microstructure and embrittlement of leaded copper alloys - Research ...

microstructure and embrittlement of leaded copper alloys - Research ... microstructure and embrittlement of leaded copper alloys - Research ...

commission.recherche.epfl.ch
from commission.recherche.epfl.ch More from this publisher
28.01.2013 Views

MICROSTRUCTURE AND EMBRITTLEMENT OF LEADED COPPER ALLOYS THÈSE N O 3217 (2005) PRÉSENTÉE À LA FACULTÉ SCIENCES ET TECHNIQUES DE L'INGÉNIEUR Institut des matériaux SECTION DE SCIENCES ET GÉNIE DES MATÉRIAUX ÉCOLE POLYTECHNIQUE FÉDÉRALE DE LAUSANNE POUR L'OBTENTION DU GRADE DE DOCTEUR ÈS SCIENCES PAR Laurent FELBERBAUM ingénieur en science des matériaux diplômé EPF de nationalité suisse et originaire de Lausanne (VD) acceptée sur proposition du jury: Prof. A. Mortensen, directeur de thèse Dr N. Eustathopoulos, rapporteur Prof. P. Stadelmann, rapporteur Dr E. Vincent, rapporteur Lausanne, EPFL 2005

MICROSTRUCTURE AND EMBRITTLEMENT OF<br />

LEADED COPPER ALLOYS<br />

THÈSE N O 3217 (2005)<br />

PRÉSENTÉE À LA FACULTÉ SCIENCES ET TECHNIQUES DE L'INGÉNIEUR<br />

Institut des matériaux<br />

SECTION DE SCIENCES ET GÉNIE DES MATÉRIAUX<br />

ÉCOLE POLYTECHNIQUE FÉDÉRALE DE LAUSANNE<br />

POUR L'OBTENTION DU GRADE DE DOCTEUR ÈS SCIENCES<br />

PAR<br />

Laurent FELBERBAUM<br />

ingénieur en science des matériaux diplômé EPF<br />

de nationalité suisse et originaire de Lausanne (VD)<br />

acceptée sur proposition du jury:<br />

Pr<strong>of</strong>. A. Mortensen, directeur de thèse<br />

Dr N. Eustathopoulos, rapporteur<br />

Pr<strong>of</strong>. P. Stadelmann, rapporteur<br />

Dr E. Vincent, rapporteur<br />

Lausanne, EPFL<br />

2005


Remerciements<br />

Ce travail de recherche a été mené au Laboratoire de Métallurgie Mécanique de<br />

l’Ecole Polytechnique Fédérale de Lausanne, avec le soutien financier de la<br />

Commisssion pour la technologie et l’innovation<br />

CTI<br />

de l’Office fédéral de la<br />

formation pr<strong>of</strong>essionnelle et de la technologie (projet 4448.2 KTS), et celui de<br />

Swissmetal®<br />

, partenaire industriel.<br />

Je tiens en premier lieu à remercier mon directeur de thèse, le pr<strong>of</strong>esseur Andreas<br />

Mortensen pour m’avoir accueilli dans son laboratoire. Je lui suis vivement<br />

reconnaissant pour sa disponibilité sans limite qu’il m’a généreusement <strong>of</strong>ferte, ainsi<br />

que pour sa passion qu’il m’a transmise et pour sa rigueur qu’il m’a enseignée.<br />

Je désire remercier également les membres de mon jury de thèse pour le temps qu’ils<br />

m’ont accordé, Dr. N. Eustathopoulos, Dr. E. Vincent, Pr<strong>of</strong>. P. Stadelmann et Pr<strong>of</strong>.<br />

H.J. Mathieu, président du jury. J’ai tout particulièrement apprécié les commentaires<br />

développés par le Dr. Eustathopoulos au cours de mon examen de thèse.<br />

®<br />

Notre collaboration avec Swissmetal a toujours été positive et source de nombreux<br />

enseignements. Je tiens à remercier particulièrement E. Vincent pour son suivi à la<br />

fois attentif et critique du projet. Je remercie aussi ses collègues Ch. Charbon, P. Isler,<br />

S. Gilléron, M. Roulin, H.-Q. Tran, P.-A. Girard, R. Flahaut et C. Bezençon.<br />

Un gr<strong>and</strong> merci au Dr. Andreas Rossoll qui m’a aidé, guidé, aiguillé et encouragé tout<br />

au long de ce projet. C’est à lui que je dois tous les calculs de simulation. Associé à<br />

Andreas Mortensen, il m’a donné d’innombrables et précieux conseils lors de nos<br />

réunions de travail.<br />

i


REMERCIEMENTS<br />

ii<br />

Je tiens à remercier les étudiants, Jeannette Frei, Patrick Beyeler, Yves Winkler,<br />

Nicolas Barbi, Alban Dubach, Mustapha Laraqui et Olivier Robyr qui ont collaboré<br />

avec moi dans le cadre de projets de semestre ou de diplôme. De par leur curiosité, ils<br />

m’ont beaucoup apporté.<br />

Avec mes complices et cammarades Ali, Jean-François, R<strong>and</strong>o, Nono, Yves, Aude,<br />

Nicolas, Charpy, Ludger, Chris, Marianna, Doris et Vincent, j’ai passé des moments<br />

de gr<strong>and</strong>e amitié: Merci!<br />

Bravo et merci à l’équipe de l’atelier mécanique emmenée par Louis-Henri Masson<br />

puis Pierre-André Despont. Merci Werner, Jean-Marc, Tristan, Yves et Eric pour<br />

votre magnifique travail. J’adresse une mention spéciale à notre ”Père Noël” du<br />

vendredi et à sa fine équipe, santé!<br />

Que Mme Zanetta et Gökhan se voient spécialement remerciés. Eux qui sont toujours<br />

là qu<strong>and</strong> il faut.<br />

Je salue et embrasse ma maman Florence, mon papa Claude, ma soeur Isaline, mon<br />

frère Milan et le chouchou Andrea. Merci de m’avoir lancé et accompagné dans cette<br />

aventure! Grazie Michela du bonheur que tu continues de m’apporter et de l’amour<br />

que tu me donnes chaque jour. Michela et Samuel, je vous aime.


Nunc est bibendum, nunc pede libero<br />

puls<strong>and</strong>a tellus, nunc Saliaribus<br />

ornare puluinar deorum<br />

tempus erat dapibus, sodales.<br />

Horace, Odes<br />

I,<br />

xxxvii, 1-4<br />

iii


12<br />

Table <strong>of</strong> contents<br />

1 Abstract 1<br />

Résumé 3<br />

2 Introduction 5<br />

2.1 Context <strong>and</strong> goals <strong>of</strong> the study 5<br />

3 Literature review 9<br />

3.1 Shape <strong>of</strong> an inclusion (trans- or intergranular) 9<br />

3.1.1 Equilibrium shape: dihedral angle 9<br />

3.1.2 Influence <strong>of</strong> stress 17<br />

3.2 Cu alloy <strong>embrittlement</strong> at intermediate temperature 26<br />

3.2.1 Phenomenology 26<br />

3.2.2 GBE 32<br />

3.2.3 LME 39<br />

3.2.4 Copper intermediate temperature <strong>embrittlement</strong> - Conclusion 49<br />

4 Equilibrium dihedral angle <strong>of</strong> Pb in Cu 51<br />

4.1 Experimental procedure 52<br />

4.1.1 Material 52<br />

4.1.2 Heat treatments 52<br />

4.1.3 Microstructure characterization 53<br />

4.2 Time for equilibration 54<br />

4.3 Dihedral angle measurement 58<br />

4.3.1 Classical 2D method 58<br />

4.3.2 New 3D method 59<br />

v


vi<br />

4.4 Results 63<br />

4.4.1 Equilibration time 63<br />

4.4.2 γSL<br />

isotropy 64<br />

4.4.3 Measured dihedral angle 65<br />

4.5 Discussion 68<br />

4.5.1 Inclusion equilibration kinetics 68<br />

4.5.2 Solid/liquid interfacial energy 71<br />

4.5.3 Dihedral angle measurement 72<br />

4.5.4 Dihedral angles in Cu-Pb 72<br />

4.6 Conclusion 77<br />

5 Influence <strong>of</strong> stress on the shape <strong>of</strong> an embedded liquid inclusion 79<br />

5.1 Theoretical predictions 80<br />

5.1.1 The Eshelby method 80<br />

5.1.2 Inclusion shape 81<br />

5.1.3 Intergranular inclusion 83<br />

5.1.4 Equilibration time 85<br />

5.2 Experimental procedures 87<br />

5.2.1 Material 87<br />

5.2.2 Interrupted creep tests 87<br />

5.3 Results 89<br />

5.3.1 Creep curves 89<br />

5.3.2 Intragranular inclusion 89<br />

5.3.3 Intergranular inclusion 91<br />

5.4 Discussion 97<br />

5.4.1 Theoretical predictions 97<br />

5.4.2 Measured inclusion shape 98<br />

5.4.3 Comparison <strong>of</strong> theory with experiment 99<br />

5.5 Conclusion 105<br />

6 Cu-<strong>embrittlement</strong>: the role <strong>of</strong> Pb 107<br />

6.1 Experimental procedures 108<br />

6.1.1 Materials 108<br />

6.1.2 Mechanical testing 109<br />

6.2 Results 112<br />

6.2.1 Ductility trough 112<br />

6.2.2 Impact energy 118<br />

6.2.3 Damage 119


6.2.4 Fractography 122<br />

6.3 Discussion 124<br />

6.3.1 Embrittlement Mechanisms 124<br />

6.4 Conclusion 139<br />

7 Quench cracking in <strong>leaded</strong> <strong>copper</strong> <strong>alloys</strong>: a case study 141<br />

7.1 experimental procedures 142<br />

7.1.1 Materials 142<br />

7.1.2 Thermal properties 142<br />

7.1.3 Mechanical analysis 144<br />

7.2 Results 146<br />

7.2.1 Mechanical properties <strong>of</strong> CN8 & NP8 between RT <strong>and</strong> 800 °C 146<br />

7.2.2 Heat transfer coefficient <strong>of</strong> different quenching media 147<br />

7.3 Implications 152<br />

7.3.1 Bar diameter 152<br />

7.3.2 Cooling medium 153<br />

7.3.3 Alloy composition 154<br />

7.4 Conclusion 155<br />

8 General conclusions 157<br />

9 Future work 159<br />

10 Appendices 161<br />

10.1 Least square method <strong>and</strong> dihedral angle measurement 161<br />

10.2 Bulk modulus <strong>of</strong> liquid lead 170<br />

10.3 Eshelby analysis 172<br />

10.4 Measured values after interrupted creep test 178<br />

11 List <strong>of</strong> References 179<br />

vii


viii


1<br />

Chapter 1<br />

Abstract<br />

Mechanisms responsible for the <strong>embrittlement</strong> <strong>of</strong> <strong>leaded</strong> "free-machining” <strong>copper</strong><br />

<strong>alloys</strong> at intermediate temperatures ( i.e.<br />

around 400 °C) are examined using a<br />

combination <strong>of</strong> microstructural <strong>and</strong> mechanical characterization.<br />

Tensile tests are conducted on pure <strong>copper</strong> <strong>and</strong> on industrial high-strength Cu-Ni-Sn<br />

<strong>alloys</strong>, comparing <strong>leaded</strong> <strong>and</strong> un<strong>leaded</strong> versions, at strain rates ranging from 10-4<br />

to<br />

-1<br />

10 s , <strong>and</strong> various temperatures between 25 °C <strong>and</strong> 800 °C. It is found that, in the<br />

<strong>leaded</strong> samples, both liquid metal <strong>embrittlement</strong> <strong>and</strong> grain boundary <strong>embrittlement</strong><br />

operate at intermediate temperature,<br />

i.e. , near the melting point <strong>of</strong> lead. Their<br />

respective importance depends mainly on the strain rate: fracture at high strain rate<br />

