You are on page 1of 10

Copyright 

c 2004 Tech Science Press cmes, vol.6, no.1, pp.105-114, 2004

Crystalline Plasticity on Copper (001), (110), and (111) Surfaces during


Nanoindentation

Haiyi Liang1 , C.H. Woo1,2 , Hanchen Huang3 , A.H.W. Ngan4 , and T.X. Yu5

Abstract: Molecular dynamics (MD) simulations are portance, besides the obvious scientific interest. Gane
performed to study crystalline plasticity during nano- and Bowden (1968) first reported that gold samples could
indentation by comparing the elastic-plastic response sustain stresses, induced by indentation on a small scale,
of three copper substrates with surfaces (001), (110), approaching the theoretical shear strength, despite the
and (111) crystallographic planes. The effects of elas- likely presence of dislocations beneath the indenter prior
tic anisotropy and crystallographic symmetry on the re- to the onset of the strain burst. With the development
duced modulus, dislocation nucleation, and subsequent of powerful instruments in recent years, such as atomic
microstructure evolution, are investigated. The reduced force microscope (AFM), and interfacial force micro-
modulus of (111) surface is found to be the largest, while scope (IFM), structures and events at atomistic scale be-
that of (001) surface is the smallest. Elastic stress dis- comes accessible and more detailed information of the
tribution calculated from finite element method (FEM) deformation mechanisms at the nanoscale have been ob-
is qualitatively consistent with the MD simulation re- tained with atomic resolution.
sults. Significant differences exist in the deformation The study of crystalline effects on plasticity has a long
behavior in the three different crystallographic orienta- history. The advent of experimental and modeling tech-
tions. The differences in the load-displacement curves niques in the nanoindentation area offers new oppor-
for the three different cases are correlated with those in tunities in the investigation of this important problem.
the corresponding evolutions of the underlying disloca- [Ghoniem and Cho(2002); Srivastava and Atluri (2002)]
tion structure. Yielding platforms exist typically in load- Indeed, the experimental works by Keily and Houston
displacement curve of Cu (001), which can be attributed (1998), and Corcoran, Colton, Lilleodden, and Gerberic
to effective resistance of dislocation locks. Load drops (1997), on gold showed that the effect of crystallography
are typically characteristic of Cu (111) and (110), due to anisotropy is an important factor responsible for the com-
a more mobile dislocation structure. plex behavior of its deformation during nanoindentation.
However, due to the many factors that may affect plastic
keyword: nanoindentation, molecular dynamics,
behavior under nanoindentation, such as the size and tip
anisotropy.
geometry of the indenter, penetration depth, surface ox-
ide layer and contamination, time-dependent characteris-
1 Introduction tics, and strain gradient effects etc., a straightforward in-
Due to the rapid development of nanotechnology in re- terpretation of nanoindentation measurements is difficult
cent years, the understanding of the mechanical prop- [Gouldstone, Koh, Zeng, Giannakopoulos, and Suresh
erties of solids in the nanoscale is of technological im- (2000), Swadener, George, and Pharr (2002), Chiu and
Ngan (2002), and Gerberich, Tymiak, Grunlan, Horste-
1 Department of Electronic and Information Engineering, Hong
meyer, and Baskes (2002)].
Kong Polytechnic University, Hong Kong SAR, China
2 Communicating author. (Fax: +852-2365-4703, To complement the experiments to better understand the
E-mail: chung.woo@polyu.edu.hk) mechanics of plastic deformation during nanoindenta-
3 Department of Mechanical, Aerospace & Nuclear Engineering,
tion, atomistic simulations have contributed a substan-
Rensselaer Polytechnic Institute, Troy, NY 12180, USA
4 Department of Mechanical Engineering, University of Hong tial portion of the knowledge of the nucleation and evo-
Kong, Hong Kong SAR, China lution of microstructure during nanoindentation, espe-
5 Department of Mechanical Engineering, Hong Kong University of cially in the initial stage. Kelchner, Plimpton, and
Science and Technology, Hong Kong SAR, China
106 Copyright 
c 2004 Tech Science Press cmes, vol.6, no.1, pp.105-114, 2004

