108
General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. Users may download and print one copy of any publication from the public portal for the purpose of private study or research. You may not further distribute the material or use it for any profit-making activity or commercial gain You may freely distribute the URL identifying the publication in the public portal If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Downloaded from orbit.dtu.dk on: May 10, 2021 Simulations of Mechanical Properties of CuZr and CuMg Metallic Glasses Paduraru, Anca Publication date: 2008 Document Version Publisher's PDF, also known as Version of record Link back to DTU Orbit Citation (APA): Paduraru, A. (2008). Simulations of Mechanical Properties of CuZr and CuMg Metallic Glasses.

Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

Users may download and print one copy of any publication from the public portal for the purpose of private study or research.

You may not further distribute the material or use it for any profit-making activity or commercial gain

You may freely distribute the URL identifying the publication in the public portal If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from orbit.dtu.dk on: May 10, 2021

Simulations of Mechanical Properties of CuZr and CuMg Metallic Glasses

Paduraru, Anca

Publication date:2008

Document VersionPublisher's PDF, also known as Version of record

Link back to DTU Orbit

Citation (APA):Paduraru, A. (2008). Simulations of Mechanical Properties of CuZr and CuMg Metallic Glasses.

Page 2: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Preface

This thesis is submitted in candidacy for the Ph.D. degree from the Technical

University of Denmark (DTU). It is based on the work carried out at the

Center for Atomic-scale Materials Design (CAMd), Department of Physics at

DTU from February 2005 to February 2008 under the supervision of Professor

Karsten W. Jacobsen and Professor Jacob Schøtz. Financial support was

provided by the EU Network on bulk metallic glass composites (MRTN-CT-

2003-504692 “Ductile BMG Composites”), Copenhagen Graduate School for

Nanoscience and Nanotechnology (CONT) and DTU.

I am grateful to my advisers Karsten Wedel Jacobsen and Jakob Schiøtz for

competent guidance throughout the years and whose patience and inspiring

discussions made this work possible. I thank Ole Holm Nielsen, Jens Jørgen

Mortensen and Marcin Dulak for help and patience in solving many technical

problems.

I thank Sergey Dobrin, Eva Fernandez, Niels Jensen, Jose Martinez, Duncan

Mowbray, and Mikkel Strange for proof reading this thesis and moral suport.

I also give many thanks to Duncan Mowbray for both his help with figures in

this thesis and our many useful discussions. Thanks to my office colleagues

and to all CAMd/CINF employees for creating a great social evironment.

I especially owe thanks to my family, for loving support and understanding in

letting me follow my dreams. They taught me the value of self-responsibility

and hard work.

i

Page 3: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under
Page 4: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Abstract

Molecular dynamic simulations have been performed to study the mechanical

properties of Cu15Mg85, respectively and Cu50Zr50 metallic glasses.

The many body interaction between atoms are described at the Effective

Medium Theory interatomic potential level of theory. In order to treat

Cu50Zr50 a new interatomic potential has been constructed. The interatomic

potential was fitted to lattice constants, elastic constants and cohesive ener-

gies of Cu, Zr and Cu50Zr50 in the B2 structure.

We tested the optimized Cu50Zr50 interatomic potential by comparing the

lattice properties of several CuZr structures to accurate Density Functional

Theory calculations and find a satisfactory agreement. Furthermore, with

the CuZr interatomic potential we are able to describe the thermodynamics

of CuZr glass during cooling in good agreement with experiments.

Two different method have been used to study the formation of shear bands

during plastic deformation: i) a simple shear deformation, and ii) nanoin-

dentation with a cylindrical indenter. For the simple shear deformation we

observe formation of shear bands with a width of 5 nm and 8 nm for Cu50Zr50

and Cu15Mg85, respectively.

The formation of shear bands during nanoindentation was found to be depen-

dent on the indentation velocity, radius of the indenter and the cooling rate.

Shear bands formation is more clear when we use low velocity of nanoin-

dentation, larger radius of the indenter and when the samples are cooled

slowly.

iii

Page 5: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under
Page 6: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Resume

De mekaniske egenskaber af Cu15Mg85 og Cu50Zr50 metalliske glasser er blevet

undersøgt ved hjælp af molekyldynamik.

Vekselvirkningerne mellem atomerne er beskrevet med et Effective Medium

Theory potential, der medtager mangelegeme-effekter i vekselvirkningerne.

Det interatomare potential blev tilpasset til gitterkonstanter, elastiske kon-

stanter og kohesivenergier for Cu, Zr og Cu50Zr50 i B2 strukturen.

Vi afprøvede det optimerede potential for CuZr ved at sammenligne krys-

talegenskaber for et antal CuZr strukturer med mere nøjagtige beregninger

baseret patthedsfunktionalteori, og finder tilfredsstillende overensstemmelse.

Potentialet er desuden i stand til at beskrive termodynamikken af en CuZr

glas under afkøling fra vskefasen med god overensstemmelse med eksperi-

mentelle data.

To forskellige metoder blev brugt til at undersøge dannelsen af forskyd-

ningsband (shear bands) under plastisk deformation: dels simpel deformation

under en forskydning, dels nanoindentation, som simulerer en hardhedsprøv-

ning med cylindriske prøvelegeme. Ved den simple forskydningsdeformation

observeres dannelse af forskydningsband med bredder pa5 og 8 nm i hen-

holdsvis Cu50Zr50 og Cu15Mg85.

Vi observerede, at dannelsen af forskydningsband ved nanoindentation afh-

nger af indentationshastigheden, prøvelegemets diameter og den hastighed,

hvormed glassen blev afkølet under sin dannelse. Dannelsen af forskyd-

ningsband er klarest ved lave hastigheder, store prøvelegemer og langsomt

afkølede metalglasser.

v

Page 7: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under
Page 8: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Contents

1 Introduction 1

1.1 Outline of this thesis . . . . . . . . . . . . . . . . . . . . . . . 2

2 Metallic Glasses 3

2.1 Glass transition temperature . . . . . . . . . . . . . . . . . . . 3

2.2 Properties and applications . . . . . . . . . . . . . . . . . . . 5

2.3 Processing of metallic glasses: Experimental methods . . . . . 8

2.4 Measuring hardness with indentation . . . . . . . . . . . . . . 11

3 Theoretical & Computational Methods 17

3.1 Methods for simulating metallic glasses . . . . . . . . . . . . . 18

3.2 Introduction to Effective Medium Theory . . . . . . . . . . . 23

3.3 Density functional theory . . . . . . . . . . . . . . . . . . . . . 27

3.4 Molecular dynamics. Equation of motion integration . . . . . 29

3.5 Energy minimization . . . . . . . . . . . . . . . . . . . . . . . 29

3.5.1 Molecular dynamics minimization . . . . . . . . . . . . 30

3.6 Langevin dynamics . . . . . . . . . . . . . . . . . . . . . . . . 30

4 EMT Interatomic Potential for CuZr 33

4.1 Fitting the EMT interatomic potential for CuZr . . . . . . . . 34

vii

Page 9: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

4.1.1 Downhill simplex method in multi-dimensions . . . . . 36

4.2 Testing the EMT interatomic potential . . . . . . . . . . . . . 37

4.2.1 Formation energies and lattice constants for CuZr crys-

talline structures. DFT versus EMT . . . . . . . . . . 37

4.2.2 MD and glass transition . . . . . . . . . . . . . . . . . 39

4.2.3 Radial distribution function and coordination numbers

for CuZr metallic glass . . . . . . . . . . . . . . . . . . 42

5 Shear Deformation 45

5.1 Simple shear deformation of CuZr and CuMg. Simulation

method and results . . . . . . . . . . . . . . . . . . . . . . . . 47

6 Nanoindentation 51

6.1 Simulation method . . . . . . . . . . . . . . . . . . . . . . . . 51

6.2 Shear band formation . . . . . . . . . . . . . . . . . . . . . . . 54

6.3 Influences on shear band formation . . . . . . . . . . . . . . . 57

6.4 Nanoindentation of crystalline samples . . . . . . . . . . . . . 63

6.5 Nanocrystals embedded In glass . . . . . . . . . . . . . . . . . 66

7 Conclusions 73

Bibliography 75

viii

Page 10: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

List of Included PapersPaper I:

An Interatomic Potential for Studying CuZr BulkMetallic Glasses

Anca Paduraru, Abdel Kenoufi, Nicholas P. Bailey and Jacob Schiøtz

Advanced Engineering Materials 9, 505 (2007)

Paper II:

Computer simulations of nanoindentation in Mg-Cuand Cu-Zr metallic glasses

Anca Paduraru, Nicholas P. Bailey, Karsten W. Jacobsen, and Jacob

Schiøtz

Submitted to Physical Review B

ix

Page 11: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under
Page 12: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Chapter 1

Introduction

The purpose of this thesis is to study the mechanical properties of metallic

glasses through atomistic simulation. A glass is a solid material obtained

from a liquid which does not crystallize during cooling. Most metals do

crystallize as they cool, with the atoms arranged in a regular spatial pattern

called a crystal lattice. A metallic glass (metallic amorphous) is obtained if

the liquid is cooled fast enough to avoid crystallization. If this happens, the

atoms are packed into a random manner similar to that in the liquid state,

lacking the long range order characteristic at crystalline materials. Due to the

lack of long range order, the plastic deformation is completely different from

that in the conventional materials. While in crystals the plastic deformation

is carried by dislocations, metallic glasses plastic deformation is correlated

with localized deformation events such as shear bands. This is believed to

be the reason why these materials have remarkable mechanical properties

[1, 2] such as high strength, fracture resistance, elastic limit and yield stress

(about 0.02Y , where Y is Young’s modulus) and corrosion resistance. Some

metallic glasses are ductile at room temperature, however, most are brittle.

To improve the ductility of these materials a better understanding of their

structure and plastic deformation is required.

We performed atomistic simulations on Cu15Mg85 and Cu50Zr50 metallic

glasses in order to study the mechanical properties by performing shear defor-

mations and nanoindentation. We use the Effective Medium Theory (EMT)

1

Page 13: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

2 Simulations of Mechanical Properties of Metallic Glasses

interatomic potential to describe the interactions between the atoms.

1.1 Outline of this thesis

An introduction to EMT and the simulation methods are presented in chapter

3. In the same chapter, a summary of other methods (particularly other

potentials) used to perform simulations on glasses is included together with

justification for our chose of potential. Chapter 4 presents the potential

we fitted for modeling the mechanical properties of CuZr metallic glasses.

Chapter 5 describes the shear deformation, which is the method we use to

measure the width of the shear bands for CuZr and CuMg metallic glasses.

Finally, chapter 6 contains a study of shear band formation and and how this

can be influenced by different parameters during simulated nanoindentation.

General conclusions follow in chapter 7.

Page 14: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Chapter 2

Metallic Glasses

A glass or an amorphous material is a disordered solid which does not pos-

sesses the long range periodicity present in a typical crystal. The definition of

a glass is made only based on its disordered structure, regardless of its chemi-

cal composition. Glass formation involves a cooling mechanism different from

crystalline materials. A glass is obtained by rapid cooling from the liquid

and is formed when the liquid becomes increasingly viscous during cooling

and fails to crystallize. An important characteristic of glass formation is the

glass transition temperature, which we will discuss in this chapter. We will

also present some experimental methods for making glasses, the mechanical

properties and their applications. Finally, we will provide an overview of

what can be done using computer simulations and why they are useful.

2.1 Glass transition temperature

The glass transition temperature is a characteristic of nanocrystalline materi-

als, which can be either organic polymers, metallic glasses or other inorganic

glasses. During cooling of a material, as the temperature decreases, the spe-

cific volume of the liquid decreases.

The slope of the volume versus temperature curve, normalized to the sam-

ple volume as shown in Fig. 2.1(a), is called the bulk thermal expansion

3

Page 15: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

4 Simulations of Mechanical Properties of Metallic Glasses

Figure 2.1: Specific volume as a function of temperature for a series of ma-

terials. (a) The liquid-to-crystalline solid transformation. A discontinuous

change in volume occurs at the melting temperature Tm. (b) The liquid-to-

glass transformation (the liquid-to crystall curve shown as reference). The

temperature at which the slope of the liquid-glass curve changes is the glass

transition temperature Tg reproduced from Schaffer et. all [3].

coefficient αv,

αv =1

V

dV

dT. (2.1)

When the cooling rate is not sufficiently high, a sample can crystallize at a

temperature Tm called the melting point. In this case, at Tm we can observe

a sudden drop in the volume versus temperature curves shown in Fig. 2.1(b).

This drop is a consequence of the change in structure, from that of a liquid to

that of a crystalline solid. Below Tm the specific volume continues to decrease

linearly with temperature. A crystal is the lowest energy state, and therefore

the most stable state for a material at low temperature. In some materials,

it is possible to avoid crystallization by cooling the liquid at a high rate,

so that the diffusion necessary to establish long range order is suppressed.

In this case, the volume of the sample decrease with the same slope as for

the liquid before the melting temperature, as shown in Fig.2.1(b), forming a

supercooled liquid. The slope of the curve for supercooled liquids decreases

at a certain temperature Tg below that of the crystal melting temperature

Tm, since a supercooled liquid cannot have a smaller volume than a crystal.

The slope of the enthalpy versus temperature curve changes as well at the

Page 16: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 2. METALLIC GLASSES 5

same temperature Tg. The temperature at which the slope of this curves

changes is called the glass transition temperature Tg.

2.2 Properties and applications

Mechanical properties of metallic glasses are different from those of con-

ventional materials. Due to lack of long range order, plastic deformations

cannot be carried out by lattice defects, as would happen in a crystal. In

the absence of a crystal slip plane, the plastic flow is concentrated into shear

bands, thus leading to high flow stresses [4]. One of the first interpretations

of glass deformation was made by Argon [5]. They found that the deforma-

tion mechanism appears to be of a very local nature. This localized strain is

produced by repeated nucleations (tiny regions of agglomeration) of shear,

which are confined within regions of free volume. According to Falk and

Langer [6], plasticity involves shear transformation zones, which operate as

centers of localized deformation. These models are supported by simulations

of deformation in metallic glasses [7, 8] and observations of localized events

[9, 10, 11].

For the purpose of developing reliable structural bulk metallic glasses, sev-

eral attempts have been made to enhance ductility [13]. A ductile material is

capable of undergoing large plastic deformations before fracture, and those

that fracture with little accompanying plastic deformation are brittle. To

enhance ductility, some composite materials have been synthesized by induc-

ing a second crystalline phase to generate multiple shear bands and to avoid

and/or delay the propagation of the shear bands through all the sample.

The ductility of glasses can also be enhanced by small changes in composi-

tion. Das et al. [12] showed (Fig. 2.2) that the presence of nanocrystals in

Cu50Zr50 enhances ductility, but not so efficiently as does the addition of 5%

Al in the composition of the glass.

Due to their structure, metallic glasses are materials with special mechanical

properties. These include high strength, high strain limit for the Hookean

elasticity, fatigue resistance, and often corrosion resistance. The fracture

Page 17: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

6 Simulations of Mechanical Properties of Metallic Glasses

Figure 2.2: Stress-strain curves of (a) Cu 50Zr50 and (b) Cu47.5Zr 47.5Al5 under

the compression test at a strain rate of 8−4s−1 ,showing a highly ”work-

hardenable” metallic glasses up to 18 % strain. The inset shows the true

stress-strain curve of alloy (b) Cu47.5Zr47.5Al5 as reproduced from Das et. al

[12].

toughness in metallic glasses is higher than in conventional glasses. In Fig.

2.3, we see that the toughest metallic glasses lie amongst the best of metals

[14].

The tensile fracture strength σf and the hardness Hv of metallic glasses have

a tendency to increase with the Young’s modulus, E, of these materials. Fig.

2.4, shows the relationship between σf , Hv and E for bulk amorphous and

conventional crystalline alloys. The amorphous alloys have both a high σf of

840-2100 MPa and a high E of 47-102 GPa [2].

We will now consider potential applications of metallic glasses as structural

materials and provide some examples of applications which are related to the

material’s properties.The high hardness of metallic glasses may be used for

tooling, particularly as knife edges. The wear resistance combined with corro-

sion resistance of metallic glasses is used in coatings for industrial equipment,

machinery and valves, in high wear, extreme temperature and high corrosion

environments. Metallic glass alloys are also used in the electronic casings in-

Page 18: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 2. METALLIC GLASSES 7

Figure 2.3: Fracture toughness and Young’s modulus for metals, alloys, ce-

ramics, glasses, polymers and metallic glasses (the gray circles represents

metallic glasses ). The contours show the toughness Gc, reproduced from

Ashby and Greer [14].

dustry and for hinge applications, see Fig. 2.5. Here they have the advantage

that they enable smaller, thinner and more durable designs [15].

Excellent wear resistance and high strength to weight ratio compared to

titanium and/or stainless steel make these materials good candidates for

medical devices. These include reconstructive devices, fractured fixations,

spinal implants, and instrumentation [15]. For medical instrumentation, they

can yield surgical blades that are sharper than steel and less expensive than

diamond, see Fig. 2.6. Metallic glasses are also used in the production of

sporting goods and leisure products such as golf clubs, tennis rackets and

baseball bats amongst others. High polish-ability, abrasion resistance and

corrosion resistance [14] make them good candidates for producing jewelry

such as rings, watches cases, and pens, see Fig.2.7.

Page 19: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

8 Simulations of Mechanical Properties of Metallic Glasses

Figure 2.4: Relationship between the tensile fracture strength σf or the Vick-

ers hardness Hv and Young’s modulus E for bulk amorphous and conventional

crystalline alloys, reproduced from Inoue [2].

.

2.3 Processing of metallic glasses: Experi-

mental methods

A glass is formed if a liquid is cooled sufficiently fast to avoid crystalliza-

tion. The most important processing step in producing metallic glasses is

the quench from the liquid state. The quench rate must be very fast (> 106

K/s) to ”freeze” the atoms in the metastable structure. Therefore, rapid

heat removal from the bulk is needed during cooling.

The gun technique developed by Duwez [16] was one of the earliest rapid

solidification method used to produce amorphous alloys. In the gun tech-

nique, a molten alloy is ejected by using a shock wave to generate very small

(∼ 1µm) droplets impacting at sonic velocity onto a chilled substrate, as

shown in Fig. 2.8.

The most often used technique for production of metallic glasses is melt

spinning, which is used to obtain wires and ribbons. This method has the

advantage of obtaining the metallic glass products in a continuous form with

Page 20: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 2. METALLIC GLASSES 9

Figure 2.5: Applications of metallic glasses as hinges due to a much higher

yield strength and hardness than any commercially cast alloys, offering flex-

ibility of design. Reproduced from [15].

cooling rates of the order of 105-106 K/s. Metallic ribbons are obtained with

this method by allowing a stream of the molten alloy to flow onto a rapidly

moving substrate. The principle of the method is shown in Fig. 2.9. By

ejecting a jet of molten alloy into a stream of water a wire of metallic glass

can be obtained, as shown in Fig. 2.9.

