plastic deformation of freestanding thin films: experiments

22
Journal of the Mechanics and Physics of Solids 54 (2006) 2089–2110 Plastic deformation of freestanding thin films: Experiments and modeling L. Nicola a , Y. Xiang b , J.J. Vlassak b , E. Van der Giessen c , A. Needleman a, a Division of Engineering, Brown University, Providence, RI 02912, USA b Division of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA c Department of Applied Physics, University of Groningen, Nyenborgh 4, 9747 AG Groningen, The Netherlands Received 19 October 2005; received in revised form 7 March 2006; accepted 8 April 2006 Abstract Experimental measurements and computational results for the evolution of plastic deformation in freestanding thin films are compared. In the experiments, the stress–strain response of two sets of Cu films is determined in the plane-strain bulge test. One set of samples consists of electroplated Cu films, while the other set is sputter-deposited. Unpassivated films, films passivated on one side and films passivated on both sides are considered. The calculations are carried out within a two- dimensional plane strain framework with the dislocations modeled as line singularities in an isotropic elastic solid. The film is modeled by a unit cell consisting of eight grains, each of which has three slip systems. The film is initially free of dislocations which then nucleate from a specified distribution of Frank–Read sources. The grain boundaries and any film-passivation layer interfaces are taken to be impenetrable to dislocations. Both the experiments and the computations show: (i) a flow strength for the passivated films that is greater than for the unpassivated films and (ii) hysteresis and a Bauschinger effect that increases with increasing pre-strain for passivated films, while for unpassivated films hysteresis and a Bauschinger effect are small or absent. Furthermore, the experimental measurements and computational results for the 0.2% offset yield strength stress, and the evolution of hysteresis and of the Bauschinger effect are in good quantitative agreement. r 2006 Elsevier Ltd. All rights reserved. Keywords: Dislocations; Thin films; Bauschinger effect; Size effects; Computer simulation ARTICLE IN PRESS www.elsevier.com/locate/jmps 0022-5096/$ - see front matter r 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.jmps.2006.04.005 Corresponding author. E-mail address: [email protected] (A. Needleman).

Upload: others

Post on 05-Jan-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

ARTICLE IN PRESS

Journal of the Mechanics and Physics of Solids

54 (2006) 2089–2110

0022-5096/$ -

doi:10.1016/j

�CorrespoE-mail ad

www.elsevier.com/locate/jmps

Plastic deformation of freestanding thin films:Experiments and modeling

L. Nicolaa, Y. Xiangb, J.J. Vlassakb, E. Van der Giessenc,A. Needlemana,�

aDivision of Engineering, Brown University, Providence, RI 02912, USAbDivision of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA

cDepartment of Applied Physics, University of Groningen, Nyenborgh 4, 9747 AG Groningen, The Netherlands

Received 19 October 2005; received in revised form 7 March 2006; accepted 8 April 2006

Abstract

Experimental measurements and computational results for the evolution of plastic deformation in

freestanding thin films are compared. In the experiments, the stress–strain response of two sets of Cu

films is determined in the plane-strain bulge test. One set of samples consists of electroplated Cu

films, while the other set is sputter-deposited. Unpassivated films, films passivated on one side and

films passivated on both sides are considered. The calculations are carried out within a two-

dimensional plane strain framework with the dislocations modeled as line singularities in an isotropic

elastic solid. The film is modeled by a unit cell consisting of eight grains, each of which has three slip

systems. The film is initially free of dislocations which then nucleate from a specified distribution of

Frank–Read sources. The grain boundaries and any film-passivation layer interfaces are taken to be

impenetrable to dislocations. Both the experiments and the computations show: (i) a flow strength

for the passivated films that is greater than for the unpassivated films and (ii) hysteresis and a

Bauschinger effect that increases with increasing pre-strain for passivated films, while for

unpassivated films hysteresis and a Bauschinger effect are small or absent. Furthermore, the

experimental measurements and computational results for the 0.2% offset yield strength stress, and

the evolution of hysteresis and of the Bauschinger effect are in good quantitative agreement.

r 2006 Elsevier Ltd. All rights reserved.

Keywords: Dislocations; Thin films; Bauschinger effect; Size effects; Computer simulation

see front matter r 2006 Elsevier Ltd. All rights reserved.

.jmps.2006.04.005

nding author.

dress: [email protected] (A. Needleman).

ARTICLE IN PRESSL. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102090

1. Introduction

The decreasing dimensions of micro-electronic and micro-mechanical devices havemotivated research on the micro- and nano-scale mechanical behavior of materials. It hasbecome clear that, at such scales, the plastic response of crystalline solids can exhibit astrong size dependence, e.g. Fleck et al. (1994), Ma and Clark (1995), Stolken and Evans(1998). Various mechanisms can lead to this size dependence. One well-appreciated sourceof size dependence is associated with plastic strain gradients and geometrically necessarydislocations, e.g. Fleck et al. (1994). A size effect can also arise even when themacroscopically applied deformation is uniform if dislocation motion is constrained due toan internal interface such as a grain boundary; leading, for example, to a grain size-dependent flow strength, i.e. the Hall–Petch effect. Recently, a strong size effect has beenseen in single crystal compression tests (Uchic et al., 2004; Dimiduk et al., 2005; Greer etal., 2005), where dislocation glide is relatively unconstrained.Motivated by these experimental observations, a variety of theoretical frameworks have

been developed to incorporate one or more material length scales into a theory of plasticflow, e.g. Gao et al. (1999), Acharya and Bassani (2000), Fleck and Hutchinson (2001),Gurtin (2002). These phenomenological theories describe size dependence arising from thepresence of geometrically necessary dislocations and, for some theories, from boundaryconstraints on plastic flow. However, neither size dependence arising from ‘‘dislocationstarvation’’ (Uchic et al., 2004; Dimiduk et al., 2005; Greer et al., 2005) nor from source-limited plasticity is modeled by current phenomenological plasticity theories. On the otherhand, a framework for solving boundary value problems has been developed, where plasticflow occurs through the collective motion of discrete dislocations (Van der Giessen andNeedleman, 1995). In this discrete dislocation plasticity framework, size dependenceemerges as a natural outcome of the boundary value problem solution. Size effects occurdue to dislocation starvation and due to limited dislocation sources, in addition to arisingfrom geometrically necessary dislocations and boundary constraints (Cleveringa et al.,1999; Shu et al., 2001; Nicola et al., 2003, 2005a; Deshpande et al., 2005).Micro-scale devices typically contain thin films, either freestanding or bonded to a

substrate. Accordingly, much attention has been directed to experimentally characterizingthe mechanical properties of thin films by a variety of techniques, e.g. (Vlassak and Nix,1992; Shen et al., 1998; Hommel and Kraft, 2001; Espinosa et al., 2003, 2004; Dehm et al.,2003; Lou et al., 2003; Phillips et al., 2004; Florando and Nix, 2005; Xiang et al., 2005),and to modeling their size-dependent behavior, e.g (Hutchinson, 2001; Haque and Saif,2003; Pant et al., 2003; Groh et al., 2003; Von Blanckenhagen et al., 2004; Ghoniem andHan, 2005; Hartmaier et al., 2005; Nicola et al., 2003, 2005a–c). Additional references canbe found in the works cited. The experiments, carried out under either thermal ormechanical loading, have shown that the inelastic response depends strongly on filmthickness and on whether or not the films are passivated. Calculations of the size-dependent response carried out using phenomenological continuum plasticity constitutiverelations and discrete dislocation descriptions of plastic flow have qualitatively reproducedaspects of the plastic response seen in the experiments. What has been lacking is a directquantitative comparison between experiment and theoretical predictions that covers abroad range of film thicknesses and both passivated and unpassivated films. Here, wepresent such a direct quantitative comparison between a series of experiments onfreestanding Cu films, Xiang et al. (2002, 2004, 2005), Xiang and Vlassak (2005a, b),

ARTICLE IN PRESSL. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–2110 2091

involving cyclic as well as monotonic loading, and the predictions of discrete dislocationplasticity.

