-47.
(15) Walters, D. A.; Ericson, L. M.; Casavant, M. J.; Liu, J.; Colbert, D.
T.; Smith, K. A.; Smalley, R. E. Appl. Phys. Lett. 1999, 74, 3803-
05.
Figure 4. Curve fits for Si and Au F-d curves to eqs 5, 6, and
10. The only adjustable parameter in the fits is the E-modulus. (A)
For a Si nanowire with a radius of 75 nm, the curve fit gives an
E-modulus of 190 GPa. (B) For a Au wire with a radius of 113
nm, an E-modulus of 75 GPa is obtained by fitting to the initial
part of the F-d curve. The extrapolated fit to the curve directly
confirms that the Au wire yields before the onset of appreciable
nonlinear behavior. Predicted F-d behavior of a hypothetical
hardened Au NW, with the same E-modulus, is shown with open
circles.
Figure 5. Plot of the E-modulus versus the nanowire radius in the
range of 50-100 nm for Si nanowires, showing that the observed
E-modulus is radius independent (within the limits of the scatter
in the data), with an average value of 158 GPa is close to the
modulus for bulk Si(111) of 169 GPa (solid line).
24
Nano Lett., Vol. 6, No. 6, 2006 1105
(16) Tombler, T. W.; Zhou, C.-W.; Alexseyev, L.; Kong, J.; Dai, H.-J.;
Lei, L.; Jayanthi, C. S.; Tang, M.-J.; Wu, S.-Y. Nature 2000, 405,
769-72.
(17) Minot, E. D.; Yaish, Y.; Sazonova, V.; Park. J. Y.; Brink, M.;
McEuen, P. L. Phys. ReV. Lett. 2003, 90, 156401.
(18) Bellan, L. M.; Kameoka, J.; Craighead, H. G. Nanotechnology 2005,
16, 1095-1099.
(19) Sharma S.; Kamins, T. I.; Williams, R. S. Appl. Phys. A: Mater.
Sci. Process. 2005, 80, 1225-29.
(20) Masuda, H.; Fukuda, K. Science 1995, 268, 1466-68.
(21) Landau, L. D.; Lifshitz, E. M. Theory of Elasticity; Pergamon:
Oxford, 1970.
(22) This analysis is valid when the spanning length of the nanowire over
the trench is much larger than its diameter. Experimentally, the trench
width is restricted since very wide trenches cause the nanowires to
droop. For aspect ratios of 10-20 (typical for our experiments), we
found these equations to be valid. Au nanowires with very large
length-to-diameter ratios (>20) can be pushed more than one diameter
before they yield, and F-d curves of these nanowires show nonlinear
behavior before onset of inelastic deformation. For the Au nanowires
studied here, this onset is obvious, but for materials that yield less
dramatically or for wire configurations that have wide range of aspect
ratios, this analysis may be exceptionally useful.
(23) Wortman, J. J.; Evans, R. A. J. Appl. Phys. 1965, 36, 153-156.
(24) Callister, W. D., Jr. Materials Science and Engineering; Wiley: New
York, 1994.
(25) Wu, B.; Heidelberg, A.; Boland, J. J.; Sader, J. E.; Sun, X.-M.; Li,
Y.-D. Nano Lett. 2006, 6, 468-472.
NL060028U
1106 Nano Lett., Vol. 6, No. 6, 2006

S-ar putea să vă placă și