In the methods described above, the cooling rates achieved is of the order

of 106 K/s. A much higher cooling rate may be achieved using pulsed laser

quenching. In this method a very thin (∼ 100 nm) surface layer is melted

by an incident beam of duration as short as a few picoseconds. The layer on

top of a large cold substrate may experience ultra rapid cooling, as fast as

1014 K/s [1].

Melt atomization is a an important technology for producing rapidly solidified

Page 21: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

10 Simulations of Mechanical Properties of Metallic Glasses

a) b)

Figure 2.6: Applications of metallic glasses as medical devices for (a) the

orthopedic field due to excellent wear resistance, biocompatiblity and high

strength to weight ratio and (b) surgical blades with lower edge degradation

and improved sharpness over steel. Reproduced from [15]

.

a)

b)c)

Figure 2.7: Applications of metallic glasses for (a) watches cases, (b) pens and

(c) jewelry (rings) due to polish-ability and abrasion and corrosion resistance.

Reproduced from [15]

.

powders. Atomization is a process whereby powder particles are formed from

the quenching and solidification of discrete droplets of liquid metal.

Page 22: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 2. METALLIC GLASSES 11

Figure 2.8: Duwez gun technique [16] of splat quenching from the melt, as

reproduced from [17].

2.4 Measuring hardness with indentation

An indentation test is an excellent way to measure the mechanical proper-

ties of very small volumes of materials. In an indentation test, a hard tip,

typically a diamond, is pressed into the sample with a known load. After

some time, the load is removed. The area of the residual indentation in the

sample is measured and the hardness, H, is defined as the maximum load, P,

divided by the residual indentation area.

Hardness is a measure of a material’s resistance to plastic deformation during

Page 23: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

12 Simulations of Mechanical Properties of Metallic Glasses

Figure 2.9: Principal methods of rapid quenching from the melt: (a) drop

smasher, using levitation-melting by induction, the pistons are pneumati-

cally accelerated and come into action when the falling drop breaks a light

beam, (b) melt-spinning, (c) pendant drop melt-extraction and (d) twin-roller

quenching device.

nanoindentation. A material’s resistance of the materials to indentation is

not necessarily the same as its resistance to abrasion. However, a hardness

measurement obtained from an indentation test can be used as an empiri-

cal test for abrasion resistance. In general, materials which possess a high

hardness also have a high resistance to abrasion.

There are various types of indentation tests available [3, 18, 19], the most

common being the Brinell test, the Vickers diamond test and the Rockwell

test.

Brinell Hardness Test is one of the oldest hardness test methods in com-

Page 24: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 2. METALLIC GLASSES 13

mon use today. The Brinell test is frequently used to determine the hardness

of steels and cast irons. By varying the test force and ball size, nearly all

metals can be tested using a Brinell test. In its original form, a Brinell test

uses a steel ball indenter of 10 mm diameter forced into the sample by a

controlled test force. The force is maintained for a specific time, normally 10

- 15 seconds. After this, the indenter is removed leaving a round indent in

the sample. The size of the indent is determined optically by measuring two

diagonals of the round indent using either a portable microscope or one that

is integrated into the load application device. The Brinell hardness number

is the ratio of the test force to the surface area of the curved indent. The

indentation is considered to be spherical with a radius equal to half the di-

ameter of the ball. The average of the two diagonals is used in the following

formula to calculate the Brinell hardness:

HB =applied load(kg)

surface area of the impression(mm2)=

2F

πD(D −√

D2 − d2)(2.2)

where D is the diameter of the indenter, d is the diameter of the impression

and F is the load as shown in Fig.2.10.

Figure 2.10: Brinell hardness test [19].

However, the Brinell test is not suitable for testing very hard materials. This

is because as the hardness of the material approaches that of the indenter,

there will be a tendency for the indenter to deform.

The Vickers Test(HV) was developed in England in 1925 and was formally

known as the Diamond Pyramid Hardness (DPH) test. The Vickers test has

Page 25: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

14 Simulations of Mechanical Properties of Metallic Glasses

two distinct force ranges, micro (10 g to 1000 g) and macro (1 kg to 100 kg), to

cover all testing requirements. All Vickers tests use the same 136◦ pyramidal

diamond type indenter, which forms a square indent. As in the Brinell test,

the indenter is pressed into the sample by an accurately controlled test force

for 10-15 s. After removing the indenter, the size of the indent is determined

optically by measuring the two diagonals of the square indent. The Vickers

hardness number is the ratio of the test force to the surface area of the indent.

The average of the two diagonals is used in the following formula to calculate

the Vickers hardness:

HD =applied load (kg)

surface area of the impression(mm2)

=Constant x test force

indent diagonal squared=

2 sin θ/2F

D2, (2.3)

where θ = 136◦ , F is the applied force in kg and D is the mean diagonal

length in mm as shown in Fig. 2.11.

Figure 2.11: Vickers hardness test [19].

The Rockwell Hardness test also uses a machine to apply a specific load

and then measure the depth of the resulting impression. The indenter may

either be a steel ball of some specified diameter or a spherical diamond-tipped

cone at a 120◦ angle with a 0.2 mm tip radius, called a brale. A minor load

Page 26: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 2. METALLIC GLASSES 15

of 10 kg is first applied, which causes a small initial penetration to seat the

indenter and remove the effects of any surface irregularities. The dial is then

set to zero and a larger load called the major load, is applied. Upon removal

of the major load, the depth reading is taken while the minor load is still

on. The hardness number may then be read directly from the scale. The

indenter and the test load determine which hardness scale that is used (A,

B, C, etc).

While lowering the indenter into the sample a load displacement curve is

recorded. The load displacement curve provides information about the macro-

scopic deformation character of the material [20].

Page 27: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under
Page 28: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Chapter 3

Theoretical and Computational

Methods

This chapter contains details about the theory and molecular dynamics meth-

ods used to perform simulations of metallic glasses. In order to study into

the mechanical properties of metallic glasses, we need large systems (on the

order of up to 1 million atoms). This makes impractical the use of quantum

theories and ab initio methods, which, although accurate are very time con-

suming, and may only simulate systems of up to about 200 atoms. To address

this problem, we will use Effective Medium Theory (EMT) [21, 22, 23, 24],

which is explained in section 2.1. EMT uses a potential which is derived from

Density Functional theory (DFT), and allows extensive molecular dynamics

simulations to be performed. The parameters of the interatomic potential

given by EMT will also be discussed. Sections 2.2 and 2.3 contain informa-

tion about how we perform the Molecular Dynamics (MD), how we integrate

the equation of motion, how we minimize the energy without using the tem-

perature, and finally how we perform MD at different temperatures.

17

Page 29: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

18 Simulations of Mechanical Properties of Metallic Glasses

3.1 Methods for simulating metallic glasses

One of the most common ways to simulate large systems is using a Lennard

Jones (LJ) potential, which is given by the expression:

ΦLJ = 4ε

[(σ

rij

)12

−(

σ

rij

)6]

. (3.1)

for the interaction potential between a pair of atoms.

The potential has an attractive tail at large distances rij, but reaches a

minimum and becomes strongly repulsive at shorter distances.

This potential has been used to simulate glassy systems as Fe83M17 (M: C, B,

P), CuY, ZrTi and model glasses [25, 26, 27, 28, 6]. Evteev et AP.[25] use this

potential to describe the interaction of atoms in Fe83M17 (M: C, B, P) metallic

glasses and to study the individual characteristics of local atomic ordering

on structure formation. Molecular dynamics has been performed to produce

the glasses in the compositions mentioned above, and radial distribution

functions were calculated as shown in Fig. 3.1. The results they obtain are

in good agreement with experiments.

The same study was made of CuY glasses [26]. However, in this case they

found poor agreement with experiments in the case of gCuCu(r), as is shown

in Fig. 3.2. The reason why the position and amplitude of the first peak for

the pair distribution functions gY Y and gCuY is in good agreement could be

because they were heavily weighted in the fitting procedure.

The Lennard Jones potential was also used to describe interactions in a model

glass and to simulate nanoindentation in order to study the shear bending

and strain localization under the indenter [28, 6].

Lennard Jones potential describes some systems well, and is useful for investi-

gations where the focus is fundamental issues, rather than studying the prop-

erties of a specific material. However, such pair potentials cannot describe

correctly the transition metals, particularly the bulk properties of metals,

e.g. elastic behavior. It is also not possible to describe atoms with different

coordination numbers using the same pair potential.

Page 30: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 3. THEORETICAL & COMPUTATIONAL METHODS 19

Figure 3.1: Pair distribution function of the amorphous alloys: Fe83C17 (1),

Fe83B17 (2) and Fe83P17 (3). Models are shown by a solid line and dots

represent X-ray pair distribution functions from experimental data for (2)

and (3) alloys, reproduced from Ecteev et al. [25].

In the literature it has been proposed that other methods may serve as alter-

natives to the Lennard Jones potential. One is the Embedded-Atoms Method

(EAM) potential [29] used, for example, to study glasses such as CuW and

CuAg [30, 31]. Based on this method, molecular dynamics has been applied

to obtain an atomic description of the glass formation process in a eutectic

C40Ag60 alloy [31]. In this case, structure and glass forming ability were

studied. Their results showed that EAM correctly predicts the glass transi-

tion. Also, the structure analysis through the calculated radial distribution

function revealed the formation of different clusters at different quenching

rates during the quenching process as shown in Fig. 3.3.

Page 31: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

20 Simulations of Mechanical Properties of Metallic Glasses

Figure 3.2: Calculated partial pair correlation functions for a Cu33Y67 metal-

lic glass from experiments (dots line) and molecular dynamics simulations

(solid line), reproduced from Frattini et al. [26].

A similar method to EAM is the Effective Medium Theory interatomic poten-

tial, which has proved to be a good method to describe the properties of close

packed metallic alloys [21, 22, 24]. This method was used by N. Bailey [32]

to describe both glass formation ability and structural properties in CuMg

glassy alloy. This method was also efficient in studying the mechanical prop-

erties of CuMg glasses and the shear band deformation mechanism [33, 34].

In order to study the shear bending, a simulated uniaxial tension test was

performed [34]. This showed that a necking instability occurred rather than

shear bending, as shown in Fig. 3.4. For initiating the shear bands a notch

was used. In the same uniaxial test using a notch it was shown that the

deformations then takes place in shear bands that lead away from the notch

at 45◦, as shown in Fig. 3.5.

Using the EMT interatomic potential, we performed simulations in order

to study the mechanical properties of CuZr metallic glasses, as discussed

Page 32: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 3. THEORETICAL & COMPUTATIONAL METHODS 21

Figure 3.3: Two dimensional cross-sectional projections of the structures

obtained at different quenching rates for a Cu40Ag60. (a) 1x1011, (b) 1x1013,

(c) 1x1014. These projections exhibit the structure of CuAg alloys. The

transparent circles represent the Ag atoms and the colored circles represent

the Cu atoms. The retention of amorphous structure requires an extremely

high cooling rate (1x1014 K/s) as reproduced from L. Qi et al. [31].

in Paper I. Thermodynamic properties were recorded during cooling of the

system and the glass transition, temperature and structure study results were

compared with experiments. In order to study the mechanical properties, we

have performed shear and nanoindentation tests. Details of these simulations

will be presented in the following chapters.

CuZr glass formation and structure have been simulated before by Duan

et al. [35]. They fitted the effective Rosato-Guillope-Legrant-type force

field parameters to data, such as lattice constants, cohesive energies and

bulk moduli obtained using a quantum mechanical model. They performed

molecular dynamics to obtain the glassy system, and observed a change in

Page 33: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

22 Simulations of Mechanical Properties of Metallic Glasses

Figure 3.4: Outline of a 14 nm×14 nm×14 nm MgCu metallic glass deformed

by uniaxial tension after 0%, 20%, 40%, 60%, 80% and 100% strain [34].

Figure 3.5: Visualization by stripe painting in a 56 nm×56 nm×18 nm MgCu

metallic glass deformed at different amounts of strain [34].

the slope of the volume versus temperature curve that corresponds to the

glass transition, as shown in Fig. 3.6. Near the glass transition temperature

they observed a change in slope of the variation of volume, enthalpy and

entropy with temperature. They calculated the radial distribution functions

and the coordination number, shown in Fig. 3.7.

The results we obtain performing simulations with a fitted EMT interatomic

potential for these properties are slightly better, and these results are dis-

cussed in chapter 3.

Similar methods have been used to describe the interactions of very large

metallic glass systems. These include quantum Sutton-Chen many-body po-

tentials for studying CuAg and CuNi glasses [38], the Finnis-Sinclair type

pair functional potential for ZrNi [39], and the same ZrNi metallic glass

Page 34: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 3. THEORETICAL & COMPUTATIONAL METHODS 23

Figure 3.6: Volume versus temperature curves for Cu46Zr54 obtained for three

different quenching rates. The glass transition temperature is where the slope

of the curve changes between 600K and 800K. Reproduced from [35].

Figure 3.7: The numbers of nearest neighbor correlations obtained from dif-

ferent techniques [35], aReference [36] , bReference [37], cReference [35]. The

experimental data was chosen by Duan at al. to compare with their simula-

tion data.

systems adapted to Hausleitne-Hafner interatomic potentials [40] analyzed

within the mode coupling theory (MCT) [41].

3.2 Introduction to Effective Medium The-

ory

The energy of an atom in an arbitrary environment is calculated within EMT

in order to calculate the interatomic forces. The energy of an atom in the

working environment is calculated first in a reference system. In our case, the

Page 35: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

24 Simulations of Mechanical Properties of Metallic Glasses

reference system is a perfect crystal. The reference system and the actual

system are related through the lattice constant, which is adjusted in such

a way that an atom in the fcc crystal is surrounded by the same average

electron density as an atom in the real system.

The total energy of the ith atom, after taking into account the previous

assumptions, will be the sum of two terms. These are the cohesive energy

Ec,i, which is the energy of the reference system, and a correction EAS,i,

called the atomic sphere correction, which estimates the difference in energy

between the two systems. The total energy of the system, E, is then:

E =∑

i

[Ec,i(ni) + ∆EAS,i]. (3.2)

The atomic sphere correction is calculated as the difference between a pair

potential for the ith and jth atoms, Vij, in the actual system and the pair

potential in the reference system, V refij . The one electron correction is in-

cluded in the atomic sphere correction by adjusting the pair potentials, so

that we obtain

E =∑

i

{Ec,i +

1

2

[∑j 6=i

Vij(rij)−ref∑j 6=i

Vij(rrefij )

]}, (3.3)

where rij and rrefij are the distances between the ith and jth atoms in the

working and reference systems, respectively.

We will begin by considering a mono-atomic system. Before we may calculate

the total energy, we must first evaluate the density. We may then express

the density ni of an atom i as:

ni =∑j 6=i

∆nj(si, rij). (3.4)

We assume that the density contribution to an atom i from a neighboring

atom ∆nj(s, r) at a distance r averaged over a neutral sphere of radius s has

the form:

∆n(s, r) = ∆n0eη1(s−s0)−η2(r−βs0), (3.5)

Page 36: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 3. THEORETICAL & COMPUTATIONAL METHODS 25

where s0 is the equilibrium neutral sphere radius, and η1 and η2 describe the

exponential decay of the density. Here,

β =√

2

(16π

3

) 13

(3.6)

is a geometric factor given by the assumption that the reference system is

an fcc crystal, and only the 12 nearest neighbor contributions are taken into

account. The nearest neighbor distance is taken to be d = βs0. As mentioned

before, the two systems are linked by adjusting the lattice constant through

the neutral sphere radius, s, to find a relation between s and the density, n.

From DFT calculations, we find that the relation between the neutral sphere

radius and the density is also an exponential, so that

n(s) = n0e−η(s−s0). (3.7)

Since the reference system is an fcc crystal, we have ∆nn0

= 12, rfcc= βs and

the exponential decay factors are related by βη2 = η + η1. Solving equation

3.7 using equation 3.5, we find the neutral sphere radius of the ith atom is:

si = s0 −1

βη2

log(σ1,i

12

), (3.8)

where σ1,i:

σ1,i =∑j 6=i

e−η2(rij−βs0). (3.9)

As shown in equation 3.3, the atomic sphere correction is the difference be-

tween the actual system pair potential and the reference system pair poten-

tial. As the reference system is the fcc crystal, its pair potential is the sum

over only the 12 nearest neighbors, so that

∆EAS = −1

2

[12V (βs)−

∑j 6=i

V (rij)

], (3.10)

and the pair potential is then

V (r) = −V0e−κ( r

β−s0). (3.11)

Page 37: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

26 Simulations of Mechanical Properties of Metallic Glasses

From equation 3.10 and 3.11, we find that the atomic sphere correction be-

comes:

∆EAS = −6V0e−κ(s−s0) +

1

2V0

∑i6=j

e−κ( rβ−s0), (3.12)

∆EAS = −6V0

[e−κ(si−s0) − σ2,i

12

], (3.13)

where σ2,i

σ2,i =∑i6=j

e−κ(rijβ−s0). (3.14)

The cohesive energy is also a function of the neutral sphere radius, which we

assume has the form

Ec,i(si) = E0f [λ (si − s0)] , (3.15)

where f = (1 + x)e−x [42].

We have shown how the EMT potential is parameterized for mono-atomic

systems, but as will be shown in the following chapters, the present work

is focused on binary alloys. For binary systems, the cohesive energy of each

atom depends again on E0, λ, and s0, which will be different for each element.

If we call the two elements A and B, then the cohesive energy of A type atoms

will depend on parameters E0A, λA, and s0A, and for B type atoms on E0B,

λB, and s0B. As mentioned in the calculations for the mono-atomic system,

the total energy of the reference system is determined by the embedding

density n which is dependent on s (equation 3.7). For the case of a binary

system, the neutral sphere radius for type A atoms has the form

sA = s0A −1

βη2A

log1

12(σ1AA + χABσ1AB) , (3.16)

where χAB = n0B/n0A is a scale factor, assuming η1A = η1B. The atomic

sphere correction for type A elements is then

∆EAAS =

∑jA 6=i

VAA(rijA) + χAB

∑jB 6=i

VAB(rijB)− 12VAA(βsiA). (3.17)

Page 38: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 3. THEORETICAL & COMPUTATIONAL METHODS 27

Previously, the parametrization of the EMT potential for the mono-atomic

and binary systems was made assuming that the reference system is an fcc

crystal and only the 12 nearest neighbors contribute to the density. To ac-

count for more than just nearest neighbor contributions, the sum σ1,i (equa-

tion 3.9) may be replaced by

σ1,i =1

γ1

∑j 6=i

e−η2(rij−βs0)Θ(rij), (3.18)

where γ1 scales σ1,i so that it still has the same value in an fcc crystal when

we extend the sum beyond nearest neighbor. Here, Θ(rij) is a cut-off function

which is chosen to have the Fermi form

Θ(rij) =1

1 + ea(r−rc), (3.19)

where a determines the sharpness of the cut-off and rc is the cut-off radius.