The experiments were carried out on polycrystalline films made either by anelectroplating or sputtering process, with thicknesses from 4:2 to 0:34mm. Films passivatedon one or both sides as well as unpassivated films were used in the experiments. Strongeffects of size and passivation were seen on the flow strength and on the Bauschinger effecton unloading.

The discrete dislocation formulation follows that of Nicola et al. (2005a) for films onstiff substrates, except that here the loading is mechanical rather than thermal. A planestrain model of the polycrystalline film is used. In this formulation, plastic flow arises fromthe collective motion of discrete edge dislocations, represented as line singularities in anelastic solid, with the long-range interactions between dislocations directly accounted for.Drag during dislocation motion, dislocation nucleation and dislocation annihilation areincorporated through a set of constitutive rules. The boundary conditions are enforced bysolving for an image field. This two-dimensional framework has predicted (Nicola et al.,2003), at least qualitatively, a range of features observed experimentally including, forexample, the bilinear hardening seen in sufficiently thin films (Florando and Nix, 2005).

In this study, three parameters entering the dislocation constitutive rules (the density ofdislocation sources, their average nucleation strength and the standard deviation of thenucleation strengths) are set to fit the measured stress–strain response of an electroplatedfilm passivated on both sides. Then, with all constitutive parameters fixed, the response, i.e.the evolution of the Bauschinger effect as well as of the flow strength, is predicted for theremaining cases, including unpassivated films as well as films passivated on one side. Thecalculations give insight into the mechanisms leading to the observed size dependence andthe predictions are compared quantitatively with the experimental observations.

2. Experiments

2.1. Methodology

The mechanical behavior of two sets of Cu films was investigated using the plane-strainbulge test (Xiang et al., 2005). In this test, silicon micromachining techniques are used tofabricate long rectangular membranes out of the film of interest. These membranes aredeformed in plane strain by applying a uniform pressure to one side of the membrane asillustrated in Fig. 1. The average stress s and strain � in the membrane are determined fromthe applied pressure, p, and the corresponding membrane deflection, d, using the followingexpressions (Xiang et al., 2005):

s ¼pða2 þ d2Þ

2dh; � ¼ �0 þ

ða2 þ d2Þ2ad

arcsin2ad

ða2 þ d2Þ� 1, (1)

where h is the film thickness, 2a the width of the membrane and e0 the residual strain in thefilm. Using this technique, it is possible to measure the plane strain stress–strain curve of athin film, as well as its residual stress. With some care, films as thin as 100 nm can be tested.

The first set of samples consisted of electroplated Cu films with thicknesses varying from1:0 to 4:2mm, but with constant microstructure. These films were fabricated byelectroplating 5:4�mm-thick Cu films on Si substrates using a commercial plating process

ARTICLE IN PRESS

Fig. 1. Perspective views of a typical bulge test sample: (a) as prepared, the Si-framed freestanding membrane is

flat and under residual tension; (b) when a uniform pressure is applied to one side of the membrane, the film is

deformed in plane-strain tension with zero strain along the longitudinal direction. Only a section of the bulge test

sample is shown.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102092

(Xiang et al., 2004). Prior to the plating process, the Si substrates were coated with a 75 nmSi3N4 film using low-pressure chemical vapor deposition, a 20 nm TaN adhesion layerusing reactive sputtering, and a Cu seed layer using magnetron sputtering. Theelectroplated films were annealed at 400 �C for 15min in vacuum to obtain a uniformand stable microstructure. The films were subsequently thinned to the desired thickness bymeans of chemical–mechanical planarization. Further details of the materials preparationprocess are given in Xiang et al. (2004).Freestanding Cu membranes were fabricated by opening long rectangular windows in

the substrate using Si micromachining techniques (Xiang et al., 2005) and by etching theexposed Si3N4 and TaN coatings. Some of the membranes were passivated by sputtercoating 20 nm Ti onto either one or both surfaces of the freestanding Cu membranes.The second set of samples consisted of sputter-deposited Cu films with thicknesses

ranging from 0:34 to 0:89mm. For each thickness, Cu was sputtered directly ontofreestanding bilayer membranes that consisted of 75 nm of Si3N4 and 20 nm of TaN.Immediately prior to the Cu deposition, the TaN surface of the membranes was sputter-cleaned in situ using an Ar plasma. After deposition, all membranes were annealed invacuum to increase the grain size of the films. Annealing temperatures were selected toensure that the membranes would not buckle during the annealing step as a consequence ofthe differential thermal expansion between the membrane and the Si frame. The resultinggrain sizes were measured using transmission electron microscopy (TEM) and are listed inTable 1. For some samples, the Si3N4/TaN layer was etched away using reactive ionetching (RIE) in order to create freestanding Cu membranes. Other samples were tested asannealed without removing the Si3N4/TaN layer.The reason for leaving the thin Si3N4/TaN layer in place is twofold: (i) The Si3N4/TaN

layer serves to passivate one of the surfaces of the Cu film by blocking dislocations fromexiting the film. Comparison of the stress–strain curves of these films with those ofunpassivated Cu films provides information on the effect of a free surface or interface onthe mechanical behavior of thin films. (ii) The Si3N4/TaN layer makes it possible to

ARTICLE IN PRESS

Table 1

Deposition method, thickness, h, and average grain size, d, of the films tested

Deposition method h (mm) d (mm) h=d

Electroplated 4:20� 0:05 1:50� 0:05 2:80� 0:131:90� 0:05 1:51� 0:04 1:26� 0:071:00� 0:05 1:50� 0:05 0:67� 0:04

Sputtered 0:89� 0:01 0:46� 0:02 1:93� 0:110:67� 0:01 0:46� 0:01 1:43� 0:050:61� 0:01 0:54� 0:02 1:14� 0:060:44� 0:01 0:39� 0:01 1:13� 0:050:35� 0:01 0:33� 0:01 1:06� 0:06

For the electroplated films, twins are counted as separate grains; the actual mean grain size, not counting twins as

separate grains, is about 2:7mm.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–2110 2093

alternate the direction of plastic flow in the Cu film during a bulge test experiment asdescribed in Xiang and Vlassak (2005a). This is realized by first loading the compositemembrane in the bulge test until the Cu flows plastically. The Si3N4/TaN layer onlydeforms elastically and a large tensile stress is built up in this layer. Upon unloading, thetensile stress in the Si3N4/TaN drives the Cu film into compression, while the overall stressin the composite membrane is kept tensile to prevent buckling of the membrane. Thestress–strain curve of the Cu film is then obtained by subtracting the elastic contribution ofthe TaN/Si3N4 coating from the stress–strain curve of the composite film. The contributionof the TaN/Si3N4 coating is readily determined by bulge testing the membrane after the Cufilm has been dissolved in dilute nitric acid. It should be noted that this techniquecircumvents the problem of buckling when testing freestanding thin films in compressionbecause the average stress in the membrane is kept tensile. As such, it is the only techniquecurrently available to test the same freestanding film in both tension and compression.