Hamilton (1998) used molecular dynamics to simulate (111), the dimension is 190 × 189 × 112 Å3 (containing
displacement-controlled indentation on the Au (111) sur- 341,334 atoms), and the top surface is (111) plane with
face, and observed dislocation loops appearing off the in- the other two side surfaces (110) and (1 12) respectively.
denter axis, when the applied shear stress was much be- In all cases, periodic boundary conditions are applied
yond the critical shear stress. Via MD simulation, Zim- along the horizontal directions. The top surface is trac-
merman, Kelchner, Klein, Hamilton, and Foiles (2001) tion free, and the atomic layers of two cutoff distances
found a significant decrease in the load to nucleate dis- at the bottom are held fixed. The temperature of the
locations due to surface step effects. Rodrı́gue, Zim- simulation cell (i.e., all the moving atoms) is kept con-
merman, González, Figuera, Hamilton, Pai, and Rojo stant at 300K through Nosé-Hoover thermostat. Follow-
(2002) simulated the formation of hillocks near inden- ing Kelchner, Plimpton, and Hamilton (1998), a repulsive
tation points, and found that the hillocks were composed potential is used to model a spherical indenter made of a
of four Shockley partials and a stair-rod dislocation. By material like tungsten, with a radius of R= 40 Å.
combining the results of atomistic and finite-element
The EAM potential of Doyama and Kogure (1999) for
modeling, Li, Vliet, Zhu, Yip, and Suresh (2002), re-
copper is used for the substrate. The parameters are
lated the concept of structural stability and the nucleation
determined from the experimental data of cohesive en-
of defects during nanoindentation in their interpretation
ergy, Born stability, elastic constants C 11 ,C12 and C44 ,
of experiments. Recently, with quasi-continuum method,
formation energy of a vacancy, and stacking fault en-
Knap and Ortiz (2003) found that, with an indenter of tip
ergy (35.7mJ/A2 ). The indentation velocity we use is
radius 70 nm, there is no force drop despite the existence
about 3 m/s, which is nine orders of magnitude higher
of profuse dislocation activity.
than the experimental ones. This excessively large in-
Although MD simulations of nano-indentation have been denting velocity is an intrinsic problem of MD. Even
reported by many authors in the literature, a systemic this 3m/s is already an order of magnitude slower than
study of the crystalline anisotropy effects on nanoinden- those used in the past, 40 m/s by Christopher, Smith and
tation and the detailed underlying dislocation evolution is Richter (2001), and the 50 m/s by Gannepalli and Mal-
still lacking. In the present paper, we perform MD simu- lapragada (2002). The sensitivity of the simulation re-
lations of nano-indentation on Cu (001), (110), and (111) sults to the indentation velocity has been discussed by
surfaces. The marked crystalline anisotropy exhibited by Liang, Woo, Huang, Ngan, and Yu (2003). The relation-
the elasticity of copper, plus its readily available and rel- ship between the applied load and the indentation depth
atively reliable inter-atomic potential, makes it a prime remains little changed, although a higher indentation ve-
candidate for the study of crystalline effects of nano- locity in MD simulation leads to a higher strength. At
plasticity via atomistic simulation. Our aim is to gain the same time, the associated dislocation structures are
insight into the relationships between the displacement- also left essentially unaffected. More importantly, the
load behavior and the concomitant development of the values will quickly converge as the velocity decreases.
dislocation structures during nanoindentation. Compari- Indeed, our results show no obvious difference at a ve-
son of the results on various crystalline surfaces reveals locity below 3 m/s. Nevertheless, one must be aware
the difference in the development of incipient plasticity that convergence within MD here does not necessarily
in non-equivalent crystallographic directions. mean that time-dependent mechanisms, such as diffusion
driven ones, have been properly included.
2 Simulation Procedure To correlate the load-displacement behavior with the dis-
Three copper substrates with (001), (110), and (111) sur- location structure evolution beneath the indenter, follow-
faces are modeled as slabs. For Cu (001), the dimension ing Zimmerman, Kelchner, Klein, Hamilton, and Foiles
is 188 × 188 × 108 Å3 (containing 324,480 atoms), the (2001), we identify dislocation structures with the help
top free surface is (001) plane, and the other two side of a slip vector defined as
surfaces are (100) and (010) respectively. For Cu (110), 1 n
ns β∑
sα = − (xαβ − Xαβ ) (1)
the dimension is 189 × 187 × 112 Å3 (containing 338,624 =α
atoms), and the top surface is (110) plane with the other
two side surfaces (110) and (001) respectively. For Cu where xαβ and Xαβ are the vectors linking atom α and
Crystalline Plasticity on Copper 107

250 250

Cu (001) Cu (001)
Cu (110)
200 Cu (111) 200
b4 b5

Force (nN)
Force (nN)

150 150 b3
b2

100 100

b1
50 50

(a) (b)
0 0
0 2 4 6 8 10 0 2 4 6 8 10

Indentation depth (Angstrom) Indentation depth (Angstrom)


250 250

Cu (110) Cu (111) d7
c5
200 200
d1
c3 d3 d5 d8
c1 c6
Force (nN)