Similarly, σ2,i becomes:

σ2,i =1

γ2

∑i6=j

e−κβ

(rij−βs0)Θ(rij). (3.20)

The scale factors γ1 and γ2 were chosen in such way to ensure Ec = Ecoh and

∆EAS = 0 for the equilibrium reference system beyond the nearest neighbor.

The given equations describe the EMT potential’s seven parameters for each

element. The way these parameters are fitted will be described in the next

chapter.

3.3 Density functional theory

Density functional theory (DFT) is a quantum mechanical method for cal-

culating the ground state energy for many-body systems. The advantage of

DFT is that it reduces a problem in terms of the many electron wave function

Ψ(r1, . . . rn) into one in terms of the electron density n(r1). This is based on

the Hohenberg Kohn theorem which states that for a given external poten-

tial the ground state energy is a unique functional of the electron density,

i.e. E0[n],

Page 39: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

28 Simulations of Mechanical Properties of Metallic Glasses

E0[n] = F [n] +

∫n(r)vext(r)dr (3.21)

where

F [n] = T [n] + Vee[n], (3.22)

T [n] is the kinetic energy, Vee describes the electron-electron interaction and

vext is the external potential. However, an exact form for F [n] in terms of

the density is unknown. Instead, we rewrite F [n] in the form:

F [n] = T [n] +1

2

∫ ∫drdr′

n(r)n(r′)

| r− r′ |−

∫drn(r)vxc[n(r)](r) (3.23)

where vxc is the exchange correlation potential which gives the correct electron-

electron interaction Vee. This potential includes corrections to the Coulomb

potential so that the Pauli exclusion principle is satisfied, electron-electron

correlations, and self interactions corrections. In practice the most common

implementation of DFT utilizes the Kohn-Sham method. Here, we assume a

form for vxc, and then solve the non-interacting Schrodinger equation

i

2∇2Ψi + v(r)Ψi = EiΨi, (3.24)

with a potential v[n(r)] = vH [n] + vxc[n] + vext(r), which depends on the

density n =∑

Ψ∗i Ψi, in a self consistent manner.

We use DFT in order to obtain parameters (lattice constants, formation ener-

gies, elastic constants), which we need to fit the EMT interatomic potential.

These parameters were calculated only for the crystalline structures that are

presented in chapter 4. The calculations were done using the GGA functional

PW91[43], a wave function cutoff of 350 eV, and a density cutoff of 450 eV.

We also used a grid of 8 k-points sampling in each direction.

The total energy of the system was calculated by varying the lattice constant

a. The lattice constant for each structure was chosen around equilibrium,

when the energy reaches a minimum. From the energy versus lattice constant

plot we calculate d2E/da2, which we used to calculate the bulk modulus

B = Vd2E

dV 2=

1

9a

d2E

da2. (3.25)

Page 40: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 3. THEORETICAL & COMPUTATIONAL METHODS 29

3.4 Molecular dynamics. Equation of motion

integration

The most time consuming part of Molecular Dynamics simulations is typi-

cally the calculation of the force acting on every atom. In our case, we have

to consider the contribution to the force on an atom i from all its neighbors.

Once we have computed the forces between the atoms, we then integrate

Newton’s equation of motion

md2ri

dt2=

∑j 6=i

F(rij). (3.26)

To integrate 3.26, we use a velocity Verlet algorithm [44], which has proved

to be one of the simplest and most efficient [45, 46]. We consider an atom i of

mass m at time t, and using the velocity Verlet algorithm we may compute

the new positions ri(t + ∆t) and velocity vi(t + ∆t) after a time step ∆t as:

ri(t + ∆t) = ri(t) + vi(t)∆t +Fi(t)

2m∆t2, (3.27)

vi(t + ∆t) = vi(t) +Fi(t + ∆t) + Fi(t)

2m∆t, (3.28)

where Fi(t) is the force that acts on the ith atom at a time t.

3.5 Energy minimization

The goal of energy minimization is to find a route from an initial configura-

tion to the nearest minimum energy conformation using as few calculations as

possible. The energy minimization is very important for an analysis of Molec-

ular Dynamics, where after a simulation is completed, the configuration is

used as the starting configuration for energy minimization. The minimiza-

tion algorithm should give the best structure for each configuration. The

energy minimization can be done using various algorithms, some of which

are very briefly described here, with special emphasis on their applicability

to metallic glasses.

Page 41: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

30 Simulations of Mechanical Properties of Metallic Glasses

3.5.1 Molecular dynamics minimization

In our simulations, we used a velocity Verlet algorithm to integrate the equa-

tions of motion and ran our MD algorithm, stopping the atoms when they

moved away from a minimum. This algorithm is implemented in our ASAP

program from the CAMd Open Software project [47] as the MDmin algorithm

[48]. First we evaluate pi(t) · Fi(t), where pi(t) is the momentum at time

t. If pi · Fi(t) > 0 it means that the atom is approaching a local minimum

for the potential energy. If pi · Fi(t) becomes zero (meaning that the atoms

is near the minimum for the potential energy) or negative then we zero the

momentum. This means that we remove the system’s kinetic energy, which

will be close to its maximum.

3.6 Langevin dynamics

The methods described previously allow us to perform Molecular Dynamics,

but in the absence of temperature. However, we need to be able to perform

our simulation at finite temperature. In order to do this, we use Langevin

Dynamics, which we consider to be a suitable thermalization algorithm for

the calculations in the present work. The idea behind Langevin dynamics

[48, 49] is to add two additional terms to Newton’s second law, which let

the system interact with a temperature reservoir. The Langevin equation is

then:

md2ri

dt2= Fi(t) + Ri(t)− γvi. (3.29)

The last term added to Newton’s law represents a frictional force γvi, where

ζ = γ/m is the friction coefficient, while the other term Ri(t) is a random

force. We require that the friction coefficient is related to the fluctuations of

the random force by the fluctuation-dissipation theorem, so that

〈R(t) ·R(t′)〉 = 2ζkBTδ(t− t′). (3.30)

In equation 3.30, the terms on the left side have been averaged over time t,

kB is Boltzmann’s constant, and T is the temperature. The energy that the

Page 42: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 3. THEORETICAL & COMPUTATIONAL METHODS 31

system loses through friction is compensated by the random force, R(t). If

the energy gets too high, then the friction term becomes dominant and the

system loses energy. The Langevin algorithm is more time-consuming than

the Verlet one, and it involves the choice of a friction coefficient, which for

this work was chosen to be ζ = 0.1.

Page 43: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under
Page 44: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Chapter 4

EMT Interatomic Potential for

CuZr

CuZr has recently been discovered to be a binary bulk metallic glass [50, 51].

Since binary alloys are easier to model than alloys with more elements, this

makes CuZr an attractive bulk metallic glass to study theoretically. The

quality of the results obtained will depend on the quality of the interatomic

potential, with simple potentials giving a less accurate description of the in-

teratomic interactions while allowing very large simulations, and more com-

plicated potentials giving a better description of the interaction, but limiting

the simulation size. Many-body potentials, such as the Embedded Atom

Method [29] and the Effective Medium Theory (EMT) [22, 21], have been

shown to give a good description of late transition metals, and their alloys

yielding crystallization in close-packed structures while still allowing simula-

tions with millions of atoms [52]. In this chapter, we will discuss an EMT

potential optimized for modeling the mechanical and thermodynamic prop-

erties of CuZr bulk metallic glass.

Previously, an interatomic force field has been fitted to CuZr by Duan et

al., but we find that the EMT force field described here provides a better

description of the structure of the metallic glass. The potential developed

here will be used to model the mechanical properties of CuZr, which are

discussed in the following chapters.

33

Page 45: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

34 Simulations of Mechanical Properties of Metallic Glasses

Property Optimized Target Difference Error

value value

Cu Shear Modulus 0.511477 0.511 0.000477 0.1%

Cu Bulk Modulus 0.889516 0.886 0.003516 0.4%

Cu C11 1.093305 1.100 -0.00695 0.6%

Cu Cohesive Energy 3.521416 3.510 0.011416 0.3%

Cu Lattice Constant 3.587869 3.610 -0.22131 6.1%

Zr Shear Modulus 0.141930 0.150 -0.008070 5.4%

Zr Bulk Modulus 0.474796 0.440 0.034796 7.9%

Zr Cohesive Energy 6.300625 6.300 0.000625 0.0%

Zr Lattice Constant 4.570 4.552 -0.017120 0.4%

CuZr Shear Modulus 0.306406 0.276 0.030406 11.0%

CuZr Bulk Modulus 0.666679 0.751 -0.084321 11.2%

CuZr Formation Energy -0.139429 -0.140 0.000571 0.4%

CuZr Lattice Constant 3.192319 3.280 -0.087681 2.7%

Table 4.1: Properties used in fitting the potential. Elastic constants are in

eV/A3, energies are in eV/atom and lattice constants are in A. All values for

CuZr refer to the B2 structure. The first five values (for pure Cu) were not

used in the fitting in this work, but are provided for reference. All values for

Zr and CuZr were obtained from DFT calculations, except the Zr cohesive

energy [53] and CuZr formation energy [54, 55], which were obtained from

experiments. See Paper I.

4.1 Fitting the EMT interatomic potential for

CuZr

In the previous chapter, we have described the functional form of EMT and

its use of seven parameters for each atomic element. Two of these parame-

ters describe the contribution of each atom to the local electron density in

its neighborhood. These are n0, which describes the magnitude of this con-

tribution, and η2, which is the inverse decay length. The EMT parameters

s0, λ and E0 describe the position of the minimum, the curvature, and the

Page 46: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 4. EMT INTERATOMIC POTENTIAL FOR CUZR 35

Parameter Cu Zr

s0 2.67 1.78

E0 -3.51 -6.3

λ 3.693 2.247

κ 4.943 3.911

V0 1.993 2.32

n0 0.0637 0.031

η2 3.039 2.282

Table 4.2: EMT parameters from Paper I.

depth of the energy function of the atoms, respectively. The pair potential

function is described by the parameters V0 and κ, which is also an inverse

decay length. In order to get the EMT interatomic potential for CuZr, we

need to fit these seven parameters for each element. The parameters for Cu

have been calculated previously [32], so that we may simply calculate the

parameters for Zr, while keeping the ones for Cu fixed. We do this fitting by

minimizing an error function

f({p}) =∑

i

qi · |CiEMT ({p})− Ctarget

i |2, (4.1)

where {p} = {s0, n0, Eo, V0, η2, λ, κ} is the list of parameters that need to

be fitted, qi=1 are normalizing constants, and CiEMT and Ci

target are the

EMT values (obtained from molecular dynamics) and target values (from

experiment or from DFT calculations), respectively. The sum goes over the

different properties used to fit the potential, such as lattice constants, cohe-

sive energies, and elastic constants for Cu, Zr and the CuZr alloy in the B2

(or CsCl) structure, as shown in table 4.1.

The target values were obtained from quantum mechanical calculations us-

ing DFT within the generalized gradient approximation (DFT-GGA), with

the Perdew-Wang 91 exchange-correlation functional [43]. Details about the

method may be found in chapter 2. Calculations were performed with the

DACAPO program from the CAMd Open Software project [47]. The usual

care was taken to ensure convergence with respect to the plane wave cutoff,

Page 47: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

36 Simulations of Mechanical Properties of Metallic Glasses

Figure 4.1: Comparison of experimental and calculated heats of formations

for binary alloys. All cases where the experimentally observed structure

corresponds to the calculated one. These are the fcc-based Cu3Au-type and

CuAu-type structures or the bcc-based CsCl-type structure. Reproduced

from Johannesson et al. [56].

number of k-points etc. The reason why we used data from DFT calculations

to fit our potential was that is has been shown previously that this calcula-

tion may accurately describe lattice properties [56], for example in Fig. 4.1

a good agreement is shown between the calculated and experimental results

for the heat of formation in the case of binary alloys.

For fitting the interatomic potential parameters, we use a downhill simplex

method [57] for optimizing the function f({p}) described above (4.1). This

method requires only evaluations of the functions and not their derivatives

when finding the minimum for more than one independent variable.

The result of the fit is shown in table 4.1, showing how well the target values

were fitted, while table 4.2 shows the resulting EMT parameters for Cu and

Zr.

4.1.1 Downhill simplex method in multi-dimensions

The Downhill Simplex method is a general algorithm for performing local

minimizations. The method only requires evaluations of the function and

Page 48: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 4. EMT INTERATOMIC POTENTIAL FOR CUZR 37

not its derivatives. This is done using a simplex.

A simplex is a geometrical figure consisting of N dimensions, N + 1 points

and all their interconnecting line segments, polygonal faces etc. A simplex

is a triangle in two dimensions and in three dimensions it is a tetrahedron.

Although this method requires a substantial amount of time to find these

minima, it is computationally efficient, stable, and easy to implement [58].

The local minimum is determined by performing a set of operations on the

simplex (4.2) according to a set of rules. The Downhill Simplex method

takes a number of steps, most of them moving the point of the simplex

where the function has the highest value through the opposite face of the

simplex to a lower point (this steps is called reflection). When it can do

so, the method expands the simplex in one direction to take larger steps.

When it can no longer reflect or expand, the method contracts itself in the

transverse direction, (this step is called contraction). When it cannot do any

of the operations described above, it means that the function is close to a

local minimum and the method contracts itself around it.

At the algorithm’s completion, the simplex is contracted around the local

minimum, as described in Fig. 4.2.

4.2 Testing the EMT interatomic potential

Before using the potential obtained in the previous sections, it is important

to test its ability to predict properties not used in the fitting procedure. Since

the intended application is CuZr metallic glasses, the potential is both tested

on ordered alloys with the same composition, and on thermodynamical and

structural properties of the metallic glass.

4.2.1 Formation energies and lattice constants for CuZr

crystalline structures. DFT versus EMT

The first test consists of comparing the formation energies and lattice con-

stants (here expressed as a volume per atom) for four different ordered alloys.

Page 49: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

38 Simulations of Mechanical Properties of Metallic Glasses

Figure 4.2: Possible operations for a step in the simplex method. Here the

simplex is a tetrahedron. The operations on a simplex can be (a) a reflection

away from a high point, (b) a relaxation and expansion away from the high

point, (c) a contraction along one dimension from the high point, or (d) a

contraction along all dimensions towards the low point.

The stable crystal structure for Cu50Zr50 is the B2 (or CsCl) structure (see

Fig. 4.3, but the other structures we will consider, the B1, L10, and L11, as

depicted in Fig. 4.3, do not exist in nature.

However, the other structures may be set up in the computer, and the ener-

gies and lattice constants can be compared to more accurate DFT calcula-

Page 50: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 4. EMT INTERATOMIC POTENTIAL FOR CUZR 39

a) NaCl (B1) Structure b) CsCl (B2) Structure

c) CuAu (L10) Structure d) CuPt (L11) Structure

Figure 4.3: Crystal Structures.

tions. The results are given in table 4.3. If the potential is able to describe

low-energy ordered structures, it is likely to correctly describe the local struc-

tures occurring in a metallic glass as well. All the structures except the L11

structure are well described. The L11 structure consists of alternating close-

packed layers of Cu and Zr. Presumably structure is far from what one will

encounter in metallic glasses, and the model’s inability to reproduce that

formation energy is of little concern.

4.2.2 MD and glass transition

A more relevant test is to examine how this potential performs for a metallic

glass. Therefore we prepared a CuZr metallic glass in order to compare

the glass transition temperature and the short-range order by the radial

distribution function to experiment.

We performed MD simulations in order to obtain the CuZr glass phase from

Page 51: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

40 Simulations of Mechanical Properties of Metallic Glasses

Structure Eform (eV/atom) V (A3/atom)

Cu50Zr50 B1 (DFT-GGA) 0.275 19.57

(EMT) 0.203 18.60

Cu50Zr50 B2 (DFT-GGA) -0.1202 17.64

experiment[54] -0.14 —

(EMT) -0.139 16.38

Cu50Zr50 L10 (DFT-GGA) -0.035 17.23

(EMT) -0.0011 17.23

Cu50Zr50 L11 (DFT-GGA) 0.250 18.01

(EMT) -0.1295 18.25

Table 4.3: Comparison between DFT and EMT results for different crystal

structures for Cu50Zr50. The B2 structure was used in the fitting of the

potential, the other structures are tests of the potential.

a) b)

Figure 4.4: CuZr glass system: a)2048 atoms and b) 42600 atoms.

the melt. We used two different system sizes, starting from body-centered

cubic crystals of 2048 and 42600 atoms with the sites randomly occupied by

Cu and Zr. The resulting glassy systems were cube-shaped with side lengths

of 3.3 nm and 9.1 nm, respectively. The systems, shown in Fig. 4.4, were

cooled from 1400 K (well above the melt temperature) in steps of 25 K, with

Page 52: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 4. EMT INTERATOMIC POTENTIAL FOR CUZR 41

0 500 1000 1500Temperature (K)

3

3,5

4

4,5

5

C p (k B)

42600 atoms2048 atoms

0 500 1000 1500

0

0,1

0,2

0,3

0,4

0,5

H (e

V/a

tom

)

Figure 4.5: Specific heat versus temperature for two different sizes of CuZr

metallic glass system during annealing. Inset: Enthalpy versus temperature

for the 2048 atoms system.

an average cooling rate of 0.54 K/ps. Each temperature step begins with a

short period of Langevin dynamics [49] to thermalize the system to the new

temperature, followed by a simulation with constant (N, P, T ) dynamics,

using a combination of Nose-Hoover and Parrinello-Rahman dynamics [59,

60]. The prethermalization with Langevin dynamics is necessary to prevent

oscillations in the temperature otherwise seen when the temperature of a

Nose-Hoover simulation is changed abruptly. During the constant (N, P, T )

dynamics we recorded the averages of different quantities of interest. These

included the pressure, volume, kinetic energy, and potential energy of the

system, which were used to perform the thermodynamics analysis [32]. The

simulations were performed using the ASAP program from the CAMd Open

Software project [47].

When plotting the enthalpy versus temperature (inset of Fig. 4.5), we observe

a change in the slope near 600–700 K which indicates the glass transition.

