an interactive micro-void shear localization mechanism in

20
Journal of the Mechanics and Physics of Solids 55 (2007) 225–244 An interactive micro-void shear localization mechanism in high strength steels Cahal McVeigh a , Franck Vernerey a , Wing Kam Liu a, , Brian Moran a , Gregory Olson b a Department of Mechanical Engineering, Northwestern University, 2145 Sheridan Road, Evanston, IL 60208-3111, USA b Department of Materials Science, Northwestern University, 2145 Sheridan Road, Evanston, IL 60208-3111, USA Received 6 April 2006; received in revised form 1 August 2006; accepted 8 August 2006 Abstract Ductility of high strength steels is often restricted by the onset of a void-sheet mechanism in which failure occurs by a micro-void shear localization process. For the first time, the micro-void shear instability mechanism is identified here by examining the interactions occurring within a system of multiple embedded secondary particles (carbides 10–100 nm), through a finite element based computational cell modeling technique (in two and three dimensions). Shear deformation leads to the nucleation of micro-voids as the secondary particles debond from the surrounding alloy matrix. The nucleated micro-voids grow into elongated void tails along the principal shear plane and coalesce with the micro-voids nucleated at neighboring particles. At higher strains, the neighboring particles are driven towards each other, further escalating the severity of the shear coalescence effect. This shear driven nucleation, growth and coalescence mechanism leads to a decrease in the load-bearing surface in the shear plane and a terminal shear instability occurs. The mechanism is incorporated mathematically into a hierarchical steel model. The simulated response corresponds to experimen- tally observed behavior only when the micro-void shear localization mechanism is considered. r 2006 Elsevier Ltd. All rights reserved. Keywords: Ductility; Microstructures; Voids and inclusions; Constitutive behaviour; Finite elements 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.2006.08.002 Corresponding author. Tel.: +847 4676502; fax: 847 4913915. E-mail address: [email protected] (W.K. Liu).

Upload: others

Post on 30-Dec-2021

11 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESS

Journal of the Mechanics and Physics of Solids

55 (2007) 225–244

0022-5096/$ -

doi:10.1016/j

�CorrespoE-mail ad

www.elsevier.com/locate/jmps

An interactive micro-void shear localizationmechanism in high strength steels

Cahal McVeigha, Franck Vernereya, Wing Kam Liua,�,Brian Morana, Gregory Olsonb

aDepartment of Mechanical Engineering, Northwestern University, 2145 Sheridan Road, Evanston,

IL 60208-3111, USAbDepartment of Materials Science, Northwestern University, 2145 Sheridan Road, Evanston, IL 60208-3111, USA

Received 6 April 2006; received in revised form 1 August 2006; accepted 8 August 2006

Abstract

Ductility of high strength steels is often restricted by the onset of a void-sheet mechanism in which

failure occurs by a micro-void shear localization process. For the first time, the micro-void shear

instability mechanism is identified here by examining the interactions occurring within a system of

multiple embedded secondary particles (carbides �10–100 nm), through a finite element based

computational cell modeling technique (in two and three dimensions). Shear deformation leads to the

nucleation of micro-voids as the secondary particles debond from the surrounding alloy matrix. The

nucleated micro-voids grow into elongated void tails along the principal shear plane and coalesce

with the micro-voids nucleated at neighboring particles. At higher strains, the neighboring particles

are driven towards each other, further escalating the severity of the shear coalescence effect. This

shear driven nucleation, growth and coalescence mechanism leads to a decrease in the load-bearing

surface in the shear plane and a terminal shear instability occurs. The mechanism is incorporated

mathematically into a hierarchical steel model. The simulated response corresponds to experimen-

tally observed behavior only when the micro-void shear localization mechanism is considered.

r 2006 Elsevier Ltd. All rights reserved.

Keywords: Ductility; Microstructures; Voids and inclusions; Constitutive behaviour; Finite elements

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

.jmps.2006.08.002

nding author. Tel.: +847 4676502; fax: 847 4913915.

dress: [email protected] (W.K. Liu).

Page 2: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244226

1. Introduction

Damage nucleation in an alloy usually involves some degree of decohesion of anembedded particle from the surrounding matrix or particle fracture (Sarkar et al., 2001;Dierickx et al., 1996). The nucleated damage subsequently grows in a manner consistentwith the stress state and the material degrades rapidly as individual damage sites begin tointeract.In the air melted high strength steel examined here, the embedded particles are generally

primary particles (titanium nitrides on the order of a micron) and secondary particles(titanium carbides and manganese sulfides on the order of 10–100 nm). It is commonto refer to the voids which nucleate at the secondary particle sites as micro-voids, todistinguish them from the larger voids which nucleate from the primary particles. Hencethere are two potential populations of voids, which exist at two distinct scales.Two ductile failure mechanisms have been identified in high strength steels:

(a)

A void coalescence process in which relatively equiaxed voids grow to impingement(associated with high triaxiality loading).

(b)

A void-sheet mechanism in which failure occurs by a micro-void shear localizationprocess (associated with low triaxiality loading).

