Atomistic Simulations of Nano-Ind

Embed Size (px)

Citation preview

  • 8/3/2019 Atomistic Simulations of Nano-Ind

    1/9

    Nanotechnology is leading to the development of materials and

    devices whose structure and function are controlled at the atomic

    level. In many applications, components experience unintentional

    or deliberate mechanical contact, and a thorough understanding of

    a materials mechanical behavior is critical for the design of

    reliable and durable systems.

    Nanoindentation is a widely used technique for probing mechanical

    properties and stability, especially of surfaces and thin films1-8. With

    the development of sensitive atomic force microscopy (AFM)3 and

    other techniques2,9, we can now measure the force Pon the indenter

    and indenter displacement hwith nanometer scale precision10,11. (For

    more details on nanoindentation, see the article by Schuh on page 32

    of this issue and reviews elsewhere12.) From the shape of such P-h

    curves, one can extract information about elastic moduli or hardness.

    In molecular dynamics (MD) simulations, one solves Newtons

    equation of motion for all the atoms in order to retrace their

    trajectories while an indenter is being pressed into the material. The

    load on the indenter Pis calculated by summing the forces acting on

    the atoms of the indenter in the indentation direction. Indentation

    depth his calculated as the displacement of the tip of the indenter

    relative to the initial surface of the indented solid. Fig.1 shows an

    example of a P-hresponse of crystalline silicon carbide (3C-SiC)

    obtained in a classical MD simulation13.

    Because of the very complex stress profile generated in the vicinity

    of the indenter, it is challenging to interpret nanoindentation

    experiments at a fundamental level. Atomistic computer simulations

    have been very helpful in unraveling the processes underlying the

    nanoindentation response14-19. The role of simulations is not

    Our understanding of mechanics is pushed to its limit when the

    functionality of devices is controlled at the nanometer scale. A

    fundamental understanding of nanomechanics is needed to designmaterials with optimum properties. Atomistic simulations can bring an

    important insight into nanostructure-property relations and, when

    combined with experiments, they become a powerful tool to move

    nanomechanics from basic science to the application area.

    Nanoindentation is a well-established technique for studying mechanical

    response. We review recent advances in modeling (atomistic and

    beyond) of nanoindentation and discuss how they have contributed to our

    current state of knowledge.

    Izabela Szlufarska

    Department of Materials Science and Engineering, University of Wisconsin, Madison, 1509 University Ave, Madison, WI 53706-1595, USA

    E-mail: [email protected]

    ISSN:1369 7021 Elsevier Ltd 2006

    Atomistic simulations

    of nanoindentation

    M AY 20 06 | VOLUME 9 | N UMBER 54 2

    mailto:[email protected]:[email protected]
  • 8/3/2019 Atomistic Simulations of Nano-Ind

    2/9

    necessarily to reproduce the exact experimental behavior, but rather

    to identify possible atomistic mechanisms in the early stages of plastic

    deformation and determine their trends as a function of the

    nanostructure and the environment (e.g. temperature or surface

    passivation)20-22. The goal is to develop robust insights into

    technologically relevant materials and ultimately design a material

    with optimum mechanical properties.

    Crystalline materials

    There are some very good review articles on the subject ofnanomechanics simulations in bulk crystalline materials23-25, where

    readers can find detailed descriptions of such phenomena as jump-to-

    contact (JC), pile-up under the indenter, nonmonotonic features in P-h

    curves, and phase transformation of the indented material. Here, we

    will provide a brief summary of these phenomena and discuss recent

    results from nanoindentation simulations of more complex materials,

    e.g. noncrystalline and nanocrystalline solids.

    The JC phenomenon has been observed by a number of groups26-30,

    and it describes the bulging up of surface atoms to meet the indenter

    tip (made of a material with higher modulus than that of the indented

    solid) before the tip makes the actual contact with the surface, i.e. at

    h< 0. Adhesive forces underlying the JC phenomenon are also

    responsible for the formation of a connective neck of atoms between

    the tip and the sample during the retraction of the indenter

    (Fig. 2)31,32. The JC phenomenon (i.e. adhesion at h < 0) manifests

    itself as a dip in the P-hcurve.