Page 53: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

42 Simulations of Mechanical Properties of Metallic Glasses

Figure 4.6: Radial distribution functions use a spherical shell of thickness dr.

From this slope the heat capacity Cp versus temperature curve may be ob-

tained, as shown in Fig. 4.5, since the specific heat is the derivative of the

enthalpy with respect to temperature. The change in the heat capacity indi-

cates the glass transition, Tg. We extracted the glass transition temperature

by taking the centered derivative of Cp(T ), and choosing the maximum [32].

Tg was estimated to be 620 K, compared to the experimental value of 670

K [12]. This good agreement is encouraging, in particular when one con-

siders that no thermodynamical data or data on glasses was used to fit the

potential.

4.2.3 Radial distribution function and coordination num-

bers for CuZr metallic glass

The radial distribution function (RDF) analysis is a standard method for

obtaining information about the structure of systems, especially for liquids

and amorphous structures. We calculate the radial distribution function [63]

as:

g(r)dr =dNA

ρA · V(4.2)

where dNA is the number of atoms of type A inside a spherical shell of

thickness dr at a distance r from a chosen atom Fig. 4.6, ρA is the density of

A atoms in the considered system, and V is the volume of the shell between

Page 54: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 4. EMT INTERATOMIC POTENTIAL FOR CUZR 43

5 10r [Angstrom]

0

1

2

3

4

5

g(r)

CuCuCuZrZrZr

Figure 4.7: Radial distribution functions g(r), for CuCu, CuZr, and ZrZr

metallic glasses.

Composition Method R(A) R(A) R(A) N N N

CuCu CuZr ZrZr CuCu CuZr ZrZr

Cu50Zr50 This work 2.56 2.75 3.13 4.44 6.24 7.59

Cu50Zr50 EXAFS[61] 2.57 2.84 3.19 4.10 5.4 7.40

Cu50Zr50 XRD[62] 2.68 2.80 3.16 4.60 4.8 7.80

Cu46Zr54 Simul. [35] 2.67 2.78 3.22 3.20 7.6/6.5 9.10

Table 4.4: Average nearest neighbor distances and coordination numbers.

The first row is this work and the next two are experimental data obtained

by two different methods. The last row is from simulations by Duan et al.

performed with another potential and a slightly different composition. Since

Cu and Zr are present in unequal amounts, the number of Zr neighbors for

a Cu atom is different from the number of Cu neighbors for a Zr atom, and

two numbers are therefore given for N(CuZr).

Page 55: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

44 Simulations of Mechanical Properties of Metallic Glasses

r and r + dr, as shown in Fig. 4.6. The volume of the shell is then given by:

V =4

3π(r + dr)3 − 4

3πr3 ≈ 4πr2dr. (4.3)

If we consider the number of particles per unit volume to be ρ, then the total

number of atoms in the shell is 4πρr2dr, and the number of atoms in the

volume element varies as r2, so that∫ ∞

0

g(r)4πr2dr = N − 1 ≈ N. (4.4)

In Fig. 4.7 we show partial RDF gCu−Cu, gCu−Zr and gZr−Zr at the glass

transition temperature. The first peak of the RDF indicates the average

distance at which we find nearest neighbor atoms, with gCu−Cu and gZr−Zr

yielding the distance between like atoms, and gCu−Zr the distance between

unlike atoms. By integrating the RDF given in eqn. 4.4 until the first

minimum, we obtain the number of nearest neighbors atoms for Cu and Zr

atoms. The results are summarized in table 4.4 with comparison with other

experimental and simulation data. The values for the coordination numbers

are in good agreement with the experimental ones.

As a final test, we have compared the bulk and shear modulus with experi-

mental values. Our calculated values were obtained at zero temperature by

applying strain in the elastic limit and relaxing the structure at each strain

step. We obtain 108.69 GPa and 29.25 GPa, compared to the experimental

values 101.2 GPa and 31.3 GPa [12], respectively, with a deviation of less

than 8%.

Page 56: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Chapter 5

Shear Deformation

Plastic deformation of metallic glasses at room temperature occurs through

the formation and evolution of shear bands in which it is localized. This

chapter contains details about simple shear simulations performed in order to

investigate the formation of shear bands and to measure their widths. Shear

bands have been seen to form above a certain shear strain in both CuZr and

CuMg under shear deformation. Details about the simulation method and

results will follow in the next section.

The width of a shear band may be measured as the characteristic width of

the displacement profile as shown in Fig. 5.1. Experimentally, shear bands

are identified as sharp surface steps when they are observed in scanning elec-

tron and atomic force microscopies [64]. Transmission electron microscopy

(TEM) is used to measure the width of the shear bands, for example the

width of a shear band measured using TEM was found to be 20 nm in a

Pd80Si20 glass [65], 10-20 nm in a Fe40Ni40B20 [66] and even ≤ 10 nm in a

Zr56.3Ti13.8Cu6.9Ni5.6Nb5.0Be12.5 [67]. But there are difficulties in characteriz-

ing the width and the structure of the shear bands using a TEM [66]. Image

interpretation is complicated by projection effects and the surface relief of

the TEM foils due to the shear offset along the band [64].

Another way to study the formation and the properties of shear bands is

through molecular dynamics simulations. Molecular dynamics simulations

45

Page 57: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

46 Simulations of Mechanical Properties of Metallic Glasses

Figure 5.1: Schematic displacement profile (solid curve) across a shear band

between two rigid blocks of metallic glass moving relative to each other (ar-

rows). Structural changes are found in the shaded band where the shear and

the shear rate are highest. The profile of the temperature rise around the

band has a width much greater than the shaded band thickness. Reproduced

from Zhang and Greer [64].

of a Mg85Cu15 model binary glass structure and deformation mechanism do

show shear localization. The simulations were carried out in this case by

using an EMT interatomic potential. The width of the shear bands could

then be measured during a simulated uniaxial deformation [34]. The shear

bands were identified by examining for each atom the residual displacement

D2min [6]. The D2

min profile of the shear band has a width of 10 nm and is

independent of the strain rate [34].

Ogata et al. [68] used a Lennard Jones potential to simulate simple shear de-

formations on Cu57Zr43 . During shearing both shear localization and shear

band nucleation were observed. Fig. 5.2 shows the inelastic displacement of

each atoms in the x direction for a 532 A×532 A×32 Asample. The average

displacement in the x-direction shows that the localization is initiated at dif-

ferent z positions and for strains larger than 8%. Once the shear localization

takes place at a z position, the z position does not change in the next steps

[68]. Above 15% , the stress-strain curves become flat because all additional

work is dispersed by sliding of the shear bands [68].

Page 58: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 5. SHEAR DEFORMATION 47

Figure 5.2: Average atomic displacements in the x direction for 9% affine

strain averaged over different z-slices for a Cu57Zr43 glass sample with geom-

etry as shown in inset. Shear localizations are seen at strains higher than

8%. The localization is initiated at different z positions. Reproduced from

Ogata et al. [68].

5.1 Simple shear deformation of CuZr and

CuMg. Simulation method and results

To study the formation of shear bands, we deformed the metallic glasses

under simple shear, using the geometry illustrated in Fig. 5.3. The sample is

divided into three parts, two outer “walls” of 3 nm thickness, and a central

region. In the central region, the dynamics of the atoms is left unperturbed,

but in the two walls the center of mass motion is constrained, so that the

two walls move with a constant speed relative to each other. The atoms in

the walls are not held rigid, they can still move with respect to each other,

Page 59: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

48 Simulations of Mechanical Properties of Metallic Glasses

Figure 5.3: Geometry of the Cu50Zr50 and Cu15Mg85 glass samples undergo-

ing shear deformation using an EMT interatomic potential.

only the sum of their velocities is held fixed. By causing one wall to move

in the y-direction, a shear deformation with constant shear rate is applied

to the system. The system has periodic boundary conditions in the y and z

directions, but free boundaries in the x direction. The shear simulation are

done at constant temperature of 50K.

During the deformation, velocity profiles (average atomic velocities as a func-

tion of x-coordinate) are calculated. Both velocity profiles and plots of the

deformed system clearly show the formation of shear bands (see Fig. 5.4 ).

The width of the shear bands, evaluated from visual observations, are 5 nm

for the CuZr glass, and 10 nm for the CuMg glass. The latter is consistent

with the simulations of Bailey et al. [34]. A better estimate can be obtained

by fitting the velocity profiles to a Fermi function 1/(exp(x/α)+1) or to the

arctan(x/α) function, as explained in PaperII. At the same time, it has been

observed before that the shear bands tend to be thinner in harder alloys [64].

In Fig. 5.4 the temperature profile during shearing shows the temperature

increases in the shear band region. We also find that the profile of raised

temperature around the band has a width greater than the shear band width

[64].

Page 60: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 5. SHEAR DEFORMATION 49

-0.20

0.20.40.60.8

1

Ave

rage

Vel

ociti

es (

Ang

strom

/ fs

)

CuMgCuZr

-300 -200 -100 0 100 200 300x length ( Angstrom )

50

100

150

200

Tem

pera

ture

( K

)

CuZr

CuMg

Figure 5.4: Velocity and temperature profiles (average atomic velocities

and temperatures as a function of x-coordinate) for Cu50Zr50 and Cu15Mg85

glasses.

Page 61: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under
Page 62: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Chapter 6

Nanoindentation

Since nanoindentation is a good method for testing mechanical properties of

materials in small volumes (see chapter 1) we performed simulated nanoin-

dentation on Cu0.50Zr0.50 and Mg0.85Cu0.15 metallic glasses in order to study

the formation of shear bands. During nanoindentation, shear bands were

seen to form under the nanoindenter, their formation and propagation being

influenced by different parameters. These were indentation velocity, radius

of the indenter, cooling rate and material compositions, see Paper II. The

simulation method and the results will be discussed in this chapter. For

comparison we performed nanoindentation on a Cu crystal which will be

discussed in this chapter as well, see Paper II.

6.1 Simulation method

In 2005 Shi and Falk [28] proposed a model for simulating nanoindentation.

They performed simulated nanoindentation on a two dimensional model bi-

nary alloy. The alloy consists of two species called S and L for small and

large with composition NL : NS = (1 +√

5) : 4. The atoms interacted via a

Lennard-Jones potential of the form:

Φij = 4ε

[(σ

rij

)12

−(

σ

rij

)6]

, (6.1)

51

Page 63: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

52 Simulations of Mechanical Properties of Metallic Glasses

where ε is the bonding energy, σ provides a length scale and the bond energy

εSS = εLL = 1/2εSL. Details about the the choice of the system and the

Lennard Jones potential parameters can be found in [28, 6]. The samples

consisted of 200 000 atoms cooled with different cooling rates [28, 69]. The

indenter was modeled by imposing a purely repulsive potential of the form:

ΦIi = ε

[rIi −RI

0.6σSL

]−12

, (6.2)

where rIi is the distance between atoms i and the center of the spherical

indenter of radius RI (=75nm) [28]. In other simulations they have used

a shaped cylindrical indenter [69]. The indenter is modeled in such a way

that it has no surface friction and is lowered into the glassy sample under

displacement control as shown in Fig. 6.1.

Figure 6.1: Illustration of the nanoindentation testing geometry. Periodic

boundary conditions (PBC) apply on the x−y and y−z planes, as reproduced

from Shi and Falk [69].

In our case, samples of metallic glass were produced by cooling Cu0.50Zr0.50

and Mg0.85Cu0.15 systems from 1400 K (above the melting temperature) down

to 40 K, with a cooling rate of 0.54 K/ps. Details about the cooling and glass

forming procedure were given previously in chapter 3. The glassy systems

obtained have dimensions of 52.9 nm × 52.9 nm × 1.6 nm for CuZr and 57.2

nm ×57.2 nm × 1.7 nm for CuMg, both corresponding to 270000 atoms.

Page 64: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 6. NANOINDENTATION 53

Before nanoindentation, the samples are repeated in the x direction, giving a

CuZr sample size of 105.9 nm × 52.9 nm × 1.6 nm and a CuMg sample size

of 114.4 nm × 57.2 nm × 1.7 nm with periodic boundary conditions along

the x and z directions, and free boundaries in the y direction.

In order to perform simulated nanoindentation, we perform molecular dy-

namics at constant temperature using Langevin dynamics [49] at constant

temperature 116K, while the indenter is modeled as a repulsive potential as

proposed by Shi and Falk [28] and described previously. The form of our

repulsive potential is:

E = c(r −R)−6, (6.3)

where r is the distance from the center of the indenter and R is the indenter

radius. The strength parameter c = 2.0 eVA6 . The indenter is lowered into

the substrate under displacement control, recording at each step the load and

the total displacement. Since the samples are very thin in the z direction,

the simulation is quasi two-dimensional, and we choose to use a cylindrical

indenter. During indentation, the lower edge of the system is held fixed by

fixing the positions of the atoms in the bottom 0.8 nm of the system. The

thickness of the system is kept constant during the indentation (plane strain),

as discussed in Paper II.

To visualize the plastic flow and the localization of the shear bands we use

the D2min method introduced by Falk and Langer [6]. According to this

method, D2min is the local deviation from an affine deformation during a time

interval of time between a pair of configurations [t−∆t, t]. Specifically,D2min

represents the residual displacement differences unaccounted for by the best

fit strain matrix for an atom. With this quantity defined for all atoms, we

are doing a coarse graining of the system, dividing it into boxes within each

the average D2min is computed.

Another method that we used to visualize the plastic flow and the localization

of the shear bands is the so called “stripe painting” method [34]. In this case,

the undeformed configuration is taken and all atoms are binned according to

their y position, and those in odd bins are labeled “black”. This designation

is maintained throughout the deformation [34].

Page 65: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

54 Simulations of Mechanical Properties of Metallic Glasses

6.2 Shear band formation

A nanoindentation simulation is performed on the CuZr sample as described

above. The cylindrical indenter with a diameter of 30 nm, is indented into

the sample with a velocity of 7 m/s, as illustrated in Fig. 6.3.

b) c)

a)

Figure 6.2: Shear band visualization using a) D2min method, b) and c)“stripe

painting” method. We highlight the shear bands location observed in b) and

c) by blue lines which match the more clearly visible shear bands seen in a).

As the shear is inhomogeneous, it is not possible to rigorously define an

equivalent shear rate.However, it corresponds roughly to a shear rate on the

order of 108 s−1 (a homogeneous shear rate of 108 s−1 would correspond to

the ends of the sample moving with a relative velocity of 10 m/s).

We visualized the plastic flow under the indenter with the D2min method and

“stripe painting” method, both described previously. As it can be observed

from Fig. 6.2, D2min is a better method to visualize the plastic flow for

this kind of deformation. However, the “stripe painting” method confirms

that atom shearing is happening, and the regions where atoms are shearing

correspond with the regions where shear bands are identified by D2min.

Page 66: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 6. NANOINDENTATION 55

10nma)

10nmc) d)

b) 10nm

10nm

0 10 20 30 40 50Displacement (Angstrom)

50

100

150

Load

(eV

/Ang

strom

)

Indentation velocity 30m/sIndentation velocity 7m/s

e)

a)

b)

c)

d)

Figure 6.3: Development of shear bands during nanoindentation on a Cu-Zr

thin sample: a-d) visualization of strain localization at a 24 A, 32 A, 42

Aand 52 Adepth of the indenter, the indentation velocity was 7 m/s; e) The

lower curve is the corresponding load displacement curve, the displacements

corresponding to the four panels are shown with arrows. It is seen that the

shear bands are forming gradually, there is no correspondence between events

on the indentation force curve and formation of shear bands. The upper curve

is a similar indentation when the indentation velocity is increased to 30 m/s,

as reproduced from Paper II.

During nanoindentation, the net force on the indenter is recorded as a func-

tion of indentation depth, see figure 6.3. The curve is not smooth, but

contains small serrations, presumably reflecting individual events during de-

Page 67: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

56 Simulations of Mechanical Properties of Metallic Glasses

formation, as discussed in Paper II. It would be tempting to interpret this

as the formation of individual shear bands, but that interpretation is not

supported by direct visualization of the shear band formation. As is seen in

figure 6.3, the shear bands are not formed at once, but grow gradually during

the nanoindentation.

The shear bands are less clearly formed than in the simulations of Shi and

Falk [28]. The simulations cannot, however, be directly compared, as the

simulations of Falk and Shi are two dimensional and use a Lennard-Jones

model potential, whereas our simulations have a finite (but small) thickness,

and use a more realistic (but also complex) potential. The more complicated

potential combined with the larger number of atoms due to the finite thick-

ness, leads to a much larger computational burden, so that our indentations

are therefore necessarily done with higher indentation velocities. There are

thus three major differences between these simulations and the simulations

by Shi and Falk. These are in dimensionality, the potential and the indenta-

tion speed, and it is not surprising that our results are significantly different.

Shi and Falk lower their indenter into the samples with a velocity of 0.3 m/s

with the samples called Sample I, Sample II and Sample III according to the

cooling rates they have obtained.

The cooling rates they use are ≈ 0.06×10−2 K/ps for Sample I, ≈ 0.03 K/ps

for Sample II. The third sample was cooled instantaneously by rescaling the

particle velocities and then allowing them to age for 100t0 (t0 = 1ps).

Figure 6.4 shows the plastic flow under the indenter for samples I, II and

III. In Sample I a few shear bands carry the plastic flow while in Sample

II, the majority of the deformation is happening in the region immediately

beneath the indenter and the plastic flow is more evenly distributed with

smaller spacing between the bands. Sample III exhibits deformation at a

finer scale then the other samples, with shear band spacing less than half

that in Sample I.

Page 68: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 6. NANOINDENTATION 57

Figure 6.4: Visualization of the magnitude of the local deviatoric shear strain

under the indenter at maximum depth in three samples produced at different

quench rates [28]. Dark regions denote regions of high strain. The scale bar

in the upper right of the image is approximately 25 nm, as reproduced from

[28].

6.3 Influences on shear band formation

In this section we study how the shear band formation is influenced by vary-

ing the parameters of the nanoindentation. This corresponds to the varying

of physical parameters during a nanoindentation experiment. We vary the

indenter velocity, its radius, the cooling rate during glass formation and the

composition of the glass. All modifications are relative to a “standard sim-

Page 69: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

58 Simulations of Mechanical Properties of Metallic Glasses

ulation” of a CuZr metallic glass produced at a cooling rate of 0.44 K/ps,

indented with a cylindrical indenter with radius 30 nm and indentation ve-

locity 7 m/s. Fig. 6.5 shows how the shear band formation is influenced by

the indentation velocity. Here, a simulation with an indentation velocity of

30 m/s is compared with the reference simulation. The shear band formation

is less obvious in the fast indentation. A clear indentation rate dependency

of the load displacement curve is also seen. The loading force - and thus

the stress in the material - is significantly higher during the fast simulation,

but no qualitative differences are seen. It can be speculated that even lower

indentation rates will lead to the formation of fewer and larger shear bands,

and possibly lead to macroscopic failure through a few very large shear bands

[70].

a) b)10nm 10nm

Figure 6.5: Influence of varying the indentation velocity on shear band for-

mation and on load displacement curves. The indentation velocities are 30

m/s (panel a) and 7 m/s (panel b). The shear bands are less developed dur-

ing the fast nanoindentation. The corresponding load displacement curves

are shown in Fig. 6.3e, as reproduced from Paper II.