Mechanism (b) is much more prevalent under shear loading. However, even under hightriaxiality loading conditions (mode-1 type loading), the micro-void-sheet mechanism caninitiate and propagate along the plane of maximum shear and severely limit ductility inhigh strength steels (Goto et al., 1999).Most of the research previously performed in the area of ductile failure has concentrated

on the behavior of primary- and secondary-particle nucleated voids under high triaxialityloading as described by mechanism (a) above. Micromechanical nucleation studies havebeen performed (Horstemeyer et al., 2003; Horstemeyer and Gokhale, 1999; Gall et al.,1999) and several useful nucleation criteria have been posed for steel alloys (Beremin,1981). Experimental studies of void growth and ductility have tended to focus on tensiletests, in which the voids are free to grow in the direction of the applied load (Manoharanand Lewandowski, 1989; Humphreys, 1989; Poza and Llorca, 1996; Verhaege et al., 1997).Factors effecting the growth of voids have also been investigated; Rice and Tracey (1969)were the first to predict the changing size of spherical voids in a deforming body. Lee andMear (1992) extended the model to account for the growth of axisymmetric ellipsoidalvoids. Hom and McMeeking (1989) investigated the growth of voids directly ahead of acrack tip and developed several void coalescence criteria which were used to predictfracture initiation. The effect of void configuration on growth and coalescence has beenmodeled by Horstemeyer et al. (2000) who discovered that coalescence effects can occurbetween voids as far apart as six void diameters.The experimental investigations and computational studies outlined above have focused

on mechanism (a). The resulting constitutive models therefore fail to predict the onset ofthe micro-void-sheet mechanism (b) which may occur over a range of loading conditionsand which is solely responsible for failure under shear loading. It is this mechanism, and inparticular its occurrence under pure shear-loading conditions, which the current paper isfocused on.

Page 3: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244 227

Some pioneering work has been performed to find the underlying mechanismresponsible for the micro-void shear localization mechanism. Low triaxiality ductilefailure was studied by Cowie et al. (1989) by investigating the shear failure of a steel samplesubjected to a ballistic impact (Fig. 1a). Experiments were performed to determine theimportance of triaxiality on the observed failure behavior (Fig. 1b). Although rate effectsplay an important role in determining the material strength, a similar shear instabilitystrain is observed under quasi-static testing conditions. Hence, the observed softeningcannot be explained in terms of (dynamic) adiabatic shear bands. An underlying failuremechanism is thought to be responsible for shear-induced ductile failure, regardless ofdeformation rate. Micrographs have shown that the fracture surface resulting from lowtriaxiality ductile failure in high strength steels shows evidence of elongated micro-voids inthe shear plane (Fig. 1c). It was also observed that the shear instability could be delayed byincreasing the interfacial strength of the secondary particle–matrix interface or by applyinga compressive hydrostatic stress; both of which delay debonding of the secondaryparticle–matrix interface and hence micro-void nucleation. Hayzelden (1987) showed thatby decreasing the number of secondary particles, the shear instability strain could beincreased and the application of a compressive stress during shear loading had less abilityto increase the instability strain. These observations strongly suggest that the shear

Fig. 1. (a) Shear band formation in a specimen that has undergone a ballistic impact. (b) The experimental set up

used to examine the triaxiality effect on material failure. (c) A micro-void which has nucleated at a pair of

secondary particles—typical of a fracture surface in shear failure (Cowie et al., 1989).

Page 4: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244228

instability originates at the secondary particle-nucleated micro-voids and is due to theirelongation and coalescence in the shear plane.Tracey and Perrone (1987) have simulated the pair-interaction between neighboring

micro-voids under various shear-loading conditions, and examined the strain localizationin the ligament between the particles. It was found that an interaction effect occurred atmicro-void spacings of three void diameters. Hutchinson and Tvergaard (1987) showedthat the presence of a particle within a micro-void is a crucial factor in the subsequentevolution of the micro-void, particularly under low and negative triaxiality loading.Tvergaard (1989) demonstrated that the presence of a secondary particle acts to prop openthe micro-void, facilitating further void growth even in shear. Tvergaard (1992)computationally examined the ability of a non-local porous ductile damage model torepresent the average behavior of an inclined row of micro-voids, leading to void-sheetfailure (1997) between the larger primary particles. Bandstra et al. (1998) and Bandstraand Koss (2001) have also used finite element micro-mechanical modeling of a HY-100steel microstructure to show that the strain localization arising between larger voids issufficient to drive nucleation of micro-voids from secondary particles in a void-sheetmechanism.Previous research has tended to investigate micro-void sheets in terms of growth

and coalescence in the plane of maximum shear during tensile loading. In that casethe dominant stress state within the void sheet is one of shear; however a significantpositive triaxiality still exists locally which explains the ease with which micro-voidsgrow and coalesce (Fig. 3a). This fails to explain how a micro-void sheet can causefailure in an alloy subjected to zero triaxiality loading conditions, as was observedexperimentally by Cowie et al. (1989) (Fig. 1), in which case the local triaxialityremains extremely low. In the current paper we simulate the experimental conditions inFig. 1b with the goal of identifying the micro-void-sheet mechanism under zero triaxiality(Fig. 3b) and the resulting shear instability i.e. a micro-void shear localization process.This is achieved by directly modeling the nucleation, growth and coalescence of interactingmicro-voids under shear-loading conditions through a computational cell modelingtechnique.In previous papers by Hao et al. (2003, 2004) a similar high strength steel was

investigated. However the authors were unable to simulate a terminal shear localizationprocess arising at the scale of the secondary particles, perhaps because the system ofparticles being examined was constructed using average particle spacing data. In thepresent work, local clusters of particles are assembled such that they successfully initiate aterminal shear localization mechanism. To the authors’ knowledge, this is the first time themicro-void shear localization softening mechanism has been identified and modeled at thescale of the secondary particles under pure shear-loading conditions. As an illustration ofthe importance of the mechanism, the cell modeling observations are incorporatedmathematically into a hierarchical steel model to predict the onset of material softeningunder pure shear-loading conditions. The simulated steel fails in a manner consistent withexperimental findings, without resorting to empiricism or introducing artificial softening inthe alloy matrix.To begin, the role of interfacial strength at the particle–matrix boundary is examined in

Section 2. The post-debonding behavior is then investigated in Section 3 in order toidentify the micro-void shear localization mechanism. The mechanism is incorporated intoa hierarchical model of high strength steel in Section 4.