Table 1 shows the film thickness h and average grain size d for all films tested. Cross-section TEM micrographs of the electroplated films, Fig. 2(a), reveal that the thinnest filmshave a grain size on the order of the film thickness with grain boundaries traversing theentire film; thick films typically have several grains through the thickness. This is notsurprising given that the average grain size of the electroplated films is approximately one-third the thickness of the thickest film. These films contain many twins, which give rise tointernal boundaries (Xiang et al., 2004, 2006). The actual mean grain size of these films isapproximately 2.7 mm, but counting the twins as separate grains gives an average grain sizeof 1:5mm as shown in Table 1, independent of film thickness. X-ray diffraction (XRD)measurements by Xiang et al. (2004) show that the crystallographic texture of theelectroplated films is independent of thickness and that it has ð1 1 1Þ, ð1 0 0Þ and ð1 1 0Þcomponents, with the plane-strain tension Taylor factor ranging between 2.97 and 3.00.The sputtered films have a mixed texture that consists mainly of ð1 1 1Þ and ð1 0 0Þ fibercomponents, with the ð1 0 0Þ component slightly weaker for the thinner films. The plane-strain tension Taylor factor of the thinnest sputtered film is 3.17 and that of the thickestsputtered film 3.11. The grain size of the sputtered films is a weak function of filmthickness. TEMmicrographs show that the sputtered films have a columnar grain structurewith equiaxed grains in the plane of the film. Fig. 2(b) presents a TEM micrograph of an

ARTICLE IN PRESS

Fig. 2. Cross-section TEM micrographs of freestanding electroplated Cu films after deformation showing:

(a) grain size on the order of the film thickness in a 1:0�mm-thick unpassivated film; (b) a region of high dislocation

density at the Cu–Ti interface in a 1:9mm passivated film.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102094

electroplated film with a Ti passivation layer that was deformed in the plane of the film. A50 nm region of high dislocation density can be observed near the Cu/Ti interface,indicating that this interface prevents dislocations from exiting the film, which hasimportant consequences for the plastic response of the film.

2.2. Experimental results

Typical stress–strain curves for the electroplated Cu films are presented in Fig. 3.Fig. 3(a) shows the effect of the Ti passivation layer on the stress–strain curve of a 1:0mmCu film, while Fig. 3(b) illustrates the effect of film thickness for films with both surfacespassivated by 20 nm Ti. For a given film thickness, the passivation layer clearly increasesthe work hardening rate of the film.Fig. 4 shows typical stress–strain curves for a 0:6�mm-thick sputter-deposited film. Films

both with and without TaN/Si3N4 passivation are shown. The film with passivation showssignificantly more work hardening than the unpassivated film. Moreover, the film withpassivation is in compression at the end of the last two unloading cycles.All unloading/reloading cycles for this film show significant hysteresis with reverse

plastic flow occurring even when the stress in the film is still tensile. By contrast, the filmwithout passivation shows little or no reverse plastic deformation when fully unloaded; i.e.,passivated films show hysteresis and a strong Bauschinger effect, while unpassivated filmsdo not.

3. Modeling

3.1. Problem formulation

The formulation and solution procedure follow that used by Nicola et al. (2005a) forfilms on stiff substrates, except that here the loading is mechanical rather than thermal. Amore complete description of the analysis procedure and further references are given inNicola et al. (2005a). As sketched in Fig. 5, the polycrystalline film is modeled as an infinitearray of planar rectangular grains of height h and width d. Plane strain conditions areassumed and the elastic anisotropy of the grains is neglected. Plasticity in the film

ARTICLE IN PRESS

ε ε

<σ 1

1>[M

Pa]

00

50

100

150

200

250

300

350

400

two surfaces passivatedone surface passivatedunpassivated

00

50

100

150

200

250

300

350

400

h=1µmh=1.9µmh=4.2µm

0.002 0.004 0.006<

σ 11>

[MPa

]0.002 0.004 0.006 0.008 0.01

(a) (b)

Fig. 3. Typical stress–strain curves for the freestanding electroplated Cu films with constant microstructure:

(a) the effect of passivation by 20 nm Ti on 1:0mm films and (b) the effect of film thickness on films passivated on

both surfaces with 20 nm Ti. All curves are offset by the equi-biaxial residual strains in the films, as represented by

the dashed lines starting from the origins. Experimental data from Xiang et al. (2004).

<σ 1

1> [

MPa

]

0-200

-100

0

100

200

300

400

500

unpassivatedone side passivated

0.0080.0060.002 0.004

ε

Fig. 4. Typical stress–strain curves for a 0:6�mm-thick sputter-deposited Cu film: the effect of the TaN/Si3N4

passivation. Both curves are offset by the equi-biaxial residual strains in the films, as represented by the dashed

line starting from the origin. Experimental data from Xiang and Vlassak (2005a).

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–2110 2095

ARTICLE IN PRESS

h

w

d

+−

εε

x1

x2 p

φ

Fig. 5. Two-dimensional model of a freestanding film passivated on both sides under tensile loading. The

calculations are carried out for a cell consisting of eight grains (w ¼ 8d in the calculations) and for films passivated

on one or both sides as well as for unpassivated films. The left most grain shows the sign convention for

dislocations on the slip system oriented at f and positive dislocations on each of the three slip systems are shown

in the center grain.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102096

originates from the motion of edge dislocations in the x1–x2 plane which are treated as linesingularities in an elastic solid. The calculations are carried out for a unit cell of width w

consisting of eight grains (w ¼ 8d). Each grain has three sets of slip planes on which edgedislocations having Burgers vector b, can nucleate and glide. The slip systems are orientedat f, fþ 60� and fþ 120� from the x1-axis. The orientation of each grain, f, is chosenrandomly from the range ½0; pÞ. Films passivated on one or both sides as well asunpassivated films are considered. Any passivation layers remain elastic, are of thickness p

and are taken to have elastic constants identical to those of the film. The grain boundariesas well as any film-passivation layer interfaces are taken to be flat and impenetrable todislocations.Tension is imposed by a prescribed displacement difference between unit cells. In

particular, for the unit cell for which 0px1pw,

uið0;x2Þ ¼ uiðw; x2Þ �UðtÞdi1, (2)

where UðtÞ is a prescribed function of time and dij is the Kronecker delta. The imposedstrain is e ¼ U=w. The top and bottom surfaces are traction-free so that

s12ðx1; 0Þ ¼ s22ðx1; 0Þ ¼ 0; s12ðx1; hþ kpÞ ¼ s22ðx1; hþ kpÞ ¼ 0; k 2 0; 1; 2 (3)

where k is the number of passivation layers (counting from the bottom x2 ¼ 0). Inaddition, to prevent rigid body motions

u1ð0; 0Þ ¼ u2ð0; 0Þ ¼ 0. (4)