Schuh et al. studied the strain rate influence on the serrated flow during

indenting for four different metallic glass samples [70]. Serrated flow was

found to be a strong function of the rate of deformation, with more rapid

indentations suppressing - and slower indentations promoting - prominent

Page 70: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 6. NANOINDENTATION 59

serrations [70]. They also identified for each alloy a critical indentation strain

rate above which load serrations apparently do not occur, as shown in Fig.

6.6.

Figure 6.6: Load displacement (P − h) curves measured on the loading por-

tion of nanoindentation experiments, for four BMG composition investigated

(Pd-40Ni-20P, Pd-30Cu-10Ni-20P, Zr-10Al-10Ni-15Cu and Zr-10Al-14.6Ni-

17.9Cu-5Ti). Curves are offset from the origin for clear viewing, and the

rates are specified in each graph, as reproduced from[70].

Jiang et al. [71] performed nanoindentation on Al90Fe5Gd5 at different de-

Page 71: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

60 Simulations of Mechanical Properties of Metallic Glasses

formation rates. Their study shows that the deformation rate affects the

number of shear bands, as shown in Fig. 6.7, and consequently the serrated

flow behavior of amorphous alloys. Their observations outside of the indenter

are consistent with other studies of strain rate dependence made in tensile

tests [72, 73]. They conclude that a lower strain rate is required for the

contribution from an individual shear band to resolved [71].

Figure 6.7: AFM illumination images of indents produced at loading rates

of (a) 100 nm/s and (b) 1 nm/s, reproduce from [71].

Tensile tests demonstrated that at low strain rates, failure occurred along a

Page 72: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 6. NANOINDENTATION 61

a) b)10nm 10nm

Figure 6.8: Influence of varying the indenter radius. The radii are 20 nm

(a) and 30 nm (b). More shear band formation is observed under the larger

indenter, as reproduced from Paper II.

single shear band, whereas many intersecting bands were involved in fracture

at high strain rates [72]. The tensile tests of Sergueeva et al. [73] showed

that at relatively low strain rates there is enough time for one or two main

shear bands to form. These may accommodate the deformation through the

production of main and secondary shear bands in a very small volume near

the fracture surface. This accommodation becomes impossible as the strain

rates increases above some critical value, and additional shear bands must

be formed in order to accommodate the strain. As a result, in this case they

observed multiple shear banding in the whole volume [73].

In spite of the indenter diameter being an order of magnitude larger than the

width of the shear band, a difference is also seen when the indenter radius

is varied. This is illustrated in figure 6.8, where the effect of reducing the

indenter radius by 33% is studied. As is seen, fewer shear bands are formed

when we compare situations with the same indentation depth. It is likely

that this difference is at least partly caused by less material being displaced

by the smaller indenter.

The effect of changing the cooling rate during the glass formation is inves-

tigated in figure 6.9, where identical nanoindentation simulations are per-

Page 73: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

62 Simulations of Mechanical Properties of Metallic Glasses

a) b)10nm 10nm

Figure 6.9: Influence of varying the cooling rate during glass formation.

Samples created with cooling rates of 2.2 K/ps (a) and 0.44 K/ps (b) were

subjected to otherwise identical nanoindentation simulations. The shear lo-

calization appears stronger in the slowly cooled sample, as reproduced from

Paper II.

formed on two samples produced with cooling rates of 2.2 K/ps and 0.44

K/ps. In this case, the difference is less pronounced, but the shear bands

appear more distinct in the slowly cooled sample. This is consistent with re-

sults by Shi and Falk [28] showing a stronger tendency for shear localization

for more gradually quenched samples, as shown in Fig. 6.4.

Finally, we investigated the consequences of replacing the CuZr glass with

a CuMg glass (Fig. 6.10). Shear localization is far more pronounced in

the CuZr sample. This is consistent with our previous findings that shear

bands are approximately twice as wide in a CuMg glass. We also observed

previously from our simulations that if the sample size and/or indenter radius

are not big enough with respect to the shear band width, the clear formation

of shear bands can not be observed. In this case, this may be the reason why

we did not see the same clear shear localization in CuMg sample as we did

in CuZr.

Page 74: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 6. NANOINDENTATION 63

a) b)10nm 10nm

Figure 6.10: Comparison of CuMg (a) and CuZr metallic glass under oth-

erwise identical nanoindentation conditions. The latter shows the strongest

shear localization, as reproduced from Paper II.

6.4 Nanoindentation of crystalline samples

For comparison, we subjected a nanocrystalline sample to indentation in a

similar manner. Nanoindentation of nanocrystalline samples has previously

been studied by Van Swygenhoven et al. [74, 75], with a particular focus on

the onset of plastic deformation under a spherical indenter. In this work, we

have chosen to mimic the indentation parameters used for the metallic glass,

so the nanocrystalline sample is quasi-two dimensional, with 〈110〉-directions

along the thin direction (the z direction). This orientation is chosen to maxi-

mize the number of possible slip systems in the crystal, which is nevertheless

still significantly lower than in a full three-dimensional simulation [76]. The

material for this simulation is copper, described with the EMT potential [22].

Details about the sample geometry and simulations can be found in Paper

II.

Figure 6.11 shows the deformation during this nanoindentation, and Fig 6.12

shows the corresponding load displacement curves. A number of notewor-

thy features are marked by letters on Fig. 6.11b. Right below the indenter

(marked “A”), the plastic deformation is severe, and several grain bound-

Page 75: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

64 Simulations of Mechanical Properties of Metallic Glasses

A

B

C D

E

F

(a)

(b)

c)

10nm

Figure 6.11: Nanoindentation of nanocrystalline copper. (a) The initial con-

figuration before indentation. Atoms in the grain interiors are not show,

atoms at stacking faults and twin boundaries are colored red, atoms in grain

boundaries, dislocation cores and at surfaces are blue. Both high-angle and

low-angle grain boundaries are seen, the former a shown as blue walls, the

latter are seen to be made of individual dislocations. A few stacking faults

created by dislocation activity during annealing are also seen. (b) After

nanoindentation, significant plastic deformation has occurred, in particular

just under the indenter, where numerous stacking faults and deformation

twins have appeared. The letters A-F mark features discussed in the text.

(c) A deformation map shows that the majority of the deformation occurs

inside the grains and is caused by dislocation activity, although some grain

boundary sliding is also seen. Reproduced from Paper II.

aries have disappeared. These are low-angle grain boundaries, which can be

regarded as dense arrays of dislocations. The high stress below the indenter

Page 76: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 6. NANOINDENTATION 65

0 10 20 30 40 50Displacement (Angstrom)

0

50

100

150

200

250

300

350

Load

(eV

/ A

ngstr

om)

a b

10nma)

10nmb)

Figure 6.12: Load displacement curve during nanoindentation of the crys-

tal. The insets marked a and b show the deformation occurring in the two

intervals between the arrows. Reproduced from Paper II.

cause these to move. Many stacking faults are seen, which are caused by

the high stress levels. This makes making the energy of the stacking fault

irrelevant compared to the elastic energy. Another example of a low angle

grain boundary disappearing as a result of dislocation migration is marked

“B”.

A few grain boundaries “reconstruct” under the applied stress by emitting an

array of partial dislocations (for example at “C”). Presumably, a lower grain

boundary energy can be obtained and some of the applied stress relieved, by

the motion of the dislocation wall at the expense of creating some stacking

faults. A similar phenomenon is seen at “D”, where a wall of full dislocations

has been emitted by the grain boundary. An extreme case is seen at “E”,

Page 77: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

66 Simulations of Mechanical Properties of Metallic Glasses

where a grain boundary has split into two separated by a region of (slightly

faulted) HCP crystal. Similar reconstructions of grain boundaries under

stress have previously been seen in computer simulations [77]. At “F” a

stacking fault has been removed by the motion of a partial dislocation, and

a new one has been created nearby by an independent event, creating the

illusion that a stacking fault has moved.

The events discussed above can also be found in the displacement map in

Fig. 6.11c. The figure shows a striking difference between the deformation

distributions in the glass and in the nanocrystalline sample: in the crystalline

sample the spatial distribution of the deformation is much more uneven, as

it is controlled by the polycrystalline structure. Deformation can only occur

at grain boundaries, or where dislocations are available, typically from grain

boundary sources.

The load-displacement curve in Fig. 6.12 shows significantly larger serrations

than for the glassy samples. This is caused by avalanches of dislocation

activity. The insets show how the deformation occurs just before and during

one of the serrations. It is clearly seen that the serration does not correspond

to a single dislocation event, but rather to an avalanche of events.

6.5 Nanocrystals embedded In glass

Cast Cu50Zr50 alloy rods with a diameter of 1 mm have been found to consist

of a glassy phase containing fine crystalline particles with a size of about

5nm [78]. These rods were found to exhibit a high yield strength under

uniaxial compression tests as shown in Fig. 6.13. A scanning tunneling

electron microscopy (STEM) image obtained by Inoue et al. after the uni-

axial compression test shows that the crystalline phase no longer consists of

spheroidal particles, as it did initially in the undeformed sample. The parti-

cle width is elongated, which suggest that the elongated particles are formed

by crystalline particle agglomeration. This means that under shear stress the

crystallites can come into contact [78, 79], as shown in Fig. 6.14.

In glassy CuZr alloys, the hardness increases with copper content [78]. It is

Page 78: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 6. NANOINDENTATION 67

Figure 6.13: True stress-strain curve for 1mm diameter rod of as-cast

Cu50Zr50 loaded in uniaxial compression. To avoid damage to the testing

machine, the test was stopped before sample fracture, at a compressive strain

of 52%. The inset shows scanning electrom microscopy (SEM) images of the

sample (a) before testing and (b) after a compressive strain of 52%. Repro-

duced from [78].

likely that copper-rich nanocrystals, ideally dislocation free, are harder than

the zirconium-enriched matrix. In such a case the size and the agglomeration

of crystallites can affect the plastic flow. As a consequence, shear bands, in

which there has already been some crystalline particles agglomeration can

become more resistant to deformation than the neighboring material [78].

We performed nanoindentation on a CuZr metallic glass obtained through the

same method explained in section 5.1. However, after cooling the samples

we embed Cu nanocrystals in the glassy matrix as shown in Figs. 6.15a)

and 6.16a). The indented samples contain nanocrystals of different size. We

will call these samples A and B consisting of 547583 and 1096945 atoms

respectively. Sample A has dimensions of 105.9 nm × 52.9 nm × 1.6 nm

containing 11 small nanocrystals of 3.5 nm × 3.5 nm × 1.6 nm. Sample B’s

geometry is 105.9 nm ×105.9nm × 1.6 nm containing 1 nanocrystal of 20

Page 79: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

68 Simulations of Mechanical Properties of Metallic Glasses

Figure 6.14: The proposed mechanism of shear-induced coalescence of

nanocrystallites. In the undeformed material (a) the crystallites are self

avoiding. In an operating shear band (b) the relatively hard crystallites can

be brought into contact in the matrix of lowered viscosity. Sintering of the

crystallites (c) can occur in a limited time, facilitated by the nanometer scale.

Reproduced from [78].

Page 80: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 6. NANOINDENTATION 69

nm × 20 nm × 1.6 nm. The indentation is performed following the same

procedure described in section 3.

Our results show that in both samples the shear bands cannot propagate

through the nanocrystals, but are instead pinned by them. A consequence

may be that the presence of the nanocrystals provides a mechanism of strain

hardening and may assist delocalization of the plastic flow around and be-

tween the nanocrystals as shown in Fig. 6.15c) and Fig. 6.16c). Though

the nanocrystals in sample A are smaller than the width of the shear bands

and they are pushed into the sample during deformation we cannot clearly

observe agglomeration of the nanocrystals (see Fig. 6.15 b) and c) as re-

ported for experimental uniaxial deformations [78, 79]. In the case of sample

B, the nanocrystal is larger than the width of the shear bands in CuZr. As a

consequence of size, the nanocrystal itself behaves as a small nanoindenter,

and beneath it we can observe activation of shear bands while it is pushed

into the glassy matrix during the deformation process.

Page 81: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

70 Simulations of Mechanical Properties of Metallic Glasses

Figure 6.15: Nanoindentation on sample A (CuZr glass sample containing

11 fine crystalline particles with a size of 3.5 nm × 3.5 nm × 1.6 nm); a)

undeformed sample, b) indented sample with a velocity of 7m/s at maximum

depth of 60A; c) visualization of strain localization under the indenter using

D2min method described previously.

Page 82: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

CHAPTER 6. NANOINDENTATION 71

Figure 6.16: Nanoindentation on sample B (CuZr glass sample containing a

crystalline particle with size of 20 nm × 20 nm × 1.6 nm); a) undeformed

sample, b) indented sample with a velocity of 7m/s at maximum depth of

60A; c) visualization of strain localization under the indenter using D2min

method described previously.

Page 83: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under
Page 84: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Chapter 7

Conclusions

The purpose of this thesis was to model and study the mechanical properties

of CuZr and CuMg metallic glasses and potentially to contribute to a better

understanding of this properties. With the hope that we have succeed to do

that we will conclude now our main results.

A new EMT interatomic potential was constructed for CuZr, with the pur-

pose of simulating CuZr bulk metallic glass. The potential was fitted to

lattice constants, cohesive energies and elastic properties of elemental Cu,

elemental Zr and Cu50Zr50 in the B2 crystal structure. The fitted poten-

tial does a reasonable job of reproducing lattice constants and energies of

other structures of Cu50Zr50. When a glassy sample of Cu50Zr50 is produced,

we find the glass transition temperature, elastic constants and structural

data in form of average coordination numbers and average nearest neighbor

interatomic distances in good agreement with experiment. Previously, an

interatomic force field has been fitted to CuZr by Duan et al. [35] but we

find that our new EMT interatomic potential described in chapter 4 provides

a better description of the structure of the metallic glass.

Using EMT interatomic potential for CuZr, we have investigated shear band

formation during homogeneous shear and during nanoindentation in CuZr

and MgCu-based metallic glasses. The studied samples are quasi two-dimen-

sional. In homogeneous shear, we observe formation of shear bands of 5 and

8 nm in thickness for CuZr and MgCu-based glasses, respectively.

73

Page 85: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

74 Simulations of Mechanical Properties of Metallic Glasses

During nanoindentation, we see a clear signature of shear band formation

under the indenter, although the shear bands are not as well developed as in

two-dimensional Lennard-Jones simulations [28]. However, a direct compar-

ison is not possible due to the computational costs of the EMT potentials

which does not allow us to use the same low velocities and cooling rates.

The shear bands are seen to form gradually during indentation. Shear band

formation is stronger if the indentation velocity is low, if the indenter is large

and if the sample is cooled slowly. In addition, shear bands were more clearly

visible clearer in CuZr than in MgCu under otherwise identical conditions,

probably due to the natural width of the shear bands in MgCu being larger.

Simulations of nanoindentation in nanocrystalline copper under otherwise

identical conditions show a significantly different deformation mechanism,

where individual dislocations strongly influence the deformation pattern.

When indenting glassy samples containing nanocrystals we observed that

the shear bands cannot propagate through the nanocrystals, being pinned

by them. This may lead to a mechanism of hardening.

Page 86: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Bibliography

[1] A. Greer, “Metallic glasses”, Sci. 267, 1947 (1995).

[2] A. Inoue, “Stabilization of metallic supercooled liquid and bulk amor-

phous alloys”, Acta Mater. 48, 279 (2000).

[3] J. Schaffer, A. Saxena, S. Antolovich, T. S. Jr., and S. Warner, The

science and design of engineering materials, 2nd ed. (McGraw-Hill In-

ternational, USA, 1999).

[4] A. L. Greer, “Metallic glasses”, Sci. 267, 1947 (1995).

[5] A. S. Argon, “Plastic deformation in metallic glasses”, Acta Metall. 27,

47 (1979).

[6] M. L. Falk and J. S. Langer, “Dynamics of viscoplastic deformation in

amorphous solids”, Phys. Rev. E 57, 7192 (1998).

[7] A. Ibenskas, C. E. Lekka, and G. A. Evangelakis, “Structural, thermo-

dynamic and mechanical properties of Zr-based binary nanowires (ZrCu

and Zr2Ni) by molecular dynamics”, Physica E 37, 189 (2007).

[8] C. E. Lekka, A. Ibenskas, A. R. Yavari, and G. A. Evangelakis, “Tensile

deformation accommodation in microscopic metallic glasses via subnan-

ocluster reconstructions”, Appl. Phys. Lett. 91, 214103 (2007).

[9] D. L. Malandro and D. J. Lacks, “Relationships of shear-induced changes

in the potential energy landscape to the mechanical properties of ductile

glasses”, J. Chem. Phys. 110, 4593 (1999).

75

Page 87: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

76 Simulations of Mechanical Properties of Metallic Glasses

[10] C. A. Schuh and A. C. Lund, “Atomistic basis for the plastic yield

criterion of metallic glass”, Nat. Mater. 2, 449 (2003).

[11] A. C. Lund and C. A. Schuh, “Yield surface of a simulated metallic

glass”, Acta Mater. 51, 5399 (2003).

[12] J. Das, M. B. Tang, K. B. Kim, R. Theissmann, F. Baier, W. H. Wang,

and J. Eckert., “Work-hardenable ductile bulk metallic glass”, Phys.

Rev. Lett. 94, 205501 (2005).

[13] D. Bae, S. Lee, J. Kwon, X. Wang, and S. Yi, “Ductile Zr-base bulk

metallic glasses”, Mat. Sci. Eng. A 449-451, 111 (2007).

[14] M. Ashby and A. Greer, “Metallic glasses as structural materials”, Sci.

76, 62 (1996).

[15] Http://www.liquidmetal.com/applications/.

[16] P. Duwez, W. Klement, and R. H. Willens, “Continuous series of

metastable solid solutions in silver-copper alloys”, J. Appl. Phys. 31,

1136 (1960).

[17] H. Jones, “A perspective on the development of rapid solidification and

nonequilibrium processing and its future”, Mat. Sci. Eng. A304-305,

11 (2001).