-1<br />

(10 s ) displays characteristic signatures <strong>of</strong> liquid metal <strong>embrittlement</strong>, while at low<br />

-4<br />

strain rate (10 s-1)<br />

the ductility trough behaviour can be attributed to grain boundary<br />

<strong>embrittlement</strong>.<br />

In <strong>leaded</strong> <strong>copper</strong> <strong>and</strong> <strong>copper</strong> <strong>alloys</strong>, lead is mostly present as discrete inclusions,<br />

frequently located along grain boundaries. The dihedral angle <strong>of</strong> intergranular lead<br />

inclusions is then the main microstructural parameter besides their size. A new<br />

technique is proposed <strong>and</strong> demonstrated for the measurement <strong>of</strong> inclusion dihedral<br />

angle in a high-purity Cu-1 wt.% Pb alloy, based on room temperature<br />

characterization <strong>of</strong> samples that were held at elevated temperature at times<br />

sufficiently long for shape equilibration <strong>of</strong> the inclusions <strong>and</strong> then rapidly quenched.<br />

Quantitative scanning electron microscopic analysis <strong>of</strong> metallographic surfaces along<br />

1


CHAPTER 1. ABSTRACT<br />

2<br />

which individual inclusions have been dissolved is used to deduce the dihedral angle<br />

using a mathematical fit <strong>of</strong> their solid/liquid interface. For a specific temperature, we<br />

show that the dihedral angle is not unique; this reflects the fact that high-angle grain<br />

boundary energies are not constant in <strong>copper</strong>.<br />

Existing micromechanical analyses <strong>of</strong> shape equilibrium for intergranular inclusions<br />

in the presence <strong>of</strong> external stress are adapted to produce an improved description at<br />

low stress <strong>and</strong> to take into account the inclusion bulk compressibility. The derivation<br />

is based on minimization <strong>of</strong> the global capillary <strong>and</strong> elastic strain energy associated<br />

with such inclusions. An adimensional parameter that depends on the applied stress,<br />

the inclusion volume, the elastic properties <strong>of</strong> the solid, <strong>and</strong> the relevant interfacial<br />

energies emerges from the analysis as the sole factor governing the shape <strong>of</strong> the<br />

inclusion. If void nucleation occurs easily within the inclusion, such that its apparent<br />

bulk modulus is nil, the inclusion become unstable, degenerating to a crack when this<br />

adimensional parameter exceeds a critical value.<br />

Measurements <strong>of</strong> the dihedral angle on samples previously subjected to a remote<br />

stress at 400 °C show that applied stress causes a reduction in the apparent dihedral<br />

angle <strong>of</strong> liquid lead inclusions. Calculated critical stresses for both <strong>leaded</strong> pure<br />

<strong>copper</strong> <strong>and</strong> <strong>leaded</strong> <strong>copper</strong>-nickel-tin <strong>alloys</strong> that fail at high strain rate by liquid metal<br />

<strong>embrittlement</strong> are calculated for individual inclusions in the metal. These calculated<br />

critical stresses correlate relatively well with the measured fracture stress, suggesting<br />

that shape instability <strong>of</strong> the inclusions cause fracture <strong>of</strong> the <strong>alloys</strong>. This result is<br />

applied towards the suggestion <strong>of</strong> general strategies for reduction <strong>of</strong> the susceptibility<br />

<strong>of</strong> <strong>leaded</strong> <strong>copper</strong> <strong>alloys</strong> to liquid metal <strong>embrittlement</strong>.<br />

A case study addressing the problem <strong>of</strong> quench cracking <strong>of</strong> <strong>leaded</strong> Cu-Ni-Sn <strong>alloys</strong><br />

on the basis <strong>of</strong> experimentation <strong>and</strong> finite element thermomechanical simulation,<br />

showing how the problem can be alleviated in production, concludes the thesis.


1<br />

Résumé<br />

Les mécanismes responsables de la fragilisation d’alliages de cuivre au plomb dits de<br />

décolletage sont examinés par le biais de leur caractérisation mécanique et<br />

microstructurale.<br />

Des essais de traction ont été menés à des taux de déformation allant de 10-4<br />

à 10 s-1,<br />

entre la température ambiante et 800 °C, sur des échantillons de cuivre pur et<br />

d’alliage industriel à base de Cu-Ni-Sn, chacun contenant ou non du plomb. Il est<br />

démontré que l’effet fragilisant du plomb aux températures intermédiaires ( i.e. aux<br />

environs de la température de fusion du plomb) est dû au mécanisme de fragilisation<br />

par les métaux liquides ainsi qu’à la ségrégation intergranulaire. L’importance<br />

relative de ces deux mécanismes dépend principalement du taux de déformation: aux<br />

-1<br />

fortes vitesses (10 s ), la rupture se produit selon le premier mécanisme, alors que le<br />

-4 -1<br />

second domine à faible taux de déformation (10 s ).<br />

Le plomb est présent dans le cuivre et ses alliages sous forme de fines inclusions<br />

rassemblées pour la plupart aux joints de grains. L’angle dièdre de ces dernières en<br />

est la caractéristique principale en plus de leur taille. Une nouvelle méthode de<br />

mesure de cet angle a été développée et appliquée à des inclusions individuelles de<br />

plomb liquide à l’équilibre dans un alliage de Cu-1% masse Pb de haute pureté.<br />

Après trempe et préparation métallographique de l’alliage, le contenu des inclusions<br />

est dissout sélectivement. Ceci permet la reconstruction numérique de l’enveloppe<br />

des inclusions. Cette dernière est approximée par deux calottes sphériques dont<br />

3


CHAPTER 1. RÉSUMÉ<br />

4<br />

l’angle d’intersection est l’angle dièdre. Pour une température d’équilibration<br />

spécifique, il est montré que l’angle dièdre n’est pas unique; ceci reflète le fait que<br />

l’énergie des joints de grains de haute énergie du cuivre n’est pas constante.<br />

Des analyses existantes de la forme d’équilibre d’inclusions sous l’influence d’une<br />

contrainte extérieure ont été adaptées afin de donner une meilleure description de la<br />

forme d’équilibre aux contraintes faibles et de prendre en compte la compressibilité<br />

des inclusions. Ces analyses se basent sur la minimisation de la somme des énergies<br />

d’origine capillaire et élastique. Il en ressort qu’un paramètre adimensionnel régit à<br />

lui seul la forme d’équilibre des inclusions; celui-ci dépend de la contrainte<br />

appliquée, de la taille des inclusions, des propriétés élastiques du solide et des<br />

énergies interfaciales mises en jeu. Si un pore germine au sein de l’inclusion, rendant<br />

nul son module de compressibilité apparent, il est montré que l’inclusion devient<br />

instable et dégénère en fissure une fois que ce paramètre adimensionnel atteint une<br />

valeur critique.<br />

Des mesures d’angle dièdre sur des échantillons de Cu-1Pb mis sous contrainte à<br />

400 °C, montrent que l’angle diminue sous l’effet de celle-ci. Pour le cuivre pur au<br />

plomb et l’alliage de Cu-Ni-Sn-Pb sujets à la fragilisation par métal liquide, les<br />

contraintes critiques pour des inclusions spécifiques de ces alliages ont été calculées.<br />

Celles-ci sont en bon accord avec les contraintes de rupture mesurées sur ces<br />

matériaux. Sur la base de ce résultat, des stratégies visant à diminuer le risque de<br />

rupture due à la fragilisation par métal liquide sont proposées.<br />

L’étude de cas d’un alliage industriel présentant des fissures de trempe conclut ce<br />

travail. Sur la base d’expériences et de calculs par éléments finis, une interprétation<br />

quantitative du phénomène est donnée. Des solutions sont proposées afin d’éviter la<br />

fissuration systématique de cet alliage lors de sa production industrielle.


2<br />

Chapter 2<br />

Introduction<br />

2.1 Context <strong>and</strong> goals <strong>of</strong> the study<br />

Small additions <strong>of</strong> lead, on the order <strong>of</strong> 1 wt. %, are frequently used to improve the<br />

machinability <strong>of</strong> <strong>copper</strong> <strong>alloys</strong>, Fig. 2-1 [1].<br />

Figure 2-1: Chips produced by drilling in a machinability test: Ø 12.7 mm, depth<br />

12.7 mm, load 86 kg, 141 rpm. (a) For a free-machinig brass Cu-34.7Zn-3.4Pb, the<br />

drilling time is 8.4 s, whereas (b) it is 59.6 s for the un<strong>leaded</strong> brass Cu-37.1Zn; from<br />

[1].<br />

Leaded high-strength <strong>copper</strong> <strong>alloys</strong>, however, have a strong tendency to crack at high<br />

temperature during processing, particularly during quenching, Fig. 2-2.<br />

The aim <strong>of</strong> the present Ph.D. thesis is to further our underst<strong>and</strong>ing <strong>of</strong> the mechanisms<br />

causing the <strong>embrittlement</strong> <strong>of</strong> <strong>leaded</strong> <strong>copper</strong> <strong>alloys</strong>.<br />

5


CHAPTER 2. INTRODUCTION<br />

6<br />

Figure 2-2: Fracture surface <strong>of</strong> a <strong>leaded</strong> Cu-Ni-Sn alloy cracked during quenching<br />

from 900 °C.<br />

Its starting point was a quench cracking problem encountered in production by the<br />

Swiss <strong>copper</strong> alloy producer, Swissmetal, with whom this project was initiated. High-<br />

strength ternary <strong>copper</strong>-nickel-tin <strong>alloys</strong>, produced at Swissmetal by the Osprey<br />

process <strong>and</strong> strengthened by spinodal decomposition <strong>and</strong> precipitation hardening,<br />

could not be produced in a <strong>leaded</strong> version because <strong>of</strong> a strong tendency for extruded<br />

bars to crack during quenching from elevated processing temperatures. Somehow, the<br />

presence <strong>of</strong> lead in these <strong>alloys</strong> rendered them very weak at elevated temperature,<br />

<strong>and</strong> it was our initial goal to explain how <strong>and</strong> why this occurred <strong>and</strong> propose<br />

remedies.<br />

To this end, an experimental set-up was designed <strong>and</strong> implemented, allowing tensile<br />

testing <strong>of</strong> cylindrical <strong>copper</strong> bars at elevated temperature over a wide range <strong>of</strong> strain<br />

-4<br />

-1<br />

rates, extending over five orders <strong>of</strong> magnitude from 10 up to 10 s . Reasons for this<br />

approach were two-fold. First, at the temperature where cracks develop in the<br />

industrial <strong>alloys</strong>, its homogenized <strong>microstructure</strong> is metastable <strong>and</strong> evolves rapidly.<br />

Consequently, a capacity for rapid straining was important. Secondly, it has been<br />

debated in the literature whether the mechanism underlying the elevated temperature<br />

<strong>embrittlement</strong> <strong>of</strong> <strong>leaded</strong> <strong>copper</strong> <strong>alloys</strong> is elevated grain boundary cavitation, liquid<br />

metal <strong>embrittlement</strong> (LME), or an accentuation <strong>of</strong> grain boundary <strong>embrittlement</strong> by<br />

segregated lead (GBE). Experimental data, such as the intergranular fracture surface<br />

in Fig. 2-2, show indeed characteristic signatures <strong>of</strong> each <strong>of</strong> these mechanisms. The<br />

strain-rate dependence <strong>of</strong> <strong>embrittlement</strong> allows to shed some light on which is the<br />

operative mechanism, so for this reason also the approach was deemed interesting.