150 c4 150
Force (nN) d2 d4 d6

100 c2 100

50 50

(c) (d)
0 0
0 2 4 6 8 10 0 2 4 6 8 10

Indentation depth (Angstrom) Indentation depth (Angstrom)

Figure 1 : Load-displacement curve of Cu (001), (110), and (111).

all its nearest neighbors β in the current and reference tallographic surfaces (001), (110), and (111) are com-
position,n s being the number of slipped neighbors. In the pared in Fig.1. F is the load applied to press the indenter.
following, atoms with slip vectors having a magnitude The indentation depth (h) is the prescribed displacement
|sα| ranging from 0.1 to 3.0 Å are displayed. The atoms of the indenter, which is related to the movement of the
are colour coded according to the magnitude of the slip spherical indenter center. We note that the actual depth is
vectors. Areas containing atoms identified with signifi- smaller than the prescribed indenter displacement due to
cant slip vectors usually denote the presence of faults in the compliance of the repulsive indenter.
the stacking, indicating the existence of faulted loops, or Fig.1a shows the strong effects of the crystallographic
the extended core of split dislocations, in which the par- anisotropy on the reduced modulus along different crys-
tials are situated at the edge of the faulted area. We find tallographic directions. According to the Hertz theory,
this a clear and convenient way of illustrating the elastic the load vs. depth behavior of an isotropic elastic media
and plastic response of the substrate during loading. follows
4 1 1 − v2sub 1 − v2ind
3 Results and Discussion F = E ∗ R1/2 h3/2 = + (2)
3 E∗ Esub Eind
3.1 Elastic regime
where E ∗ is the reduced modulus, a combination of the
The nanoindentation load-displacement curves, i.e., elastic modulus E sub and Eind and Poisson’s ratios v sub
force (F) vs. indentation depth (h), of copper on crys- and vind of the substrate and the indenter respectively, and
108 Copyright 
c 2004 Tech Science Press cmes, vol.6, no.1, pp.105-114, 2004

(001) (110) (111)

(a) MD simulation results


(001) (110) (111)

(b) Finite element method results

Figure 2 : Stress σz under indenter, calculated from (a) MD simulation and (b) Finite element method. For each
case, the left picture is side-view, and the right is top view. Only atoms having |σ z | > 4GPa are plotted.

R is the indenter radius. lation and experiment [Lilleodden, Zimmerman, Foiles,


Using values of F andh read off directly from Fig.1a, and Nix (2003), Kiely and Houston (1998)]. We note
the reduced modulus can be calculated using equation that the model of Vlassak and Nix (1994) considered the
(2). For h= 2.0Å, we obtain values of E ∗ of 75.7GPa, elastic response of substrate under a pyramidal indenter
101.8GPa, and 123.4GPa for Cu (001), (110), and (111), during unloading, in the presence of full plasticity. In the
respectively. present paper, on the other hand, we consider the early
part of the loading stage, with a fully elastic substrate
We note that, strictly speaking, equation (2) is only
under spherical indenter, using the Hertz theory. Despite
applicable to an isotropic medium. The foregoing re-
the difference of elastic anisotropy effects in MD simu-
sults are thus only qualitatively valid for the anisotropic
lation, experiments, and theoretical method, all studies
copper substrates presently under consideration. As-
show that the (111) surface has the largest reduced mod-
suming a tungsten-like indenter with E ind =400GPa,
ulus while (001) surface is the softest.
vind =0.28, and Young’s modulus for the copper substrate
of E001 =77.5GPa, E110 =137.0GPa, and E111=191.9GPa, Apart from the effects on the reduced modulus, the elas-
calculated from the Doyama-Kogure potential, and Pois- tic anisotropy also influences the stress distribution un-
son’s ratio v sub =0.3, the reduced modulus from equation der the indenter. In Fig.2a, we show the top- and the
(2) are calculated to be E001 ∗ ∗
=71.2GPa, E110 =111.7GPa, side-views of the spatial dependence of the σ z component

and E111 =141.9GPa. These agree reasonably well with (along indentation direction), using atomic stress tensor
the values obtained directly from the load-displacements [Egam, Maeda, and Vitek (1980)], in the elastic regime
curves. for each of the three indentation directions. The stress
distributions exhibit strong anisotropy, showing 90 ◦ rota-
Using the theoretical method proposed by Vlassak and
tional symmetry in Cu (001), 180 ◦ rotational symmetry
Nix (1994), the reduced modulus for copper (001), (110),
in Cu (110), and 120 ◦ rotational symmetry in Cu (111).
and (111) are 108.8GPa, 114.2GPa, and 115.9GPa re-
While for an isotropic media, the stress field under the
spectively, with elastic constants from Doyama-Kogure
indenter is axisymmetric, in comparison.
potential. The theoretical predication shows much less
anisotropic effect than our simulation. Stronger elas- Three-dimensional analysis using finite element method
tic anisotropy effects were also observed in MD simu- (FEM) for a stiff frictionless spherical indenter press-
Crystalline Plasticity on Copper 109