Although the boundary conditions, Eqs. (2)–(4), completely specify a boundary valueproblem, the highly nonuniform distribution of plastic deformation leads to extensivebending of the film. The actual films do not exhibit this behavior due to the constraintimposed by grains located in front of and behind the plane of deformation considered inthe two-dimensional analysis. We mimic this constraint by imposing

u2 x1;h

2þ kp

� �¼ 0; k 2 0; 1 (5)

along the mid-plane of the film.The solution of the boundary value problem is based on using superposition (Van der

Giessen and Needleman, 1995) of the singular elastic fields of individual dislocations

ARTICLE IN PRESSL. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–2110 2097

together with the solution of a boundary value problem for nonsingular image fields(Nicola et al., 2005a). The finite element method is used to obtain the image field solutions.In this approach, the long-range elastic interactions between dislocations are accounted fordirectly, and short-range interactions are added through a set of constitutive rules of thetype suggested by Kubin et al. (1992) for dislocation glide, dislocation annihilation anddislocation nucleation.

The simulations start from a dislocation free state. Dislocation sources are randomlydistributed on the slip planes with each source characterized by a nucleation strength tnuc,chosen out of a Gaussian distribution of strengths. A source becomes active when thePeach–Koehler force on dislocation I, given by

f ðIÞ ¼ nðIÞi sij þ

XJaI

sðJÞij

!bðIÞj , (6)

exceeds btnuc for a time span tnuc. Here, nðIÞi is the normal to the slip plane of dislocation I

and bðIÞj its Burgers vector. Also, ~sðJÞij is the singular stress field of dislocation J and sij is the

image stress field. When the nucleation criterion is met, a dislocation dipole is nucleated:two dislocations of opposite sign are introduced on the slip plane at a distance Lnuc whichis such that the attractive stress field that the dislocations exert on each other isequilibrated by tnuc, so that

Lnuc ¼E

4pð1� n2Þb

tnuc, (7)

where E is Young’s modulus and n is Poisson’s ratio.The glide velocity V

ðIÞgln along the slip plane of dislocation I is taken to be linearly related

to the Peach–Koehler force through the drag relation

VðIÞgln ¼

1

Bf ðIÞ, (8)

where B is the drag coefficient.When two opposite signed dislocations come closer to each other than the critical

distance Lann they annihilate and are removed from the calculation. Dislocations that exitthe film through a free surface leave a step at the surface.

The deformation history is calculated incrementally. At time t, the stress anddisplacement fields are known at each material point along with the positions of alldislocations. An increment of the applied displacement U is prescribed and calculation ofthe state at time tþ dt involves: (i) determining the Peach–Koehler forces on thedislocations; (ii) determining the rate of change of the dislocation structure caused by themotion of dislocations, the generation of new dislocations and their mutual annihilationand (iii) determining the stress and strain state for the updated dislocation arrangement.

3.2. Choice of parameters

The Burgers vector is taken to be b ¼ 0:25 nm, Young’s modulus is E ¼ 110GPa andPoisson’s ratio is n ¼ 0:34, values that are representative of copper. The elastic constants ofany passivation layers are taken to be identical to those of the film. The annihilationdistance, Lann, is fixed at 6b and the drag coefficient at B ¼ 10�4 Pa s. Dislocation sources

ARTICLE IN PRESSL. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102098

are evenly distributed among the eight grains in the unit cell and are randomly positionedon the slip planes. Parallel slip planes are spaced at 200b.The values of the source density, the average nucleation strength tnuc and its standard

deviation were set by fitting the measured stress–strain response of the electroplated filmwith h ¼ 4:2 mm that is passivated on both sides. The flow strength is mainly sensitive tothe source strength while the hardening rate is mainly sensitive to the source density. Toaccount for the constraint on slip due to the twins, the grain size in calculations for theelectroplated films is taken to be d ¼ 1:5mm (see Table 1).Curves of average stress in the film, hs11i, versus imposed strain, e, are shown in Fig. 6,

with

hs11i ¼1

h

Zh

s11ðw;x2Þdx2, (9)

where the integral over the film thickness excludes any passivation layer. Note that fromequilibrium hs11i is independent of x1. To facilitate comparison between the computed andexperimental stress–strain curves, the experimental curves from Fig. 3(b) are also plotted.The calculations in Fig. 6, which shows the response for three values of film thickness, werecarried out using a source density of 15mm�2 and a mean nucleation strength tnuc ¼

ε

<σ 1

1> [

MPa

]

00

50

100

150

200

250

300

350

h=1µm

h=1.9µm

h=4.2µm

model exp

0.002 0.004 0.006

Fig. 6. Curves of average stress, hs11i, given by Eq. (9) versus applied strain e for films with both surfaces

passivated. The 0.2% yield strength, sy, is obtained from the intersection of the dash-dotted line (elastic slope)

with the computed stress–strain curve. The grain size is taken as d ¼ 1:5mm to account for the constraint on slip

imposed by the twins and, for comparison purposes, the experimental stress–strain data from Fig. 3(b) are also

plotted. The values of the dislocation nucleation parameters (specified in the text) were chosen to match the

experimental stress–strain response of the h ¼ 4:2mm film.

ARTICLE IN PRESSL. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–2110 2099

100MPa with a standard deviation of 20MPa. With tnuc ¼ 100MPa, from Eq. (7)Lnuc ¼ 0:0247mm. The films are passivated on both sides by elastic coatings of thicknessp ¼ 20 nm and subjected to a displacement rate of _U ¼ 3� 104 mm=s (this very high strainrate, _e ¼ 2500 s�1, is used to limit the computation time).

The 0.2% offset yield strength, denoted by sy, is obtained from the intercept of the dash-dotted line in Fig. 6, which has slope E=ð1� n2Þ, with the computed stress–strain curves.The values of sy are 175, 219 and 245MPa for the 4.2, 1.9 and 1 mm films, respectively. Thecorresponding experimental values are 179, 204 and 237MPa. The initial portions of allthree computed and experimental stress–strain curves agree quite well and a good fit isobtained for the values of sy. However, the hardening becomes linear for the calculationswith h ¼ 1:9 and 1mm while the experimental curves continue to exhibit parabolichardening. The lack of an accurate hardening description as deformation proceeds mostlikely arises from idealizations of the analysis. In particular: (i) the dislocation constitutiverules used in the calculations do not account for effects such as the creation of dislocationsources and obstacles during the deformation history (the extended dislocation constitutiverules developed by Benzerga et al. (2004) provide a more accurate hardening description)and (ii) the grain boundaries and interfaces are idealized as being impenetrable todislocations and the possibility of grain boundaries and interfaces acting as sources and/orsinks for dislocations is not accounted for.

The same dislocation constitutive parameters are employed in all calculations reportedon here; for unpassivated as well as passivated films, and both for the sputtered and theelectroplated films.