Page 5: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244 229

2. Role of interfacial strength

2.1. Decohesion relation

Cowie et al. (1989) determined that nucleation of micro-voids in a standard 4340 steeloccurred by debonding at the secondary particle–matrix interface. The size and propertiesof the chromium (Cr)-based secondary particles in the standard 4340 steel are similar tothose examined in the modified 4340 steel used in the present study and the same nucleationmechanism is expected. The modification used to introduce a titanium (Ti) component tothe 4340 steel is described by Shabrov et al. (2004). Micro-void nucleation occurs at bothtitanium carbide (TiC) and fine manganese sulfide (MnS) secondary particles in themodified 4340 steel. However, only the titanium carbide (TiC) particles are modeled here.The bonding behavior at a titanium carbide–iron (TiC–Fe) interface has been examined byLee et al. (2005) through a first principals approach; an atomic scale potential wassubsequently derived. Hao et al. (2003, 2004) used this potential to derive a microscale

interfacial potential, which accounted for the adverse effects of defects which inevitablyexist over a microscale interface. This scaling-up procedure allows for the development of adecohesion relation suitable for modeling the titanium carbide–iron interface in a highstrength steel. The interfacial behavior used in this study is shown schematically in Fig. 2

Δ u

Deformed Configurationparticle/matrix interfaceInitial Configuration of

t2= t2 (Δ u) = −t1

t1= t1 (Δ u)

Γ1

Γ1Γ1 = Γ2

Nor

mal

Tra

ctio

n x1

0e9

Pa

Normal Separation x10e-8 m

1.5

1

0.5

01.510.50

(a)

(b)

Fig. 2. (a) Interfacial debonding is modeled through a traction–separation cohesive relation implemented at the

particle–matrix interface. (b) At any position on the separating surfaces, the traction is computed as a function of

the normal separation between the two surfaces during deformation.

Page 6: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244230

and is implemented in the form of a traction–separation relation during finite elementanalysis.Although primary particle fracture is known to contribute to primary void formation,

secondary particle fracture is not considered as a micro-void nucleation mechanism here.Also, it is assumed that any initial internal stresses present in the particle due to latticemisfit strain will relax rapidly once the surrounding alloy matrix undergoes plasticdeformation. Given the large strain nature of the phenomenon being modeled, such initialstresses are neglected in the current model.

2.2. Finite element implementation

The role of interfacial strength is examined through a computational cell modelingapproach in which a single particle is modeled in a periodically repeating unit cell.It is assumed that a continuum-based approach can capture the mechanics of thedecohesion process; the only physical length scale in the simulation is introduced throughthe decohesion relation at the interface. Implementation of the interfacial behavior isachieved as follows. The expression for the virtual internal power is given in terms of thepower arising within the body and the internal power due to the traction on the interfacialsurfaces:

dPint ¼

ZOsijdDij dOþ

ZG1

t1i dvi dGþZG2

t2i dvi dG, (1)

where sij is the Cauchy stress, t1i is the traction on the internal surface G1, t2i is the tractionon the internal surface G2, dvi is the virtual velocity and Dij is the symmetric part of thevelocity gradient. G1 and G2 represent the surfaces found at the particle–matrix interface.Initially, these surfaces are coincident and the tractions are zero. As deformation occurs,the traction at a position t on the internal surface is computed in terms of the subsequentnormal separation between two initially coincident points as shown in Fig. 2b. The equaland opposite nature of the traction acting at points which were originally coincidentensures that the continuity condition (Belytschko et al., 2000),

njsji

� �1� njsji

� �2

� �¼ 0 (2)

remains valid across the interface and momentum balance is satisfied in the body. nj

represents the unit normal to the internal surface.The system is discretized and solved quasi-statically using commercially available finite

element analysis software (ABAQUS Explicitr). Elastic deformation of both the particlesand the alloy matrix is described by a hypoelastic response. J2 mises plasticity is usedto describe the plastic deformation of the alloy matrix, within a large deformationformulation. A constrained shear stress periodic boundary condition is applied such thatthe top and bottom horizontal surfaces remain parallel to the x-axis, while being free tomove in the y-direction. The left and right vertical surfaces are periodic with each other.These simple boundary conditions are sufficient for simulating the behavior of an infiniteseries of particles in a shear plane, with extremely low local stress triaxiality.The average shear stress and strain response of the cell is calculated as a volume

average; the divergence theorem is used to convert the volume average to a moremanageable integration over the cell surface in terms of the measurable surface tractionsand displacements. A ramped velocity is applied uniformly to the top and bottom surfaces,

Page 7: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244 231

in the x-direction. In the three-dimensional (3D) analysis, the front and back surfaces arealso constrained to deform in a periodic manner (Fig. 5).

The material properties used during cell modeling are illustrated in Fig. 3c. The alloymatrix material is modeled by experimentally obtained tensile data for a modified 4340high-strength steel (Hao et al., 2003). The secondary particles are considered to be perfectlyelastic, with an elastic modulus consistent with titanium carbide (TiC).

As a first approximation, a two-dimensional (2D) plane strain cell model is created. Inreality, this equates to a cylindrical particle embedded in a unit cell. A 3D model is thenexamined. For comparison, in the 3D example, the particle volume fraction is first set equalto the 2D particle area fraction.

2.3. Decohesion behavior: results and discussion

2.3.1. 2D analysis of interfacial decohesion

Fig. 4 illustrates the plastic strain contours in the sheared unit cell during deformation,after decohesion. A particle area fraction of 4% is used to magnify the decohesion effect.The effect of the interfacial debonding energy is shown in Fig. 4. The presence of a bondedparticle results in a stronger average response, compared to the case where no bondingexists. However, even in the ‘no bonding’ case, the average behavior of the cell is that oflinear hardening.