2.1 CONTEXT<br />

AND GOALS OF THE STUDY<br />

Given the existing controversy on the underlying mechanism <strong>of</strong> <strong>copper</strong><br />

<strong>embrittlement</strong> by lead, it was also decided to study the more “basic” <strong>and</strong><br />

microstructurally simpler system <strong>of</strong> high-purity <strong>copper</strong> with <strong>and</strong> without lead<br />

additions. The goal <strong>of</strong> this part <strong>of</strong> the thesis work was to contribute, using our<br />

capacity for tensile testing over a wide range <strong>of</strong> strain rates coupled with a novel<br />

approach to the characterization <strong>of</strong> dihedral angles in this system, to a discrimination<br />

<strong>of</strong> which mechanism operates, <strong>and</strong> when.<br />

The thesis is structured in a series <strong>of</strong> chapters that go from the more general to the<br />

more specific. The next chapter, Chapter 3, provides a review <strong>of</strong> the literature on the<br />

two basic questions that underlie this thesis, namely the shape <strong>and</strong> evolution under<br />

stress <strong>of</strong> liquid inclusions, <strong>and</strong> the general phenomenon <strong>of</strong> intermediate temperature<br />

<strong>embrittlement</strong> <strong>of</strong> <strong>copper</strong> <strong>and</strong> its <strong>alloys</strong>.<br />

Chapter 4 is dedicated to the description <strong>of</strong> a new method for dihedral angle<br />

measurement. Results are presented for a high-purity Cu-1 wt.% Pb alloy annealed at<br />

temperatures ranging from 400 to 950 °C.<br />

The influence <strong>of</strong> an applied remote stress on the shape <strong>of</strong> the liquid inclusions is then<br />

studied in chapter 5. Theoretical predictions are confronted with experimental results<br />

obtained on an industrial Cu-1Pb alloy subjected to interrupted creep tests. Both<br />

measured <strong>and</strong> predicted dihedral angle values are shown to decrease as a result <strong>of</strong> the<br />

application <strong>of</strong> a remote stress.<br />

Chapter 6 is focused on the intermediate temperature tensile behaviour <strong>of</strong> <strong>leaded</strong> <strong>and</strong><br />

un<strong>leaded</strong> <strong>copper</strong>. It is shown that, in fact, several <strong>embrittlement</strong> mechanisms operate:<br />

GBE dominates at low strain rate, whereas LME is dominant at high strain rates.<br />

Finally, the starting point <strong>of</strong> this work is presented as an illustration <strong>of</strong> the<br />

manifestation <strong>of</strong> LME. This is given in chapter 7, where transient thermal stresses are<br />

shown to be responsible for the systematic cracking <strong>of</strong> a water-quenched <strong>leaded</strong> high-<br />

strength Cu-Ni-Sn alloy.<br />

The thesis concludes with an overview <strong>of</strong> the main points demonstrated by this work,<br />

together with suggestions for future work. Appendices list the Mathcad<br />

7


CHAPTER 2. INTRODUCTION<br />

8<br />

spreadsheet computer codes that have been built for implementation <strong>of</strong> the analyses<br />

presented in Chapters 4 <strong>and</strong> 5.


3<br />

Chapter 3<br />

Literature review<br />

3.1Shape <strong>of</strong> an inclusion (trans- or intergranular)<br />

3.1.1Equilibrium shape: dihedral angle<br />

It was shown by Smith in a classical paper that phase arrangements in equilibrium<br />

<strong>microstructure</strong>s are not fortuitous [2]: surfaces arrange themselves in order to<br />

minimize their total energy. Provided that the interfacial energies are isotropic, an<br />

intragranular inclusion will be spherical, whereas in intergranular inclusion will<br />

adopt a lenticular shape, Fig. 3-1 [2]. Equilibrium holds when:<br />

where<br />

γ 12 γ 23 γ 13<br />

= =<br />

sin φ sin φ sin φ<br />

3<br />

Eq.3-1<br />

γxy<br />

is the interfacial energy between the phases x <strong>and</strong> y,<br />

<strong>and</strong> φz<br />

the dihedral<br />

angle corresponding to phase<br />

1<br />

2<br />

z.<br />

Figure 3-1: Interface equilibrium between three immiscible liquids <strong>and</strong> mechanical<br />

analogy with triangle <strong>of</strong> forces; from [2].<br />

9


CHAPTER 3. LITERATURE<br />

REVIEW<br />

10<br />

In the specific case <strong>of</strong> a liquid intergranular inclusion, Fig. 3-2, from Eq. 3-1 <strong>and</strong><br />

elementary trigonometry, we have in the absence <strong>of</strong> torque components that can arise<br />

with very low angle grain boundaries [3]:<br />

where γgb,<br />

<strong>and</strong><br />

respectively.<br />

When<br />

γ = 2γ<br />

gb SL<br />

⎛ φ<br />

cos<br />

⎞<br />

⎝ 2⎠<br />

γ<br />

SL<br />

Eq.3-2<br />

are the grain boundary, <strong>and</strong> solid-liquid interface energies<br />

b<br />

Figure 3-2: Intergranular, lenticular in shape, liquid inclusion.<br />

φ is zero, the liquid wets perfectly the grain boundaries. If φ is lower than 60°,<br />

symmetric grain boundary triple lines are wetted. Finally, if<br />

are spherical <strong>and</strong> do not interact with grain boundaries, [2].<br />

a<br />

φ/2<br />

φ is 180° the inclusions<br />

Figure 3-3: Influence <strong>of</strong> the dihedral angle on the shape <strong>of</strong> an inclusion located at the<br />

intersection <strong>of</strong> three grains: the cut is perpendicular to the triple line; from [2].<br />

The theoretical shape <strong>of</strong> secondary phase inclusions located at symmetric four-grains<br />

junctions (for tetrakaidecahedral ideal grains) was calculated by Wray [4]. A uniform<br />

γ SL<br />

γ SL<br />

γ gb


3.1 SHAPE<br />

OF AN INCLUSION ( TRANS-<br />

OR INTERGRANULAR)<br />

dihedral angle was considered. The grain boundary coverage <strong>of</strong> the secondary phase,<br />

βA,<br />

its volumic concentration,<br />

β<br />

V<br />

<strong>and</strong> the dihedral angle,<br />

φ,<br />

are shown to be<br />

interdependent. Six different morphologies <strong>of</strong> the secondary phase inclusions exist<br />

<strong>and</strong> can be predicted from two out <strong>of</strong> these three parameters, Fig. 3-4.<br />

corresponds to a perfect tetraedron. It is shown that even if<br />

φ = 70.53°<br />

φ > 60°, the secondary<br />

phase can form an interconnected network provided that its volume fraction is<br />

sufficiently high. Knowing both the evolution <strong>of</strong><br />

φ with temperature <strong>and</strong> the phase<br />

diagram then leads to the easy determination <strong>of</strong> βV,<br />

<strong>and</strong> thus to the secondary phase<br />

morphology, for a specific alloy composition at a specific temperature.<br />

Figure 3-4: The shape <strong>of</strong> secondary phase inclusions at a four-grains junction is given<br />

as a function <strong>of</strong> its volume fraction , <strong>and</strong> its equilibrium dihedral angle; from [4].<br />

Given its importance in the <strong>microstructure</strong>s <strong>of</strong> materials, <strong>and</strong> given the information it<br />

conveys on interfacial <strong>and</strong> grain boundary energy values, there has long been high<br />

interest in measuring φ with good precision, e.g. , Ref. [5].<br />

A difficulty encountered in practice when measuring<br />

φ is that many materials are not<br />

transparent to light. Therefore, unlike contact angles for which the sessile drop<br />

technique provides an elegant <strong>and</strong> geometrically unambiguous three-dimensional<br />

(3D) measurement technique, dihedral angles must <strong>of</strong>ten be observed in two<br />

dimensions, along polished metallographic planes that cut r<strong>and</strong>omly through the<br />

sample <strong>and</strong> its inclusions. In practice, apparent two-dimensional (2D) dihedral angles<br />

are measured either directly at the tip <strong>of</strong> a surface groove (”external” angle), at the<br />

apex <strong>of</strong> intergranular inclusions (”internal” angle), or alternatively for regular lens-<br />

11


CHAPTER 3. LITERATURE<br />

REVIEW<br />

12<br />

shaped inclusions such as that depicted in Fig. 3-2 by using a simple equation linking<br />

the apex angle with the width,<br />

2a,<br />

<strong>and</strong> thickness, 2b,<br />

<strong>of</strong> a 2D lens [6]:<br />

cos<br />

Eq.3-3<br />

Since metallographic planes cut the inclusions r<strong>and</strong>omly, the problem arises <strong>of</strong><br />

φ<br />

2 2<br />

⎛ ⎞<br />

⎝<br />

2 2<br />

2⎠<br />

=<br />

a − b<br />

a + b<br />

converting such 2D angles into their real 3D value<br />

φ defined above. To this end, a<br />

statistical treatment is <strong>of</strong>ten used on a sufficiently large number <strong>of</strong> 2D measurements.<br />

Early measurements <strong>of</strong> this type were made by Smith [2]. Assuming a single dihedral<br />

angle, the mode (i.e., the most frequently observed value) in a distribution <strong>of</strong> 2D<br />

angles was considered to be the true dihedral angle. 250 angles were measured,<br />

giving an estimated (statistical) precision <strong>of</strong> ± 5° [2]. Also assuming a single dihedral<br />

angle, Harker <strong>and</strong> Parker calculated the theoretical distribution <strong>of</strong> apparent dihedral<br />

angles measured on a r<strong>and</strong>om section [7]. Based on these calculations, Rieger <strong>and</strong><br />

Van Vlack reported that the measured apparent angles should follow a normal<br />

distribution. The median <strong>of</strong> this distribution then corresponds to the value <strong>of</strong> the true<br />

dihedral angle within 1° [8]. These authors also reported that 25 (or 100) angle<br />

measurements are sufficient to reach a precision <strong>of</strong> ± 10° (± 5°) within a 96%<br />

confidence interval in the most difficult situation where φ = 90° [8].<br />

Such a precision was not deemed satisfactory by Stickels <strong>and</strong> Hucke, who calculated<br />

confidence intervals for the population median in samples <strong>of</strong> size varying between 25<br />

<strong>and</strong> 1000 [9]. These authors also quantified the effects <strong>of</strong> (i) a non-unique dihedral<br />

angle, (ii) measuring errors, <strong>and</strong> (iii) apparent particle size on the distribution <strong>of</strong><br />

apparent φ values. It was concluded that the distribution is not necessarily normal but<br />

its median is almost not influenced by these factors (< 1° error). The median thus still<br />

corresponds to the true dihedral angle.<br />

In an extension <strong>of</strong> this study, the influence <strong>of</strong> a given distribution <strong>of</strong> the true 3D<br />

dihedral angle on the distribution <strong>of</strong> the apparent 2D dihedral angles measured along<br />

a r<strong>and</strong>om section was also quantified [10]. A visible effect, namely a relative<br />

broadening <strong>of</strong> the latter distribution, was shown to result once the true dihedral angle<br />

φ varies by more than 10°. A method for the determination <strong>of</strong> the true dihedral angle


3.1 SHAPE<br />

OF AN INCLUSION ( TRANS-<br />

OR INTERGRANULAR)<br />

distribution based on distribution discretization <strong>and</strong> on single angle measurement<br />

along a polished section was also developped [11]. A mathematical treatment was<br />

then used to demonstrate that the mean value <strong>of</strong> angles measured on a polished<br />

section is the actual average value <strong>of</strong> the true dihedral angle [12]. In the same article,<br />

De H<strong>of</strong>f proposes a new way to measure apparent dihedral angles on a<br />

metallographic section, based on the use <strong>of</strong> two stereological measurements: tangent<br />

<strong>and</strong> triple-line counts. Individual edge angles then do not have to be measured on the<br />

microsection, which eases the procedure operationally [12].<br />

Figure 3-5: Distribution <strong>of</strong> apparent dihedral angles measured on a 2-D cut. N=150<br />

measurements were performed on a polished section <strong>of</strong> a Cu-10Pb alloy heat treated at<br />