    As a matter of fact, JC is only one of many atomistic events that

    can leave a signature on a computer-generated P-hcurve. Another

    example is shown inFig. 1a, where the nonmonotonic features of the

    P-hcurve are correlated with discrete dislocation bursts in the indented

    material33 (Fig. 3). Similar pop-in events have been observed in

    experimentally determined P-hcurves34-37.

    Atomistic simulations have also shed light on the solid-state phase

    transformations that take place in material in the vicinity of the

    indenter38. For instance, Kallman etal.39 observed a localized

    crystalline-to-amorphous transition in Si at temperatures close to the

    melting point, which is consistent with experiments35,40. These

    simulations reveal a dependence of the yield strength of Si on

    structure, rate of deformation, and temperature. Solid-state

    amorphization has also been observed in nanoindentation simulationsof 3C-SiC13. Defect-stimulated growth and coalescence of dislocation

    Fig. 1 (a) Force P versus depth h of the indenter obtained in an MD simulation of crystalline SiC. The load drops in the P-h response correspond to discrete plasticevents in the indented material. (b) Plot of P against h during the unloading phase where the indenter is pulled out from the sample in 0.5 increments. Up toh = 1.83 , the deformation is entirely elastic, i.e. unloading from that depth produces a curve (squares) that retraces the loading curve (diamonds). The onset of

    plastic deformation happens at h = 2.33 , reflected in the hysteresis of the second unloading curve (circles). (Reprinted with permission from13

    . 2004 AmericanInstitute of Physics.)

    Atomistic simulations of nanoindentation REVIEW FEATUR

    M AY 20 06 | VOLUME 9 | N UMBER 5

    Fig. 2 MD simulation of indentation of solid Au with a Ni indenter. Atomicpositions during loading and unloading simulations are shown from top left tobottom right. During unloading a connective neck is formed by Au (yellow)atoms. (Reprinted with permission from32. 1995 Nature Publishing Group.)

    (a) (b)

  • 8/3/2019 Atomistic Simulations of Nano-Ind

    3/9

    REVIEW FEATURE Atomistic simulations of nanoindentation

    M AY 20 06 | VOLUME 9 | N UMBER 54 4

    loops are found to be the atomistic mechanisms underlying the

    crystalline-to-amorphous transition. In simulations by Walsh et al.41,

    amorphization has been identified as a primary deformationmechanism in indentation of Si3N4. During the indentation,

    amorphization was arrested by cracking at the indenter corners and

    piling up of material along the indenter sides (Fig. 4). The pile-up

    material itself has an amorphous structure.

    Structural changes, such as amorphization, can be identified in

    simulations by means of the radial distribution function, which can be

    directly compared with X-ray diffraction experiments. Other commonly

    used methods to monitor the evolution of simulated indentation

    damage are bond angle distribution (Fig. 4); local variation in potential

    energy, pressure, and shear stress; visual inspection of the computer-

    generated atomic structure; and changes in local topologies

    (topological changes can be analyzed, for example, by means of

    shortest-path ring statistics33,39,42).

    Amorphous and quasicrystallinematerialsAmorphous solids constitute a separate class of materials whose

    mechanical properties are of great fundamental and technological

    interest. Because amorphous materials lack a topologically ordered

    network, analysis of deformations and defects presents a significant

    challenge. Various models have been proposed to describe defects in

    such structures43,44. For example, Gilman45 has conjectured that an

    analog to a crystal dislocation exists in noncrystalline solids, and hasdescribed defects as dislocation lines with variable Burgers vectors.

    Because of the presence of these inhomogenities frozen into the entire

    material, the nanoindentation damage cannot be readily identified by

    computational techniques used for crystalline solids. The prevailing

    theory of plasticity in metallic glasses involves localized flow events in

    shear transformation zones (STZ). An STZ is a small cluster of atoms

    that can rearrange under applied stress to produce a unit of plastic

    deformation46-49.

    Even though these theories provide an essential physical insight, a

    truly atomistic model of plastic flow in amorphous materials is still

    lacking. A few atomistic simulations have been performed to tackle this

    problem. For example, Sinnott et al.50 undertook the difficult task of

    simulating nanoindentation of amorphous carbon (a-C:H) material with

    an sp3 bonded indenter. The simulations reveal that indentation has

    little effect on the hybridization of the carbon atoms or the randomly

    distributed stresses within the material. They also show that while

    penetrating the amorphous solid, the tip deforms only slightly via

    shear. This is in contrast to indentation simulations of crystalline

    diamond, where the tip deformation includes significant shear and

    twist components.