ing onto substrates corresponding to Cu (001), (110), large load drop they found for (111) Au is essentially the
and (111), is performed with the ANSYS software. The same as our case.
stresses σz are shown in Fig.2b, in comparison with the Instead of yielding platforms as in Cu (100), the most
MD results in Fig.2a. Despite the difference between the prominent feature of the load-displacement curve of Cu
MD and the FEM models, in the indenters used and the (110) are large load drops (Fig.1c), at points #c1 (h ∼5.3
assumption regarding the elastic linearity, the qualitative Å and F ∼155.0 nN), #c3 (h ∼6.4 Å and F ∼161.5 nN)
resemblance between FEM and MD results shows the and #c5 (h ∼7.6 Å and F ∼203.2nN). These occur due
consistency of both calculations. It also suggests that, to a sudden relaxation of the internal stress supporting
in the elastic regime, the discrete lattice can still be de- the indenter, such as caused by a sudden burst of dislo-
scribed satisfactorily as a continuum, despite the highly cations from a source. It is also noted that the relaxation
concentrated local stresses. One may even speculate that here is associated only with a relatively small advance-
the potential sites of dislocation nucleation in a single ment of the indenter. Between the points #c3 and #c4
crystal can be predicated by FEM, by analyzing for elas- there is a yielding platform preceded by a small stress re-
tic instability [Li, Vliet, Zhu, Yip, and Suresh (2002)]. laxation, in which relatively large indenter advancement
Since the FEM results can be scaled arbitrarily, for a large occurs, probably caused by dislocation glide driven by
indenter, the stressed volume increases accordingly, and the applied shear against some barriers. The steadily ris-
the nucleation site also moves deeper into the substrate. ing parts of the curve between points #c2 and #c3, #c4
The increased volume may be sufficiently large to ac- and #c5, and after point #c6 exhibit the typical quasi-
commodate the simultaneous generation of several dis- elastic behaviour, corresponding to the accumulation of
locations once the critical stress is attained. internal stresses, caused by the formation of a relatively
inactive dislocation structure, such as locks, at #c2 and
3.2 Incipient plasticity #c6.
In the following, we describe the different load- The load-displacement curve of Cu (111) in Fig.1d bears
displacement relations of the indentation in the three much more similarity to Cu (110) than to Cu (100). The
crystallographic orientations (Fig.1). The concomi- outstanding features are also large load drops with little
tant microstructure development and deformation mech- deformation at points #d1 (h ∼5.1 Å and F ∼176.3nN)
anisms will be considered in the next subsection. In all and #d7 (h ∼7.6 Å and F ∼222.1nN), caused by the re-
cases, the load-displacement relations are characterized laxation of an overstressed crystal through a sudden burst
by a smooth steady elastic region, beyond which inflex- of dislocation activities. The region between points #d2
ion points signaling the occurrence of plasticity can be and #d6 behaves in a way similar to Cu (110) between
seen. #c3 and #c4, namely a yielding stage that involves a small
stress relaxation, but a relatively large indenter advance-
For Cu (001) (Fig.1b), there are five inflexion points pro-
ment, probably caused by dislocation glide under a con-
duced by minor yielding due to steady inelastic deforma-
stant driving shear. Between points #d6 and #d7, the
tion under the indenter. Between the inflexion points,
curve rises again quasi-elastically, with a minor inflex-
the curve shows a relatively stable and smooth quasi-
ion, corresponding to the accumulation of internal stress,
elastic behavior, corresponding to a steady buildup of
caused by a relatively inactive dislocation structure.
internal stress, consistent with a relatively inactive dis-
location structure. The magnitude of the inelastic dis- In the forgoing, it is noted that, independent of the
placements (or yielding platforms) at points #b1, #b3, crystallographic direction of the indentation, three types
and #b5 (Fig.1b) are 0.06, 0.13 and 0.60 Å respectively, of features can be identified in the load displacement
indicating an increasing magnitude of the underlying dis- curves: (1) a quasi-elastic regime; (2) sudden load drops;
location activities, as the indenter advances. Lilleodden and (3) displacement under constant load (simply called
et al. (2003) found small load drops for (001) Au in- strain burst). This is in complete agreement with ex-
stead of yielding platforms. This is somewhat different perimental observation in which discrete load jump or
from our results, but may be attributed to the difference strain burst with intermittent elastic responses is typical
in the hardness of indenter, atom potentials, and simula- [Suresh, Nieh, and Choi (1999), Kiely, Jaraush, Hous-
tion conditions adopted in either case. Nevertheless, the ton, and Russell (1999)]. Despite the commonality, com-
110 Copyright 
c 2004 Tech Science Press cmes, vol.6, no.1, pp.105-114, 2004

