Mechanical Properties of Nanostructures: Synonyms

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Dr. Laurent Pizzagalli, Dr. Sandrine Brochard, Dr. Julien Godet and Dr.

Celine Gerard
Mechanical Properties of Nanostructures SpringerReference

Mechanical Properties of Nanostructures


Synonyms
Elastic and plastic behavior of materials with nanometric dimensions; Properties related to applied stress or strain

Definition
In the context of materials science, the usual definition of a nanostructure implies that at least one of its dimension ranges
from few nanometers to at most few hundreds. Then, according to the numbers of these dimensions, it could be
(non-exhaustive list) a nanofilm, a membrane, a nanoplate, a multilayer (one dimension), a nanowire, a nanopillar, a
nanobridge, a nanotube (two dimensions), or a nanoparticle, a nanograin (three dimensions). The response of these
systems to an applied stress (or equivalently to a deformation) depends on their mechanical properties. Those include the
elastic regime, for which changes induced in the material are reversible, and plastic regime, for which they are not.

Concepts
Nanostructures Versus Bulk Materials

The main difference between macroscale materials and nanostructures lies in the significant role of surfaces (or
interfaces) for the latter. Atoms located at surfaces have typically fewer neighboring atoms than in bulk, leading to marked
changes of the local electronic structure. Structural modifications usually occur, thanks to surface relaxation or
reconstruction, both mechanisms leading to surface energy minimization, but also adding an extra contribution to surface
stress. Besides, surfaces are often reactive, and most of nanostructure surfaces are covered with passivants or oxide
layers, except in a well-controlled environment.
Surfaces and interfaces can have a strong impact on mechanical properties of nanostructures. In the elastic regime, the
moduli that link stress to strain can change compared to their bulk counterparts, due to the surface stress. The latter will
either facilitate or hinder the deformation of the nanostructure, thus decreasing or increasing the elasticity coefficients. In
the plastic regime, deformation mechanisms can be different in bulk and in nanostructures. For instance, the dislocation
starvation process has been proposed to explain the first stages of plastic deformation in nanostructures [1]. Also,
surfaces act as nucleation centers for dislocations [2]. This is also true for boundaries between nanograins, which can
emit or absorb dislocations [3] or favor other deformation mechanisms like twinning for instance [4]. Furthermore surface
diffusion may play an important role in the plastic deformation of thin nanowires.
In addition to surfaces/interfaces, the mechanical behavior of nanostructures is influenced by reduced dimensions. In fact,
the number of preexisting defects which can act as sources for plastic deformation mechanisms scales with the volume of
the nanostructure. This source exhaustion is often proposed as an explanation for the hardening in small nanostructures,
as reported in many cases [5]. Reduced dimensions can also hinder common bulk mechanisms. At last, quantum
confinement effects can become significant in nanostructures with a dimension of few nanometers. For instance they
have been shown to influence elastic moduli [6].
Another phenomenon associated to nanostructures is phase transformation. In fact, in certain systems, phases
metastable in bulk can be stabilized in nanostructures due to the large surface (or interface) to bulk ratio. This has
obvious consequences on the mechanical properties. Finally, a special set of nanostructures includes systems which are
characterized by a specific structure, as nanotubes or fullerenes, for example, inducing specific mechanical properties.
As a consequence of all the aforementioned points, the general response of nanomaterials to a mechanical solicitation
can be much different than their bulk counterparts. This is spectacularly exemplified in the case of covalent materials like
semiconductors. Although they are usually brittle at room temperature in bulk state, small systems can exhibit ductility
instead [7, 8]. Overall, the mechanical behavior is greatly changed by the reduction of dimensions [9].

Molecular Modeling Versus Experiments

In molecular modeling, investigated system is described at an atomistic level, i.e., as a collection of atoms. Molecular

http://www.springerreference.com/index/chapterdbid/435247 20 Apr 2015 14:02


Springer-Verlag Berlin Heidelberg 2015 1
Dr. Laurent Pizzagalli, Dr. Sandrine Brochard, Dr. Julien Godet and Dr. Celine Gerard
Mechanical Properties of Nanostructures SpringerReference

modeling is complementary to experiments for several reasons. First, it allows to probe the mechanical properties of very
small nanostructures, in a dimension range hardly attainable experimentally. Then, in the simulations, it is possible to
follow the trajectories of all atoms in the system, which provides a wealth of information. Activated mechanisms can be
thoroughly analyzed. Also, the time resolution is of the order of a femtosecond, much lower than in common mechanical
testing experiments. This point is particularly important in the plastic regime, where plasticity mechanisms such as
dislocation nucleation or fracture propagation can occur at very short timescale. Finally, molecular modeling allows a full
control of the initial system and of the mechanical test. The effect of specific features or conditions, even beyond
experimental realizations, can then be finely examined.