    Fig. 3 Atomic configurations during indentation of crystalline SiC obtained by MD simulations. The cut of the sample shows the atoms (a) before and (b) after thefirst load drop (compare with the P-h curve inFig. 1). The region marked by yellow rectangles reveals that the load drop is correlated with slipping of atomic layers.(c) and (d) Simultaneously, dislocations are nucleated from under the indenter. Atoms whose local topological network deviates from the perfect crystallographicorder in SiC are shown. (Reprinted with permission from 33. 2005 American Physical Society.)

    (a) (b)

    (c) (d)

  • 8/3/2019 Atomistic Simulations of Nano-Ind

    4/9

    A series of simulations of nanoindentation of amorphous silicon

    carbide (a-SiC)51 point to a noticeable localization of damage in the

    vicinity of the indenter, however the localization is less pronounced

    than in the case of 3C-SiC. As shown inFig. 5, the P-hcurve for a-SiC

    exhibits irregular, discrete load drops similar to 3C-SiC (Fig. 1). Here,

    the load drops correspond to breaking of the local arrangements of

    atoms, in analogy to the slipping of atomic layers in 3C-SiC shown in

    Fig. 3. Simulations also show that, even at indentation depths hsmaller

    than those at which the material yields plastically, the materials

    response is not entirely elastic. Instead, the amorphous structure, which

    is metastable by nature, supports a small inelastic flow related to

    relaxation of atoms through short migration distances.

    In metals, there is experimental evidence that the most stable bulk

    metallic glasses may exhibit a local quasicrystal order52. Additional

    insight has been provided in recent MD simulations by Shi and Falk 53.

    The authors show that in a two-dimensional metallic alloy with

    quasicrystalline medium-range order, the deformation localization can

    arise as the result of the breakdown of stable quasicrystal-like atomic

    configurations. Indentation simulations are of great interest for

    elucidating the science underlying plasticity in amorphous structures.

    The advent of these simulations is particularly timely because of the

    growing potential for structural applications of amorphous materials54,55.

    Nanostructured materialsIt has been demonstrated in experiments56-58 and simulations59-61 that

    nanocrystalline materials exhibit unusual mechanical behavior when

    compared with their polycrystalline counterparts. For example,

    normally brittle ceramics are shown to have very high hardness, high

    fracture toughness, and superplastic behavior, as the grain size is

    refined to the nanometer regime62,63. (For more detailed descriptions

    of these and other properties of nanocrystalline materials, see the

    article by Van Swygenhoven on page 24 of this issue.)

    Atomistic simulations of nanoindentation REVIEW FEATUR

    M AY 20 06 | VOLUME 9 | N UMBER 5

    Fig. 4 MD simulation of nanoindentation of Si3N4. Slices of the material normal to the indenter reveal cracking in the tensile regions at the indenter corners (left).Atomic configuration near the crack in the vicinity of the indenter (right). The structure of the deformed material was determined by calculating bond angledistribution. (a) Bond angle distribution for bulk crystalline (yellow) and amorphous (green) Si3N4. (b) Local bond angle distribution for region I (yellow) andII (green). Comparison of (a) and (b) indicates that region I is largely crystalline and region II resembles amorphous structure. (Reprinted with permission from41. 2000 American Institute of Physics.)

    (a)

    (b)

  • 8/3/2019 Atomistic Simulations of Nano-Ind

    5/9

    REVIEW FEATURE Atomistic simulations of nanoindentation

    M AY 20 06 | VOLUME 9 | N UMBER 54 6

    Fig. 6 MD simulation of indentation of nc-Au showing interactions between dislocations and GBs. (a)-(c) Atomic configurations during loading, and (d)-(f)corresponding stress distribution. During the simulation, dislocations are emitted from under the indenter and propagate through the grains until they becomeabsorbed by the GBs. A dislocation is represented by two red lines (two parallel planes20) that mark a stacking fault left behind a propagating partial dislocation.The yellow arrow in (d) marks the region at the GB where a leading partial dislocation arrives. (Reprinted with permission from 67. 2004 Elsevier.)