(a) h =4.8 Å (b) h =6.4 Å (c) h =6.9 Å


[001]

(111) (1 11) (111)


faulted loop (111)
[110]
[1 10] lock-1 lock-2
bow-out lock-3
on (111)
(d) h =8.3 Å (e) h =8.9 Å

(111)

(111)
(111)
(111)
(111)
(111)

Figure 3 : Microstructure evolution beneath the contact surface on Cu (001)

parison of the load-displacement curves in Fig.1b-d re- contact surface. The Shockley partial pair grows to inter-
veals significant differences in the deformation behav- sect with the surface, and with each other forming a stair-
ior during nanoindentation in the three different crystal- rod dislocation and a Lomer-Cottrell lock (lock-1) from
lographic directions. While all three types of features the dislocation reaction 16 a[112] + 16 a[11 2] = 13 a[110]
are present in the case of Cu (111) and (110), indicating (a is the lattice constant of copper). After the lock is
the evolution of a strongly active dislocation structure, formed, the crystal hardens, as reflected by the resump-
quasi-elastic behavior dominates the load-displacement tion of loading. The presence of lock-1 restricts the plas-
relationship of Cu (001), interrupted only by a few small tic flow of the material inside the wedged-shape region
strain bursts, suggesting a less active movement of under- bounded by the two {111} slip planes, and therefore
lying dislocation. To achieve a mechanistic understand- hardens the material inside the wedge. The load trans-
ing of the foregoing stress-strain relationship in terms of mitted by the indenter through the wedge creates a stress
the underlying dislocation structure, the microstructure concentration at its tip, which generates the nucleation of
development during the indentation is considered in the a pair of faulted loops at both ends of the Lomer-Cottrell
following subsection. lock, resulting in the minor yield at point #b2 (Fig.1b).
When the formation of the faulted loops completes, the
3.3 Microstructure development quasi-elastic behavior after point #b2 (Fig.1b) resumes.
When the contact area gets bigger as the loading contin-
3.3.1 Cu (001) ues, i.e., between points #b1 and #b3 (Fig.1b), the in-
During the early part of the loading curve, i.e., prior to crease of local stress forces the unlocked non-edge seg-
point #b1 in Fig.1b, no dislocations are identified, which ments on one of the Shockley partial pair to bow out,
suggests that the substrate is deforming elastically. As expanding the faulted area on the (111) plane (Fig.3b,
the indentation depth increases beyond ∼4.8 Å (point 3c). Just at point #b3 in Fig.1b, two new Lomer-Cottrell
#b1 in Fig.1b), a pair of Shockley partials, on (111) and locks (lock-2 and lock-3, in Fig.3c) come into being, per-
(111) slip planes (Figs. 3a), are nucleated beneath the pendicular to lock-1. The processes of Shockley partial
Crystalline Plasticity on Copper 111

(a) h 5.4Å (b ) h 7.0 Å (c ) h 8.2 Å

2
1
(11 1) locks 5
6
(111)
[110]
(1 11)
5
6
(111)
[110] [001]

Figure 4 : The side view of microstructure evolution beneath the contact surface on Cu (110)

[110] (c ) h 8.2 Å
(a) h 5.4Å (b ) h 7.0 Å
[001]
5 6
5 6
2 1

(11 1) (111)
4 7 8
4 3 3
3

locks

Figure 5 : The bottom view of microstructure evolution beneath the contact surface on Cu (110)