Methodologies
Different Theoretical Frameworks

There are two main classes of molecular modeling frameworks. In the first one, commonly called "classical," the
electronic structure is not calculated and only the forces between atoms are considered. These are typically computed
from empirical interatomic potentials, which are usually fitted on known bulk materials properties. This is often the
selected approach for modeling nanostructures, since it allows to describe systems whose dimensions are within the
reach of experiments. With this approach, multicomponent systems can be more difficult to deal with, since appropriate
potentials must be available to model all different atomic interactions.
In the second class of molecular modeling frameworks, the electronic structure is calculated and taken into account into
the modeling. There are several ways to do it with different levels of approximation. In the case of nanostructure
modeling, tight-binding and density functional theory are possible choices. However, although providing accurate and
reliable results, only small nanostructures with dimensions of the order of several nanometers can be dealt with these
approaches. As a consequence, to date most of the molecular modeling investigations have been performed using
interatomic potentials.

Technical Setup

Prior to the numerical simulation step, one has to build the initial system by setting the positions in space of all needed
atoms. This allows for a total control of the geometry and composition of the nanostructure. For instance, the Fig. 1a
shows a typical setup for the compression of a nanoparticle. An important issue for the system setup concerns the
boundary conditions used in the simulation. The most common choice is periodic boundary conditions (PBC). It can be
required depending on the selected theoretical framework. For instance, PBC are enforced when plane-waves density
functional theory simulations are made. This has obvious consequences on the modeling setup. PBC can be used to
model a nanostructure with one or several infinite dimensions (as one of the examples will show in the following). When
PBC are used, it is important to check possible artificial size effects due to the interaction of the system with its replicas.
Note that these potential size effects can often be dealt with by including a sufficient amount of void around the system,
which then exhibits surfaces, simply by increasing the size of the simulation cell.

http://www.springerreference.com/index/chapterdbid/435247 20 Apr 2015 14:02


Springer-Verlag Berlin Heidelberg 2015 2
Dr. Laurent Pizzagalli, Dr. Sandrine Brochard, Dr. Julien Godet and Dr. Celine Gerard
Mechanical Properties of Nanostructures SpringerReference

Fig. 1
Atomistic simulations of the flat punch compression of gold nanoparticles (Courtesy of D. Mordehai [10]). (a) Ball-and-stick
representation of the system setup (Au nanoparticle in gold and flat punch indenter atoms in gray) for a 4.9 nm high nanoparticle. (b)
Top view of the nanoparticle, before and after complete load. (c) Top view of the nanoparticle, before and after the first dislocation
nucleation events (i)-(iv) during compression (only atoms in free surfaces and in the defect cores are shown; atoms in stacking faults
are shaded in gray)

The mechanical properties of a given system can be obtained as its response to a mechanical load. Three different
classes of methods can be used to apply this load. In the first one (Fig. 2a, d), not only the nanostructure is modeled with
atoms but also the devices that transmit the load in a real experiment (such as plates or an indenter). This method is
probably the closest to experiments and allows for taking into account the additional mechanical response of the devices
or possible adhesion effects. Besides, devices can also be modeled as non-attractive or infinitely rigid. The mechanical
test can be strain or stress controlled, depending whether the position of devices is fixed, or an applied stress is used.
The drawback of this approach is the extra computational cost associated with the atoms used for modeling the devices.
A second option is to take advantage of PBC (Fig. 2b, e). In that case, strain- or stress-controlled testing is done by
simply setting the computational cell dimensions or applying the appropriate stress on the cell faces. With this method,
load-transmitting devices are not considered. This is best suited to determine intrinsic nanostructure properties such as
elastic constants, except for nanoparticles for which void is necessary present around the whole system. Finally, in the
third method, the load is transmitted to the nanostructure by either applying a force field on selected atoms (stress
controlled) or by fixing the positions of these atoms (strain controlled). The latter is also called fixed boundary conditions
and can be useful in specific cases. For instance, it does not rely on the use of PBC (as the first method), and additional
atoms are not required.