[18] V. Jhon, Introduction to engineering materials, 4th ed. (Palgrave, USA,

New York, 2003).

[19] Http://www.instron.us/wa/applications.

[20] R. Vaidynathan, M. Dao, G. Ravichandran, and S. Suresh, “Study of

mechanical deformation in bulk metallic glasses through instrumented

indentation”, Acta Mater. 49, 3781 (2001).

[21] K. W. Jacobsen, J. K. Nørskov, and M. J. Puska, “Interatomic interac-

tions in the effective-medium theory”, Phys. Rev. B 35, 7423 (1987).

[22] K. Jacobsen, P. Stoltze, and J. Nørskov, “A semi-empirical effective

medium theory for metals and alloys”, Surf. Sci. 366, 394 (1996).

Page 88: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

BIBLIOGRAPHY 77

[23] N. Chetty, K. Stokbro, K. W. Jacobsen, and J. K. Nørskov, “Ab initio

potential for solids”, Phys. Rev. B 46, 3798 (1992).

[24] K. W. Jacobsen, “Bonding in metallic systems: An effective-medium

aproach”, Coments on Cond. Mat. Phys 14, 130 (1988).

[25] A. Evteev, A. Kosilov, and E. Levtchenko, “Atomic mechanism of for-

mation and structure relaxation of Fe 83M 17 (M : C, B, P) metallic

glasses”, Acta Mater. 51, 2665 (2003).

[26] R. Frattini and R. D. Valle, “Molecular dynamics simulation of glassy

CuY”, Phys. Rev. B 6, 3620 (1995).

[27] F. Albano, N. Lacevic., M. Falk, and S. Glotzer, “Relating metallic glass

mechanical properties to liquid structure”, Mater. Sci. Eng. 375, 671

(2004).

[28] Y. Shi and M. L. Falk, “Structural transformation and localization

during simulated nanoindentation of nanocrystalline metal film”, Appl.

Phys. Lett. 86, 011914 (2005).

[29] M. S. Daw and M. I. Baskes, “Embedded-atom method: derivation and

application to impurities, surfaces, and other defects in metals”, Phys.

Rev. B 29, 6443 (1984).

[30] H. Gong, L. Kong, W. Lai, and B. Liu, “Glass forming ability de-

termened by an n-body potential in a highly immiscible CuW system

through molecular dynamics simulations”, Phys. Rev. B 68, 1442011

(2003).

[31] L. Qi, H. Zhang, and Z. Hu, “Molecular dynamics simulation of glass

formation in binary metals: Cuag using eam”, Intermetal. 12, 1192

(2004).

[32] N. P. Bailey, J. Schiøtz, and K. W. Jacobsen, “Simulation of cu-mg

metallic glass: Thermodynamics and structure”, Phys. Rev. B 69,

144205 (2004).

Page 89: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

78 Simulations of Mechanical Properties of Metallic Glasses

[33] N. P. Bailey, J. Schiøtz, and K. W. Jacobsen, “Atomistic simulation

of Mg-Cu metallic glasses: mechanical properties”, Mater. Sci. Eng. A

387, 996 (2004).

[34] N. P. Bailey, J. Schiøtz, and K. W. Jacobsen, “Atomistic simulation

study of the shear band deformation mechanism in Mg-Cu metallic

glass”, Phys. Rev. B 73, 064108 (2006).

[35] D. et al., “Molecular dynamics study of the binary Cu46Zr54 metal-

lic glass motivated by experiments: Glass formation and atomic level

structure).”, Phys. Rev. B 71, 224208 (2005).

[36] H. Chen and Y. Waseda, “Structure of glassy ZrCu and NbNi alloys”,

Phys.Status Solidi A 51, 593 (1979).

[37] A. Sadoc, Y. Calvayrac, A. Quivy, M. Harmelin, and A. Flank, “Study

of the local structure of CuZr amorphous alloys by EXAFs. effect of a

structural relaxation”, J. Non.Cryst. Solids 65, 109 (1984).

[38] Y. Qi, T. Cagin, Y. Kimura, and W. G. III, “Molecular dynamics simula-

tions of glass formation and crystalizationin binary liquidmetals: CuAg

and CuNi”, Phys. Rev. B 59, 3527 (1999).

[39] T. Aihara, Y. Kawazoe, and T. Masumoto, “Molecular dynamics simu-

lations for binary amorphous ZrNi alloys”, J. Non-Cryatalline Sol. 205,

875 (1996).

[40] C. Hausleitner and J. Hafner, “Structural and electronic properties of

Y-based metallic glasses”, J. Phys.: Condensed Matter 4, 9557 (1992).

[41] H. Teichler, “Mode coupling theory and the glass transition in molecular

dynamics simulated NiZr”, Phys. Rev. Lett. 76, 62 (1996).

[42] J. Rose, J. Ferrante, and J.R.Smith, “Universal binding energy curves

for metals and bimetallic interfaces”, Phys. Rev. Lett. 47, 675 (1981).

[43] J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson,

D. J. Singh, and C. Fiolhais, “Atoms, molecules, solids, and surfaces:

Page 90: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

BIBLIOGRAPHY 79

Applications of the generalized gradient approximation for exchange and

correlation”, Phys. Rev. B 46, 6671 (1992).

[44] L. Verlet, “Computer “experiments” on clasical fluids I. Thermodynam-

ical properties of Lenard Jones molecules”, Phys. Rev. 159, 98 (1967).

[45] D. Frenkel and B. Smit, Understanding molecular simulations, 2nd ed.

(Academic Press, London, 2002).

[46] M. Ferrario, G. Ciccotti, and K. Binder, Computer simulations in con-

densed Matter: From Materials to Ch emical Biology, 1st ed., vol. 1

(Springer, Berlin, 2006).

[47] The CAMd Open Software project, http://www.camp.dtu.dk/Software.

[48] P. Stoltze, Simulation methods in atomic scale material physics, 1st ed.

(Polyteknisk Forlag, Lyngby, 1997).

[49] M. Allen and D. Tildesley, Computer simulations of liquids, 1st ed.

(Clarendon Press, Oxford, 1987).

[50] D. H. Xu, B. Lohwongwatana, G. Duan, W. L. Johnson, and C. Garland,

“Bulk metallic glass formation in binary cu-rich alloy series Cu100−xZrx

(x = 34, 36, 38.2, 40 at.) and mechanical properties of bulk Cu64Zr36

glass”, Acta Mater. 52, 2621 (2004).

[51] R. Ray, B. C. Giessen, and N. J. Grant, “New non-crystalline phases in

splat cooled transition metal alloys”, Scripta Metall. 2, 357 (1968).

[52] J. Schiøtz and K. W. Jacobsen, “A maximum in the strength of

nanocrystalline copper”, Sci. 301, 1357 (2003).

[53] Http://www.webelements.com/.

[54] A. I. Zaitsev and N. E. Zaitseva, “Thermodynamics and amorphization

of the Cu-Zr melt”, Doklady Physical Chemistry 386, 214 (2002).

[55] I. Ansara, A. Pasturel, and K. Buschow, “Density of states, composi-

tional short range order and stability of amorphous CuxZr1−x alloys”,

Phys. Status. Solidi. 69, 7441 (1982).

Page 91: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

80 Simulations of Mechanical Properties of Metallic Glasses

[56] G. Johannesson, T. Bligaard, A. Ruban, H. Skriver, K. Jacobsen, and

J. Nørskov, “Combined electronic structure and evolutionary search ap-

proach to materials design”, Phys. Rev. Lett. 88, 255506 (2002).

[57] W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery,

Numerical recipes in C++, 2nd ed. (Cambridge University Press, Cam-

bridge, 2002).

[58] N. P. Bailey, (private comunication).

[59] S.Melchionna, “Constrained systems and statistical distribution”, Phys.

Rev.E 61, 6165 (2000).

[60] S.Melchionna, G. Ciccoti, and B. L. Holian, “Hoover npt dynamics for

systems varying in shape and size”, Mol. Phys. 78, 533 (1993).

[61] Y. Bubanov, V. Schvetsov, and A. Siderenko, “Atomic structure of

binary amorphous alloys by combined EXAFS and x-ray scattering”,

Physica B 208, 375 (1995).

[62] M. Bionducci, G. Licheri, G. Navarra, and B. Bouchet-Fabre, “Deter-

mination of the partial structure factors of amorphous CuZr by anoma-

lous x-ray scattering and reverse monte carlo”, J. Phys. Sci. 51, 224208

(1996).

[63] D. McQuarrie, Statistical mechanics, 1st ed. (Harper and Row, New

York, 1973).

[64] Y. Zhang and A. Greer, “Thickness of shear bands in metallic glasses”,

Appl. Phys. Lett. 89, 071907 (2006).

[65] T. Masumoto and R. Maddin, “The mechanical properties of palladium

20 a/o silicon alloy quenched from the liquid state”, Acta Metall. 19,

725 (1971).

[66] P. Donovan and W. Stobbs, “The structure of shear bands in metallic

glasses”, Acta Metall. 29, 1419 (1981).

Page 92: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

BIBLIOGRAPHY 81

[67] E. Pekarskaya, C. Kim, and W. Jhonson, “In situ transmission elec-

tron microscopy studies of shear bands in a bulk metallic glass based

composite”, J. Mater. Res. 16, 2513 (2001).

[68] S. Ogata, F. Shimizu, J. Li, M. Wakeda, and Y. Shibutani, “Atomistic

simulation of shear localization in Cu-Zr bulk metallic glass”, Inter-

metallics 14, 1033 (2006).

[69] Y. Shi and M. Falk, “Stressed-induced structural transformation and

shear banding during simulated nanoindentation of a metallic glass”,

Acta Mater. 55, 4317 (2007).

[70] C. A. Schuh and T. G. Nieh, “A nanoindentation study of serrated flow

in bulk metallic glasses”, Acta Mater. 51, 87 (2003).

[71] W. H. Jiang and M. Atzmon, “Rate dependence of serrated flow in a

metallic glass”, J. Mater. Res. 18, 755 (2003).

[72] T. Mukai, T. Nieh, Y. Kawamura, A. Inoue, and K. Higashi, “Dynamic

response of a Pd40Ni40P20 bulk metallic glass in tension”, Scr. Mater.

46, 43 (2002).

[73] A. Sergueeva, N. Mara, J. Kuntz, E. Lavernia, and A. Mukherjee, “Shear

band formation and ductility in bulk metallic glasses”, Philos. Mag 85,

2671 (2005).

[74] H. V. Swygenhoven, P. Derlet, and A. Hasnaoui, “Atomistic modeling

of strength of nanocrystalline metals”, Adv. Eng. Mat. 5, 345 (2003).

[75] D. Feichtinger, P. Derlet, and H. V. Swygenhoven, “Atomistic simula-

tions of spherical indentations in nanocrystalline gold”, Phys. Rev. B

67, 24113 (2003).

[76] V. Yamakov, D. Wolf, M. Salazar, S. Phillpot, and H. Gleiter, “Length-

scale effects in the nucleation of extended dislocations in nanocrystalline

Al by molecular-dynamics simulation”, Acta Mater. 49, 2713 (2001).

[77] J. Schiøtz, T. Rasmussen, K. Jacobsen, and O. Nielsen, “Mechanical de-

formation of nanocrystalline materials”, Phil. Mag. Lett. 74, 339 (1996).

Page 93: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

82 Simulations of Mechanical Properties of Metallic Glasses

[78] A. Inoue, W. Zhang, T. T. ans A.R. Yavari, and A. Greer, “Unusual

room-temperature compressive plasticity in nanocrystal-toughened bulk

copper-zirconium glass”, Phil. Mag. Lett. 85, 221 (2007).

[79] K. Hajlaoui, A. Yavari, J. Das, and G. Vaughan, “Ductilization of BMG

by optimization of nanoparticle dispersion”, J. Alloys Compd. 434-435,

6 (2007).

Page 94: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Paper I

Page 95: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under
Page 96: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

DOI: 10.1002/adem.200700047

An Interatomic Potential for Studying CuZr Bulk Metallic Glasses**By Anca Paduraru,* Abder Kenoufi, Nicholas P. Bailey and Jacob Schiøtz

Binary alloys capable of forming metallic glasses havebeen discovered recently.[1–8] The mechanical properties ofBMGs are remarkably different from the ones of ordinary me-tallic alloys due to the atomic level disorder in the glassystate. Unlike crystalline materials plastic deformation in me-tallic glasses cannot be caused by lattice defects but takesplace through atomic-scale deformation events and mayfurthermore involve localization through formation of shearbands. For understanding the origin of their mechanicalproperties it is important to get the basic understanding offundamental theoretical problems through atomistic simula-tions.

Molecular dynamics (MD) treats atomic systems accordingto Newtonian mechanic laws. Atoms are point particles, in-teracting through an interatomic potential, describing the en-ergy of an atom as a function of the positions of all atoms in aneighboring region of space. The time evolution of the systemis obtained by numerically integrating Newton’s second law.Often the interatomic potential is a classical potential, noquantum mechanical description of the material is attempted,

but the functional form of the potential may be derived fromquantum mechanical arguments.[9] MD is able to treat sys-tems with millions of atoms, and permits the average calcula-tions of transport (diffusion, thermal conductivities, viscosi-ty), or mechanical quantities (elastic constant, plastic yield),and also the modeling of complex phenomena (shear bandlocalization, fracture appearance, neutronic cascades).

The quality of the results will depend on the quality of theinteratomic potential, simple potentials giving a less accuratedescription of the interatomic interactions while allowingvery large simulations, more complicated potentials may givea better description of the interaction, but limit the simulationsize. Many-body potentials such as the Embedded AtomMethod[10] and the Effective Medium Theory (EMT)[11,12] havebeen shown to give a good description of the late transitionmetals crystallizing in close-packed structures, and theiralloys, while still allowing simulations with millions ofatoms.[13] In this paper, we create an EMT potential optimizedfor modeling the mechanical and thermodynamic propertiesof CuZr bulk metallic glass.

CuZr was recently discovered to be a binary bulk metallicglass.[6,14] Since binary alloys are easier to model than alloyswith more elements, this makes CuZr an attractive bulk me-tallic glass to study theoretically. Previously, an interatomicforce field has been fitted to CuZr by Duan et al.,[28] but wefind that the EMT force field described here provides a betterdescription of the structure of the metallic glass. The potentialdeveloped here will be used to model the mechanical proper-ties of CuZr, to be published elsewhere.

Fitting an Effective Medium Theory Potential

The functional form of EMT is described by Jacobsen etal.[11] and will not be repeated here, as it is rather complex.The energy of each atoms is the sum of two terms, the cohesiveenergy and the atomic-sphere correction. The cohesive energyapproximates the energy of an atom with the energy of anatom in a perfect face-centered cubic (FCC) crystal with thesame electron density. Each atom contributes to the local elec-tron density in its neighborhood with an exponentially decay-ing density. Two EMT parameters describe this contributionto the electron density: n0 describes its magnitude, and g2 isthe inverse decay length. The energy of an atom is then afunction of the electron density at its positions, this functionhas a minimum determining the equilibrium lattice constant.The EMT parameters s0,k and E0 describes the position of thisminimum, the curvature, and the depth, respectively. Theatomic-sphere correction handles departures from the refer-ence FCC structure. It is described as the difference between

CO

MM

UN

ICATIO

NS

ADVANCED ENGINEERING MATERIALS 2007, 9, No. 6 © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 505

–[*] Dr. A. Paduraru, Dr. A. Kenoufi

Center for Atomic-scale Materials Physics (CAMP)NanoDTU, Department of PhysicsTechnical University of DenmarkDK-2800 Lyngby, DenmarkE-mail: [email protected]

Dr. N. P. BaileyDepartment of Mathematics and Physics (IMFUFA)DNRF Center “Glass and Time”Roskilde UniversityP.O. Box 260, DK-4000 Roskilde, Denmark

Dr. J. SchiøtzDanish National Research Foundations Center for IndividualNanoparticle Functionality (CINF)NanoDTU, Department of PhysicsTechnical University of DenmarkDK-2800 Lyngby, DenmarkE-mail: [email protected]

[**] The authors thank prof. Karsten W. Jacobsen for his contribu-tions to this project. This work was supported by the EUNetwork on bulk metallic glass composites(MRTN-CT-2003-504692 “Ductile BMG Composites”) and by the DanishCenter for Scientific Computingthrough Grant No. HDW-1101-05. “Glass and Time” and Center for Individual Nano-particle Functionality (CINF) are sponsored by The DanishNational Research Foundation.

Page 97: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

a pair potential in the actual structure and the same pairpotential in the reference structure. This pair potential is alsoan exponentially decaying function, the EMT parameters V0

and j describe the strength and inverse decay length. Theatomic-sphere correction term is responsible for a system’sresistance to shear.

EMT does not contain special parameters for describingthe interactions between unlike atoms in an alloy, insteadthese interactions are described by combining the potentialparameters for the elements. For example, the cohesive ener-gy of a Cu atom surrounded by Zr atom will depend on theCu parameter describing how the energy of the Cu atomdepends on the local electron density (s0, k and E0), and onthe Zr parameter describing how this density decays near theZr atom (n0 and g2).

We fit the seven parameters relating to Zr, while keepingthe Cu parameters fixed to the values obtained in a previouswork on the CuMg metallic glass.[15] We do the fitting byminimizing an error function

f��p�� ��

i

qi � �CiEMT ��p�� � Ctarget

i �2 (1)

where {p}={s0,n0,E0,V0,g2,k,j} is the list of parameters thatneed to be fitted, qi are normalizing constants, Ci

EMT andCi

target are the EMT values (obtained from molecular dy-namics) and target values (from experiment or from DFTcalculations), respectively. The sum goes over the differentproperties used to fit the potential, such as lattice constants,cohesive energies and elastic constants for Cu, Zr and theCuZr alloy in the B2 (or CsCl) structure, see table 1.

The target values were obtained from quantum mechani-cal calculations using Density Functional Theory within thegeneralized gradient approximation (DFT-GGA), using thePerdew-Wang 91 exchange-correlation functional[19] and ul-trasoft pseudopotentials. Calculations were performed withthe dacapo program from the CAMP Open Software pro-ject.[20] The usual care was taken to ensure convergence withrespect to plane wave cutoff (350 eV), number of k-points(8 × 8 × 8) etc.

For fitting the interatomic potential parameters we use aNelder Mead simplex method[21] for optimizing the Func-tion 1 described above. The method requires only evaluationsof the functions and not their derivatives for finding the mini-mum of more than one independent variable.

The result of the fit is shown in Table 1, showing how wellthe target values were fitted, and Table 2 showing the result-ing EMT parameters for Cu and Zr.