    Fig. 5 P-h response of a-SiC indented in an MD simulation. The curve exhibits a series of load drops, similar in nature to those in Fig. 1for crystalline SiC. Here, theload drops are irregularly spaced as a result of the lack of long-range order in amorphous networks, and correspond to breaking of the local atomic arrangements.The maximum indentation load reached in the a-SiC is lower than in the analogous simulation of crystalline SiC. (Reprinted with permission from 51. 2005American Institute of Physics.)

    (a) (d)

    (b) (e)

    (c) (f)

  • 8/3/2019 Atomistic Simulations of Nano-Ind

    6/9

    Atomistic simulations of nanoindentation REVIEW FEATUR

    M AY 20 06 | VOLUME 9 | N UMBER 5

    Mechanical properties of nanocrystalline materials are controlled on

    the atomic level, and therefore atomistic simulations can bring an

    invaluable physical insight to experiments and ultimately enable the

    design of materials with optimum properties.

    Since nanocrystalline materials have an increased volume fraction

    of grain boundaries (GBs)64,65, it is essential to understand the

    interactions of GBs with dislocations emitted from under the

    indenter. This question has been addressed by Feichtinger et al.66,

    who performed MD nanoindentation simulations of nanocrystalline

    Au (nc-Au) with grains 12 nm in diameter. In the simulation,

    dislocation nucleation within the grain occurs at the onset of plastic

    deformation at an indenter depth hsimilar to that in a perfect crystal.

    The GBs act as an efficient sink for partial and full dislocations and

    intergranular sliding is observed. A decrease in Youngs modulus is

    also seen as the grain size is refined to 5 nm in diameter. Recently,

    the same group reported another simulation67, where the indenter

    size is smaller than the grain size, which shows that the GBs can notonly act as a sink for dislocations, but can also reflect or emit

    dislocations (Fig. 6).

    In contrast to metals, GBs in nanocrystalline ceramics form a thicker

    GB phase that is highly disordered and has a fairly uniform thickness 62.

    These soft GBs essentially determine the materials mechanical

    response. Recent MD simulations of nanoindentation of nc-SiC show

    how the coexistence of brittle grains and soft amorphous GBs results in

    unusual deformation mechanisms68. The simulated material was

    sintered from crystalline grains of 8 nm in diameter and consists of

    about 19 million atoms. As the indenter depth hincreases, a crossover

    is observed from a cooperative deformation mechanism involving

    multiple grains to a decoupled response of individual grains, e.g. grain

    rotation and sliding, and intragranular dislocation activity (Fig. 7). The

    crossover is also reflected in a switch from deformation dominated by

    crystallization to deformation dominated by disordering, as explained

    in the caption of Fig. 8. In the early stages of plastic deformation, the

    soft (amorphous) GB phase screens the crystalline grains from

    deformation, thus making nc-SiC more ductile than its coarse-grained

    counterpart. Fracture toughness (a measure of how much energy it

    takes to propagate a crack), measured experimentally in

    nanocomposites with an nc-SiC matrix and diamond inclusions58,

    exceeds that of a polycrystalline matrix by about 50%. Increasedfracture toughness does not necessarily lead to a lower value of

    hardness. Recent experiments of nanoindentation of nc-SiC with grain

    sizes of 5-20 nm show quite the opposite trend. Liao et al.63 report nc-

    SiC to be superhard, i.e. to have a hardness of 30-50 GPa, which is

    larger than that of crystalline SiC. The hardness value of 39 GPa

    Fig. 7 MD simulation of indentation of nc-SiC. There is a crossover from cooperative response of grains at smaller indentation depths to a decoupled response ofindividual grains at larger depths. (a) Localization of the deformation after the crossover, where the color corresponds to the displacement of the center of mass ofeach grain. (b) Mean displacement of all grains where the crossover at hCR can be clearly seen (coupled at h < hCR and decoupled at larger h). (c)-(e) Examples ofdiscrete plastic events inside the grains, such as sliding at the GB (c2) or propagation of a dislocation (black arrow in (e)) along the stacking fault line (yellow line in(d) and (e)). (Reprinted with permission from68. 2005 American Association for the Advancement of Science.)