http://www.springerreference.com/index/chapterdbid/435247 20 Apr 2015 14:02


Springer-Verlag Berlin Heidelberg 2015 3
Dr. Laurent Pizzagalli, Dr. Sandrine Brochard, Dr. Julien Godet and Dr. Celine Gerard
Mechanical Properties of Nanostructures SpringerReference

Fig. 2
Possible ways to load a nanostructure (yellow) in a molecular modeling simulation. (a) and (d): the load is transmitted by an "indenter"
(brown) made of atoms. (b) and (e): the load is transmitted by the simulation cell. (c) and (f): the load is directly applied on the
nanostructure. In (a-c), the mechanical testing is strain controlled by changing the dimension h, eventually rescaling atoms positions,
whereas it is stress controlled in (d-f). Simulations are usually performed in strain-controlled mode, easier to use than the
stress-controlled mode

Molecular Dynamics or 0 K Force Relaxation

Simulations are often performed at finite temperature using molecular dynamics. In this framework, the dynamical
evolution of all atoms in the system can be calculated by solving the Newton equations of motion, interatomic forces being
obtained with one of the aforementioned formalisms. In order to compare with experiments, simulations are typically done
in the canonical ensemble, using a thermostat to keep the temperature to the desired value. Generally, a simulation starts
by a thermalization step, followed by an increasing applied strain (or stress) on the nanostructure. Important quantities
such as the energy variation, the positions of atoms, or the local stress tensor can be monitored during the test for further
analyses.
Besides, zero temperature simulations can also be performed (also known as molecular mechanics). For instance, this is
done to compute elastic constants or to investigate the effect of stress on the plastic deformation without the stochastic
influence of thermal activation. In that case, forces are simply relaxed at a given strain (or stress), leading to a
configuration of minimum energy. To be complete, one can also mention that advanced simulation techniques such as
chain of states methods have also been applied for determining the mechanical properties of nanostructures [ 11].

Examples
Molecular modeling has been used to investigate the mechanical properties of many nanostructures. Most of these were
metallic systems or nanotubes, for which there are a lot of experimental data. But the literature also includes several
studies of nanostructures made of covalent or ionic materials, such as semiconductors or oxides. Furthermore, although

http://www.springerreference.com/index/chapterdbid/435247 20 Apr 2015 14:02


Springer-Verlag Berlin Heidelberg 2015 4
Dr. Laurent Pizzagalli, Dr. Sandrine Brochard, Dr. Julien Godet and Dr. Celine Gerard
Mechanical Properties of Nanostructures SpringerReference

most of the studied systems are quite ideal, there is an effort to consider more realistic or complex nanostructures in
molecular modeling simulations. Those could be core/shell systems, for instance, where the nanostructures are
composed of two different phases or materials.
In this entry, three examples of recent works focusing on the molecular modeling of the plasticity properties of
nanostructures are presented, in order to better illustrate the concepts developed in the previous sections.

Metallic Nanoparticles
The first example concerns the plastic deformation of facetted gold nanoparticles [10]. The latter can be experimentally
obtained as stand-alone nanostructures on a sapphire substrate by using a solid-state dewetting process and are next
compressed by a flat punch indenter. Equivalent computational systems have been built (Fig. 1a), the nanoparticle being
located between two infinitely rigid gold plates, which are then brought closer during a molecular dynamics simulation, in
a strain-controlled mode, with a constant rate of 1 m s1. An embedded atom model potential is used to model interatomic
interactions. The procedure used here corresponds to the case (a) in Fig. 2.
After complete compression of the nanoparticle, a flake-like structure is obtained (Fig. 1b), in excellent agreement with the
combined experiments. Nevertheless, the true power of the simulations is revealed when focusing on the plasticity onset.
Experimentally, one can only see the termination of the elastic regime due to a sudden and large strain burst. Instead,
atomistic simulations allowed to fully characterize the very first stages of plasticity. Hence it is revealed that plastic
deformation first occurred by the nucleation of leading Shockley partial dislocations from the upper vertices of the
nanoparticles (Fig. 1c).
Experiments as well as simulations both showed that the yield stress associated with the onset of plasticity is size
dependent and well reproduced by a power law (Fig. 3). The exponents are found to be 0.77 and 0.74 in experiments
and simulations, then in excellent agreement. This proves the ability of molecular modeling to accurately describe a
complex process such as the plastic deformation of a metallic nanostructure.