As expected, a fully bonded hard particle acts to resist deformation in the shear plane itlies in. Once initiated, debonding rapidly propagates along the interface, in a mannersimilar to a fracture process. Physically speaking, a micro-void has been nucleated at theinterface. Even in pure shear strain loading conditions, some limited growth of this micro-void can be seen in Fig. 4 due to elastic deformation of the matrix. The central planerapidly switches from being the strongest plane, to a weak plane containing a void propped

0.0

2.0

4.0

6.0

8.0

1.0

1.2

1.4

1.6

0 0.2 0.4 0.6 0.8

Shear Strain

Tru

e S

hea

r S

tres

s (G

Pa)

200

600

600

H(matrix)

E(TiN)

E(TiC)

0.56

1.06

E(matrix) GPa

GPa

GPa

GPa

Y0(matrix) GPa

=====

E

H

0Y

22σ

11σ

22σ sτ

(a)

(b) (c)

Fig. 3. Experimental data showing elasto-plastic response of alloy matrix material. The particles (TiC and TiN)

are considered to behave elastically.

Page 8: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESS

Fig. 4. The 2D plane strain periodic unit cell (plastic strain contour) and the corresponding average cell

constitutive behavior with different interfacial strengths. The boundary conditions lead to a constrained shear

stress state—the top surface remains flat but can translate up and down.

C. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244232

open by a particle. Elastically stored strain energy in the surrounding areas immediatelyunloads into the weak plane and a shear band rapidly propagates between particles. This isaccompanied by a drop in the average flow stress; the magnitude of the drop is related tothe interfacial strength.Although higher debonding strengths delay the onset of instability, the resulting drop in

stress also increases. Regardless of interfacial strength, once the debonding instabilityoccurs the cell ‘recovers’ it’s load carrying ability and resumes linear hardening.

2.3.2. 3D analysis of interfacial decohesion

Fig. 5 illustrates the plastic strain contours in the sheared unit cell during deformation,after decohesion has occurred. The volume fraction is equal to 4%. The effect of theinterfacial debonding energy is shown in Fig. 6. The 3D decohesion and instabilitybehavior is similar to the 2D case; the same explanation is valid.From Fig. 6 it is again clear that the interfacial strength plays an important role in

determining the transient decohesion instability behavior. However, the average behaviorreturns to a stable linear hardening response thereafter. Hence, it is concluded that theterminal micro-void shear localization mechanism cannot be captured by examiningthe decohesion of a single periodically arranged secondary particle as in Fig. 5. Althoughthe model captures the nucleation process which triggers the shear localization mechanism,it fails to capture the key interactions between secondary particles which lead to a terminalshear instability. With this in mind, the post decohesion behavior is now examined for asystem of multiple secondary particles.

Page 9: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESS

dcba

e

0.0

2.0

4.0

6.0

8.0

1.0

1.2

1.4

1.6

0 0.2 0.4 0.6 0.8

Shear Strain

Tru

e S

hea

r S

tres

s (G

Pa)

Interface Normal Peak Stress (Pa)

ba 0

1.0E91.5E92.0E92.5E9

cde

Fig. 6. The average cell constitutive behavior with different interfacial strengths.

Fig. 5. Accumulated plastic strain arising in a unit cell. Limited void growth is visible.

C. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244 233

3. Identification of a multi-particle shear instability mechanism

3.1. 2D cluster models

The periodically repeating particle model in the previous section offers limited insightinto the physics of a real system of embedded particles under shear-loading conditions.

Page 10: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244234

To construct a more realistic model of the possible interactions between particles duringshear, a 2D plane strain analysis of a random particle cluster is undertaken. The particlevolume fraction is 2% and particle radii are in the range 30–50 nm. The alloy matrix isagain modeled using the experimental linear hardening data. Approximately 40 elementswere used along the periphery of each particle to avoid mesh distortion effects at largedeformations. Fig. 7 shows such an arrangement and the resulting average constitutiveresponse of the cell.The plane of shear localization can be observed in Fig. 7. In this plane, highlighted in

Fig. 8, the top of one particle is aligned with the bottom of another in the shear plane.Upon loading, the embedded particles debond from the surrounding matrix andthe resulting micro-voids initially grow at approximately 451 to the shear plane. As theparticles are driven towards each other, the micro-voids begin to turn toward and growalong the shear plane. The tail-like elongated voids propagate towards each other, in a

Fig. 7. Shear loading of a 2D unit cell containing a particle cluster. The shear failure plane occurs between two

optimally placed particles. The average response is shown.

Page 11: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESS

Fig. 8. A periodic arrangement of the top section of the unit cell showing the interaction between particles.

b

aCoalescence

Growth

Fig. 9. The micro-void shear localization mechanism.

C. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244 235

similar manner to the propagation of a shear crack i.e. the void surface area increases butthe void volume does not increase appreciably. A contact algorithm was used to ensurethat opposite faces of the micro-voids did not penetrate each other during growth alongthe shear plane. Eventually, the ligament between the voids neck i.e. coalescence occurs. Asdeformation increases the particles themselves are driven towards each other, increasingthe severity of the coalescence.

This shear controlled nucleation, growth and coalescence process explains the instabilityobserved in the average constitutive response (Fig. 7). The resulting reduction in loadcarrying cross section on the shear plane is similar to the necking observed in a uniaxialtensile specimen. With reference to Fig. 9, the average true shear stress in a cell is given by

S ¼ t=a, (3)

where t is the applied load on the surface whose area is given by a. In the shear plane, theaverage true tensile stress in the necking region is similarly given by

s ¼ t=b, (4)

where b is the load carrying surface area in the shear plane. As the force must balance inevery plane, the relationship between the average shear stress and the surface area is thensimply

S � sb=a. (5)

As the load-bearing surface area b decreases during shear driven coalescence, the averageshear stress S sustained by the cell diminishes. Eqs. (3–5) illustrate the softening responseobserved during shear loading, in terms of the reduction in load-bearing surface areacaused by elongated ‘void tail’ growth and shear driven coalescence. This interaction isnow examined briefly in three dimensions.