Testing the Potential

Before using the potential obtained in the previous sec-tions, it is important to test its ability to predict properties notused in the fitting procedure. Since the intended application

Paduraru et al./An Interatomic Potential for Studying CuZr Bulk Metallic Glasses

CO

MM

UN

ICATI

ON

S

506 http://www.aem-journal.com © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim ADVANCED ENGINEERING MATERIALS 2007, 9, No. 6

Table 1. Properties used in fitting the potential. Elastic constants are in Å/eV3,energies are in eV/atom and lattice constants are in Å. All values for CuZr refers to theB2 structure. The first five values (for pure Cu) were not used in the fitting in thiswork, but are provided for reference. All values for Zr and CuZr are obtained withDFT calculations, except the Zr [16] and CuZr cohesive [17,18] energies, which areexperimental.

Property Optimized value Target value Difference

Cu Shear Modulus (eV/Å3) 0.511 0.511 0.000

Cu Bulk modulus (eV/Å3) 0.889 0.886 0.003

Cu C11 (eV/Å3) 1.09 1.10 –0.01

Cu Cohesive Energy (eV) 3.52 3.51 0.01

Cu Lattice Constant (Å) 3.58 3.61 –0.03

Zr Shear Modulus (eV/Å3) 0.141 0.150 –0.009

Zr Bulk modulus (eV/Å3) 0.474 0.440 0.034

Zr Cohesive Energy (eV) 6.30 6.30 0.00

Zr Lattice Constant (Å) 4.57 4.55 –0.02

CuZr Shear Modulus (eV/Å3) 0.306 0.276 0.03

CuZr Bulk modulus (eV/Å3) 0.667 0.751 –0.084

CuZr Formation Energy (eV) –0.139 –0.140 0.001

CuZr Lattice Constant (Å) 3.19 3.28 –0.09

Table 2. EMT parameters.

Parameter Cu Zr

s0 (Å) 1.41 1.78

E0 (eV) –3.51 –6.3

k (Å–1) 3.693 2.247

j (Å–1) 4.943 3.911

V0 (eV) 1.993 2.32

n0 (Å–3) 0.0637 0.031

n2 (Å–1) 3.039 2.282

Table 3. Comparison between DFT and EMT results for different crystal structures forCu50Zr50. The B2 structure was used in the fitting of the potential, the other structures aretests of the potential.

Structure E (eV/atom) V (Å /atom)

Cu50Zr50B1 (DFT-GGA) 0.275 19.57

(EMT) 0.203 18.60

Cu50Zr50B2 (DFT-GGA) –0.1202 17.64

experiment[17] –0.14 —

(EMT) –0.139 16.38

Cu50Zr50L10 (DFT-GGA) –0.035 17.23

(EMT) –0.0011 17.23

Cu50Zr50L11 (DFT-GGA) 0.250 18.01

(EMT) –0.1295 18.25

Page 98: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

is CuZr metallic glasses, the potential is both tested on or-dered alloys with the same composition, and on thermodyna-mical and structural properties of the metallic glass.

The first test consists of comparing the formation energiesand lattice constants (here expressed as a volume per atom)for four different ordered alloys. These structures do not existin nature; the B2 (or CsCl) structure is the stable crystal struc-ture for Cu50Zr50. However, the other structures can be set upin the computer, and the energies and lattice constants can becompared to more accurate DFT calculations. The results aregiven in Table 3. If the potential is able to describe low-ener-gy ordered structures, it is likely to correctly describe thelocal structures occurring in a metallic glass as well. All thestructures except the L11 structure are well described. TheL11 structure consists of alternating close-packed layers of Cuand Zr, this structure is presumably far from what one willencounter in metallic glasses, and the inability to reproducethat formation energy is not too worrying.

A more relevant test is to examine how this potential per-forms when used on a metallic glass. We therefore prepared aCuZr metallic glass in order to compare the glass transitiontemperature and the short-range order as measured by theradial distribution function to experiment.

We performed MD simulations in order to obtain the CuZrglass phase from the melt. We used two different systemssizes, starting from body-centered cubic crystals with 2048and 42600 atoms, and the sites randomly occupied by Cu andZr. The resulting glassy systems were cube-shaped with sidelengths of 3.3 nm and 9.1 nm, respectively. The systems werecooled from 1400 K (well above the melt temperature) insteps of 25 K, with an average cooling rate of 0.54 K/ps. Eachtemperature step begins with a short period of Langevin dy-namics[22] to thermalize the system to the new temperature,followed by a simulation with constant (N,P,T) dynamics,using a combination of Nóse-Hoover and Parrinello-Rahmandynamics.[23,24] The prethermalization with Langevin dy-namics is necessary to prevent oscillations in the temperature

otherwise seen when the temperature of a Nosé-Hoover sim-ulation is changed abruptly. During the constant (N,P,T) dy-namics we recorded averages of different quantities of inter-est like pressure, volume, kinetic and potential energy of thesystem, in order to perform the thermodynamics analysis.[15]

The simulations were performed using the ASAP programfrom the CAMP Open Software project.[20]

When plotting the enthalpy versus temperature (inset ofFig. 1), we observe a change in the slope near 600–700 Kwhich indicates the glass transition. From this slope the heatcapacity vs. temperature (Fig. 1) can be determined. Thechange in the heat capacity indicates the glass transition, Tg.We extracted the glass transition temperature by taking thecentered derivative of Cp(T) and choosing the maximum.[15]

Tg was estimated to be 620 K, compared to the experimentalvalue of 670 K.[25] This good agreement is encouraging, inparticular when one considers that no data on glasses and nothermodynamical data at all was used to fit the potential.

The radial distribution function analysis (RDF) is a stan-dard method for obtaining information about the structure ofthe systems, especially for liquids and amorphous structures.In Figure 2 we show partial RDF functions gCu–Cu, gCu–Zr,gZr–Zr at the glass transition temperature. The first peaks ofthe RDF indicate the average distance at which we find near-est neighbor atoms (gCu–Cu and gZr–Zr yields the distance be-tween like atoms, gCu–Zr the distance between unlike atoms).By integrating the RDF functions until the first minimum, weobtain the number of nearest neighbors atoms for Cu and Zratoms. The results are summarized in Table 4 with compari-son with other experimental and simulations data. The valuesfor the coordination numbers and the interatomic distancesare in good agreement with the experimental ones, althoughthe interatomic distances are slightly underestimated, and theCuZr coordination number is slightly overestimated. The dif-ferences are, however, no larger than the differences betweenthe different experiments.

Paduraru et al./An Interatomic Potential for Studying CuZr Bulk Metallic Glasses

CO

MM

UN

ICATIO

NS

ADVANCED ENGINEERING MATERIALS 2007, 9, No. 6 © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.aem-journal.com 507

Fig. 1. Specific heat vs. temperature for two different sizes of CuZr metallic glass sys-tem during annealing. Inset: Enthalpy versus temperature for the 2048 atoms system.

Fig. 2. Radial distribution function for CuZr metallic glass.

Page 99: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

As a final test, we have compared the bulk and shear mod-ulus with experimental values. Our calculated values wereobtained at zero temperature by applying strain in the elasticlimit and relaxing the structure each strain step. We find108.69 GPa and 29.25 GPa, respectively, compared to the ex-perimental values 101.2 GPa and 31.3 Gpa.[25] The deviationis less than 8 %.

Conclusion

An EMT interatomic potential was fitted for CuZr, withthe purpose of simulating the CuZr bulk metallic glass. Thepotential was fitted to lattice constants, cohesive energies andelastic properties of elemental Cu, elemental Zr and Cu50Zr50

in the B2 crystal structure. The fitted potential does a reason-able job of reproducing lattice constants and energies of otherstructures of Cu50Zr50. When a glassy sample of Cu50Zr50 isproduced, we find a glass transition temperature, elastic con-stants and structural data in form of average coordinationnumbers and average nearest neighbor interatomic distancesin good agreement with experiment.

–[1] H. S. Chen, Y. Waseda, Phys. Status Solidi A 1979, 51,

593.[2] M. Widom, K. J. Strandburg, R. H. Swendsen, Phys.

Rev. Lett. 1987, 58, 706.[3] A. L. Greer, Science 1995, 267, 1947.[4] A. Inoue, Acta Mater. 2000, 48, 279.

[5] D. Wang, Y. Li, B. B. Sun, M. L. Sui, K. Lu, E. Ma, Appl.Phys. Lett. 2004, 84, 4029.

[6] D. H. Xu, B. Lohwongwatana, G. Duan, W. L. Johnson,C. Garland, Acta Mater. 2004, 52, 2621.

[7] M. B. Tang, D. Q. Zhao, M. X. Pan, W. H. Wang, Chin.Phys. Lett. 2004, 21, 901.

[8] A. Inoue, W. Zhang, J. Saida, Mater. Trans. 2004, 45,1153.

[9] N. Chetty, K. Stokbro, K. W. Jacobsen, J. K. Nørskov,Phys. Rev B 1992, 46, 3798.

[10] M.S. Daw, M. I. Baskes, Phys. Rev. B 1984, 29, 6443.[11] K. W. Jacobsen, P. Stoltze, J. K. Nørskov, Surf. Sci. 1996,

366, 394.[12] K. W. Jacobsen, J. K. Nørskov, M. J. Puska, Phys. Rev. B

1987, 35, 7423.[13] J. Schiøtz, K. W. Jacobsen, Science 2003, 301, 1357.[14] R. Ray, B. C. Giessen, N. J. Grant, Scr. Metall. 1968, 2,

357.[15] N. P. Bailey, J. Schiøtz, K. W. Jacobsen, Phys. Rev. B

2004, 69, 144205.[16] http://www.webelements.com/[17] A. I. Zaitsev, N. E. Zaitseva, Phys. Chem. 2002, 386, 214.[18] I. Ansara, A. Pasturel, K. Buschow, Phys. Status. Solidi.

(A) 1982, 69, 447.[19] J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson,

M. R. Pederson, D. J. Singh, C. Fiolhais, Phys. Rev. B1992, 46, 6671.

[20] The CAMP Open Software project, http://www.camp.dtu.dk/Software

[21] W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P.Flannery, “Numer. recipes in C++”, Sec. Ed. CambridgeUniversity Press, 2002, 413.

[22] M. P. Allen, D. J. Tildesley, Comput. Simulation of Liq.Claredon Press, Oxford, 1987.

[23] S. Melchionna, Phys. Rev. E 2000, 61, 6165.[24] S. Melchionna, G. Ciccoti, B. L. Holian, Mol. Phys. 1993,

78, 533.[25] J. Das, M. B. Tang, K. B. Kim, R. Theissmann, F. Baier,

W. H. Wang, J. Eckert., Phys. Rev. Lett. 2005, 94, 205501.[26] Yu. A. Bubanov, V. R. Schvetsov, A. F. Siderenko, Phys.

B 1995, 208–209, 375.[27] M. Bionducci, G. Licheri, G. Navarra, B. Bouchet-Fabre,

J. Phys. Sci. 1996, 51, 71.[28] G. Duan, D. Xu, Q. Zhang, G. Zhang, T. Cagin, W. L.

Johnson, W. A. Goddard, Phys. Rev. B 2005, 71, 224208.

______________________

Paduraru et al./An Interatomic Potential for Studying CuZr Bulk Metallic Glasses

CO

MM

UN

ICATI

ON

S

508 http://www.aem-journal.com © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim ADVANCED ENGINEERING MATERIALS 2007, 9, No. 6

Table 4. Average nearest neighbor distances and coordination numbers. The first line isthis work, the two next are experimental data with two different methods. The last line issimulations by Duan et al. with another potential and a slightly different composition.Since Cu and Zr are present in unequal amounts, the number of Zr neighbors to a Cu atomis different from the number of Cu neighbors to a Zr atom, and two numbers are thereforegiven for N(CuZr).

Composition Method R(Å)CuCu

R(Å)CuZr

R(Å)ZrZr

NCuCu

NCuZr

NZrZr

Cu50Zr50 This work 2.56 2.75 3.13 4.44 6.24 7.59

Cu50Zr50 EXAFS[26] 2.57 2.84 3.19 4.1 5.4 7.4

Cu50Zr50 XRD[27] 2.68 2.80 3.16 4.6 4.8 7.8

Cu46Zr54 Simul.[28] 2.67 2.78 3.22 3.2 7.6/7.5 9.1

Page 100: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Paper II

Page 101: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under
Page 102: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

Computer simulations of nanoindentation in Mg-Cu and Cu-Zr metallic glasses

A. Paduraru,1 N. P. Bailey,2 K. W. Jacobsen,1 and J. Schiøtz3

1Center for Atomic-scale Materials Design (CAMD), Department ofPhysics, Technical University of Denmark, DK-2800 Lyngby, Denmark

2Department of Mathematics and Physics (IMFUFA), DNRF Center “Glassand Time”, Roskilde University, P.O. Box 260, DK-4000 Roskilde, Denmark

3Danish National Research Foundations Center for Individual Nanoparticle Functionality(CINF), Department of Physics, Technical University of Denmark, DK-2800 Lyngby, Denmark∗

(Submitted to Phys. Rev. B: January 30, 2008)

The formation of shear bands during plastic deformation of Cu0.50Zr0.50 and Mg0.85Cu0.15 metallicglasses are studied using atomic-scale computer simulations. The atomic interactions are describedusing realistic many-body potentials within the Effective Medium Theory. The metallic glassesare deformed both in simple shear and in a simulated nanoindentation experiment. Plastic shearlocalizes into shear bands with a width of approximately 5 nm in CuZr and 8 nm in MgCu. Insimple shear, the shear band formation is very clear, whereas only incipient shear bands are seen innanoindentation. The shear band formation during nanoindentation is sensitive to the indentationvelocity, indenter radius, and the cooling rate during the formation of the metallic glass. Forcomparison, a similar nanoindentation simulation was made with a nanocrystalline sample, showinghow the presence of a polycrystalline structure leads to a different and more spatially distributeddeformation pattern, where dislocation avalanches play an important role.

PACS numbers: 62.20.F-, 61.43.Dq

I. INTRODUCTION

Since the discovery of metallic glasses,1 there has beena great interest in understanding and improving theirmechanical properties.2,3 The main obstacle for applica-tions of bulk metallic glasses is their brittleness, due tointense localization of the plastic deformation into shearbands. This localization is presumably due to the lackof work hardening in amorphous alloys. This differencebetween the macroscopic mechanical behavior of amor-phous and crystalline alloys is caused by the fundamen-tally different atomic-scale deformation mechanisms ofthe two classes of materials. In crystalline alloys, plas-tic deformation is carried by dislocations, which multiplyand eventually entangle in a deforming crystal, leadingto work hardening.4 In an amorphous alloy, dislocationsare not possible due to the lack of a crystal structure,and deformation proceeds through localized deformationevents.5–9 Since these events do not lead to local strength-ening, deformation may localize into shear bands.

For these reasons, much work has been done to studydeformation and shear band formation in metallic glasseswith atomic-scale simulation methods.9–16 Most of thiswork has been done with Lennard-Jones model sys-tems, sometimes in two dimensions to make the systemsizes larger for the same computational burden. WhileLennard-Jones systems can capture much of the essentialphysics of a deforming metallic glass, the description ofthe atomic interactions is too simple to give a realisticdescription of a specific metallic glass. For such a pur-pose, many-body potentials such as the Embedded AtomMethod17 or the Effective Medium Theory (EMT)18,19

must be used, at the price of a significant increase in thecomputational burden.

We have previously shown that computer simulationscan capture the formation of shear bands in MgCu metal-lic glass, deformed under tension in plane strain.15 Inthis paper, we study the formation of shear bands un-der nanoindentation of a CuZr glass and a MgCu glass.CuZr was recently discovered to be a binary bulk metal-lic glass.20,21 Since binary alloys are easier to model thanalloys with more elements, this makes CuZr an attractivebulk metallic glass to study theoretically.

II. CREATING VIRTUAL METALLIC GLASS

SAMPLES

We perform quasi-two-dimensional molecular dynam-ics simulations on binary alloys in order to simulatenano-indentation. The materials are Cu0.50Zr0.50 andMg0.85Cu0.15, described with EMT potentials19 recentlyfitted to describe these materials.22,23 Samples of metal-lic glass were produced by cooling the systems from1400K (well above the melting temperature) to the fi-nal temperature in steps of 25K, with an average cool-ing rate of 0.54K/ps, following the procedure publishedpreviously.23 In all cases periodic boundary conditionswere used along all three axes during the cooling pro-cedure, for the subsequent deformation simulations theperiodicity along one axis was later removed.

Samples with two sizes were made. Small samples con-taining 62 400 atoms were produced to be used for simpleshear deformation. The sample dimensions were 52.9 nm× 12.2 nm × 1.6 nm for CuZr and 57.0 nm × 13.1 nm ×1.7 nm for MgCu. These samples were cooled to 50 K.

Much larger samples were needed for the nanoindenta-tion simulations: Samples were created with dimensions

Page 103: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

2

FIG. 1: The setup used for simple shear deformations. Theshear is applied by forcing the walls to move as indicated bythe arrows. There are periodic boundary conditions in the y

and z directions.

of 52.9 nm × 52.9 nm × 1.6 nm for CuZr and 57.2 nm× 57.2 nm × 1.7 nm for MgCu, both corresponding to270 000 atoms. These samples were cooled to 116 K. Be-fore nanoindentation, the samples were replicated in thex direction, giving, e.g., a CuZr sample size of 105.9 nm× 52.9 nm × 1.6 nm, with periodic boundary conditionsalong the x and z directions, and free boundaries in they direction.

III. DEFORMATION IN SIMPLE SHEAR

We deformed the metallic glasses under simple shear,using a geometry illustrated in Fig. 1. The sample isdivided into three parts, two outer “walls” of 3 nm thick-ness, and a central part. In the central part, the dynamicsof the atoms is left unperturbed, but in the two walls thecenter of mass motion is constrained, so the two wallsmove with a constant relative speed. The atoms in thewalls are not held rigid; they can still move with respectto each other, since only the mass-weighted sum of theirvelocities is held fixed. By causing one wall to move inthe y-direction, a shear deformation with constant shearrate is applied to the system. The system has periodicboundary conditions in the y and z directions, but freeboundaries in the x direction. The shear simulations aredone at a temperature of 50 K.