Fig. 3
Compressive stress at the point of nucleation of the first dislocation versus particle height in the simulations (Adapted from Fig. 5 in
Mordehai et al. [10], courtesy of D. Mordehai). A scaling effect can be clearly seen and is best fitted by a power law with an exponent
of 0.740.08

Metallic Nanowires
A second example concerns the plastic deformation of gold nanowires with a 110 orientation. Molecular dynamics
simulations have been performed for nanowires of lengths 200 nm and diameters of at most 50 nm, including several
millions of Au atoms, and compared to deformation experiments on slightly larger systems [ 12]. The flat punch indenter

http://www.springerreference.com/index/chapterdbid/435247 20 Apr 2015 14:02


Springer-Verlag Berlin Heidelberg 2015 5
Dr. Laurent Pizzagalli, Dr. Sandrine Brochard, Dr. Julien Godet and Dr. Celine Gerard
Mechanical Properties of Nanostructures SpringerReference

and the supporting surface (SrTiO3 in experiments) are modeled by few infinitely rigid gold layers. The setup is then quite

similar to the previous example, albeit the used EAM potentials are different and the deformation rate of 20 m s 1 is
higher.
In both experiments and simulations, the nanowire is initially compressed until a significant amount of plastic deformation
is obtained, first due to the nucleation of perfect dislocations in form of prismatic loops in both cases. This stage also
allows to form a strong adhesive contact between the testing devices and the nanowire. In a second step, the nanowires
are deformed in tension and yield by the formation of twins (Fig. 4). The simulation reveals that these twins result from the
successive nucleation of leading partial dislocations in adjacent planes. Reversing again the deformation direction leads
to a detwinning process in the nanowire in compression (Fig. 4), due to the successive nucleation of trailing partial
dislocations in the twinned layers. This quite impressive result can be explained by the differences in resolved shear
stresses for leading and trailing partial dislocations according to the deformation direction. As a result, it is demonstrated
that these gold nanowires combine an ultrahigh strength together with a very high ductility (strains of at most 30 % were
obtained).

Fig. 4
Snapshots from 300 K molecular dynamics simulations of a gold nanowire, showing the size increase (decrease) of the twinned
regions during tension (compression). The left representation shows atoms belonging to twins as gray balls and the other as pink balls
. The right representation shows later stages, with the presence of partial dislocations as red lines and twin boundaries in green
(Adapted from Fig. 6 in Lee et al. [12], courtesy of E. Bitzek, S. Lee, and S.H. Oh)

Overall, the virtual mechanical tests performed with these large-scale molecular dynamics simulations lead to structures
in excellent agreement with in situ transmission electron microscopy observations. The activated plasticity mechanisms
also appear to be similar in simulations and experiments. This confirms the reliability of these simulations for the
investigation of mechanical properties of nanostructures. This study also emphasizes how molecular modeling can be
helpful for a fine analysis of the plasticity mechanisms.

Silicon Nanowires
As a last example, the molecular modeling of the mechanical properties of a semiconductor nanostructure is presented. J.
Gunol and coworkers have recently investigated the behavior in compression of silicon nanowires with a 123
orientation, using either conjugated gradient minimization or finite temperature molecular dynamics [13]. In this work, the
systems are smaller than in previous examples, with characteristic dimensions of about several tens of nanometers, but
several different surface states and nanowire shapes are considered. In particular, the effect of an amorphous coating of
the surface is analyzed, in order to obtain insights about the influence of an oxidized nanowire surface, or amorphized
following preparation by ion beam implantation for instance.
Conversely to the examples described above, no indenters are used in this study, and compression is obtained by scaling
the simulation cell along the nanowire orientation (case (b) in Fig. 2). The chosen interatomic potential is specifically
designed for modeling silicon in high pressure high shear conditions.
The simulations revealed that plasticity at 0 K occurred at large compressive strains, about 15-18 % depending on the
surface state and nanowire shape. Dislocations are nucleated from the surface, with an increased probability in case of
slightly disordered surfaces, and preferentially at edges in rhombic nanowires. However, these features all vanish when
an amorphous Si coating is present. In this case, the activated plasticity mechanism remains dislocation nucleation, but it