3.3. Numerical results

Both monotonic and cyclic loading calculations were carried out for a range of grainsizes corresponding to those in the experiments. For the sputtered films, calculations werecarried out with both h ¼ d (square grains) and with d fixed at 0:5mm to give an indicationof the effect of grain aspect ratio on the response. In this section, results are presented toillustrate the qualitative features that emerge from the calculations; in particular, theeffects of passivation, grain size and grain aspect ratio. We emphasize that allresults presented here and subsequently use the fixed set of material parameters given inSection 3.2.

Fig. 7 illustrates the effect of passivation on the film stress–strain response. Computedcurves of hs11i versus e are shown for films with no passivation, with one side passivatedand with both sides passivated. For all three films h ¼ 1mm. Consistent with theexperimental data, shown for comparison purposes in Fig. 7, the flow strength increaseswith an increasing number of passivation layers. However, for the passivated films thecomputations give linear hardening for strains greater than � 0:003 while the experimentscontinue to show parabolic hardening. Also, the computational results somewhatoverestimate the effect of the passivation. For example, the experiments give 0.2% offsetyield strength values of 196, 210 and 233MPa, while the corresponding computed valuesare 182, 214 and 245MPa (the computation with two surfaces passivated in Fig. 7 has adifferent realization of the statistically distributed dislocation sources than the curve withh ¼ 1mm shown in Fig. 6 and the corresponding stress–strain curves in the two figuresdiffer slightly). Although the stress level is lower, the calculation for the unpassivated filmin Fig. 7 exhibits strain hardening similar to that seen in the experiments. With the

ARTICLE IN PRESS

ε

<σ 1

1> [

MPa

]

00

50

100

150

200

250

300

350

two surfaces passivated

one surface passivated

unpassivated

model exp

0.002 0.004 0.006

Fig. 7. Curves of average stress, hs11i, versus applied strain e showing the effect of the presence of passivation

layer(s) on the film response. The films have h ¼ 1mm and d ¼ 1:5mm. For comparison purposes, the

experimental stress–strain data from Fig. 3(a) are also plotted.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102100

dislocation constitutive rules used, hardening arises from dislocations remaining inside thefilm due to being pinned either at the interface between the film and any passivation layersor at grain boundaries. For the unpassivated film in Fig. 7, where h ¼ 1mm andh=d ¼ 0:67, many of the dislocations exit through the film surfaces, so that a high densityof dislocations has not developed over the range of strain considered. The calculationsunderestimate the 0.2% yield strength of the unpassivated film and overestimate the 0.2%yield strength of the film with two surfaces passivated. Thus, the increase in hardening dueto passivation is overestimated by the modeling. This suggests that the discrepancybetween the computed and observed stress–strain behavior mainly arises from theidealization of grain boundaries and interfaces as being completely impenetrable todislocations.The effect of passivation on the dislocation structure that develops and on the stress

distribution is shown in Fig. 8. Superposed on the contours of s11 is the distribution ofdislocations and sources. For the unpassivated film, Fig. 8(c), relatively few dislocationspile up at the grain boundaries. With passivation, dislocations pile up at the film-passivation layer interface, but also the number of dislocations piled up at grainboundaries increases. The significantly higher stress level in the passivated films, as well asthe large grain-to-grain variation, can be seen in Fig. 8. Large grain-to-grain variations (inthe number of slip lines per grain) are also seen in the experiments, Fig. 2, although it ispossible that different diffraction conditions for different grains may be responsible for atleast some of the grain-to-grain variation. It is also worth noting that an absence of

ARTICLE IN PRESS

Fig. 8. Distributions of s11 and dislocations at e ¼ 0:005 for films with h ¼ 1mm that are (a) passivated on both

sides; (b) passivated on one side; or (c) unpassivated. The solid circles mark the locations of the dislocation

sources, the þ and � symbols represent dislocations (with the sign convention of Fig. 5). All distances are in mm.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–2110 2101

dislocations, as in the grain in Fig. 8(c) lying between x1 ¼ 6 and 7:5mm, does not imply alack of dislocation activity: dislocations nucleated in this grain have exited through the freesurfaces. The dislocation structure seen in Figs. 8(a) and (b) that develops due to therestriction on dislocation glide by the film-passivation layer (or layers) and by the grainboundaries is what gives rise to having a back stress that inhibits subsequent nucleation atthe dislocation sources, which in turn leads to increased stress levels.

The predicted stress–strain response is sensitive to the grain aspect ratio, as illustratedfor unpassivated films in Fig. 9. Curves of average stress, hs11i, versus applied strain, e, areshown for two grain aspect ratios with the film thickness fixed at h ¼ 1mm. In onecalculation, the grain size is at the reference value d ¼ 1:5mm, so that h=d ¼ 0:67; in theother calculation d ¼ 0:3mm so that h=d ¼ 3:33 (note that even with d ¼ 0:3mm, the grainsize is still more than an order of magnitude greater than Lnuc). With the larger value of thegrain aspect ratio, h=d, the number of active slip planes terminating at grain boundariesincreases, leading to an increase in the number of dislocations piled up at the grainboundaries and, as a consequence, to hardening.

The effects of grain aspect ratio and of passivation on hysteresis and on the Bauschingereffect are shown in Fig. 10. Fig. 10(a) shows results for the reference values h ¼ 1mm andd ¼ 1:5mm so that h=d ¼ 0:67, while in Fig. 10(b), h ¼ 0:6mm, d ¼ 0:5mm givingh=d ¼ 1:2. In Fig. 10(a), the unpassivated film exhibits nearly no hysteresis and noBauschinger effect, whereas the film passivated on one side shows a Bauschinger effect andhysteresis loop that increases with increasing pre-strain. Reverse plastic flow starts whenthe film is still in tension. The high back stress from the dislocation pile-ups provides thedriving force for dislocation activity during the early stages of unloading. First, dislocationpile-ups collectively stretch out, with dislocations gliding back on their slip planes andsubsequently new dislocations (with an opposite sign from those generated during loading)

ARTICLE IN PRESS

ε

<σ 1

1> [

MPa

]

00

100

200

300

400

500

0.002 0.004 0.006 0.008 0.01

d=0.3µm

d=1.5µm

Fig. 9. Average stress, hs11i, versus applied strain e for two unpassivated films with h ¼ 1mm; h=d ¼ 0:67 for one

and h=d ¼ 3:33 for the other.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102102

are nucleated. On reloading, there is a significant knee in the stress–strain curve due to thedislocations present in the film.The evolution of the hysteresis and the Bauschinger effect in Fig. 10(a) reproduces the

trends seen in Fig. 4: (i) for the film passivated on one side the width of the hysteresis loopand the Bauschinger effect increase with increasing pre-strain while (ii) no hysteresis orBauschinger effect is evident for the unpassivated film. The agreement between theory andexperiment is remarkable; the evolution of the size and shape of hysteresis loops is difficultto model within a phenomenological plasticity framework. Discrete dislocation plasticity isapparently able to capture the generation of the internal stress that is responsible for thehysteresis. In Fig. 10(b), with a higher grain aspect ratio, h=d ¼ 1:2, some hysteresis is seeneven for an unpassivated film. This is due to increased piling-up of dislocations against thegrain boundaries of higher aspect ratio grains. It is worth noting that in the experiments,the films are pre-strained during deposition so that there are initial stresses due to thedislocation structure whereas in the calculations the initial state is taken to be stress free.Nevertheless, the same features are seen in the experimental and computational results.In Fig. 10(b), for the passivated film with h=d ¼ 1:2, the hardening is higher and a larger

hysteresis loop develops than for the unpassivated film. This suggests that hardening inthin films with small grain size occurs mainly because the back stress leads to thesuppression of dislocation nucleation. When the film thickness and grain size of the films isreduced, there are insufficient sources available on favorable slip planes to nucleatedislocations (this effect has been identified in Nicola et al., 2003, 2005c to be responsible