820 °C. The median <strong>of</strong> the distribution is 57.5°; from [13].<br />

Methods proposed in the literature for the measurement <strong>of</strong> dihedral angles along<br />

inclusions that intersect a grain boundary are thus <strong>of</strong>ten statistical; values reported<br />

are generally gathered using this approach, Fig. 3-5 [13]. Such methods are, however,<br />

cumbersome <strong>and</strong> tributary to the methods <strong>and</strong> assumptions used in data analysis.<br />

Another approach is to conduct precise three-dimensional measurements <strong>of</strong><br />

individual dihedral angles, as is done in the sessile drop method for wetting angles.<br />

Stereoscopic <strong>and</strong> statistical mathematical data analysis are then avoided, easing<br />

interpretation <strong>and</strong> increasing the reliability <strong>of</strong> the data. Two methods are generally<br />

used to this end.<br />

One is to create <strong>and</strong> examine the external angle at grain boundary grooves, formed<br />

along the lines <strong>of</strong> intersection <strong>of</strong> a grain boundary with a free surface. If the free<br />

13


CHAPTER 3. LITERATURE REVIEW<br />

14<br />

surface <strong>of</strong> a sample is wetted by the liquid metal, the liquid-filled grain boundary<br />

groove shape is governed by the dihedral angle φ <strong>of</strong> Eq. 3-2. Three-dimensional<br />

effects are resolved if the orientation <strong>of</strong> the grain boundary along the free surface is<br />

known (a 2D cut perpendicular to the grain boundary can then be obtained). Mullins’<br />

treatment <strong>of</strong> the shape <strong>of</strong> grain boundary grooves [14] can then be used to extrapolate<br />

the liquid/solid interface surface to the bottom <strong>of</strong> the groove, so as to deduce the<br />

value <strong>of</strong> φ from the shape <strong>of</strong> the groove [15].<br />

Figure 3-6: (a) AFM image <strong>of</strong> a grain boundary groove in Ni-43.6 at.% Al after<br />

annealing in vacuum at 1400 °C for 1h. (b) Typical depth pr<strong>of</strong>ile <strong>of</strong> a grain boundary<br />

groove. Angles θ 1 <strong>and</strong> θ 2 are determined by parabolic interpolation; from [15].<br />

A second technique for the direct measurement <strong>of</strong> φ is transmission electron<br />

microscopy (TEM) on samples containing inclusions that are sufficiently small to be<br />

embedded within an electron-transparent metallographic sample. With sufficient<br />

contrast between matrix <strong>and</strong> inclusion, the inclusion can be clearly distinguished.<br />

With adequate tilting <strong>of</strong> the sample <strong>and</strong> taking other precautions, the true dihedral<br />

angle <strong>of</strong> a single inclusion can then be measured, Fig. 3-7.


3.1 SHAPE OF AN INCLUSION (TRANS- OR INTERGRANULAR)<br />

Figure 3-7: Pb nanometric-inclusions within an Al matrix observed by TEM; from<br />

[16].<br />

Gabrisch et al. made in-situ TEM measurements <strong>of</strong> dihedral angles using this<br />

method, reporting the influence <strong>of</strong> (i) faceting, (ii) free surfaces, <strong>and</strong> (iii) projection<br />

errors on the measurement <strong>of</strong> the dihedral angle on a single inclusion [16]. The<br />

inclusions are typically nanoscopic (100 nm or less), <strong>of</strong>ten causing their shape to be<br />

influenced by size effects, such as ”magic size” phenomena observed with inclusions<br />

a few nanometres wide [17, 18]. This approach is also limited to temperatures below<br />

that at which inclusions start migrating along the grain boundaries towards the<br />

sample free surface; with Al/Pb this temperature is around 500 °C [16, 19].<br />

The Cu-Pb system is a good model system for such capillarity studies since lead<br />

melts at a much lower temperature than <strong>copper</strong>. Moreover the solubility <strong>of</strong> lead in<br />

solid <strong>copper</strong> is almost nil, while that <strong>of</strong> <strong>copper</strong> in liquid lead is limited. It has been<br />

studied by some authors [20, 21], who were mainly concerned with the growth<br />

kinetics <strong>of</strong> grain boundary grooves according to the extensive treatment by Mullins,<br />

see e.g. [14].<br />

Two major contributions on the subject are owed to C. S. Smith in the late fourties [2,<br />

22], <strong>and</strong> N. Eustathopoulos more recently [5, 13, 23-28]. These authors were mainly<br />

concerned with intergranular lead inclusions, <strong>and</strong> reported specifically the evolution<br />

<strong>of</strong> the dihedral angle with temperature, Fig. 3-8.<br />

15


CHAPTER 3. LITERATURE REVIEW<br />

16<br />

Figure 3-8: Evolution <strong>of</strong> the dihedral angle with temperature for the Cu-Pb system;<br />

from [13] <strong>and</strong> citing [22].<br />

The dihedral angle depends on one h<strong>and</strong> on the temperature but also on the chemical<br />

composition <strong>of</strong> both liquid <strong>and</strong> solid metals, see e.g. [29]. It was deduced from<br />

thermodynamic analysis that the observed evolution <strong>of</strong> the dihedral angle with<br />

temperature indicates that the temperature coefficients <strong>of</strong> both γ SL <strong>and</strong> γ gb are<br />

negative, as schematically shown on Fig. 3-9 (c) [30].<br />

Figure 3-9: Grain boundary groove morphologies (a) below, <strong>and</strong> (b) above T w , the<br />

wetting transition temperature. (c) Schematic evolution <strong>of</strong> γ gb <strong>and</strong> γ SL with temperature;<br />

from [30].<br />

More specifically, the evolution <strong>of</strong> γ SL with temperature was quantified according to<br />

a simple relationship involving the temperature, solubility, heat <strong>of</strong> fusion <strong>and</strong> molar<br />

volume <strong>of</strong> the two atomic species. From these results, <strong>and</strong> from the φ(T) data<br />

presented in Fig. 3-8, it was concluded that there is a linear decrease <strong>of</strong> γ gb with<br />

increasing temperature. This in turn was shown to imply that there is no, or a very<br />

limited, segregation <strong>of</strong> Pb at both the solid-liquid interface <strong>and</strong> the Cu grain<br />

boundaries [27].


3.1 SHAPE OF AN INCLUSION (TRANS- OR INTERGRANULAR)<br />

The influence <strong>of</strong> a third element was also studied [2, 28]. The addition <strong>of</strong> nickel to<br />

the Cu-Pb system has no influence on its capillary equilibrium since Ni do not<br />

segregate to the solid-liquid interface nor to the grain boundaries, whereas silver or<br />

bismuth additions do influence φ due to strong interfacial segregation Fig. 3-10.<br />

Figure 3-10: Dihedral angle <strong>of</strong> the liquid phase in Cu-Pb-Bi alloy as a function <strong>of</strong> the<br />

liquid phase composition. Increasing the Bi content results in a lower dihedral angle;<br />

from [2].<br />

3.1.2Influence <strong>of</strong> stress<br />

It has been suggested by Smith that the application <strong>of</strong> a tensile stress may cause the<br />

full penetration <strong>of</strong> grain boundaries by a liquid metal when in the absence <strong>of</strong> stress φ<br />

is somewhat greater than zero [2]. This was proposed to result from the formation <strong>of</strong><br />

a liquid film (φ = 0°) <strong>and</strong> its propagation, both being accelerated if a tensile stress is<br />

applied.<br />

Complete grain boundary coverage by a liquid metal can occur even if φ is positive<br />

[4], see the dashed line on Fig. 3-11 (b). Knowing both the evolution <strong>of</strong> the dihedral<br />

angle with temperature <strong>and</strong> the phase diagram for a specific system leads to the easy<br />

determination <strong>of</strong> φ <strong>and</strong> β V , thus the secondary phase morphology for a particular alloy<br />

composition at a precise temperature. The different domains <strong>of</strong> each morphology can<br />

accordingly be sketched on the phase diagram.<br />

As an illustration, the Al-Sn phase diagram is presented, Fig. 3-11. Fracture stress<br />

contours are also drawn. These are parallel to the estimated complete coverage line<br />

17


CHAPTER 3. LITERATURE REVIEW<br />

18<br />

(dashed line: β A =1). It has been argued that this suggests that the morphology <strong>of</strong> a<br />

secondary liquid phase is influenced by stress [4].<br />

(a) (b)<br />

Figure 3-11: (a) Phase diagram <strong>of</strong> the Al-Sn system [31], <strong>and</strong> (b) same diagram with<br />

the domains <strong>of</strong> morphology <strong>of</strong> the secondary phase, see Fig. 3-4. The dotted line<br />

corresponds to βA =1. Fracture stress contours are also added; from [4].<br />

A quantification <strong>of</strong> the influence <strong>of</strong> an applied stress on φ was performed by Stickels<br />

on Ni-2 wt.% Pb samples. Hydrostatic pressure [32], as well as uniaxial stress in both<br />

tension <strong>and</strong> compression [33] were applied. Dihedral angles were measured on 2-D<br />

polished sections according to the method described in [8]. The influence <strong>of</strong> stress on<br />

the dihedral angle in <strong>leaded</strong> <strong>copper</strong> was also reported [34] <strong>and</strong> discussed, Fig. 3-12<br />

[35].


3.1 SHAPE OF AN INCLUSION (TRANS- OR INTERGRANULAR)<br />

Figure 3-12: Influence <strong>of</strong> tensile stress on the distribution <strong>of</strong> observed dihedral angle.<br />

Material was uniaxially stressed at 650 °C, <strong>and</strong> subsequently quenched with the load<br />

still on. The median <strong>of</strong> each distribution is indicated; from [35].<br />

In all instances, the application <strong>of</strong> stress results in a slight decrease <strong>of</strong> φ. According to<br />

Stickels et al., this is rationalized by the fact that the dihedral angle can no longer be<br />

expressed in terms <strong>of</strong> interfacial tensions alone. Citing J.W. Gibbs: “The sum <strong>of</strong> the<br />

product <strong>of</strong> the volumes <strong>of</strong> the masses by their pressures, diminished by the sum <strong>of</strong> the<br />

products <strong>of</strong> the area <strong>of</strong> the surface <strong>of</strong> discontinuity by their tensions, must be a<br />

maximum“. The equilibrium particle shape, <strong>and</strong> hence the apparent dihedral angle<br />

are therefore dependent on both the total interfacial energy <strong>and</strong> the total strain energy<br />

[33].<br />

Cu-1Pb<br />

650 °C<br />

The equilibrium shape <strong>of</strong> an intragranular void in an elastic solid subjected to a<br />

remote stress has been calculated by considering both interfacial <strong>and</strong> stored elastic<br />

energies [36]. The latter energy accounts for the stress field perturbation due to the<br />

introduction <strong>of</strong> a void into a stressed solid. Assuming that the void is ellipsoidal in<br />

shape, <strong>and</strong> that the solid is linear elastic, this can be calculated analytically using the<br />

Eshelby formalism [37, 38].<br />

12 MPa<br />

φ = 81°<br />

21 MPa<br />

φ = 75°<br />

0 MPa<br />

φ = 93°<br />

7 MPa<br />

φ = 86°<br />

Minimization <strong>of</strong> the interfacial energy alone leads to a spherical shape, whereas the<br />

elastic interaction strain energy is minimized if the pore collapses to a crack [36].<br />

Once a remote stress is applied, a pore adopts either an ellipsoidal shape or collapses<br />

to a crack. There is a critical parameter, Λ c , above which the pore shape becomes<br />

19


CHAPTER 3. LITERATURE REVIEW<br />

20<br />

unstable. This parameter parallels the Griffith criterion for elastic crack propagation;<br />

however, here Λ c predicts crack nucleation, with:<br />

2<br />

σ ⋅ a0<br />

Λ=<br />

E ⋅γ<br />

Eq.3-4<br />

where σ is the remote stress, a0 , the initial pore diameter, E, the Youngs’ modulus,<br />

<strong>and</strong> γ, the interfacial energy. The Griffith criterion is:<br />

2<br />

2 2 1 σ a<br />

K1c= Y σ a = Eγ<br />

⇔ =<br />

Y E γ<br />

Eq.3-5<br />

Both Λc <strong>and</strong> the pore shape are shown to depend on the stress state, Fig. 3-13 (a).<br />

Pores namely remain spherical as Λ is increased to Λ c =1.78 if the solid is subjected<br />

to a hydrostatic stress; however, under uniaxial stress pores are ellipsoidal in shape<br />

provided that Λ < Λ c = 0,91[36].<br />

The same trends are found for the equilibrium shape <strong>of</strong> a grain boundary void, also<br />

assumed to be ellipsoidal [39]. The equilibrium shape results from the competition<br />

between elastic energy stored in the solid, <strong>and</strong> both grain boundary <strong>and</strong> surface<br />

energies. The initial shape <strong>of</strong> the intergranular void, which depends on the γ gb /γ ratio,<br />

therefore influences Λ c . Under uniaxial loading, <strong>and</strong> for γ gb /4γ = 1/11 (corresponding<br />

to φ = 159°), Λ c is about 0.7, Fig. 3-13 (b) [39], whereas it is 0.91 in the case <strong>of</strong> an<br />

intragranular void (φ = 180°) Fig. 3-13 (a) [36].