http://www.springerreference.com/index/chapterdbid/435247 20 Apr 2015 14:02


Springer-Verlag Berlin Heidelberg 2015 6
Dr. Laurent Pizzagalli, Dr. Sandrine Brochard, Dr. Julien Godet and Dr. Celine Gerard
Mechanical Properties of Nanostructures SpringerReference

now always occurs at the interface between the a-Si layer and the c-Si core for a strain of about 15 %. Structural analysis
reveals that the dislocation formation is initiated at interface sites characterized by a specific topology (Fig. 5c). Also, only
those sites concentrating enough stress can be dislocation sources (Fig. 5d).

Fig. 5
Atomistic simulation of the mechanical test of a Si nanowire at 0 K (Figure adapted from Gunol et al. [ 13]). (a) 3D view of the
nanowire for a compressive strain of 15.5 % strain, just before yielding (atoms colored according to the coordination number). ( b)
Same nanowire state, but only the slice including the site of plasticity onset is shown (marked by the red arrow). (c) Zoom on this
plasticity mechanism for different deformation states. (d) Shear strain map just before yielding (the white arrow shows the site A,
characterized by a stress concentration high enough to nucleate a dislocation, unlike the site B)

This work shows how simulations can be useful for obtaining information on nucleation sites, even when those are
located deep below the surface, and to investigate many different configurations (shape, surface state). Besides, while
simulations were used as complement to experiments in the previous cases, their predictive power is illustrated in the
present case.

Limits and Issues


Atomistic simulations rely on several approximations and suffer from certain technical limitations. For instance, although
an accurate description can be obtained using first-principles approaches, they are computationally too demanding to
deal with nanostructure sizes usually found in experiments. This is why the vast majority of the available simulation works
make use of classical interatomic potentials. The reliability of these potentials is an important issue. In fact, they are fitted
on several properties of the materials and usually tested for low energy configurations, not far from stable states.
Conversely, the atomistic modeling of nanostructures first requires a good description of surfaces, which include

http://www.springerreference.com/index/chapterdbid/435247 20 Apr 2015 14:02


Springer-Verlag Berlin Heidelberg 2015 7
Dr. Laurent Pizzagalli, Dr. Sandrine Brochard, Dr. Julien Godet and Dr. Celine Gerard
Mechanical Properties of Nanostructures SpringerReference