ARTICLE IN PRESS

ε

<σ 1

1> [

MPa

]

00

100

200

300

400

500

ε<

σ 11>

[M

Pa]

00

200

400

600

800

1000

h=1µm, d=1.5µm

one side passivatedunpassivated

one side passivatedunpassivated

h=0.6µm, d=0.5µm

0.005 0.01 0.015 0.005 0.01 0.015

(a) (b)

Fig. 10. Average stress, hs11i versus applied strain e for two films with: (a) h ¼ 1mm and d ¼ 1:5mm (h=d ¼ 0:67);(b) h ¼ 0:6mm and d ¼ 0:5mm (h=d ¼ 1:2). In each plot, one of the films is passivated on one side.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–2110 2103

for the enhanced hardening rate in very thin single crystal films on stiff substrates).Because of the high back stress, plastic flow occurs more readily when the sign of theloading is reversed.

4. Comparison of experimental and computational results

The computational results and experimental observations for the effect of film thicknesson the yield strength are shown in Figs. 11 and 12. In these figures the 0.2% offset yieldstrength, sy, is plotted versus 1=h. We recall that all computations were carried out usingthe same values of the three dislocation constitutive parameters—the density of dislocationsources, their average nucleation strength and the standard deviation of the nucleationstrengths—which were chosen to fit the experimental stress–strain response for anelectroplated film with h ¼ 4:2mm and passivated on both sides. The experimental resultsfor both the passivated (one-sided passivation, Fig. 11) and unpassivated (Fig. 12) filmsshow a 1=h scaling. A similar scaling is seen for the computational results for the largegrained electroplated films, but there is more scatter in the computational results for thesmall grain size sputtered films than in the experimental results and, particularly for theunpassivated films, the agreement between the experimental and computational results isnot as good as for the films with larger grains. One contribution to this is that allcalculations are carried out for eight grains in the unit cell analyzed. Since the dislocationsource density is the same for all films, thinner films and films with a smaller grain size,which have a smaller volume of material in the computational cell, have fewer dislocationsources and are therefore more affected by statistical variations. For some grain sizes,calculations were carried out for several realizations of the source distribution and,particularly for small grain sizes where there may be only one dislocation source per grain,the variation in yield strength is large. This is seen particularly in Fig. 12 for h=d ¼ 1 andh ¼ 0:3mm where the value of sy varies between 287 and 560MPa. Nevertheless, in

ARTICLE IN PRESS

σ y [M

Pa]

00

200

400

600

800

1000 electroplated (d=1.5µm)

sputtered

simulations (d=0.5µm)

simulations (h/d=1)

simulations (d=1.5µm)

0.5 1 1.5 2 2.5 3 3.5

1/h [µm-1]

Fig. 11. The 0.2% offset yield strength, sy, plotted versus the reciprocal of the film thickness 1=h for films

passivated on one side. The filled symbols show experimental values while computational results are denoted with

open symbols.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102104

addition to a clear agreement between computation and experiment for the trend of thevariation, considering the idealizations in the modeling, the quantitative agreement inFigs. 11 and 12 is remarkably good. This agreement suggests that the main contribution tothe size dependence in the experiments comes from the constraint on slip by grainboundaries (including the twin boundaries in the electroplated films) and by anypassivation layers.In both the experiments and the calculations four distinct slopes in the sy versus 1=h

curves are seen (Figs. 11 and 12): (i) for the passivated electroplated films; (ii) for thepassivated sputtered films; (iii) for the unpassivated electroplated films and (iv) for theunpassivated sputtered films. The difference in crystallographic texture between theelectroplated and sputtered films leads to only a small difference (about 5%) in the plane-strain tension Taylor factors so that the change in slope in the experiments cannot beattributed to a change in texture. Since the same material properties are used in thecalculations for the electroplated films and for the sputtered films, the agreement betweenthe experimental and computed values indicates that the different slopes seen in theexperiments are not due to a difference in material properties between the electroplatedand sputtered films. In the calculations, the difference between the electroplated films(hX1mm) and the sputtered films (ho1mm) is the difference in grain size; the electroplatedfilms are taken to have the fixed grain size 1:5 mm, whereas the grain size for the sputteredfilms is taken to be either 0:5mm or the film thickness h. Hence, the agreement between the

ARTICLE IN PRESS

0 1 1.50

200

400

600

800

1000 electroplated (d=1.5µm)

sputtered

simulations (h/d=1)

simulations (d=1.5µm)

σ y [

MPa

]

0.5 2 2.5 3 3.5

1/h [µm-1]

simulations (d=0.5µm)

Fig. 12. The 0.2% offset yield strength, sy, versus the reciprocal thickness 1=h for unpassivated films. The filled

symbols show experimental values while the open symbols show computational results.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–2110 2105

calculations and the experiments indicates that the rather abrupt change in slope at � 1mmin Figs. 11 and 12 is mainly associated with a grain size difference between the twosets of films.

The plots of dislocation density at a stage of deformation where hs11i is approximatelysy versus film thickness in Fig. 13 give insight into the means by which grain size affects theyield strength. In the computations for the electroplated films (hX1mm), the dislocationdensity at yield is comparatively low and increases slightly with decreasing film thickness.On the other hand, for the sputtered films (ho1mm) the dislocation density at yield ismuch higher but tends to decrease with decreasing film thickness. There is a lowerdislocation density in the computations modeling the larger grain size electroplated filmsbecause many of the nucleated dislocations are able to exit the film through anunpassivated surface. For the smaller grain size sputtered films, the grain boundariesprohibit many of the dislocations from gliding to the free surface, so the dislocation densityis higher. In these films, increasing sy corresponds to decreasing dislocation density due toa combination of two effects: (i) the hindering of slip by the grain boundaries andinterfaces leads to dislocation pile-ups that induce a back stress which inhibits furtherdislocation nucleation (Nicola et al., 2005a) and (ii) as the grain size is decreased, thenumber of dislocation sources per grain decreases. Passivation affects the magnitude of thedislocation density but the variation with film thickness in Fig. 13 exhibits similar trendsfor the passivated and unpassivated films.