Page 12: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244236

3.2. 3D cluster models

As a first attempt to reproduce the same mechanism in three dimensions, a narrow cell ismodeled as shown in Fig. 10. This is similar to the 2D case, however some material doesexist on either side of the particle in the 3D case. A regular unit cell containing a cluster ofparticles concentrated in the shear plane was also modeled (Fig. 11). The particle density isrepresentative of a heterogeneous cluster of particles i.e. a region which is more susceptibleto shear failure. It is not representative of the average particle fraction in a typical steelmicrostructure. In both cases the material at either side of the particles stabilizes the cell,delaying and reducing the effect of the shear localization compared to the 2D simulations.However the shear instability is still terminal and the underlying shear driven growth andcoalescence mechanism is the same in 2D and 3D analyses. The 3D analysis is useful forverifying the 2D results, particularly as the goal of the current project is to model real 3Dmicrostructures for the purposes of optimized materials design.In order to gauge the importance of relative particle position, a study is performed in

which the initial positions of two particles are translated in the shear direction andtransverse direction. Consider a cell sheared in the x–z plane. Both particles lie on the x–y

plane and are translated in the x direction and y direction, respectively, as shown in Fig. 12along with the corresponding constitutive behavior. The curves indicate that the optimumparticle placement which maximizes the micro-void localization effect occurs when theparticles lie within one diameter of each other parallel to the shear plane and slightly oneither side of the shear plane in the y direction. This is consistent with the work of

Fig. 10. Cross section of a deforming narrow 3D cell. Void tails grow into the shear plane.

Page 13: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESS

Fig. 11. A (translucent) 3D cell containing a cluster of particles in the shear plane. Void growth in the shear plane

is evident when viewed from above. The dark shadows represent the free surfaces of the tail-like micro-voids.

0.0E+00

2.0E+08

4.0E+08

6.0E+08

8.0E+08

1.0E+09

1.2E+09

0 0.5 1 1.5 2 2.5

Shear Strain

0 0.5 1 1.5 2 2.5

Shear Strain

Tru

e S

hea

r S

tres

s (P

a)

0.0E+00

2.0E+08

4.0E+08

6.0E+08

8.0E+08

1.0E+09

1.2E+09

1.4E+09

1.6E+09

Tru

e S

hea

r S

tres

s (P

a)

x

y

z

Fig. 12. The effect of the relative position of two interacting particles and the shear instability mechanism

observed.

C. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244 237

Horstemeyer et al. (2000) and Tracey and Perrone (1987) who found void pair interactionsarising at spacings of six and three diameters, respectively. The effect of particle spacingsuggests that clustering plays an important role in micro-void localization in high strengthsteels.

Page 14: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244238

3.3. Constitutive relation for micro-void shear localization

The softening effect described by Eqs. (3–5) was explained in terms of shear drivenmicro-void tail growth and coalescence in the shear plane. In particular, the coalescenceindicates the onset of a terminal shear instability. This coalescence mechanism is nowincorporated mathematically into a Gurson-type damage model (Gurson, 1977; Tvergaardand Needleman, 1984) and calibrated through a computational cell modeling technique.The resulting material model represents the average response of an alloy matrix containinga population of TiC secondary particles over a range of triaxiality loading conditions.The Gurson potential is chosen to represent the damage-dependent yield surface:

F ¼seqsy

� �2

þ 2q1f cosh q2

smsy

� �� 1þ q3f

2� �

¼ 0, (6)

where f represents the effect of porosity and acts to contract the yield surface, seq is theequivalent or Mises stress, sy is the flow stress in the matrix, sm is the hydrostatic stressand q1, q2 and q3 are material constants.In ductile fracture models, this porosity parameter is often calculated as a function of

nucleation and growth terms as

_f ¼ _f nuc þ_f gro, (7)

where

_f gro ¼ 1� fð Þtrace Dpð Þ (8)

and Dp is the plastic part of the rate of deformation. _f nuc and_f gro are the porosity rate due

to nucleation of voids and growth of existing voids. Void coalescence effects are generallyintroduced by adding a third term, _f coa, once a critical value of porosity, fc, has beenreached. This term describes a rapid increase in porosity as voids impinge upon each other:

_f ¼ _f nuc þ_f gro þ

_f coa ¼_f nuc þ

_f gro þ C1_f nuc þ

_f gro

� , (9)

where

C1 ¼0 when f � f co0;

3 when f � f cX0:

((10)

Eqs. (6)–(10) are valid only for high triaxiality loading i.e. they describe volumetric growthand impingement of micro-voids. On the other hand, it is well known that high strength steelsamples loaded in pure shear exhibit a plastic instability due to micro-void coalescence,through the mechanism identified in the previous section. Although developed in the contextof high triaxiality, the Gurson potential of Eq. (6) remains a useful tool to incorporate a lowtriaxiality coalescence effect in a homogenized material model through the porosityparameter, f. In the previous sections, shear driven coalescence was observed to be a functionof the shear deformation. The simplest way to incorporate the effect of shear driven micro-void coalescence is to modify the existing coalescence term (Eq. (9)) to include a dependenceon the accumulated plastic strain, �̄. The total rate of porosity becomes

_f ¼ _f nuc þ_f gro þ

_f coa ¼_f nuc þ

_f gro þ C1_f nuc þ

_f gro

� þ C2

_̄�. (11)

Page 15: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244 239

The term C2_̄� does not represent a physical value of porosity but ensures the detrimental

effect of low triaxiality coalescence exists in the potential F. A criterion similar to that inEq. (10) is required to determine when coalescence occurs. Two extremes cases exist:

(a)

In zero triaxiality loading, porosity remains low and coalescence occurs at some criticalvalue of plastic strain �̄c.

(b)

In high triaxiality loading, porosity increases rapidly and coalescence occurs at somecritical porosity fc.