under-coordinated atoms. Moreover, mechanical testing of nanostructures in the plastic regime involves highly strained
systems and distorted atomic configurations. For instance, the description of crack tip formation and propagation is one of
the most difficult phenomena to reproduce with interatomic potentials, for it involves bonds that are highly stretched until
the rupture. Without any references to compare with, it is not always possible to assess the reliability of a given potential
in a specific situation. Nevertheless, it seems that for fcc metals, an accurate and reliable description can be obtained
using embedded atom method potentials, as shown in the aforementioned examples. For other materials such as bcc
metals or covalent systems, no such consensus exists. Then, if an original mechanism is revealed by using a given
interatomic potential, one may often question whether the mechanism is real or an artificial product of the use of this
potential. One way to tackle this issue is to try to reproduce the same mechanism in a much smaller systems, using more
accurate and reliable calculations [14]. Another approach is to employ different interatomic potentials for the same
investigation. Obtaining the same outcome in all cases surely strengthen the conclusions.
Another issue is related to the size difference between systems investigated in experiments and simulations. It is possible
to perform classical molecular dynamics simulations including tens and even hundreds of millions of atoms on nowadays
computers (but at the expense of the timescale, see below). This corresponds to system dimensions in the range 50-100
nm at most, thus large enough for many of the experimentally investigated systems. Other systems in the submicronic
scale remain hardly accessible with current atomistic simulations. However, the size issue is not very severe. In fact, on
the one hand, atomistic modeling using interatomic potentials scales linearly as the number of considered atoms, and the
continuous increase of available computational power should allow larger systems in the future. On the other hand, the
comparison with experiments is increasingly possible because of impressive recent experimental developments. It is now
more and more feasible to fabricate systems truly in the nanoscale, which can be mechanically tested (and sometimes
monitored in situ). For instance it has been possible to measure the mechanical strength of nanowires with diameters of
few nanometers [15].
At last, the timescale issue is probably the most critical one in dynamical molecular modeling simulations. A standard
timestep in these simulations is about 1 femtosecond, because of physical arguments and numerical stability constraints.
Then to reach a simulation duration of only 1 ns, 106 iterations are required. While this is not problematic for small
systems, it severely conflicts with the above size issue when one wants to consider multimillion atom systems.
Mechanical properties of materials are experimentally probed by deformation apparatus working at rates in the range 10
4
101 s1, typically. However, because of the timescale issue, molecular dynamics simulations of large systems are
performed at best at deformation rates of 107 s1, thus many orders of magnitude higher. Such a huge difference is
probably not an issue in the elastic regime but can be dramatic in the plastic regime. For instance the nucleation of the
first dislocation, setting the onset of plasticity in the nanostructure and defining the yield point, is a rare event in the
dynamical evolution of the system. Very high deformation rates mean less time for this unfrequent event to occur. As a
consequence, yield strains (or stresses) are much higher in numerical simulations than in experiments. Fortunately it is
possible to apply post-corrections to bring numbers in closer agreement [ 16]. Nevertheless, it cannot be stated with
certainty that plasticity mechanisms activated in the simulations and in experiments at different stresses/strains are
always the same, since counterexamples exist [17]. There have been several attempts to improve this timescale issue.
For instance, few orders of magnitude can be gained by using specific techniques such as hyperdynamics, parallel replica
dynamics, or temperature accelerated dynamics [18]. Another approach relies on the use of static saddle point searching
methods such as ART, dimer, or nudged elastic band to investigate mechanisms [ 11, 19], but the comparison with
experiments is much less straightforward than with molecular dynamics.

Cross-References
Ab Initio DFT Simulation of Nanostructures
Mechanical Properties of Nanostructures
Nanomechanical Properties of Nanostructures
Nano-Pillars
Plasticity Theory at Small Scales
Size-Dependent Plasticity of Single Crystalline Metallic Nanostructures

References

http://www.springerreference.com/index/chapterdbid/435247 20 Apr 2015 14:02


Springer-Verlag Berlin Heidelberg 2015 8
Dr. Laurent Pizzagalli, Dr. Sandrine Brochard, Dr. Julien Godet and Dr. Celine Gerard
Mechanical Properties of Nanostructures SpringerReference