3.1 SHAPE OF AN INCLUSION (TRANS- OR INTERGRANULAR)<br />

(a) from [36] (b) from [39]<br />

Figure 3-13: Stability conditions <strong>of</strong> an (a) intragranular, <strong>and</strong> (b) intergranular pore<br />

within a stressed elastic matrix. m is a shape parameter that varies from 0 to 1 as the<br />

shape varies from a sphere to a crack. Solid lines represent stable equilibrium, whereas<br />

dashed line represent unstable equilibrium. Note that λ=γgb /4γ (i.e. φ = 159°) <strong>and</strong><br />

ω=σ1/σ3 in (b), therefore ω=1, <strong>and</strong> ω=0 correspond to hydrostatic <strong>and</strong> uniaxial stress,<br />

respectively.<br />

Raj proposed independently a similar model, aiming to calculate the stress to be<br />

applied to induce the formation <strong>and</strong> unstable spreading <strong>of</strong> a liquid film from a stable<br />

intergranular liquid inclusion [40]. The volume <strong>of</strong> the inclusion, V is assumed to<br />

remain constant. The elastic strain energy is also computed with the aid <strong>of</strong> the<br />

Eshelby formalism [37]. Here, the inclusion is approximated by a penny-shaped<br />

crack. An adimensionnal parameter, W is defined as:<br />

W =<br />

with ν, the Poisson’s ratio, <strong>and</strong> G, the elastic shear modulus.<br />

Eq.3-6<br />

With an initially spherical inclusion, ν = 0.33, <strong>and</strong>, for an elastic solid the relation:<br />

we have:<br />

( ) ⎛<br />

2<br />

1−νσ3V⎞ ⎜ ⎟<br />

π G γ ⎝ 2π⎠<br />

E<br />

G =<br />

21+ ν<br />

( )<br />

SL<br />

1/ 3<br />

2<br />

σ ⋅ a0<br />

W(<br />

φ0<br />

= 180° )≈0.<br />

36 = 0. 36 Λ<br />

E ⋅γ<br />

Eq.3-7<br />

Eq.3-8<br />

21


CHAPTER 3. LITERATURE REVIEW<br />

22<br />

It is shown that W describes both stable <strong>and</strong> unstable equilibrium configurations.<br />

Similarly to Λ c [36], a critical W* is defined, above which stable lenticular inclusions<br />

cannot exist, Fig. 3-14. On this figure, where the case φ 0 = 60° is considered, we read<br />

W* = 0.06. This corresponds to Λ c = 0.57.<br />

Figure 3-14: Stability conditions <strong>of</strong> an intergranular liquid inclusion imbedded in an<br />

elastic material subjected to a normal remote stress. Here, the dimensionless parameter<br />

W is plotted as a function <strong>of</strong> the apparent half-dihedral angle,θ 0 .Under stress-free<br />

conditions, the equilibrium dihedral angle φ is here 2θ e =60°. The liquid inclusion<br />

adopts a stable shape for the values <strong>of</strong> the apparent dihedral angle 2θ 0 ranging from<br />

60° to about 24°. Under this latter value, the inclusion is not stable <strong>and</strong> evolves to a<br />

crack; from [40].<br />

Following Ref. [36], a further paper refines the calculation <strong>of</strong> the elastic stored<br />

energy with the aid <strong>of</strong> a complex variable method [41] applicable to 2-D (cylindrical)<br />

problems. This was motivated by the fact that the stress state at the apex <strong>of</strong> a pore<br />

evolving to a crack is largely underestimated when assuming an ellipsoidal shape.<br />

However, since this new analysis is 2-dimensional; it is not applicable to real (3-D)<br />

inclusions. Nevertheless the same trends are shown, Λ c values only being lower. In<br />

the case <strong>of</strong> uniaxial loading, Λ c = 0.27 [41].<br />

The shape <strong>of</strong> the pores changes via surface diffusion. The shape evolution was also<br />

computed by Wang et al. [41] taking into account this specific path for matter<br />

transport, Fig. 3-15. A characteristic relaxation time is defined:


3.1 SHAPE OF AN INCLUSION (TRANS- OR INTERGRANULAR)<br />

4<br />

a0⋅kT τ =<br />

Ω<br />

⋅γ ⋅Dsδs<br />

Eq.3-9<br />

where Ω is the atomic volume, Ds the surface self-diffusion coefficient, <strong>and</strong> δs the<br />

thickness <strong>of</strong> the atomic layer participating in surface diffusion. Above Λ c = 0.27, the<br />

pore collapses very rapidly to a crack, Fig. 3-15 bottom sequence.<br />

Figure 3-15: Two sequences <strong>of</strong> evolving pores under uniaxial stress (stress axis is<br />

vertical). The stress state <strong>of</strong> the sequence on the top leads to a stable equilibrium shape:<br />

ΛΛ c ; from [41].<br />

Crack nucleation at the surface <strong>of</strong> a stressed elastic solid in contact with an agressive<br />

environment was also addressed by Suo et al., Fig. 3-16 (a) [42]. The equilibrium<br />

groove geometry is determined by energy minimization. The contributions <strong>of</strong> the<br />

surface energy, the grain boundary energy, the energy associated with the evaporation<br />

<strong>and</strong> condensation <strong>of</strong> matter <strong>and</strong> the elastic stored energy are considered. The surface<br />

pr<strong>of</strong>ile is approximated by a cycloid. This enables the analytical calculation <strong>of</strong> the<br />

stored elastic energy. The dimensionless parameter Λ is defined similarly to the<br />

previous analysis, save for the fact that the grain size, λ, is considered instead <strong>of</strong> the<br />

pore radius:<br />

23


CHAPTER 3. LITERATURE REVIEW<br />

24<br />

2<br />

σ ⋅ λ<br />

Λ=<br />

E ⋅γ<br />

Eq.3-10<br />

Below Λ c , grain boundary grooves approach their invariant shape, whereas above Λ c<br />

grooves sharpen into a crack. Λ c is found to decrease linearly from four to zero as the<br />

interfacial energy ratio, γ gb /γ varies from zero to two, corresponding to an apparent<br />

dihedral angle <strong>of</strong> 180 <strong>and</strong> 0° respectively. We then have:<br />

( )<br />

2<br />

⎛ σ ⋅ λ⎞<br />

⎜ ⎟ = 2 2γ<br />

−γ<br />

gb<br />

⎝ E ⎠ c<br />

Eq.3-11<br />

Note that this linear relationship has no fundamental physical grounding: it is namely<br />

tributary to the assumed cycloid geometry used to model the surface pr<strong>of</strong>ile; however<br />

the trend is meaningful [42]. Notice that a groove can form in a stressed single crystal<br />

provided that Λ>π. In that case, crack nucleation is predicted if Λ oversteps the value<br />

<strong>of</strong> four, Fig. 3-16 (b) [42].<br />

(a) (b)<br />

Figure 3-16: (a) Semi-infinite elastic polycrystal, in contact with an aggressive<br />

environment, subject to a stress parallel to the surface [42]. (b) Stability conditions <strong>of</strong> a<br />

groove for different γgb /γ ratio (i.e. different initial φ); from [42].<br />

The above analyses [36, 39, 41, 42] derive the global void shape based on global<br />

equilibrium, <strong>and</strong> assume a chosen regular shape for the void, inclusion or groove.<br />

The dihedral angle, measured at the very apex <strong>of</strong> an intergranular inclusion or void,<br />

or at the tip <strong>of</strong> a grain boundary groove, is on the other h<strong>and</strong> not expected to be<br />

influenced by an applied or residual stress, since very local stresses are always<br />

relieved by diffusion [43], <strong>and</strong> since capillary phenomena dominate at atomic-


3.1 SHAPE OF AN INCLUSION (TRANS- OR INTERGRANULAR)<br />

dimensions. Namely these scale with r 2 , whereas elastic energy scales with r 3 , where<br />

r is the distance from the triple line. It was moreover shown that at the triple line, the<br />

singularity associated with elastic fields is insufficient for these to overcome<br />

interfacial energy. Equilibrium angles governed by capillarity can therefore not be<br />

modified by elastic effects [44].<br />

A constant dihedral angle, uninfluenced by stress, is accordingly assumed in a study<br />

<strong>of</strong> the equilibrium shape <strong>of</strong> intergranular creep cracks [45].<br />

In a more recent numerical analysis <strong>of</strong> the shape <strong>of</strong> grain boundary voids, two models<br />

are proposed <strong>and</strong> compared. The first assumes a constant φ, whereas it is considered<br />

in the second model that the dihedral angle is not always constant, but adopts over<br />

time a stationary value [46]. It is shown that the second model allows much more<br />

accurate predictions <strong>of</strong> the void shrinkage process. The authors claim that this<br />

suggests that φ is not always constant. This discrepancy with the conclusions <strong>of</strong> Rice<br />

et al. [43, 45] probably arises from the fact that a numerical approach can not reflect<br />

the actual phenomena at the atomic-scale due to discrete meshing.<br />

As a conclusion, applied stress does influence the overall shape <strong>of</strong> a grain boundary<br />

groove, intergranular void or inclusion. Therefore the ”macroscopic dihedral angle”<br />

that reflects the energetic optimization <strong>of</strong> the overall inclusion shape is affected by<br />

stress, even though, as shown by Chuang <strong>and</strong> Rice, the ”local dihedral angle” defined<br />

at the atomic length scale remains unaffected by stress.<br />

25


CHAPTER 3. LITERATURE REVIEW<br />

26<br />

3.2 Cu alloy <strong>embrittlement</strong> at intermediate temperature<br />

3.2.1 Phenomenology<br />

”Intermediate Temperature Embrittlement” is frequently observed with <strong>copper</strong> <strong>and</strong><br />

<strong>copper</strong>-based <strong>alloys</strong>. Such <strong>embrittlement</strong> is manifest in a sharp decrease in ductility<br />

(generally measured as tensile elongation, reduction in area, or impact energy<br />

absorption) as the testing temperature <strong>of</strong> the material is gradually increased by a few<br />

hundred degrees above room temperature. The domain <strong>of</strong> reduced ductility begins in<br />

the vicinity <strong>of</strong> 300 ˚C, <strong>and</strong> frequently extends over a limited temperature range,<br />

above which the ductility increases again. This type <strong>of</strong> behavior has been termed<br />