The computed value of sy for the unpassivated electroplated films (hX1mm) is nearlyindependent of film thickness and is consistent with the resolved value of sy being

ARTICLE IN PRESS

0 2 2.5 30

50

100

150

200

250

0

50

100

150

200

250sputtered (d=0.5µm)electroplated (d=1.5µm)

sputtered (d=0.5µm)electroplated (d=1.5µm)

ρ [µ

m-2

]

ρ [µ

m-2

]

1/h [µm-1] 1/h [µm-1]

0.5 1 1.5 3.5 0 2 2.5 30.5 1 1.5 3.5

(a) (b)

Fig. 13. Calculated dislocation density at approximately sy versus the reciprocal of the film thickness, 1=h. For

the calculations modeling the sputtered films (ho1mm), only films with a fixed grain size, d, of 0:5mm are shown.

(a) Films passivated on one side. (b) Unpassivated films.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102106

approximately equal to tnuc (a mean nucleation strength of 100MPa corresponds tosy ¼ 231MPa; the values for the electroplated films in Fig. 12 are somewhat below this butwith a standard deviation of 20MPa, sources weaker than the mean are available). Valuesof sy in Figs. 11 and 12 for both the electroplated and sputtered films can be much abovethe stress required for dislocation nucleation. The strengthening mechanism seen herediffers from that for ‘‘dislocation starvation’’ which also gives rise to increasing strengthwith decreasing dislocation density, but where the increase in flow strength with decreasingsize stems from dislocations more readily exiting from small single crystal specimens andwhere the nucleation stress acts as an upper bound to the flow strength, as seen in thecalculations by Deshpande et al. (2005).As sketched in the inset to Fig. 14, the Bauschinger effect for unpassivated films and

films passivated on one side is quantified in terms of the difference between the actualresidual strain ep and the residual strain, eq, that would have been present if the unloadingwere purely elastic. The Bauschinger effect measure is then

eq � ep

ey,

where ey ¼ syð1� n2Þ=E.Due to experimental difficulties with unloading films at small pre-strains, in the

experiments the unloading from some of the earlier cycles is not complete and the value ofðeq � epÞ is taken at the lowest load in the cycle, which underestimates the value of theBauschinger strain at lower values of e0. In Fig. 14, this measure of the Bauschinger effectis plotted versus pre-strain e0. In both the experiments and computations, the unpassivatedfilms show nearly no Bauschinger effect while for the passivated films there is a significantBauschinger effect that increases with increasing pre-strain, e0. For passivated films, thepredicted Bauschinger effect tends to be on the high side of what is seen in the experiments.This is consistent with the calculations somewhat overestimating the back stress because of

ARTICLE IN PRESS

ε0/εy

(εq-

ε p)/

ε y

0 0.4 2

0

0.2

h=0.34; h=0.61; h=0.89 µm (unpassivated)h=0.60 µm (unpassivated)h=0.34 µmh=0.44 µmh=0.61 µmh=0.67 µmh=0.89 µmh=0.30 µmh=0.60 µm

εεp

εq

ε0

σ

0.8 1.2 1.6 2.4

h=0.80 µm

0.4

0.6

Fig. 14. Bauschinger effect versus applied pre-strain e0 in unpassivated films and films passivated on one side. The

Bauschinger effect is quantified in terms of ðeq � epÞ=ey, where ep is the residual strain, eq is the residual strain

corresponding to a hypothetical purely elastic unloading, ey is the plane strain yield strain and e0 is the pre-strain,as sketched in the inset. The filled symbols show experimental values while computational results are denoted with

open symbols. The grain size of the simulated films is d ¼ 0:5mm, the grain size of the films in the experiments is

listed in Table 1.

L. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–2110 2107

the idealization that grain boundaries and interfaces are completely impenetrable todislocations.

5. Concluding remarks

A quantitative comparison has been carried out between experimental measurementsand discrete dislocation plasticity predictions for the thickness dependence of the tensilestress–strain response of freestanding passivated (on one or both sides) and unpassivatedCu films. Films with thicknesses ranging from 0:3 to 4:2mm and having a columnar grainstructure were considered. The values of the dislocation source density, average sourcestrength and standard deviation were chosen to fit the experimentally measuredstress–strain response of a film passivated on both sides and with a thickness of 4:2mm.

With no further adjustable parameters, the discrete dislocation plasticity calculations: (i)reproduce the trends seen in the experiments and (ii) are in remarkably good quantitativeagreement with the experimental measurements for the 0.2% offset yield strength, and forthe evolution of hysteresis and of the Bauschinger effect. For example,

The yield strength of unpassivated films with sufficiently large grains is nearlyindependent of film thickness. � For sufficiently thin unpassivated films and for films passivated on one side, the yield

strength increases with decreasing film thickness as h�1. The slope of the variation withh�1 is much greater for the passivated films than for the unpassivated films.

ARTICLE IN PRESSL. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102108

In the passivated films, there is hysteresis and a Bauschinger effect, and the size of thehysteresis loop increases with increasing pre-strain. The evolution of such hysteresisloops is difficult to predict using a phenomenological plasticity framework. In theunpassivated films, the hysteresis is small or absent and there is little or no Bauschingereffect.

In the calculations for passivated films, linear hardening eventually occurs whereasparabolic hardening is seen to continue in the experiments. This difference is most likelydue to the interfaces and grain boundaries taken to be completely impenetrable todislocations. It is also possible that the use of simple dislocation constitutive rules that donot account for effects such as cross-slip and the dynamic creation of dislocation sourcesand obstacles plays a role.The calculations indicate that the increase in yield strength with decreasing film

thickness mainly arises from the restriction on slip due to any passivation layers and grainboundaries leading to a dislocation structure with a back stress that inhibits dislocationnucleation.

Acknowledgments

LN and AN are grateful for support from the Materials Research Science andEngineering Center at Brown University (NSF Grant DMR-00520651). YX and JV wouldlike to acknowledge financial support from the National Science Foundation (NSF GrantDMR-0133559) and from the Materials Research Science and Engineering Center of theNational Science Foundation under NSF Grant DMR-0213805. The work by LN andEVdG is partly funded by the Netherlands Institute for Metals Research.

References

Acharya, A., Bassani, J.L., 2000. Lattice incompatibility and a gradient theory of crystal plasticity. J. Mech. Phys.

Solids 48, 1565–1595.

Benzerga, A.A., Brechet, Y., Needleman, A., Van der Giessen, E., 2004. Incorporating three-dimensional

mechanisms into two-dimensional dislocation dynamics. Modelling Simul. Mater. Sci. Eng. 12, 159–196.

Cleveringa, H.H.M., Van der Giessen, E., Needleman, A., 1999. A discrete dislocation analysis of bending. Int. J.

Plasticity 15, 837–868.

Dehm, G., Balk, T.J., Edongue, H., Arzt, E., 2003. Small-scale plasticity in thin Cu and Al films. Microelectron.

Eng. 70, 412–424.

Deshpande, V.S., Needleman, A., Van der Giessen, E., 2005. Plasticity size effects in tension and compression of

single crystals. J. Mech. Phys. Solids 53, 2661–2691.

Dimiduk, D.M., Uchic, M.D., Parathasarathy, T.A., 2005. Size-affected single-slip behavior of pure nickel

microcrystals. Acta Mater. 53 (15), 4065–4077.

Espinosa, H.D., Prorok, B.C., Fischer, M., 2003. A methodology for determining mechanical properties of

freestanding thin films and MEMS materials. J. Mech. Phys. Solids 51, 47–67.