    (a) (b)

    (c)

    (d) (e)

  • 8/3/2019 Atomistic Simulations of Nano-Ind

    7/9

    REVIEW FEATURE Atomistic simulations of nanoindentation

    M AY 20 06 | VOLUME 9 | N UMBER 54 8

    obtained in the MD simulation described above is consistent with theseexperimental measurements.

    Challenges in modeling nanoindentationDespite the great advantages of simulating nanoindentation, the

    technique faces some serious challenges. The first is the limited time

    scales accessible to simulations because of limited computational

    resources. For example, the slowest time scales available to MD give

    ~1 m/s indentation velocities69, while state-of-the-art AFM systems

    can only operate at up to 0.001 m/s. Indentation measurements are

    limited to slower rates of the order of ~25 m/s. As a result, there is a

    significant disparity between simulated strain rates and those

    attainable by experiments. The rationale for modeling nanoindentation

    with MD is that for the simulated solids, all the above speeds are far

    below the speed of sound in materials (for example, the speed of sound

    in SiC is 11 000 m/s). For this reason, MD simulations are able to

    dissipate any reflected waves that arise from the motion of the

    indenter (this can be done, for example, by coupling the equations of

    motion to Nose-Hoover thermostats70,71). There is good reason to

    believe, therefore, that despite the time-scale problem, simulations can

    provide understanding of atomistic mechanisms and qualitative trends

    in nanoindentation response. However, this hypothesis needs to be

    scrupulously tested.

    Limitations in computational resources affect not only the time

    scale, but also the system dimensions available to simulations. Small

    system dimensions can introduce unrealistic boundary conditions that,

    in turn, will artificially alter processes such as dislocation dynamics. For

    example, Li et al.72 studied nucleation and propagation of dislocations

    in indented solid Al by means of MD simulations. Because the bottom

    surface of the sample remained unconstrained, a nucleated dislocation

    loop was able to move all the way to the bottom and leave the sample.

    On the other hand, in most MD simulations of nanoindentation

    reported in literature, the bottom layer of the sample is frozen and

    dislocation motion through the material is inhibited. The simplest

    solution to this problem is to have a good understanding of the

    implications of given boundary conditions and be aware of the

    limitations in the conclusions drawn. The case for MD is not lostbecause even small systems are well suited to studying the effects of

    boundary conditions on plasticity, and deformation mechanisms in

    small nanostructures are becoming of increasing interest to

    experiments and applications. Also, because of the fast development of

    computer technology as well as new algorithmic optimization methods,

    MD simulations are now possible for systems consisting of billions of

    atoms, i.e. for system dimensions in the submicron regime. At this

    length scale, the artifacts of the boundary conditions can be avoided by

    smart simulation techniques, such as efficient dissipation of any energy

    reflected from the system boundaries by strategically distributed

    thermostats.

    In order to extend simulations beyond the micrometer regime,models are being developed that combine direct atomistic simulations

    with continuum methods. For example, a quasicontinuum model has

    been developed by Tadmor et al.73,74 and applied to study

    nanoindentation75,76. In this approach, a continuum finite element (FE)

    is employed to characterize the mechanical response of the material,

    i.e. the positions of the majority of atoms are constrained and

    determined by the displacement of the nearby node. In contrast to the

    standard FE method, in the quasicontinuum approach the constitutive

    response of the system is determined from atomistic calculations based

    on interatomic potentials. Combined FE and MD simulations of

    nanoindentation have also been performed by other groups. Li

    et al.72,77,78 performed direct FE simulations in which large strain

    constitutive relations are derived from an interatomic potential (Fig. 9).

    Unlike the quasicontinuum method, this approach remains fully

    continuum. A review of simulation work based on FE is beyond the

    scope of this article but can be found elsewhere79,80.

    Another on-going challenge for MD simulations is the availability of

    reliable interatomic potentials. Parameters of a (classical) potential

    function are usually fitted to reproduce empirical data as well as

    Fig. 8 (a) Atomic configuration of nc-SiC with white grains and yellow GBs.At lower indentation depths h, deformation of the material is dominated byrecrystallization (blue atoms). At depths h > hCR, deformation is dominatedby disordering (red atoms). (b) Percentage of disordered atoms in thematerial as a function of h reflects the crossover. (Reprinted with permissionfrom68. 2005 American Association for the Advancement of Science.)