”ductility trough behavior” or ”intermediate temperature <strong>embrittlement</strong>”; an<br />

illustration from the literature is given in Fig. 3-17 [47].<br />

Figure 3-17: Illustration <strong>of</strong> the ductility trough for a 200 ppm Pb containing Cu. The<br />

elongation, ∆l/l 0 , Reduction <strong>of</strong> area at fracture, RA, <strong>and</strong> fracture stress, R m are plotted<br />

as a function <strong>of</strong> test temperature, T; from [47].<br />

Such a sharp reduction in ductility has, in metal forming, the implication that many<br />

<strong>copper</strong> <strong>alloys</strong> are susceptible to cracking, during deformation or afterwards, as a<br />

result either <strong>of</strong> residual stresses or <strong>of</strong> transient stresses <strong>of</strong> thermal origin. This is<br />

generaly termed ”hot shortness”, or more accurately for <strong>copper</strong>, ”warm shortness”.<br />

An engineering-oriented review <strong>of</strong> the problem is given in Ref. [48].<br />

Ductility trough behavior is documented even for pure <strong>copper</strong>; however, the<br />

phenomenon can be far stronger in <strong>alloys</strong>. The intermediate-temperature<br />

<strong>embrittlement</strong> <strong>of</strong> <strong>copper</strong> <strong>and</strong> its <strong>alloys</strong> has been linked with the onset <strong>of</strong> grain


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

boundary sliding above about 300 ˚C [49]: this causes voiding at grain boundaries<br />

<strong>and</strong> shifts the fracture mode towards intergranular fracture. This is observed in OFHC<br />

(> 99.95% pure) <strong>copper</strong>, <strong>and</strong> in <strong>copper</strong> alloyed with known non-embrittling elements<br />

such as Ni or Zn [49-51]. Extensive creep data gathered on CuOFHC at intermediate<br />

temperature, shown on Fig. 3-18, thus report mostly intergranular fracture [52].<br />

Figure 3-18: Fracture-mechanism map for OFHC <strong>copper</strong> tested in tension. The map<br />

shows four fields corresponding to four modes <strong>of</strong> failure. Solid symbols mean that the<br />

fracture was identified as intergranular. Shading indicates a mixed mode <strong>of</strong> fracture;<br />

from [52].<br />

Fracture mechanisms occuring at intermediate temperature, i.e. above 300 °C, are<br />

schematically shown on Fig. 3-19 [52].<br />

Figure 3-19: Schematic view <strong>of</strong> the main fracture mechanisms occuring at intermediate<br />

temperature; from [52].<br />

27


CHAPTER 3. LITERATURE REVIEW<br />

28<br />

Voids nucleate at strain concentration sites such as intersecting sliding boundary <strong>and</strong><br />

slip planes, Fig. 3-20 [53, 54] or at secondary intergranular particles [55]. They are<br />

found to interlink once a critical level <strong>of</strong> grain boundary sliding is attained [56]. Note<br />

that grain boundary sliding is not a prerequisite for the occurence <strong>of</strong> such ductility<br />

loss but exacerbates it [57].<br />

Figure 3-20: A mechanism for microcavitation in the grain boundary due to interaction<br />

<strong>of</strong> primary <strong>and</strong> secondary slips with a sliding grain boundary; from [54].<br />

In pure <strong>copper</strong>, the loss in ductility increases as the strain rate is decreased [49, 51],<br />

as one would expect for a time-dependent damage phenomenon such as grain<br />

boundary voiding. This <strong>embrittlement</strong> ceases at temperatures above that at which<br />

grain boundary mobility, <strong>of</strong>ten accompanied by dynamic recrystallization, becomes<br />

operative [49, 51, 58].<br />

General grain boundaries are more prone to grain boundary sliding than special ones.<br />

This is one <strong>of</strong> the motivations for the grain boundary engineering concept proposed<br />

by Watanabe [54, 59]. The ductility trough was accordingly reported to widen <strong>and</strong><br />

deepen as the grain boundary energy increases [57, 60-62].<br />

Ductility trough behavior can be amplified by two main causes (i) <strong>embrittlement</strong><br />

caused by the segregation, prior to fracture, <strong>of</strong> specific atomic species to grain<br />

boundaries, <strong>and</strong> (ii) metal induced <strong>embrittlement</strong> (MIE), which includes liquid metal<br />

<strong>embrittlement</strong> but also vapour <strong>and</strong> solid metal induced <strong>embrittlement</strong>, reviewed in<br />

Ref. [63]. The difference between these two classes <strong>of</strong> mechanisms is that, in MIE,


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

fracture is caused by the migration <strong>of</strong> another embrittling metal towards crack tips<br />

via the crack. Grain boundary <strong>embrittlement</strong>, on the other h<strong>and</strong>, is caused by<br />

migration, prior to fracture, <strong>of</strong> impurity atoms from the bulk metal to the grain<br />

boundaries.<br />

Putting together data from the literature, the ductility trough <strong>of</strong> <strong>copper</strong> <strong>alloys</strong> has the<br />

following characteristics:<br />

• The trough widens, both at its lower <strong>and</strong> at its upper limits (which respectively<br />

decrease <strong>and</strong> increase) <strong>and</strong> also deepens (i.e., the measured ductility decreases<br />

further) as the percentage <strong>of</strong> embrittling impurities named above increases, Fig. 3-21<br />

[47, 50, 53, 64-68].<br />

Figure 3-21: Influence <strong>of</strong> the lead content in a Cu-Pb dilute alloy on the magnitude <strong>of</strong><br />

the ductility trough; from [47].<br />

• The reverse effect has been noted with the addition <strong>of</strong> certain elements: Zr [69, 70],<br />

<strong>and</strong> Mg, B, Y, Ce, La, Ca <strong>and</strong>, but to a lesser extent, P [70-73], <strong>and</strong> Ti[70]: such<br />

additions have been documented to reduce the trough extent. Foulger also reports that<br />

Mg <strong>and</strong> Li improve the ductility <strong>of</strong> cupronickels at 900 ˚C [48]. We note that<br />

elements Li, Mg, B, Y, Ce, La, <strong>and</strong> Ca all belong to columns II <strong>and</strong> III <strong>of</strong> the periodic<br />

table. The role <strong>of</strong> Zr, Y, Ce <strong>and</strong> La is explained as being due to sulfur scavenging [69,<br />

70, 72, 73]. A 1995 patent also cites B <strong>and</strong> Mn as additions decreasing crack<br />

formation (in both casting <strong>and</strong> deformation) <strong>of</strong> nickel containing <strong>copper</strong> <strong>alloys</strong> [74].<br />

• As the strain rate increases the trough narrows <strong>and</strong> becomes more shallow; this is<br />

found with both pure <strong>and</strong> alloyed <strong>copper</strong> [47, 49, 51, 70, 71, 75-77]. Such behavior is<br />

classical for grain boundary <strong>embrittlement</strong> [78].<br />

29


CHAPTER 3. LITERATURE REVIEW<br />

30<br />

Figure 3-22: Influence <strong>of</strong> the strain rate on the magnitude <strong>of</strong> the ductility trough <strong>of</strong><br />

OFHC <strong>copper</strong>; from [77].<br />

• The recorded ductility trough becomes deeper as the alloy grain size increases [71,<br />

79-81]. This can be rationalized since (i) while the overall grain boundary surface<br />

diminishes, the impurity coverage per area increases, (ii) the dislocation pile-ups are<br />

larger <strong>and</strong> hence produce higher stresses, (iii) grain boundary sliding is favoured <strong>and</strong><br />

(iv) the intergranular voids are larger [82].<br />

• The upper (brittle to ductile transition) limit <strong>of</strong> the ductility trough is associated<br />

with the onset <strong>of</strong> grain boundary mobility, <strong>of</strong>ten manifest as dynamic<br />

recrystallization, which thus restores the ductility [47, 49, 58, 64, 67, 70]. This<br />

influence <strong>of</strong> dynamic recrystallization is frequent [52, 83].<br />

Grain boundary mobility is highly dependant on the material purity [84]. The<br />

occurence <strong>of</strong> dynamic recrystallization is hence shifted to a lower temperature in an<br />

ultra-high purity material. This was documented for <strong>copper</strong> by Fujiwara in a study <strong>of</strong><br />

the hot-ductility <strong>of</strong> ultra-high purity <strong>copper</strong> [68]. The ductility trough behaviour is<br />

even found to disappear for 8 nines purity Cu, since dynamic recrystallization occurs<br />

at a temperature as low as 250 °C, Fig. 3-23. It must be emphasized that the strain<br />

rate used by Fujiwara was very low, namely 4.2 10 -5 s -1 , as compared to those on<br />

Fig. 3-22. At such very low strain rate, conventional purity <strong>copper</strong> exhibits a clear<br />

ductility trough behaviour, see solid circles for 3N-Cu (three-nines <strong>copper</strong>) on Fig. 3-<br />

23.


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

Figure 3-23: Ductility trough <strong>of</strong> <strong>copper</strong> <strong>of</strong> different purity. The reduction in area was<br />

measured after tensile testing in vacuum at a strain rate <strong>of</strong> 4.2 10 -5 s-1 ; from [68].<br />

• There is a time-dependency <strong>of</strong> the <strong>embrittlement</strong> phenomenon. This is related to the<br />

need for the embrittling species to migrate to the grain boundary or the site <strong>of</strong><br />

fracture [69, 80, 85-90]. The dependence <strong>of</strong> <strong>embrittlement</strong> on annealing time <strong>and</strong><br />

temperature then shows the classical ”nose” shape on a TTT diagram [80].<br />

31


CHAPTER 3. LITERATURE REVIEW<br />

32<br />

3.2.2 GBE<br />

The ideal cohesive stress <strong>of</strong> a material can be schematically derived from its curve <strong>of</strong><br />

potential energy vs. interatomic distance. Segregation influences the latter, <strong>and</strong><br />

therefore can reduce or increase the interatomic cohesion. The strengthening effect <strong>of</strong><br />

boron in nickel-base <strong>alloys</strong> is one classical example <strong>of</strong> cohesion enhancement.<br />

Ab initio calculations allow to draw potential-distance curves, <strong>and</strong> interfacial stress<br />

vs. separation curves [91, 92]. This is illustrated in Fig. 3-24 for a NiAl-Mo interface<br />

with <strong>and</strong> without C, S, or O [91]. The two latter elements are shown to reduce the<br />

cohesion <strong>of</strong> the interface, whereas carbon is a cohesion enhancer.<br />

(a) (b)<br />

Figure 3-24: (a) Calculated potential-separation, <strong>and</strong> (b) interfacial stress-separation<br />

curves for an NiAl-Mo interface with <strong>and</strong> without C, S, or O; from [91].<br />

Charge transfer is <strong>of</strong>ten associated with the occurence <strong>of</strong> GBE. This can be quantified<br />

by ab initio calculations. It is suggested that a localized charge on the impurity<br />

induces directional bonds <strong>and</strong> weakens the metal-metal bonds as less charge is<br />

available for these. On the other h<strong>and</strong>, cohesion enhancers form homopolar bonds,<br />