In the range of triaxialities between these two extreme cases, some combination of bothmechanisms occurs. Hence it is necessary to formulate a coalescence criterion in terms ofthe porosity and plastic strain at the onset of coalescence, fc and �̄c, respectively. A cellmodeling technique is used to quantitatively examine this relationship. Two interactingsecondary particles are examined in a cell subjected to various average triaxiality loadingconditions. Following the work of Kouznetsova et al. (2002) and Arsenlis and Parks(1999), periodic boundary conditions are applied to achieve triaxialities ranging from pureshear stress to a significant degree of tensile hydrostatic stress and shear. A particle fractionof 1% and spacing of five particle diameters is chosen (particle diameter ¼ 30 nm). FromSection 3.2, this is a favorable arrangement for the onset of shear driven localization and istypical of a cluster of secondary particles in high strength steel.

For each loading condition, the critical porosity fc and critical plastic strain �̄c arerecorded at the instability point in the curve of shear stress v accumulated plastic strain. �̄cand fc are plotted in Fig. 13. In high triaxiality (cell a in Fig. 13) coalescence occurs at aporosity of approximately 0.12; a value often used in standard damage laws to indicate theonset of coalescence. In pure shear (cell d in Fig. 13) coalescence occurs at a high value ofplastic strain of approximately 1.

The plotted data points are fitted to a simple linear trend line (Fig. 13),

f c ¼ A�̄c þ B, (12)

where

A ¼ �0:1222 and B ¼ 0:1236:

The condition for the onset of coalescence under a range of stress states is therefore

f þ 0:1222�̄� 0:123640. (13)

The final expression for C1 and C2 in Eq. (11) is then given by

C1 ¼ 0 and C2 ¼ 0 when f þ 0:1222�̄� 0:1236o0,

C1 ¼ 3 and C2 ¼ 1 when f þ 0:1222�̄� 0:1236X0. ð14Þ

In other words, _f coa becomes active only when coalescence occurs (Eq. (13) is satisfied). Thecoalescence term C2 in Eq. (11) can be determined from the slope of the stress–strain curve afterthe instability has occurred. It is noted that the resulting material constants are valid for themicrostructure simulated in Fig. 13 only. In reality, the constants A, B, C1 and C2 are related tothe size and spacing of the secondary particles. This relationship has not been examined here.

The resulting homogenized damage law is illustrated in Fig. 14 under different loadingconditions. The damage law captures softening over a range of loading conditions i.e. itworks in pure shear but covers all stress states of interest to tensile fracture. At high

Page 16: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESS

Coalescence

No Coalescence

Cell Model Computations

Linear Model

ab

c

d

Increasing Triaxiality

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0 0.2 0.4 0.6 0.8 1 1.2

Critical Plastic Strain

Crit

ical

Por

osity

0 0.2 0.4 0.6 0.8 10

2

4

6

8

10

12

14

16x 108

Average Plastic Strain

Ave

rage

She

ar S

tres

s (P

a)

a b

c

d

(a) (b) (c) (d)

Fig. 13. Cell models are subjected to various triaxiality loading conditions. The plastic strain contours at

instability (the onset of coalescence) are shown for four cases (a–d). Critical plastic strain and critical void volume

fraction at instability (onset of coalescence) are recorded–data points corresponding to the contour plots (a)–(d)

are highlighted in the graph.

0.00E+00

5.00E+07

1.00E+08

1.50E+08

2.00E+08

2.50E+08

3.00E+08

3.50E+08

4.00E+08

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

Shear Strain

Tru

e S

hea

r S

tres

s (P

a)

Increasing

Triaxiality

Pure Shear

hf

hf

sf

h

s

fT

f=

T = 0.0

T = 0.2

T = 0.5

T = 0.9T = 0.7

T = 1

Fig. 14. The modified damage law can capture the effects of shear driven coalescence. As expected, at high

triaxiality, hydrostatic void growth and coalescence dominates. As triaxiality decreases, shear driven coalescence

begins to dominate.

C. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244240

Page 17: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244 241

triaxialities, volumetric void growth and coalescence dominate. Under low triaxiality andpure shear loading, the porosity remains low and the micro-void shear driven coalescenceidentified in the previous section dominates.

It should be noted that the idea of using a periodically repeating unit cell to capturestrain localization effects is not unique to this work. However, in reality one suchlocalization band will ultimately ‘win’ and strain localization will concentrate in that bandonly. A periodic array of localization bands, as is predicted when using the periodic unitcell assumption, is therefore unrealistic. As such, the current model may overestimate thestrain at which instability occurs and underestimate the catastrophic nature of post-instability softening.

4. Hierarchical model of high strength steel

The key advantage of the modified damage model developed in the previous section overa standard Gurson damage model is that it has been developed from cell models whichhave successfully captured micro-void coalescence under loading conditions ranging fromzero to high triaxiality. Hence it is valid in pure shear but covers all stress states of interestto tensile failure also. The modified damage law is now illustrated within the context of ahierarchical model of high strength steel, similar to that used by Tvergaard (1992).However, here we are interested in the steel response when loaded in pure shear. Highstrength steels are known to exhibit a shear instability at relatively low shear strains(approximately 15%) under shear-loading conditions (Cowie et al., 1989).

A plane strain cell model is constructed which captures the interactions between theprimary (TiN) and secondary (TiC) particles in an alloy matrix. The cell is loaded underpure shear stress conditions. As it would be too expensive to explicitly model populations ofembedded particles over two very different length scales, a hierarchical approach is used:

The TiN large primary particles are modeled explicitly. � The TiC secondary particles and alloy matrix are modeled through (a) the modified

damage model formulated in the previous section and (b) a standard Gurson damage law.

The volume fraction of primary particles is 3% with particle diameters of 1 mm. Theauthors recognize that this particle fraction is artificially high. The volume fraction ofsecondary particles is set at 1%. The size of the finite elements is approximately equal tothe secondary particle spacing in the cell. This is consistent with other work based onGurson damage models and ensures that when localization occurs, the size of the resultingshear band has a physical meaning related to the necking region between two particles.