bow-out and formation of new locks contribute to the mi- (Fig.1b). The large excursions beyond point #b5 corre-
nor yield at point #b3 (Fig.1b). spond to the usual cold-working process, in which dislo-
After the formation of lock-2 and lock-3, the yielding is cations are generated, interact, and form locks.
interrupted and the load displacement curve resumes its Despite the weaker strength in [100] direction, the re-
quasi-elastic behavior between points #b3 and #b4. At peated formation of locks thwarted attempts of the crys-
this point, different parts of the substrate deform via dif- tal to relax the stresses accumulated as the indentation
ferent mechanisms, some elastically, some via the dislo- proceeds. As a result, the load-sustaining capability to
cation bow out on the (111) plane, and some via the nu- nanoindentation in this direction is comparable to the
cleation and growth of Shockley partials. The resultant other directions (Fig.1b-d).
effect produces the quasi-elastic behavior on the load-
displacement curve between points #b3 and #b4 (Fig.1b). 3.3.2 Cu (110)
Beyond point #b4 (Fig.1b), the expansion of the con- The development of microstructure during indentation on
tact area drives the bow out of the non-edge segments Cu (110) is shown in Fig.4a-c, the bottom views of which
on lock-2 and lock-3 on the (1 11) faulted plane (Fig.3d), are shown in Fig.5a-c. As the indentation depth increases
then producing major plastic deformation at point #b5 beyond ∼5.4 Å (point #c1 in Fig.1c), a Shockley par-
112 Copyright 
c 2004 Tech Science Press cmes, vol.6, no.1, pp.105-114, 2004

tial embryo starts first to grow on plane #1 (Fig.4a, 5a), be one cause of strain burst. Additionally, this prismatic
which triggers the subsequent unsymmetrical nucleation dislocation can be regarded as geometrically necessary
of the other three partials on plane #2, #3, and #4 (Fig.4a, dislocation (GND), accommodating those atoms pushed
5a). Planes #1 and #3 are two adjacent (111) planes, and into substrate by indenter (Gao and Huang, 2003).
planes #2 and #4 are two adjacent (111) planes. Dur-
ing the same period, two other partials nucleate on (111) 3.3.3 Cu (111)
and (111) planes, intersect with the partials on planes #1
and #2 respectively, and form two Lomer-Cottrell locks The development of microstructure during indentation on
(Fig.4a, 5a). Unlike the Lomer-Cottrell lock in Fig.3a, Cu (111) is shown in Fig.6a-d. As the indentation depth
the lock here is not effective enough to prevent the fur- increases beyond the elastic regime at ∼5.1 Å (point
ther movement of dislocations. The nucleation and de- #d1 in Fig.1d), a Shockley partial embryo starts to ap-
velopment of the above structure leads to a sudden load pear on (111) (plane #1), followed by the nucleation and
drop at #c1 (Fig.1c), and the resumption of quasi-elastic development of Shockley partial on (111) (plane #2 in
behavior at #c2 is due to the formation of the locks. From Fig.6a). The two partial dislocations intersect with each
point #c2 to #c3 (Fig.1c), the indentation strain is accom- other to form a Lomer-Cottrell lock (pointed by arrow in
modated elastically, with a corresponding accumulation Fig.6a). The dislocation structure in Fig.6a is the same
of the stress. There is not much evolution of the disloca- as that shown by Lilleodden, Zimmerman, Foiles, and
tion structure between #c2 and #c3 (Fig.4a). Nix (2003), with stacking fault near the free edges, and
a perfect lattice near the top of the intersection of plane
As the stress continues to accumulate, the dislocation
#1 and #2. Similar to Cu (110), the current lock can only
structure starts to change at point #c3, leading to the
offer weak resistance to the movement of existing dislo-
load drop at this point. From point #c3 (∼6.5 Å) to #c4
cations. As a result, the associated dislocation activity is
(∼7.0 Å) in Fig.1c, the dislocations becomes very ac-
responsible for the large load drop between #d1 and #d2
tive, and the original dislocation structure as represented
in Fig.1d.
by Fig.4a change dramatically. Dislocation on plane #3
grows much larger than that on plane #4, while those on The wavy yielding region, from point #d2 to #d6 in
planes #1 and #2 shrink back and almost vanish. New Fig.1d, involves a sequence of minor but distinct local
locks are formed preventing further movement. Most no- rearrangement of the dislocation structure. At point #d4
tably, two new dislocations, nucleated on (111) (plane (Fig.6b, ∼5.8 Å), a new partial dislocation nucleates on
#5) and (111) (plane #6) respectively, intersecting and (111) (plane #3), which subsequently intersects with for-
locked with each other, stem downwards from under the mer dislocations on planes #1 and #2 to form a stable
(110) surface (Fig.4b, 5b). tetrahedron defect structure beneath the indenter. In com-
parison, a pyramidal defect structure has been observed
The load-displacement relation between points #c4 to
in a MD simulation with a truncated pyramid inden-
#c5 (Fig.1c) is essentially elastic, similar to the elastic
ter (Gannepalli and Mallapragada 2002). Similar to the
behavior between points #c2 and #c3 (Fig.1c). The shape
wedged-shape dislocation structure in Fig.3a, this tetra-
of underlying dislocation structure again remains almost
hedron structure here serves as a source of stress concen-
unchanged from point #c4 to #c5 in Fig.1c.
tration. Dislocation loops (for instance, partial disloca-
At point #c5 in Fig.1c, a large load drop occurs as two tions on (111),(111) and (111) planes, namely planes #4,
new dislocations nucleate on (111) (plane #7) and (111) #5 and #6 in Fig.6b, 6c) are forced out along the “walls”
(plane #8), and intersect with the previous dislocations of the tetrahedron (Fig.6b, 6c). These intensive disloca-
on planes #5 and #6 to form a prismatic dislocation loop tion activities are completed at point #d6 in Fig.1d (∼6.2
gliding away from the surface (Fig.4c, 5c) and stop at Å, Fig.6c). The intensive local atom movement leads to
a point where stress due to the indentation is sufficiently continuous stress relief. Therefore, the substrate under-
diminished. Here, we point out that there is one prismatic goes large deformation between #d2 and #d6 without a
dislocation at the bottom end of the square pipe, the left significant build-up of stress. It is worth pointing out that
part of which is the trace this dislocation left. As en- in this process, dislocations nucleated out of the tetrahe-
visioned by Suresh, Nieh, and Choi (1999), the periodi- dron will grow or shrink according to local stress varia-
cal punching of prismatic dislocation into substrate might tion.
Crystalline Plasticity on Copper 113