<strong>and</strong> in turn do not draw charge <strong>of</strong>f the metal atoms, Fig. 3-25 [93]. EELS spectra can<br />

provide experimental evidence for charge transfer associated with GBE [94-96];<br />

however, its detection may fail [97]. Note that minimal charge transfer was also<br />

calculated at Fe grain boundaries embrittled by segregated S or P, in which case


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

charge transfer cannot account for the well-known <strong>embrittlement</strong> effect <strong>of</strong> these<br />

elements in iron [98].<br />

Figure 3-25: Ab initio calculations <strong>of</strong> the charge density around an interstitial sulphur<br />

or boron atom in a tetrahedral site in nickel. Charge is localized in the former case,<br />

whereas it is more equally distributed in the second case. This suggests that grain<br />

boundary cohesion is dependent on charge transfer, since sulphur embrittles nickel,<br />

whereas boron acts as cohesion enhancer; from [93].<br />

The amount <strong>of</strong> segregation to a grain boundary has been reported to be related to the<br />

grain boundary energy. Namely twin Σ3 boundaries in <strong>copper</strong> were never found to<br />

exhibit bismuth segregation [97, 99]. Also, the least embrittled boundaries in a Cu-<br />

0.1Bi alloy are those <strong>of</strong> lowest energies [100]. This illustrates another motivation for<br />

the concept <strong>of</strong> grain boundary engineering [101].<br />

Grain boundary <strong>embrittlement</strong> has been amply documented for <strong>copper</strong> <strong>and</strong> its <strong>alloys</strong>.<br />

Cited embrittling elements are many: Bi (a strong embrittler subject <strong>of</strong> many studies)<br />

[48, 53, 94, 95, 99, 100, 102, 103], Te, [48, 104, 105], S [48, 69, 106], Sn [48, 85, 86,<br />

107], O [87, 108-110], Sb [48, 107, 110], Pb <strong>and</strong> Se [48, 64], as well as C <strong>and</strong> in<br />

some instances (but not always) P [48]. Gavin ranks their potency in the order <strong>of</strong><br />

(most to least potent) Bi, Te, Pb, Se, S [64]. More generally, metal grain boundary<br />

embrittling elements tend to belong to columns IV to VI <strong>of</strong> the periodic table [93,<br />

111]; this is indeed true <strong>of</strong> all elements mentioned in what precedes. Embrittling<br />

species also tend to be more electronegative than the host metal [111].<br />

33


CHAPTER 3. LITERATURE REVIEW<br />

34<br />

Figure 3-26: Compilation <strong>of</strong> sublimation enthalpies <strong>and</strong> atom size for different<br />

elements. Elements below the dashed line reduce the fracture energy <strong>of</strong> <strong>copper</strong>; from<br />

[112].<br />

Tabulated values <strong>of</strong> sublimation enthalpies <strong>and</strong> atom sizes allow to rank the potency<br />

<strong>of</strong> <strong>embrittlement</strong> or cohesion enhancement <strong>of</strong> segregating elements [112]. This<br />

simple theory, based on pair bonding, is shown to be equivalent to the general<br />

thermodynamic treatment formulated by Rice <strong>and</strong> Wang [113]. Embrittlement is<br />

predicted if the surface segregation free energy, ∆g(segreg) surf is higher (in absolute<br />

value) than the grain boundary segregation free energy, ∆g(segreg) gb . Consequently,<br />

2γ int , the ”ideal work <strong>of</strong> reversibly separating an interface against atomic cohesion” is<br />

reduced, since [113]:<br />

( )<br />

2 2 0<br />

γ int = γ int − X ∆g( segreg) −∆g(<br />

segreg)<br />

gb gb surf<br />

Eq.3-12<br />

with 2γ int 0 , the ideal work <strong>of</strong> separation in absence <strong>of</strong> segregation, <strong>and</strong> Xgb , the<br />

concentration <strong>of</strong> the segregant per unit area <strong>of</strong> interface.


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

One necessary condition for intergranular crack propagation to occur instead <strong>of</strong> crack<br />

tip blunting is that [113]:<br />

2γ int < Gdisl Eq.3-13<br />

with Gdisl , the energy to nucleate a dislocation at the crack tip. Moreover for<br />

intergranular rupture to occur, the stress must reach locally σ m , the cohesive stress, in<br />

order to nucleate a crack. The segregation <strong>of</strong> impurities is known to reduce both 2γ int<br />

<strong>and</strong> σ m , <strong>and</strong> in turn to induce GBE [113].<br />

Note that the term between parentheses in Eq. 3-12 corresponds to the reversible<br />

work <strong>of</strong> moving a solute atom from an adsorption site along a free surface to a<br />

specific site along a grain boundary. This value is clearly positive in the case <strong>of</strong> the<br />

system Cu-Bi, indicating its propensity to GBE. Indeed, it is known that bismuth<br />

segregates strongly, both to the surface <strong>and</strong> to the grain boundaries in <strong>copper</strong>. After<br />

surface cleaning by sputtering, a Bi film reforms at the surface by spontaneous<br />

diffusion from the general grain boundaries [114]. This experimental observation is<br />

not surprising since grain boundary <strong>embrittlement</strong> is commonly illustrated by the Bi-<br />

Cu pair [48, 53, 94, 95, 99, 100, 102, 103]. Grain boundary <strong>embrittlement</strong> can thus be<br />

ascribed to the difference in segregation <strong>of</strong> embrittling elements at surfaces <strong>and</strong> grain<br />

boundaries.<br />

Both thermodynamics <strong>and</strong> kinetics have to be considered in quantitative analyses <strong>of</strong><br />

the phenomenon; these are reviewed in [88]. Bi segregation to <strong>copper</strong> grain<br />

boundaries is a good illustration <strong>of</strong> these two facets <strong>of</strong> the phenomenon [115, 116].<br />

The effect <strong>of</strong> impurity content, temperature, <strong>and</strong> type <strong>of</strong> interface on the equilibrium<br />

segregation are presented in Fig. 3-27. These curves are drawn according to the<br />

classical thermodynamical treatment by McLean [88], Eq. 3-14. For grain boundary<br />

segregation in a dilute solution, one has:<br />

Xgb<br />

X − X<br />

max gb gb<br />

θ<br />

= = X<br />

1 − θ<br />

bulk<br />

⎛ −∆g(<br />

segreg)<br />

exp⎜<br />

⎝ RT<br />

gb<br />

⎞<br />

⎟<br />

⎠<br />

Eq.3-14<br />

35


CHAPTER 3. LITERATURE REVIEW<br />

36<br />

with X gb <strong>and</strong> X bulk , the concentration <strong>of</strong> the impurity at the grain boundary <strong>and</strong> in the<br />

bulk respectively. X max gb is the saturation value <strong>of</strong> X gb , <strong>and</strong> θ is the grain boundary<br />

coverage.<br />

It follows from Eq. 3-14 that the segregation increases with the alloy impurity<br />

content, as one <strong>of</strong> course expects; however, only the amount <strong>of</strong> impurity that is<br />

actually in solution has to be taken into account. Since the solubility changes, <strong>and</strong><br />

generally increases with temperature, it cannot be stated for a dilute alloy that<br />

increasing the temperature necessarily results in a decrease <strong>of</strong> the segregation level.<br />

(a) (b) (c)<br />

Figure 3-27: (a) The equilibrium fractional segregation,θ is plotted as a function <strong>of</strong><br />

temperature, T for several impurity contents in solution Xi . (b) The equilibrium<br />

segregation amount is plotted as a function <strong>of</strong> the impurity contents at several<br />

temperatures. (c) The equilibrium segregation at the surface is compared to that at a<br />

grain boundary (joint de grain) for a specific temperature. In the case where<br />

Xi =10 ppm, the fractional surface coverage, θ S , is much higher than that <strong>of</strong> the grain<br />

boundary, θ J ; from [117].<br />

A proposed mechanism for GBE is schematically illustrated on Fig. 3-28 [117]. The<br />

term between parentheses in Eq. 3-12 is taken positive; this corresponds to the<br />

situation depicted in Fig. 3-27 (c). A microcrack opens at a grain boundary triple<br />

junction, segregants enrich the free surface, <strong>and</strong> favour further grain boundary<br />

decohesion [117].


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

Figure 3-28: Proposed mechansim for GBE. (a) intergranular segregation at a triple<br />

junctions, (b) microcrack formation <strong>and</strong> surface enrichment, (c) crack propagation, <strong>and</strong><br />

further surface enrichment, from [117].<br />

It has been stated above that GBE is caused by the migration, before fracture, <strong>of</strong><br />

impurity atoms from the bulk metal to the grain boundaries. This is true in the ideal<br />

case <strong>of</strong> instantaneous decohesion (constant X gb). Decohesion under fully equilibrium<br />

conditions governed by thermodynamics (constant chemical potential µ) is on the<br />

other h<strong>and</strong> characterized by a change in X gb , the impurity concentration at the grain<br />

boundary [113]. It was shown that the actual work <strong>of</strong> separation 2γ int is [118]:<br />

2γ int( const Xgb ) > 2γ int ><br />

2γ<br />

int(<br />

const µ )<br />

Eq.3-15<br />

During interfacial separation, the impurity penetrates into the interface <strong>and</strong> reduces<br />

its cohesion. The kinetics <strong>of</strong> impurity migration were analyzed for slow separation<br />

(constant µ), fast separation (constant X gb ) <strong>and</strong> transient regimes [119]. A high<br />

sensitivity to the strain rate is reported: <strong>embrittlement</strong> is more severe at lower strain<br />

rate. This parallels the widely documented dynamic <strong>embrittlement</strong> phenomenon [109,<br />

119-123], defined as resulting from the stress-induced diffusion <strong>of</strong> an embrittler to<br />

the tip <strong>of</strong> an intergranular crack. This in turn reduces the strength <strong>of</strong> the grain<br />

boundary <strong>and</strong> promotes intergranular fracture. Decohesion occurs when a critical<br />

combination <strong>of</strong> stress <strong>and</strong> impurity concentration is reached, Fig. 3-29 [86].<br />

37


CHAPTER 3. LITERATURE REVIEW<br />

38<br />

Figure 3-29: Schematic view <strong>of</strong> the dynamic <strong>embrittlement</strong> process; from [86].<br />

An influence <strong>of</strong> stress on grain boundary segregation was reported in [124]. It was<br />

namely found in a martensitic stainless steel that phosphorus segregates to a higher<br />

extend to grain boundaries lying normal to applied tensile stress, <strong>and</strong> also to grain<br />

boundaries parallel to a compressive stress [124]. It was also found in one semi-<br />

quantitative study that <strong>embrittlement</strong> can be obtained as a result <strong>of</strong> prior exposure<br />

under applied stress to temperatures above the testing temperature [80]. Anisotropic<br />

segregation <strong>of</strong> sulphur in a low-alloy steel was also found to result from the<br />

application <strong>of</strong> an applied stress. The enrichment was maximal at grain boundaries<br />

lying perpendicularly to the applied tensile stress. Note that after 25 h, the grain<br />

boundary coverage returned to a homogeneous equilibrium value. This suggests that<br />

the stress-induced redistribution <strong>of</strong> sulphur on the grain boundaries is a transient<br />

process that may lead to grain boundary decohesion [125]. This redistribution is due<br />

to a diffusional process characterized by an activation energy <strong>of</strong> 98 kJ/mol (the<br />

activation energy <strong>of</strong> the grain boundary diffusion <strong>of</strong> sulphur in iron ranging from 97<br />

to 135 kJ/mol [126]) <strong>and</strong> a stress exponent close to unity, as in diffusional creep<br />

[127]. The segregation <strong>of</strong> some known embrittling elements, such as oxygen or<br />

sulfur, to <strong>copper</strong> grain boundaries can also be enhanced by an applied tensile stress<br />

[89, 121].