Espinosa, H.D., Prorok, B.C., Peng, B., 2004. Plasticity size effects in free-standing submicron polycrystalline

FCC films subjected to pure tension. J. Mech. Phys. Solids 52, 667–689.

Fleck, N.A., Hutchinson, J.W., 2001. A reformulation of strain gradient plasticity. J. Mech. Phys. Solids 49,

2245–2271.

Fleck, N.A., Muller, G.M., Ashby, M.F., Hutchinson, J.W., 1994. Strain gradient plasticity: theory and

experiment. Acta Metall. Mater. 42, 475–487.

Florando, J.N., Nix, W.D., 2005. A microbeam bending method for studying stress–strain relations for metal thin

films on silicon substrates. J. Mech. Phys. Solids 53, 619–638.

ARTICLE IN PRESSL. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–2110 2109

Gao, H., Huang, Y., Nix, W.D., Hutchinson, J.W., 1999. Mechanism-based strain gradient plasticity. J. Mech.

Phys. Solids 47, 1239–1263.

Ghoniem, N.M., Han, X., 2005. Dislocation motion in anisotropic multilayer materials. Philos. Mag. 85,

2809–2830.

Greer, J.R., Oliver, W.C., Nix, W.D., 2005. Size dependence of mechanical properties of gold at the micron scale

in the absence of strain gradients. Acta Mater. 53, 1821–1830, erratum Acta Mater. 54, 1705.

Groh, S., Devincre, B., Kubin, L.P., Roos, A., Feyel, F., Chaboche, J.-L., 2003. Dislocations and elastic

anisotropy in heteroepitaxial metallic thin films. Philos. Mag. Lett. 83, 303–313.

Gurtin, M.E., 2002. A gradient theory of single-crystal viscoplasticity that accounts for geometrically necessary

dislocations. J. Mech. Phys. Solids 50, 5–32.

Haque, M.A., Saif, M.T.A., 2003. Strain gradient effect in nanoscale thin films. Acta Mater. 51, 3053–3061.

Hartmaier, A., Buehler, M.J., Gao, H., 2005. Two-dimensional discrete dislocation models of deformation in

polycrystalline thin metal films on substrates. Mater. Sci. Eng. A 400, 260–263.

Hommel, M., Kraft, O., 2001. Deformation behavior of thin copper films on deformable substrates. Acta Mater.

49, 3935–3947.

Hutchinson, J.W., 2001. Strain gradient plasticity theory revisited. Materials Science for the 21st Century, vol. A.

Society of Materials Science, Japan, pp. 307–315.

Kubin, L.P., Canova, G., Condat, M., Devincre, B., Pontikis, V., Brechet, Y., 1992. Dislocation microstructures

and plastic flow: a 3D simulation. Solid State Phenom. 23, 455–472.

Lou, J., Shrotriya, P., Buchheit, P., Yang, D., Soboyejo, W.O., 2003. An investigation of the effects of thickness

on mechanical properties of LIGA nickel MEMS structures. J. Mater. Sci. 38, 4137–4143.

Ma, Q., Clark, D.R., 1995. Size dependent hardness of silver single crystals. J. Mater. Res. 10,

853–863.

Nicola, L., Van der Giessen, E., Needleman, A., 2003. Discrete dislocation analysis of size effects in thin films.

J. Appl. Phys. 93, 5920–5928.

Nicola, L., Van der Giessen, E., Needleman, A., 2005a. Size effects in polycrystalline thin films analyzed by

discrete dislocation plasticity. Thin Solid Films 479, 329–338.

Nicola, L., Van der Giessen, E., Gurtin, M.E., 2005b. Effect of defect energy on strain-gradient predictions of

confined single-crystal plasticity. J. Mech. Phys. Solids 53, 1280–1294.

Nicola, L., Van der Giessen, E., Needleman, A., 2005c. Two hardening mechanisms in single crystal thin films

studied by discrete dislocation plasticity. Philos. Mag. A 85, 1507–1518.

Pant, P., Schwarz, K.W., Baker, S.P., 2003. Dislocation interactions in thin FCC metal films. Acta Mater. 51,

3243–3258.

Phillips, M.A., Spolenak, R., Tamura, N., Brown, W.L., MacDowell, A.A., Celestre, R.S., Padmore, H.A.,

Batterman, B.W., Arzt, E., Patel, J.R., 2004. X-ray microdiffraction: local stress distributions in

polycrystalline and epitaxial thin films. Microelectron. Eng. 75, 117–126.

Shen, Y.-L., Suresh, S., He, M.Y., Bagchi, A., Kienzle, O., Ruhle, M., Evans, A.G., 1998. Stress evolution in

passivated thin films of Cu on silica substrates. J. Mater. Res. 13, 1928–1937.

Shu, J.Y., Fleck, N.A., Van der Giessen, E., Needleman, A., 2001. Boundary layers in constrained

plastic flow: comparison of nonlocal and discrete dislocation plasticity. J. Mech. Phys. Solids 49,

1361–1395.

Stolken, J.S., Evans, A.G., 1998. A microbend test method for measuring the plasticity length scale. Acta Mater.

46, 5109–5115.

Uchic, M.D., Dimiduk, D.M., Florando, J.N., Nix, W.D., 2004. Sample dimensions influence strength and crystal

plasticity. Science 305 (5686), 986–989.

Van der Giessen, E., Needleman, A., 1995. Discrete dislocation plasticity: a simple planar model. Modelling

Simul. Mater. Sci. Eng. 3, 689–735.

Vlassak, J.J., Nix, W.D., 1992. A new bulge test technique for the determination of Young’s modulus and

Poisson’s ratio of thin films. J. Mater. Res. 7, 3242–3249.

Von Blanckenhagen, B., Arzt, E., Gumbsch, P., 2004. Discrete dislocation simulation of plastic deformation in

metal thin films. Acta Mater. 52, 773–784.

Xiang, Y., Vlassak, J.J., 2005a. Bauschinger effect in thin metal films. Scr. Mater. 53, 177–182.

Xiang, Y., Vlassak, J.J., 2005b. Length-scale and Bauschinger effects in freestanding thin films, submitted for

publication.

Xiang, Y., Tsui, T.Y., Vlassak, J.J., 2006. The mechanical properties of freestanding electroplated Cu thin films.

J. Mater. Res. 21 (6), in press.

ARTICLE IN PRESSL. Nicola et al. / J. Mech. Phys. Solids 54 (2006) 2089–21102110

Xiang, Y., Chen, X., Vlassak, J.J., 2002. The mechanical properties of electroplated Cu thin films measured by

means of the bulge test technique. Mater. Res. Soc. Symp. Proc. 695, 189–194.

Xiang, Y., Vlassak, J.J., Perez-Prado, M.T., Tsui, T.Y., McKerrow, A.J., 2004. The effects of passivation layer

and film thickness on the mechanical behavior of freestanding electroplated Cu thin films with constant

microstructure. Mater. Res. Soc. Symp. Proc. 795, 417–422.

Xiang, Y., Chen, X., Vlassak, J.J., 2005. The plane-strain bulge test for thin films. J. Mater. Res. 20, 2360–2370.