    37.6

    37.4

    37.2

    37.0

    36.8

    36.6-5 0 5 10 15 20 25 30

    (b)

    Crossover depthhCR = 14.5

    Regimes 1 & 2

    Regimes 3 & 4

    Indenter depth, h [A]

    Percentage

    ofdisorderedatoms

    (b)

    (a)

  • 8/3/2019 Atomistic Simulations of Nano-Ind

    8/9

    accurate quantum mechanical calculations. Current, state-of-the-art

    empirical potentials can account for bond formation and breaking,change in hybridization, charge transfer, etc.81,82. In spite of these

    developments, there is not one single analytic potential that is capable

    of describing all possible properties that might be of interest in a

    particular material. Furthermore, fitting an accurate potential is a

    difficult and time-consuming process.

    An approach that bypasses the need for an interatomic potential is

    based on combined first-principle and FE calculations. For example,

    Hayes et al.83 have recently simulated the nanoindentation of Al by

    means of the orbital-free density functional theory (OFDFT) local

    quasicontinuum (LC) method. In this OFDFT-LC model, the

    quasicontinuum approach is adopted but the atomic-scale calculations,

    based previously on empirical potentials, are now replaced with fast

    and inexpensive first-principles theory. This method is well suited to

    study phenomena such as initial dislocation formation; however, it is

    not capable of treating intermediate length scales (e.g. GBs in

    nanocrystalline materials). It is clear that with improving computer

    technology and the development of new algorithms, first-principles-

    based calculations will play an increasingly important role in

    nanoindentation modeling.

    Fully atomistic simulation of large systems involving many millions

    of atoms creates another nontrivial challenge, i.e. to seek patterns andextract information from such massive, multivariable datasets. For

    example, a single nanocrystalline ceramic can contain thousands of

    randomly oriented crystalline grains surrounded by intergranular

    regions with various levels of topological disorder. The change in a

    grains crystallographic structure and chemical ordering during

    indentation is a complex phenomenon that depends on many variables.

    Tracking deformations in a stand-alone amorphous material presents a

    challenge in itself, let alone as a part of a complex nanocrystalline

    material. Seeking patterns in such structures requires an extensive,

    hands-on analysis.

    In order to analyze the complicated profiles of indentation damage

    in structurally advanced materials, there is an urgent need to develop

    more efficient data-mining techniques. Such developments can be

    fostered by interdisciplinary collaborations between materials and

    computer scientists.

    OutlookFor the design of materials with superior mechanical properties, a

    mutual feedback process between experiments and simulations is

    Atomistic simulations of nanoindentation REVIEW FEATUR

    M AY 20 06 | VOLUME 9 | N UMBER 5

    Fig. 9 Combined MD and FE simulations of indentation of Cu. (a) P-h curves obtained from MD (red) and FE (blue) calculations are in good agreement.MD configurations at the beginning of the simulation (b) and (c) after several nucleation events. In (c), a shear band (dashed line) is formed. (d) The initialnucleation event modeled by the FE method, where the color corresponds to the Mises stress. Good agreement is found between MD and FE regarding the predictednucleation site, slip plane, and Burgers vector. (Reprinted with permission from72. 2002 Nature Publishing Group.)

    (b)(a)

    (c) (d)