In Fig. 15, the larger TiN primary particles debond with the matrix at quite low strains,consistent with the inverse relationship which exists between particle size and interfacialstrength. The deformation is insufficient to drive void growth and coalescence of theprimary voids (the voids created by the debonding of primary particles). However,deformation does localize between the stable primary voids as shown in Fig. 15. Thematerial in the localized region consists of an alloy matrix with embedded TiC secondaryparticles (represented by the modified damage law developed in the previous section). Thelocalized strain in this region drives micro-void nucleation. Although the local triaxialityremains low, the modified damage model predicts micro-void coalescence in the localizedregion leading to an overall shear instability in the cell. The standard Gurson model fails to

Page 18: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESS

Fig. 15. A system of primary particles is modeled explicitly. The matrix material is modeled using (a) a regular

damage model (Gurson) and (b) the modified damage modeling which captures shear driven void coalescence. The

average response is given.

C. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244242

capture the low triaxiality coalescence of micro-voids and no plastic instability is observedin the average cell response in that case.The average response of the cell is defined by two key events (a) debonding of the

primary particles and (b) debonding of the secondary particles. Shear driven micro-voidcoalescence does not occur until the primary particles and secondary particles debondfrom the matrix. Hence, the average softening response in Fig. 15 is controlled by theinterfacial strength of the primary and secondary particles—as is observed experimentally.

5. Conclusion

Amicro-void shear localization mechanism has been identified in high strength steels duringzero triaxiality loading, using a computational cell modeling technique. Previous modelingefforts have been performed within the framework of micro-void sheets at skew angles undertensile loading in which a hydrostatic component of stress was available to encourage micro-void growth and coalescence. To the authors’ knowledge, this micro-void shear localizationmechanism has not been identified or modeled before under pure shear-loading conditions.The mechanism is captured by modeling the interactions between a population of micro-

voids nucleated through the decohesion of TiC secondary particles under shear-loadingconditions. The nucleated micro-voids subsequently grow into the shear plane inan elongated ‘tail like’ manner. As deformation increases, these micro-voids coalesce inthe shear plane. At higher shear strains, the secondary particles are driven towardsneighboring particles in the shear plane. The micro-voids which are being propped open by

Page 19: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244 243

these particles are forced to coalesce as the ligaments between them neck. This mechanismis consistent with experimental observations in which instability strain is delayed byincreasing interfacial strength of the TiC secondary particles and thereby delaying thedebonding and coalescence.

A hierarchical cell model has shown that the localized strain field between debondingprimary particles is sufficient to initiate the micro-void localization mechanism betweenprimary particles, again under pure shear stress loading conditions. The instability strainresulting from this void-sheet mechanism occurs at a realistic strain of 0.15.

The finite element method worked sufficiently well in the current investigation toidentify the mechanics of the shear localization process. However, it is noted that the largedeformation observed in the vicinity of the shear plane could be more accurately simulatedusing a mesh free method such as the reproducing kernel particle method (RKPM)(Liu et al., 1995a, b, 1997). The authors also recognize that gradient effects, which have notbeen modeled here, may play an important role at the scale of the secondary particles. Inparticular, the high gradients at the periphery of a micro-void are likely to generate highlevels of geometrically necessary dislocations which cause extra hardening in the alloymatrix and possibly resist void growth to some extent. We hope to address this issue byusing a gradient-based model such as that used by Fleck and Hutchinson (1997) torepresent the alloy matrix in future work.

Identification of the key micromechanics controlling the micro-void-sheet mechanism isa crucial step towards the goal of modeling the behavior of real 3D microstructures for thepurposes of optimized alloy design.

Acknowledgments

The authors gratefully acknowledge the support of the Office of Naval Research underthe D3D Digital Structure consortium program and the helpful advice of Professor DavidParks.

References

Arsenlis, A., Parks, D.M., 1999. Crystallographic aspects of geometrically-necessary and statistically-stored

dislocation density. Acta Mater. 47 (5), 1597–1611.

Bandstra J.P., Koss D.A., 2001. Modeling the ductile fracture process of void coalescence by void-sheet

formation. In: Materials Science and Engineering A 12th International Conference on the Strength of

Materials ICSMA-12, pp. 490–495.

Bandstra, J.P., Goto, D.M., Koss, D.A., 1998. Ductile failure as a result of a void-sheet instability: experiment and

computational modeling. Mater. Sci. Eng. A: Struct. Mater. Properties Microstruct. Process. A 249 (1, 2), 46–54.

Belytschko, T., Liu, W.K., Moran, B., 2000. Nonlinear Finite Elements for Continua and Structures. Wiley,

New York, 600pp.

Beremin, F.M., 1981. Cavity formation from inclusions in ductile fracture of A508 steel. Metall. Trans. 12A,

723–731.

Cowie, J.G., Azrin, M., Olson, G.B., 1989. Microvoid formation during shear deformation of ultrahigh strength

steels. Metall. Trans. A (Phys. Metall. Mater. Sci.) 20A (1), 143–153.

Dierickx, P., Verdu, C., Reynaud, A., Fougeres, R., 1996. A study of physico-chemical mechanisms responsible

for damage of heat treated and as-cast ferritic spheroidal graphite cast irons. Scripta Metall. 34, 261–268.

Fleck, N.A., Hutchinson, J.W., 1997. Strain gradient plasticity. Adv. Appl. Mech. 33, 296–358.

Gall, K., Yang, N., Horstemeyer, M.F., McDowell, D.L., Fan, J., 1999. Debonding and fracture of Si particles

during the fatigue of a cast Al–Si alloy. Metall. Mater. Trans. A: Phys. Metall. Mater. Sci. 30A (12),

3079–3088.