During the deformation, velocity profiles (averageatomic velocities as a function of x-coordinate) are cal-culated. Both velocity profiles and plots of the deformedsystem clearly show the formation of shear bands (seeFig. 2). Visual inspection indicates that the width of theshear bands is 5 nm for the CuZr glass, and almost 10nm for the MgCu glass; the latter is consistent with thesimulations of Bailey et al15. A better estimate can beobtained by fitting the velocity profiles to a Fermi func-tion 1/ (exp(x/α) + 1) or to the arctan(x/α) function,the former gives a slightly better fit as the tails of thearctan function are too large. This gives a width pa-rameter α of 1.51 nm for CuZr and 1.93 nm for CuMg,corresponding to shear band widths of 6 and 8 nm, re-spectively, as the majority of the variation of the fermi

CuZr

CuMg

FIG. 2: Shear deformation on CuZr and MgCu thin sampleand velocity profile with respect to the length. The width ofthe shear bands were found to be 5–6 nm for CuZr and 8 nmfor MgCu.

function occurs in the interval from −2α to 2α. The fitsto the arctan function give α = 1.23 nm and 2.03 nmcorresponding to shear band widths of 5 and 8 nm forCuZr and CuMg. In all cases the fits were done withfour free parameters: the width, an amplitude and twotranslations.

For comparison, simulations were also done withoutthe walls, but with full periodic boundary conditions inall three directions. The deformation was made by chang-ing the shape of the computational box, i.e. changing oneof the three vectors describing the periodicity of the sys-tem. The results were essentially the same.

IV. NANOINDENTATION

To simulate nanoindentation, we perform moleculardynamics using Langevin dynamics24 at a constant tem-perature of 116 K. During the simulation, an indenter ispushed into the sample at constant velocity. Since thesamples are very thin in the z direction, the simulationis quasi two-dimensional, and we choose to use a cylin-drical indenter (see Fig. 3). The indenter is modeled as arepulsive potential, as proposed by Shi and Falk.25 Theform of the repulsive potential is

E = c (rxy − R)−6

(1)

where rxy is the distance in the x-y plane from the centerof the indenter to the atom, and R the indenter radius.

The strength parameter c = 2.0 eVA6. The indenter

is lowered into the substrate under displacement controlrecording at each step the load and the total displace-ment. During indentation, the lower edge of the systemis held fixed by fixing the positions of the atoms in the

Page 104: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

3

FIG. 3: The geometry used during nanoindentation. Thecylindrical indenter is modeled as a repulsive potential. Thereare periodic boundary conditions in the x and y directions(left-right and out of the plane of the paper). The uppersurface is free, the lower surface is held fixed by fixing thepositions of all atoms in a 0.8 nm thick layer.

bottom 0.8 nm of the system. The thickness of the systemis kept constant during the indentation (plane strain).

To visualize the plastic flow and the localization of theshear bands we use the D2

min method introduced by Falkand Langer.11 D2

min quantifies the local deviation fromaffine deformation during the interval of time between apair of configurations.

A. Formation of shear bands under the

nanoindenter

A nanoindentation simulation is performed on theCuZr sample as described above. The cylindrical inden-ter with a radius of 30 nm is indented into the samplewith a velocity of 7 m/s, as illustrated in Fig. 3. Asthe shear is inhomogeneous, it is not possible to rigor-ously define an equivalent shear rate, but it correspondsroughly to a shear rate of the order of 108s−1 (a homoge-neous shear rate of 108s−1 would correspond to the endsof the sample moving with a relative velocity of 10 m/s).

During nanoindentation, the net force on the inden-ter is recorded as a function of indentation depth, seeFig. 4. The curve is not smooth, but contains small ser-rations, presumably reflecting individual events duringdeformation. It would be tempting to interpret this asthe formation of individual shear bands, but that inter-pretation is not supported by direct visualization of theshear band formation. As it is seen in Fig. 4, the shearbands are not formed at once, but grow gradually duringthe nanoindentation. Just like in the simulations of Shiand Falk,16 the curved shape of the emerging shear bandscorresponds to the expected directions of the maximallyresolved shear stress, as predicted by slip line theory.26

10nma)

10nmc) d)

b) 10nm

10nm

0 10 20 30 40 50

Displacement (Angstrom)

50

100

150L

oad

(eV

/Ang

stro

m)

Indentation velocity 7 m/sIndentation velocity 30 m/s

e)

a)

b)

c)

d)

FIG. 4: Development of shear bands during nanoindentationon a Cu-Zr thin sample : a–d) visualization of strain local-ization at 24A, 32A, 42A, and 52A depth of the indenter, theindentation velocity was 7 m/s; e) The lower curve is the cor-responding load displacement curve, the displacements cor-responding to the four panels are shown with arrows. It isseen that the shear bands are forming gradually, there is nocorrespondence between events on the indentation force curveand formation of shear bands. The upper curve is a similarindentation when the indentation velocity is increased to 30m/s.

The shear bands are less clearly formed than in thesimulations of Shi and Falk.27 The simulations cannot,however, be directly compared as the simulations of Falkand Shi are two dimensional and use a Lennard-Jonesmodel potential, whereas our simulations have a finite(but small) thickness, and use a more realistic (but alsocomplex) potential. The more complicated potentialcombined with the larger number of atoms due to the

Page 105: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

4

a) b)10nm 10nm

FIG. 5: Influence of varying the indentation velocity on shearband formation and on load displacement curves. The inden-tation velocities are 30 m/s (panel a) and 7 m/s (panel b).The shear bands are less developed during the fast nanoin-dentation. The corresponding load displacement curves areshown in Fig. 4e.

finite thickness leads to a much larger computation bur-den, and our indentations are therefore necessarily donewith higher indentation velocities. There are thus threemajor differences between these simulations and the sim-ulations by Shi and Falk: the dimensionality, the poten-tial and the indentation speed, and it is not surprisingthat the results are somewhat different.

B. Influences on the shear band formation.

In this section we study how the shear band formationis influenced by varying the parameters of the nanoin-dentation, corresponding to varying physical parametersduring a nanoindentation experiment. We vary the in-denter velocity, its radius, the cooling rate during glassformation and the composition of the glass. All modifi-cations are relative to a “standard simulation” of a CuZrmetallic glass produced at a cooling rate of 0.44 K/ps,indented with a cylindrical indenter with radius 30 nmand indentation velocity 7 m/s.

Fig. 5 shows how the shear band formation is influ-enced by the indentation velocity. A simulation withindentation velocity of 30 m/s is compared with the ref-erence simulation. The shear band formation is less ob-vious in the fast indentation. A clear indentation ratedependence of the load displacement curve is also seen;the loading force — and thus the stress in the material— is significantly higher during the fast simulation, butno qualitative differences are seen. It can be speculatedthat even lower indentation rates will lead to formation offewer and larger shear bands, and possibly lead to macro-scopic failure through a few very large shear bands.28

Despite the indenter diameter being an order of magni-tude larger than the width of the shear band, a differenceis also seen when the indenter radius is varied. This isillustrated in Fig. 6, where the effect of reducing the in-denter radius by 33% is studied. As is seen, fewer shearbands are formed, when we have chosen to compare situ-ations with the same indentation depth. It is likely that

a) b)10nm 10nm

FIG. 6: Influence of varying the indenter radius. The radiiare 20 nm (a) and 30 nm (b); more shear band formation isobserved under the larger indenter.

a) b)10nm 10nm

FIG. 7: Influence of varying the cooling rate during glassformation. Samples created with cooling rates of 2.2 K/ps(a) and 0.44 K/ps (b) were subjected to otherwise identicalnanoindentation simulations. The shear localization appearsstronger in the more slowly cooled sample.

this difference is at least partly caused by less materialbeing displaced by the smaller indenter.

The effect of changing the cooling rate during theglass formation is investigated in Fig. 7, where identicalnanoindentation simulations are performed on two sam-ples produced with cooling rates of 2.2 K/ps and 0.44K/ps. In this case the difference is less pronounced, butthe shear band appears more distinct in the slowly cooledsample. This is consistent with results by Shi and Falk25

showing a stronger tendency for shear localization forslowly annealed systems.

Finally, we investigated the consequences of replacingthe CuZr glass with a MgCu glass (Fig. 8). Shear local-ization is far more pronounced in the CuZr sample. Thisis consistent with our previous findings that shear bandsare approximately twice as wide in MgCu glass (see alsoFig. 2.

V. NANOINDENTATION OF

NANOCRYSTALLINE SAMPLES

For comparison, we subjected a nanocrystalline sam-ple to indentation in a similar manner. Nanoindentationof nanocrystalline samples has previously been studiedby Van Swygenhoven et al.,29,30 with a particular fo-

Page 106: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

5

a) b)10nm 10nm

FIG. 8: Comparison of MgCu (a) and CuZr metallic glassunder otherwise identical nanoindentation conditions. Thelatter shows the strongest shear localization.

cus on the onset of plastic deformation under a spher-ical indenter. In this work we have chosen to mimicthe indentation parameters used for the metallic glass, sothe nanocrystalline sample is quasi-twodimensional with〈110〉-directions along the thin direction (the z direction).This orientation is chosen to maximize the number ofpossible slip systems in the crystal, which is neverthelessstill significantly lower than in a full three-dimensionalsimulation.31 The material for this simulation is copper,described with the EMT potential.19

The nanocrystalline sample was made by a randomVoronoi construction. Random grain centers were cho-sen, and the part of space closer to a given grain centerthan to any other grain center was filled with a randomlyrotated crystal, as published previously.32

The sample has the same dimensions as the CuZrmetallic glass sample, and similarly to the glassy sam-ple, it is made by repeating a system with half the sizein the x direction. The average grain diameter is 15 nm.The annealing is, however, only done from a temperatureof 464 K (kT = 0.04 eV), staying well below the meltingpoint to preserve the nanocrystalline structure. The in-dentation is done to a depth of around 5 nm, comparableto the simulations with metallic glass. After the simu-lation, the local atomic structure is analyzed with theCommon Neighbor Analysis,33 identifying atoms whichare locally coordinated as face-centered cubic (grain inte-riors), hexagonal closed-packed (stacking faults and twinboundaries) or otherwise (grain boundaries and disloca-tion cores).

Figure 9 shows the deformation during this nanoinden-tation, and Fig. 10 the corresponding load displacementcurves. A number of noteworthy features are marked byletters on Fig. 9b.

Right below the indenter (marked “A”), the plas-tic deformation is severe, and several grain boundarieshave disappeared. These are low-angle grain boundaries,which can be regarded as dense arrays of dislocations.The high stress below the indenter causes these to move.Many stacking faults are seen; they are caused by thehigh stress levels, making the energy of the stacking faultirrelevant compared to the elastic energy. Another exam-

FIG. 9: Nanoindentation of nanocrystalline copper. (a) Theinitial configuration before indentation. Atoms in the graininteriors are not show, atoms at stacking faults and twinboundaries are colored red, atoms in grain boundaries, dis-location cores and at surfaces are blue. Both high-angle andlow-angle grain boundaries are seen, the former are shownas blue walls, the latter are seen to be made of individualdislocations, for example the one marked “1”. A few stack-ing faults created by dislocation activity during annealing arealso seen. (b) After nanoindentation, significant plastic de-formation has occured, in particular just under the indenter,where numerous stacking faults and deformation twins haveappeared. The letters A – F mark features discussed in thetext. (c) A deformation map shows that the majority of thedeformation occurs inside the grains and is caused by dislo-cation activity, although some grain boundary sliding is alsoseen.

ple of a low angle grain boundary disappearing as a resultof dislocation migration is marked “B”.

A few grain boundaries “reconstruct” under the ap-plied stress by emitting an array of partial dislocations(for example at “C”). Presumably a lower grain bound-ary energy can be obtained and some of the applied stressrelieved by the motion of the dislocation wall, at the ex-pense of creating some stacking faults. A similar phe-nomenon is seen at “D”, where a wall of full dislocationshas been emitted by the grain boundary, and an extremecase at “E”, where a grain boundary has split into two

Page 107: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

6

0 10 20 30 40 50Displacement (Angstrom)

0

50

100

150

200

250

300

350L

oad

(eV

/ A

ngst

rom

)a b

10nma)

10nmb)

FIG. 10: Load displacement curve during nanoindentation ofthe crystal. The insets marked a and b show the deformationoccurring in the two intervals between the arrows.

separated by a region of (slightly faulted) HCP crystal.Similar reconstructions of grain boundaries under stresshave previously been seen in computer simulations.34 At“F” a stacking fault has been removed by the motion ofa partial dislocation, and a new one has been creatednearby by an independent event, creating the illusionthat a stacking fault has moved.

The events discussed above can also be found in thedisplacement map in Fig. 9c. The figure shows a strik-ing difference between the deformation distributions inthe glass and in the nanocrystalline sample: in the crys-talline sample the spatial distribution of the deformationis much more uneven, as it is controlled by the polycrys-talline structure. Deformation can only occur at grainboundaries, or where dislocations are available, typicallyfrom grain boundary sources.

The load-displacement curve in Fig. 10 shows signifi-cantly larger serrations than for the glassy samples. Thisis caused by avalanches of dislocation activity.35 The in-sets show how the deformation occurs just before andduring one of the serrations, it is clearly seen that the ser-ration does not correspond to a single dislocation event,

but rather to an avalanche of events.

VI. SUMMARY AND CONCLUSION

We have investigated shear band formation during ho-mogeneous shear and during nanoindentation in CuZr-and MgCu-based metallic glasses, using realistic inter-atomic potentials within the Effective Medium Theory.In homogeneous shear, shear bands of 5 and 8 nm thick-ness form in the CuZr- and MgCu-based glasses, re-spectively. In nanoindentation, we see a clear signa-ture of shear band formation under the indenter, al-though the shear bands are not as well developed as intwo-dimensional Lennard-Jones simulations. The shearbands form gradually during indentation. Shear bandformation is stronger if the indentation velocity is low, ifthe indenter is large and if the sample was cooled slowly.In addition, shear bands were clearer in CuZr than inMgCu under otherwise identical conditions, probably dueto the natural width of the shear bands in MgCu beinglarger.

Simulations of nanoindentation in nanocrystalline cop-per under otherwise identical conditions show a signifi-cantly different deformation mechanism, where individ-ual dislocations strongly influence the deformation pat-tern.

Acknowledgments

This work was supported by the EU Network onbulk metallic glass composites (MRTN-CT-2003-504692“Ductile BMG Composites”) and by the Danish Cen-ter for Scientic Computing through Grant No. HDW-1101-05. The Center for Atomic-scale Materials Designis supported by the Lundbeck Foundation. “Glass andTime” and Center for Individual Nanoparticle Function-ality (CINF) are supported by The Danish National Re-search Foundation. The authors would like to thank Dr.Abdel Kenoufi and Prof. James P. Sethna for fruitful dis-cussions.

∗ Electronic address: [email protected] W. Klement, R. H. Willens, and P. Duwez, Nature 187,

869 (1960).2 A. L. Greer, Science 267, 1947 (1995).3 A. Inoue, Acta Mater. 48, 279 (2000).4 J. P. Hirth and J. Lothe, Theory of Dislocations (Krieger,

Malabar, Florida, 1992), 2nd ed.5 A. S. Argon, Acta Metall. 27, 47 (1979).6 V. A. Khonik, A. T. Kosilov, V. A. Mikhailov, and V. V.

Sviridov, Acta Mater. 46, 3399 (1998).7 A. Ibenskas, C. E. Lekka, and G. A. Evangelakis, Physica

E 37, 189193 (2007).8 C. E. Lekka, A. Ibenskas, A. R. Yavari, and G. A. Evan-

gelakis, Appl. Phys. Lett. 91, 214103 (2007).9 N. P. Bailey, J. Schiøtz, A. Lemaıtre, and K. W. Jacobsen,

Phys. Rev. Lett. 98, 095501 (2007).10 D. Deng, A. S. Argon, and S. Yip, Philos. Trans. R. Soc.

London Ser. A 329, 613 (1989).11 M. L. Falk and J. S. Langer, Phys. Rev. E 57, 7192 (1998).12 F. Varnik, L. Bocquet, J. L. Barrat, and L. Berthier, Phys.

Rev. Lett. 90, 095702 (2003).13 N. P. Bailey, J. Schiøtz, and K. W. Jacobsen, Mater. Sci.

Eng. A 387-389, 996 (2004).14 Y. F. Shi and M. L. Falk, Phys. Rev. Lett. 95, 095502

(2005).15 N. P. Bailey, J. Schiøtz, and K. W. Jacobsen, Phys. Rev.

Page 108: Simulations of Mechanical Properties of CuZr and CuMg ... · Center for Atomic-scale Materials Design (CAMd), Department of Physics at DTU from February 2005 to February 2008 under

7

B 73, 064108 (2006).16 Y. F. Shi and M. L. Falk, Acta Mater. 55, 4317 (2007).17 M. S. Daw and M. I. Baskes, Phys. Rev. B 29, 6443 (1984).18 K. W. Jacobsen, J. K. Nørskov, and M. J. Puska, Phys.

Rev. B 35, 7423 (1987).19 K. Jacobsen, P. Stoltze, and J. Nørskov, Surf. Sci. 366,

394 (1996).20 D. H. Xu, B. Lohwongwatana, G. Duan, W. L. Johnson,

and C. Garland, Acta Mater. 52, 2621 (2004).21 R. Ray, B. C. Giessen, and N. J. Grant, Scripta Metall. 2,

357 (1968).22 N. P. Bailey, J. Schiøtz, and K. W. Jacobsen, Phys. Rev.

B 69, 144205 (2004).23 A. Paduraru, A. Kenoufi, N. P. Bailey, and J. Schiøtz, Adv.

Eng. Mater. 9, 505 (2007).24 M. P. Allen and D. J. Tildesley, Computer simulation of

liquids (Clarendon Press, Oxford, 1987).25 Y. Shi and M. L. Falk, Appl. Phys. Lett. 86, 011914 (2005).26 R. Hill, The mathematical theory of plasticity (Clarendon

Press, Oxford, 1950).27 Y. F. Shi and M. L. Falk, Thin Solid Films 515, 3179

(2007).28 C. A. Schuh and T. G. Nieh, Acta Mater. 51, 87 (2003).29 H. Van Swygenhoven, P. M. Derlet, and A. Hasnaoui, Adv.

Eng. Mater. 5, 345 (2003).30 D. Feichtinger, P. M. Derlet, and H. Van Swygenhoven,

Phys. Rev. B 67, 024113 (2003).31 V. Yamakov, D. Wolf, M. Salazar, S. R. Phillpot, and

H. Gleiter, Acta Mater. 49, 2713 (2001).32 J. Schiøtz, T. Vegge, F. D. Di Tolla, and K. W. Jacobsen,

Phys. Rev. B 60, 11971 (1999).33 D. Faken and H. Jonsson, Comput. Mater. Sci. 2, 279

(1994).34 J. Schiøtz, T. Rasmussen, K. W. Jacobsen, and O. H.

Nielsen, Phil. Mag. Lett. 74, 339 (1996).35 F. F. Csikor, C. Motz, D. Weygand, M. Zaiser, and S. Zap-

peri, Science 318, 251 (2007).