  • 8/3/2019 Atomistic Simulations of Nano-Ind

    9/9

    REVIEW FEATURE Atomistic simulations of nanoindentation

    M AY 20 06 | VOLUME 9 | N UMBER 55 0

    REFERENCES

    1. Doerner, M. F., and Nix, W. D.,J. Mater. Res. (1986) 1, 601

    2. Oliver, W. C., and Pharr, G. M.,J. Mater. Res. (1992) 7, 1564

    3. Gerberich, W. W., et al., Acta Mater. (1996) 44, 3585

    4. Gouldstone, A., et al., Acta Mater. (2000) 48, 2277

    5. Page, T. F., et al.,J. Mater. Res. (1992) 7, 450

    6. Pollock, H. M., In: Friction, Lubrication, and Wear Technology, Blau, P. J., (ed.)ASM Metals Handbook (1992) 18, 419

    7. Thurn, J., and Cook, R. F.,J. Mater. Res. (2004) 19, 124

    8. Jungk, J. M., et al.,J. Mater. Res. (2004) 19, 2821

    9. Joyce, S. A., and Houston, J. E., Rev. Sci. Instrum. (1991) 62, 710

    10. Corcoran, S. G., et al., Phys. Rev. B(1997) 55, R16057

    11. Kiely, J. D., et al., Phys. Rev. Lett. (1998) 81, 4424

    12. Cheng, Y.-T., et al.,J. Mater. Res. (2004) 19, 1

    13. Szlufarska, I., et al., Appl. Phys. Lett. (2004) 85, 378

    14. Landman, U., and Luedtke, W. D.,J. Vac. Sci. Technol. B (1991) 9, 414

    15. Tomagnini, O., et al., Surf. Sci. (1993) 287-288, 1041

    16. Zimmerman, J. A., et al., Phys. Rev. Lett. (2001) 87, 165507

    17. de la Fuente, O. R., et al., Phys. Rev. Lett. (2002) 88, 036101

    18. Choi, Y., et al.,J. Appl. Phys. (2003) 94, 6050

    19. Lilleodden, E. T., et al.,J. Mech. Phys. Solids(2003) 51, 901

    20. Harrison, J. A., et al., Surf. Sci. (1992) 271, 57

    21. Harrison, J., et al., Mat. Res. Soc. Symp. Proc. (1992) 239, 57322. Lund, A. C., and Schuh, C., Acta Mater. (2005) 53, 3193

    23. Harrison, J., et al., In CRC Handbook of Micro/Nanotribology, Bhushan, B., (ed.),CRC Publishers, (1999), 525

    24. Heo, S. J., et al., In Nanotribology and Nanomechanics: An Introduction, Bhushan,B., (ed.) Springer-Verlag, (2005), 623

    25. Sinnott, S. B., In Handbook of Nanostructured Materials and Nanotechnology,Nalwa, H., (ed.), Academic Press, San Diego, CA, (2000), 2