1. Greer, J.R., Nix, W.D.: Nanoscale gold pillars strengthened through dislocation starvation. Phys. Rev. B 73,
245410 (2006)
2. Nix, W.D., Lee, S.W.: Micro-pillar plasticity controlled by dislocation nucleation at surfaces. Philos. Mag. 91,
1084 (2011)
3. Van Swygenhoven, H., Derlet, P.M., Frseth, A.G.: Stacking fault energies and slip in nanocrystalline metals.
Nat. Mater. 3, 399 (2004)
4. Chen, M., Ma, E., Hemker, K.J., Sheng, H., Wang, Y., Cheng, X.: Deformation twinning in nanocrystalline
aluminum. Science 300, 1275 (2003)
5. Uchic, M.D., Dinmiduk, D.M., Florando, J.N., Nix, W.D.: Sample dimensions influence strength and crystal
plasticity. Science 305, 986 (2004)
6. Cherian, R., Gerard, C., Mahadevan, P., Cuong, N.T., Maezono, R.: Size dependence of the bulk modulus of
semiconductor nanocrystals from first-principles calculations. Phys. Rev. B 82, 235321 (2010)
7. Michler, J., Wasmer, K., Meier, S., stlund, F., Leifer, K.: Plastic deformation of gallium arsenide micropillars
under uniaxial compression at room temperature. Appl. Phys. Lett. 90, 043123 (2007)
8. Kang, W., Saif, M.T.A.: In situ study of size and temperature dependent brittle-to-ductile transition in single
crystal silicon. Adv. Funct. Mater. 23, 713 (2013)
9. Gerberich, W.W., Michler, J., Mook, W.M., Ghisleni, R., stlund, F., Stauffer, D.D., Ballarini, R.: Scale effects
for strength, ductility, and toughness in "brittle" materials. J. Mater. Res. 24, 898 (2009)
10. Mordehai, D., Lee, S.W., Backes, B., Srolovitz, D.J., Nix, W.D., Rabkin, E.: Size effect in compression of
single-crystal gold microparticles. Acta Mater. 59, 5202 (2011)
11. Zhu, T., Li, J., Samanta, A., Leach, A., Gall, K.: Temperature and strain-rate dependence of surface
dislocation nucleation. Phys. Rev. Lett. 100, 025502 (2008)
12. Lee, S., Im, J., Yoo, Y., Bitzek, E., Kiener, D., Richter, G., Kim, B., Oh, S.H.: Reversible cyclic deformation
mechanism of gold nanowires by twinning-detwinning transition evidenced from in situ TEM. Nat. Commun. 5,
3033 (2014)
13. Gunol, J., Godet, J., Brochard, S.: Plasticity in crystalline-amorphous core-shell Si nanowires controlled by
native interface defects. Phys. Rev. B 87, 045201 (2013)
14. Godet, J., Brochard, S., Pizzagalli, L., Beauchamp, P., Soler, J.M.: Dislocation formation from a surface step in
semiconductors: an ab initio study. Phys. Rev. B 73, 092105 (2006)
15. Kizuka, T., Takatani, Y., Asaka, K., Yoshizaki, R.: Measurements of the atomistic mechanics of single
crystalline silicon wires of nanometer width. Phys. Rev. B 72, 035333 (2005)
16. Ryu, S., Kang, K., Cai, W.: Predicting the dislocation nucleation rate as a function of temperature and stress.
J. Mater. Res. 26, 2335 (2011)
17. Ph, S.H., Legros, M., Kiener, D., Dehm, G.: In situ observation of dislocation nucleation and escape in a
submicrometre aluminium single crystal. Nat. Mater. 8, 95 (2009)
18. Voter, A.F., Montalenti, F., Germann, T.C.: Extending the time scale in atomistic simulation of materials. Annu.
Rev. Mater. Res. 32, 321 (2002)
19. Pizzagalli, L., Pedersen, A., Arnaldsson, A., Jnsson, H., Beauchamp, P.: Theoretical study of kinks on screw
dislocation in silicon. Phys. Rev. B 77, 064106 (2008)

http://www.springerreference.com/index/chapterdbid/435247 20 Apr 2015 14:02


Springer-Verlag Berlin Heidelberg 2015 9
Dr. Laurent Pizzagalli, Dr. Sandrine Brochard, Dr. Julien Godet and Dr. Celine Gerard
Mechanical Properties of Nanostructures SpringerReference

Mechanical Properties of Nanostructures

Dr.
Department of Physics and Mechanics of Materials, Institut P0, CNRS - Universite de
Laurent
Poitiers - ENSMA, Chasseneuil Cedex, France
Pizzagalli
Dr.
Department of Physics and Mechanics of Materials, Institut P0, CNRS - Universite de
Sandrine
Poitiers - ENSMA, Chasseneuil Cedex, France
Brochard
Dr.
Department of Physics and Mechanics of Materials Institut Pprime, CNRS - Universite
Julien
de Poitiers - ENSMA BP 30179, Chasseneuil Cedex, France, Poitiers,, France
Godet
Dr. Department of Physics and Mechanics, Department of Physics and Mechanics of
Celine Materials Institut P0, CNRS - Universite de Poitiers - ENSMA BP 30179, Chasseneuil
Gerard Cedex, France, Lyon, France

DOI: 10.1007/SpringerReference_435247
URL: http://www.springerreference.com/index/chapterdbid/435247
Part of: Encyclopedia of Nanotechnology

Editor: Bharat Bhushan


PDF
created April, 20, 2015 14:02
on:

Springer-Verlag Berlin Heidelberg 2015

http://www.springerreference.com/index/chapterdbid/435247 20 Apr 2015 14:02


Springer-Verlag Berlin Heidelberg 2015 10

You might also like