Page 20: An interactive micro-void shear localization mechanism in

ARTICLE IN PRESSC. McVeigh et al. / J. Mech. Phys. Solids 55 (2007) 225–244244

Goto, D.M., Koss, D.A., Jablokov, V., 1999. Influence of tensile stress states on the failure of HY-100 steel.

Metall. Mater. Trans. A: Phys. Metall. Mater. Sci. 30A (11), 2835–2842.

Gurson, A.L., 1977. Continuum theory of ductile rupture by void nucleation and growth and flow rules for

porous ductile media. J. Eng. Mater. Technol. 99 (1), 2–15.

Hao, S., Moran, B., Liu, W.K., Olson, G.B., 2003. A hierarchical multi-physics model for design of high

toughness steels. J. Comput.-Aided Mater. Des. 10 (2), 99–142.

Hao, S., Liu, W.K., Moran, Vernerey, F., Olson, G.B., 2004. Multi-scale constitutive model and computational

framework for the design of ultra-high strength, high toughness steels. Comput. Methods Appl. Mech. Eng.

193 (17–20), 1865–1908.

Hayzelden, C., 1987. Modeling of Interactions in void nucleation and growth. Sagamore Army Mater. Res. Conf.

Proc. Innovations Ultrahigh-Strength Steel Technol. 34, 383–424.

Hom, C.L., McMeeking, R.M., 1989. J. Mech. Phys. Solids 37 (3), 395–415.

Horstemeyer, M.F., Gokhale, A.M., 1999. Void-crack nucleation model for ductile metals. Int. J. Solids Struct. 36

(33), 5029–5055.

Horstemeyer, M.F., Matalanis, M.M., Sieber, A.M., Botos, M.L., 2000. Micromechanical finite element

calculations of temperature and void configuration effects on void growth and coalescence. Int. J. Plast. 16 (7,

8), 979–1015.

Horstemeyer, M.F., Ramaswamy, S., Negrete, M., 2003. Using a micromechanical finite element parametric study

to motivate a phenomenological macroscale model for void/crack nucleation in aluminum with a hard second

phase. Mech. Mater. 35 (7), 675–687.

Humphreys, F.J., 1989. In-situ SEM studies of the fracture of aluminium particulate composites. In: Proceedings

of EMAG MICRO 89. IOP Publishing Ltd., London, pp. 465–468.

Hutchinson, J.W., Tvergaard, V., 1987. Modeling of interactions in void nucleation and growth. Sagamore Army

Mater. Res. Conf. Proc.: Innovations Ultrahigh-Strength Steel Technol. 34, 347–356.

Kouznetsova, V., Geers, M.G.D., Brekelmans, W.A.M., 2002. Multi-scale constitutive modelling of

heterogeneous materials with a gradient-enhanced computational homogenization scheme. Int. J. Num.

Methods Eng. 54 (8), 1235–1260.

Lee, B.J., Mear, M.E., 1992. Axisymmetric deformation of power law solids containing a dilute concentration of

aligned spheroidal voids. J. Mech. Phys. Solids 40 (8), 1805–1836.

Lee, J.H., Shishidou, T., Zhao, Y.J., Freeman, A.J., Olson, G.B., 2005. Strong interface adhesion in Fe/TiC.

Philos. Mag. 85 (31), 3683–3697.

Liu, W.K., Jun, S., Zhang, Y.F., 1995a. Reproducing kernel particle methods. Int. J. Numer. Methods Fluids 20,

1081–1106.

Liu, W.K., Jun, S., Li, S.F., Adee, J., Belytschko, T., 1995b. Reproducing kernel particle methods for structural

dynamics. Int. J. Numer. Methods Eng. 38, 1655–1679.

Liu, W.K., Li, S.F., Belytschko, T., 1997. Moving least-square reproducing kernel methods. 1. Methodology and

convergence. Comput. Methods Appl. Mech. Eng 143, 113–154.

Manoharan, M., Lewandowski, J.J., 1989. In situ deformation studies of an aluminium metal–matrix composite

in a scanning electron microscope. Scripta Metall. 23, 1801–1804.

Poza, P., Llorca, J., 1996. Fracture toughness and fracture mechanisms of Al–Al2O3 composites at cryogenic and

elevated temperatures. Mater. Sci. Eng. 206A (2), 183–193.

Rice, J.R., Tracey, D.M., 1969. On the enlargement of voids in triaxial stress fields. J. Mech. Phys. Solids 17 (3),

201–217.

Sarkar, J., Kutty, T.R.G., Konlon, K.T., Wilkinson, D.S., Embury, J.D., Lloyd, D.J., 2001. Tensile and bending

properties of aa5754 aluminum alloys. Mater. Sci. Eng. A 316, 52–59.

Shabrov, M.N., Sylven, E., Kim, S., Sherman, D.H., Chuzhoy, L., Briant, C.L., Needleman, A., 2004. Void

nucleation by inclusion cracking. Met. Mater. Trans. A 35 (A), 1745–1755.

Tracey, D.M., Perrone, P.J., 1987. Modeling of interactions in void nucleation and growth. Sagamore Army

Mater. Res. Conf. Proc. Innovations Ultrahigh-Strength Steel Technol. 34, 315–330.

Tvergaard, V., 1989. Numerical study of localization in a void-sheet. Int. J. Solids Struct. 25 (10), 1143–1156.

Tvergaard, V., 1992. Numerical analysis of 3D localization failure by a void-sheet mechanism. Eng. Fract. Mech.

41 (6), 787–803.

Tvergaard, V., Needleman, A., 1984. Analysis of the cup-cone fracture in a round tensile bar. Acta Metall. 32 (1),

157–169.

Verhaege, B., Louchet, F., Brechet, Y., Massoud, J.P., 1997. Damage and rupture mechanisms in an

austenoferritic duplex steel. Acta Mater. 45, 1811–1819.