    26. Smith, J. R., et al., Phys. Rev. Lett. (1989) 63, 1269

    27. Pethica, J. B., and Oliver, W. C., Mat. Res. Soc. Symp. Proc. (1989) 130, 13

    28. Landman, U., et al., Science(1990) 248, 454

    29. Rafii-Tabar, H., et al., Mater. Res. Soc. Symp. Proc. (1992) 239, 313

    30. Rafii-Tabar, H., and Kawazoe, Y.,Jpn. J. Appl. Phys. (1993) 32, 1394

    31. Landman, U., et al., Wear(1992) 153, 3

    32. Bhushan, B., et al., Nature(1995) 374, 607

    33. Szlufarska, I., et al., Phys. Rev. B(2005) 71, 174113

    34. Bradby, J. E., et al.,J. Mater. Res. (2004) 19, 380

    35. Clarke, D. R., et al., Phys. Rev. Lett. (1988) 60, 2156

    36. Pharr, G. M., et al., Scripta Metall. (1989) 23, 1949

    37. Pharr, G. M., et al.,J. Mater. Res. (1991) 6, 1129

    38. Cheong, W. C. D., and Zhang, L. C., Nanotechnology(2000) 11, 173

    39. Kallman, J. S., et al., Phys. Rev. B(1993) 47, 7705

    40. Minowa, K., and Sumino, K., Phys. Rev. Lett. (1992) 69, 320

    41. Walsh, P., et al., Appl. Phys. Lett. (2000) 77, 4332

    42. Rino, J. P., et al., Phys. Rev. B(2004) 70, 045207

    43. Rivier, N., Philos. Mag. A (1979) 40, 859

    44. Sheng, H. W., et al., Nature(2006) 439, 419

    45. Gilman, J. J.,J. Appl. Phys. (1973) 44, 675

    46. Argon, A. S., and Kuo, H. Y., Mater. Sci. Eng. (1979) 39, 101

    47. Schuh, C., and Lund, A. C., Nat. Mater. (2003) 2, 449

    48. Falk, M. L., Phys. Rev. B(1999) 60, 7062

    49. Falk, M. L., and Langer, J. S., Phys. Rev. E(1998) 57, 7192

    50. Sinnott, S. B., et al.,J. Vac. Sci. Technol. A (1997) 15, 936

    51. Szlufarska, I., et al., Appl. Phys. Lett. (2005) 86, 021915

    52. Shi, Y., and Falk, M. L., Phys. Rev. Lett. (2005) 95, 095502

    53. Shi, Y., and Falk, M. L., Appl. Phys. Lett. (2005) 86, 011914

    54. Schroers, J., and Johnson, W. L., Phys. Rev. Lett. (2004) 93, 255506

    55. Lu, Z. P., et al., Phys. Rev. Lett. (2004) 92, 245503

    56. Siegel, R. W., Nanostructures of Metals and Ceramics, In Nanomaterials:Synthesis, Properties and Applications, Edelstein, A. S., and Cammarata, R. C.,(eds.), Institute of Physics, Bristol, UK, (1996), 201

    57. Zhang, S., et al., Surf. Coat. Technol. (2003) 167, 113

    58. Zhao, Y., et al., Appl. Phys. Lett. (2004) 84, 1356

    59. Li, J., and Yip, S., Comp. Model. Eng. Sci. (2002) 3, 229

    60. Schitz, J., et al., Nature(1998) 391, 561

    61. Yip, S., Nature(1998) 391, 532

    62. Chen, D., et al.,J. Am. Ceram. Soc. (2000) 83, 2079

    63. Liao, F., et al., Appl. Phys. Lett. (2005) 86, 171913

    64. Keblinski, P., et al., Phys. Rev. Lett. (1996) 77, 2965

    65. Keblinski, P., et al., Acta Mater. (1997) 45, 987

    66. Feichtinger, D., et al., Phys. Rev. B(2003) 67, 024113

    67. Hasnaoui, A., et al., Acta Mater. (2004) 52, 2251

    68. Szlufarska, I., et al., Science(2005) 309, 911

    69. Belak, J., and Stowers, I. F., In Fundamentals of Friction: Macroscopic andMicroscopic Processes, Singer, I. L., and Pollock, H. M., (eds.), Kluwer AcademicPublishers, Dordrecht, (1992), 511

    70. Hoover, W. G., Phys. Rev. A (1985) 31, 1695

    71. Nose, S., Mol. Phys. (1984) 52, 255

    72. Li, J., et al., Nature(2002) 418, 307

    73. Tadmor, E. B., et al., Philos. Mag. A (1996) 73, 1529

    74. Shenoy, V. B., et al.,J. Mech. Phys. Solids(1999) 47, 611

    75. Knap, J., and Ortiz, M., Phys. Rev. Lett. (2003) 90, 226102

    76. Smith, G. S., et al., Acta Mater. (2001) 49, 4089

    77. Zhu, T., et al.,J. Mech. Phys. Solids(2004) 52, 691

    78. Van Vliet, K. J., et al., Phys. Rev. B(2003) 67, 104105

    79. Mackerle, J., Modell. Simul. Mater. Sci. Eng. (2005) 13, 935

    80. Mackerle, J., Eng. Comp. (2003) 21, 23

    81. van Duin, A. C. T., et al.,J. Phys. Chem. A (2003) 107, 3803

    82. Brenner, D. W., et al.,J. Phys.: Condens. Matter(2002) 14, 783

    83. Hayes, R. L., et al., Multiscale Model. Simul. (2005) 4, 359

    critical. Because of the transferability of simulation tools and the wide

    variety of application areas, it is also essential to create a platform for

    collaboration among scientists from multiple disciplines. New structural

    applications are being extensively explored for amorphous and

    nanostructured materials, e.g. superhard coatings, sporting goods, high-

    speed machining and tooling, and biomaterial implants. If the

    mechanical properties of these complex materials are to be exploited in

    industrial applications, a thorough understanding of their mechanical

    response (e.g. through indentation) is of vital importance.

    Acknowledgments

    The author gratefully acknowledges support from the US National ScienceFoundation grant DMR-0512228. I am also thankful to D. Stone and D. Morganfor helpful comments on the manuscript.