(a) h 5.3 Å (b ) h 5.8 Å

[111]
2 1
(111)
3

[01 1] (11 1) (1 11)


[2 1 1] 4

(c ) h 6.2 Å lock (d ) h 7.9 Å

3 7
8
6
9
5

Figure 6 : Microstructure evolution beneath the contact surface on Cu (111)

At the stable stage from point #d6 to #d7 (Fig.1d), the ement method and molecular dynamics are qualitatively
dislocation structures grow a little, without essential dif- the same, despite the distinctly discrete nature of lattice
ference in shape from that in Fig.6c. At point #d8 in at such small scales.
Fig.1d (∼7.9 Å), a large load drop occurs, which can be In all three cases, the load-displacement relations are
attributed to intensive dislocation activities, i.e. nucle- characterized by repetition of a smooth region (indicat-
ation and development of partial dislocations on planes ing a stable crystalline structure) followed by yielding
#7∼#9, around the tetrahedron (Fig.6d). It is interesting platform or load drop (signaling intensive dislocation ac-
to note that a dislocation nucleates horizontally on (111) tivities).
(plane #9).
For Cu (001), due to the effective resistance of Lomer-
Cottrell lock to further movement of dislocation, only
4 Summary yielding platform tends to appear. For Cu (110) and
(111), despite the existence of profuse lock formation,
In this paper, molecular dynamics simulations of nano-
dislocations are subjected to much weaker resistance
indentation are performed to study the elastic anisotropy
than for Cu (001), which results in frequent load drop.
effects on the elastic-plastic deformation on copper
In addition, a geometrically necessary dislocation is ob-
(001), (110) and (111) surfaces. The evolution of the un-
served for Cu (110) responsible for a large force drop.
derlying dislocation structure is correlated with the load-
displacement relation.
Acknowledgement: The work described in this paper
Due to elastic anisotropy, the reduced modulus of (111) was supported by grants from the Research Grants Coun-
surface is found to be the largest, while that of (001) the cil of the Hong Kong Special Administrative Region
smallest. The normal stress distribution beneath the in- (PolyU 1/99C, PolyU 5309/03E and PolyU 5312/03E).
denter shows different rotational symmetry. It is inter-
esting to find that stress tensors calculated from finite el-
114 Copyright 
c 2004 Tech Science Press cmes, vol.6, no.1, pp.105-114, 2004

References Rev. B, Vol. 57, No. 19, pp.12588-12594.


Kiely, J. D.; Jaraush, K. F.; Houston, J. E.; Russell, P.
Chiu, Y. L.; Ngan, A.H.W. (2002): Time-dependent E., (1999): Initial stages of yield in nanoindentation, J.
characteristics of incipient plasticity in nanoindentation Mater. Res., Vol. 14, No. 6, pp.2219-2227.
of a Ni3 Al single crystal, Acta Mater., Vol. 50, pp.1599- Knap, J.; Ortiz, M. (2003): Effect of indentation-radius
1611. size on Au (001) nanoindenation, Phys. Rev. Lett., Vol.
Corcoran, S. G.; Colton, R. J.; Lilleodden, E. T.; Ger- 90, No. 22, pp.226102-226105.
beric, W. W. (1997): Anomalous plastic deformation at Li, J.; Van Vliet, K. J.; Zhu, T.; Yip, S.; Suresh, S.
surfaces: nanoindentation of gold single crystals, Phys. (2002): Atomistic mechanisms governing elastic limit
Rev. B, Vol. 55, No. 24, pp.16057-16060. and incipient plasticity in crystals, Nature, Vol. 418,
Christopher, D.; Smith, R.; Richter, A. (2001): Atom- pp.307-310.
istic modeling of nanoindentation in iron and silver, Nan- Liang, H. Y.; Woo, C. H.; Huang, H. C.; Ngan, A.
otechnology, Vol. 12, pp.372-383. H. W.; Yu, T. X. (2003): Dislocation Nucleation in the
Doyama, M.; Kogure, Y. (1999): Embedded atom po- Initial Stage During Nanoindentation, Philos. Mag., Vol.
tentials in fcc and bcc metals, Comp. Mat. Sci., Vol. 14, 83, No.31-34, pp.3609-3622.
pp.80-83. Lilleodden, E. T.; Zimmerman, J. A.; Foiles, S. M.;
Egami, T.; Maeda, K.; Vitek, V. (1980): Structural de- Nix, W. D. (2003): Atomistic simulations of elastic de-
fects in amorphous solids - a computer simulation study, formation and dislocation nucleation during nnaoinden-
Phil. Mag. A, Vol. 41, No. 6, pp.883-901. tation, J. Mech. Phys. Solids, Vol. 51, No.5, pp.901-920.
Gane, N.; Bowden, F. P. (1968): Microdeformation of Rodrı́guez de la Fuente, O.; Zimmerman, J. A.;
solid, J. Appl. Phys., Vol. 39, No. 3, pp.1432-1435. González, M. A.; de la Figuera, J.; Hamilton, J. C.;
Gannepalli. A.; Mallapragada. S. K. (2002): Atom- Pai, W. W.; Rojo, J. M. (2002): Dislocation emis-
istic studies of defect nucleation during nanoindentation sion around nanoindentation on a (001) fcc metal surface
of Au (001), Phys. Rev. B., Vol. 66, pp.104103-104111. studied by scanning tunneling microscopy and atomistic
simulations, Phys. Rev. Lett., Vol. 88, No. 3, pp.36101-
Gao, H. J.; Huang, Y. G. (2003): Geometrically nec-
36104.
essary dislocation and size-dependent plasticity, Script.
Mater., Vol. 48, pp.113-118. Srivastava, D.; Atluri, S. N. (2002): Computational
nanotechnology: A current perspective. CMES: Com-
Gerberich, W. W.; Tymiak, N. I.; Grunlan, J. C.;
puter Modeling in Engineering & Sciences, vol. 3, no. 5,
Horstemeyer, M. F.; Baskes, M. I. (2002): Interpreta-
pp. 531-538.
tions of indentation size effects, J. Appl. Mech., Vol. 69,
pp.433-441. Suresh, S.; Nieh, T. G.; Choi, B. W. (1999): Nano-
indentation of copper thin films on silicon substrates,
Ghoniem, N. M.; Cho, K. (2002): The emerging role
Script. Mater., Vol. 41, No. 9, pp.951-957.
of multiscale modeling in nano- and micro-mechanics of
materials. CMES: Computer Modeling in Engineering & Swadener, J. G.; George, E. P.; Pharr, G. M. (2002):
Sciences, vol. 3, no. 2, pp. 147-174. The correlation of the indentation size effect measured
with indenters of various shapes, J. Mech. Phys. Solids,
Gouldstone, A.; Koh, H. J.; Zeng, K. Y.; Gian-
Vol. 50, pp.681-694.
nakopoulos, A. E.; Suresh, S. (2000): Discrete and
continuous deformation during nanoindentation of thin Vlassak, J. J.; Nix, W. D. (1994): Measuring the elastic
films, Acta Mater., Vol. 48, pp.2277-2295. properties of anisotropic materials by means of indenta-
tion experiments,J. Mech. Phys. Solids, Vol. 42, No. 2,
Kelchner, C. L.; Plimpton, S. J.; Hamilton, J. C.
pp.1223-1245.
(1998): Dislocation nucleation and defect structure dur-
ing surfaces indentation, Phys. Rev. B, Vol. 58, No.17, Zimmerman, J. A.; Kelchner, C. L.; Klein, P. A.;
pp.11085-11088. Hamilton, J. C.; Foiles, S. M. (2001): Surface step ef-
fects on nanoindentation, Phys. Rev. Lett., Vol. 87, No.
Kiely, J. D.; Houston, J. E. (1998): Nanomechanical
16, pp.165507-165510.
properties of Au (111), (001), and (110) surfaces, Phys.

You might also like