3.2.3 LME<br />

3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

A second mechanism that can cause or increase the severity <strong>of</strong> ductility trough<br />

behaviour in <strong>copper</strong> <strong>and</strong> its <strong>alloys</strong> is liquid metal <strong>embrittlement</strong> [78, 128-130].<br />

LME differs from grain boundary <strong>embrittlement</strong> processes in that the embrittling<br />

metal migrates during fracture to the site <strong>of</strong> crack nucleation <strong>and</strong> growth. Given its<br />

importance, the phenomenon has been extensively studied; several reviews exist on<br />

the subject [128, 129, 131-139], as well as dedicated conferences [140].<br />

LME refers to the ductility loss in normally ductile metals when stressed while in<br />

contact with a liquid metal [137]. From a phenomenological st<strong>and</strong>point, it has the<br />

following characteristics:<br />

• The phenomenon can be revealed <strong>and</strong> studied using both tensile <strong>and</strong> fracture<br />

toughness testing. In tensile testing, it leads to the appearance <strong>of</strong> a ductility trough<br />

[128, 132, 133, 139]. Moreover the stress-strain curves <strong>of</strong> the embrittled <strong>and</strong> the non-<br />

embrittled material superpose well until fracture, Fig. 3-30.<br />

Figure 3-30: Polycrystalline pure Al embrittled by various Hg solutions. Note the<br />

apparent correlation between Pauling electronegativity (in brackets) <strong>of</strong> solute element<br />

<strong>and</strong> severity <strong>of</strong> <strong>embrittlement</strong>, from [131].<br />

In LME, the lower-end (ductile to brittle) ductility transition is generally situated at<br />

the melting point <strong>of</strong> the embrittling metal; however, many exceptions exist to this<br />

observation. These exceptions occur either because <strong>of</strong> embrittler undercooling [Roth,<br />

1980 #303] (when it takes the form <strong>of</strong> small inclusions within the solid metal) or<br />

because solid metal induced <strong>embrittlement</strong> (SMIE) is also active.<br />

39


CHAPTER 3. LITERATURE REVIEW<br />

40<br />

• LME varies strongly in severity with the material system at h<strong>and</strong>; this has been<br />

termed the ”specificity” <strong>of</strong> LME, Fig. 3-31 [141, 142]. Some common features tend<br />

to be exhibited by documented embrittled/embrittler metal pairs: low mutual<br />

solubility, lack <strong>of</strong> mutual intermetallics, similar electronegativity, dissimilar heat <strong>of</strong><br />

fusion [128, 131-133]; however, all rules that have been proposed so far suffer<br />

exceptions [133]. Copper <strong>and</strong> its <strong>alloys</strong> are embrittled by liquid Sn, Pb, Bi, Pb-Bi, Sn-<br />

Pb, Pb-Ag, Hg, Ga, Ga-In, Ga-Hg, In, In-Hg, Zn, Li, Na, Cd, Sb, Th, Na [132, 139].<br />

Quoting a Russian study, Nicholas ranks the potency <strong>of</strong> some these embrittlers in the<br />

order (most to least potent) Bi, Pb, Cd, Sn, Ga [132].<br />

Figure 3-31: Illustration <strong>of</strong> the LME susceptibility, the LME couples are marked by a<br />

cross. ”P” st<strong>and</strong>s for nominally pure element, ”A” for alloy, ”C” for commercial purity,<br />

<strong>and</strong> ”L” for laboratory purity, from [141]. We thus read that both pure <strong>and</strong> alloyed<br />

<strong>copper</strong> <strong>of</strong> commercial purity are embrittled by liquid lead.<br />

It is <strong>of</strong>ten stated that LME is far more general than published data may imply, to the<br />

point where even the notion <strong>of</strong> any ”specificity” <strong>of</strong> the phenomenon has been<br />

disputed, particularly for LME [133, 137, 142]. A main reason why the occurrence <strong>of</strong><br />

LME is judged wider than data suggest is that observation <strong>of</strong> the phenomenon is<br />

highly dependent on testing conditions (temperature, strain rate, presence <strong>of</strong> a notch


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

etc), in particular because <strong>of</strong> the need for good wetting between the embrittling liquid<br />

<strong>and</strong> the solid metal: this is <strong>of</strong>ten difficult to produce in tests [132, 133, 143, 144].<br />

• For LME to appear, the applied stress must generate at least localized plastic<br />

deformation in the solid metal [128, 132, 133, 136, 139]. Crack initiation is<br />

furthermore driven by the presence <strong>of</strong> stress-concentrating obstacles to slip [139].<br />

• The onset <strong>of</strong> liquid spreading is <strong>of</strong>ten associated with LME failure [34, 65, 145];<br />

this agrees with the fact that liquid metals facilitate LME crack propagation [66].<br />

• LME bears strong similarities with other environmental <strong>embrittlement</strong> phenomena,<br />

including stress-corrosion cracking or hydrogen-assisted cracking [135, 139]. In<br />

particular, the relation between crack velocity <strong>and</strong> the crack stress intensity factor is<br />

similar in LME <strong>and</strong> these other environmental <strong>embrittlement</strong> phenomena: a threshold<br />

value K th LME below which cracks do not grow <strong>and</strong> slightly above which crack<br />

growth increases steeply with increasing K until a plateau is reached, the plateau<br />

extending until K reaches the critical stress intensity factor for rapid fracture, Fig. 3-<br />

32 [135].<br />

Figure 3-32: Crack velocity as a function <strong>of</strong> the stress-intensity factor for aluminium<br />

<strong>and</strong> titanium <strong>alloys</strong> tested in liquid metal <strong>and</strong> aqueous environment; from [135].<br />

In LME, there is though the significant difference that the plateau cracking velocity is<br />

very high (on the order <strong>of</strong> a meter per second [132, 136, 146]) because supply <strong>of</strong><br />

liquid metal to a moving crack tip is comparatively rapid, being limited only by<br />

viscous flow, which is rapid given the low viscosity <strong>of</strong> liquid metals [147].<br />

Additionally, all these environmental <strong>embrittlement</strong> phenomena have common<br />

41


CHAPTER 3. LITERATURE REVIEW<br />

42<br />

fractographic signatures [135]. LME is, however, characterized by a lower tendency<br />

for crack branching [139].<br />

• Fracture by LME is generally intergranular in polycrystalline materials [128, 129,<br />

132, 133, 139]; however, this is not always so <strong>and</strong> even single crystals or amorphous<br />

<strong>alloys</strong> are sensitive to the phenomenon (e.g., [133, 135]).<br />

Figure 3-33: Transgranular <strong>and</strong> intergranular fracture surfaces <strong>of</strong> Cu-1%Pb alloy (a)<br />

room temperature, <strong>and</strong> (b) 350 °C impact tested respectively; from [130].<br />

A clear correlation exists between the grain boundary energy <strong>and</strong> the crack extension<br />

force in LME. This was shown for symmetric Al bicrystals in contact with<br />

liquid Hg-Ga: maximum crack extension forces are required with low-energy grain<br />

boundaries, Fig. 3-34 [148].


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

Figure 3-34: (top) 3.5 µm/s crack extension force as a function <strong>of</strong> tilt angle for pure Al<br />

bicrystals in contact with liquid Hg-3 at.% Ga [148], <strong>and</strong> (bottom) corresponding grain<br />

boundary energy; from [149].<br />

It has therefore been argued that, since segregation reduces the grain boundary<br />

energy, prior impurity segregation to grain boundaries should reduce the severity <strong>of</strong><br />

LME. Indeed, a beneficial effect <strong>of</strong> segregated antimony was reported for pure<br />

<strong>copper</strong> subjected to creep in liquid bismuth [146].<br />

Observations made on the system Cu + liquid Pb-Bi suggest that the resistance to<br />

LME increases with increasing γ SL [150]. The intermediate temperature fracture<br />

stress <strong>and</strong> ductility are reported to increase with decreasing bismuth content [151],<br />

while γ SL increases accordingly to Morgan quoted in [150].<br />

• By <strong>and</strong> large, the response <strong>of</strong> the system to metallurgical <strong>and</strong> test variables is the<br />

same as that which is observed in low-temperature plasticity <strong>and</strong> fracture <strong>of</strong> brittle<br />

metallic materials [129, 133, 139]. Embrittlement is, in other words, promoted by all<br />

factors that tend to increase the intensity <strong>of</strong> local stress concentration caused by slip,<br />

particularly - but not only - at grain boundaries [133].<br />

• Following this general trend, the severity <strong>of</strong> <strong>embrittlement</strong> increases as the strain<br />

rate increases [78, 128, 132, 133, 139]. In this regard, LME differs from grain-<br />

43


CHAPTER 3. LITERATURE REVIEW<br />

44<br />

boundary induced ductility trough behaviour, in particular as observed for <strong>copper</strong> <strong>and</strong><br />

its <strong>alloys</strong> (see above).<br />

Several models have been proposed for LME. None <strong>of</strong> these was found to provide a<br />

universally applicable theory. The most relevant models were recently reviewed in<br />

[137-139]. These <strong>and</strong> most recent models are briefly summarized below. It is in any<br />

case generally accepted that the reduction in surface energy due to the presence <strong>of</strong> a<br />

wetting liquid is a necessary but not sufficient condition for <strong>embrittlement</strong> [137].<br />

According to Robertson, an elastic crack propagates in LME through dissolution <strong>of</strong><br />

the solid atoms at the crack tip, diffusion <strong>of</strong> these through the liquid, <strong>and</strong> redeposition<br />

on the stress-free <strong>and</strong> flat walls <strong>of</strong> the crack [152].<br />

The adsorption induced reduction in cohesion model was proposed independently in<br />

1963 by Stol<strong>of</strong>f <strong>and</strong> Johnston [153], <strong>and</strong> Westwood <strong>and</strong> Kamdar [154]. These authors<br />

assume that the stress to extend a crack is reduced if an embrittler element is<br />

adsorbed at the crack tip Fig. 3-35 (a), the adsorbed atom inducing a decrease <strong>of</strong> the<br />

cohesive stress <strong>of</strong> the material, Fig. 3-35 (b). This model resembles thus strongly the<br />

dynamic <strong>embrittlement</strong> model proposed by McMahon et al., compare Figs. 3-35<br />

<strong>and</strong> 3-29.<br />

(a) (b)<br />

Figure 3-35: (a) Schematic illustration <strong>of</strong> an adsorbed B-atom at the crack tip <strong>of</strong> a<br />

material subjected to tensile stress. (b) potential-separation curves as well as derived<br />

cohesive stress for both cases where a B-atom is adsorbed (U(a) B & σ(a) B ) or not; from<br />

[154].


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

According to Lynch, adsorption <strong>of</strong> a third element results rather in a reduction <strong>of</strong> the<br />

shear strength <strong>of</strong> atoms at the crack tip, increasing dislocation activity there.<br />

Enhanced dislocation emission then induces a more localized plastic zone, such that<br />

cracking occurs with less energy dissipation, Fig. 3-36 [135]. Arguments for this<br />

theory are largely derived from the presence <strong>of</strong> ductile fractographic signatures <strong>of</strong><br />

metals embrittled by LME.<br />

Figure 3-36: Mechanisms <strong>of</strong> crack growth by nucleation, growth <strong>and</strong> coalescence <strong>of</strong><br />

voids (a) in absence or (b) in presence <strong>of</strong> adsorbed species at the crack tip; from [155].<br />

The GALOP model was proposed recently by Glickman [150, 156], GALOP st<strong>and</strong>ing<br />

for grooving accelerated by local plasticity. It is proposed that grain boundary crack<br />

propagation is an iterative process where minute Mullins grooves [14] controlled by<br />

diffusion in the liquid are periodically blunted due to the presence <strong>of</strong> stress at the<br />

groove tips, Fig. 3-37 [150]. The diffusion length thus remains very short. This<br />

induces in turn a much higher crack propagation rate in comparison with classical<br />

Mullins grooving, where the rate-limiting diffusion length equals the total groove<br />

depth [150]. A strong dependence on the equilibrium dihedral angle is then predicted<br />

for the average groove velocity, V ∞ [156]:<br />

V ∞ ∝<br />

tan 3<br />

⎛ φ<br />

⎞<br />

⎝ 2⎠<br />

Eq.3-16<br />

45


CHAPTER 3. LITERATURE REVIEW<br />

46<br />

Figure 3-37: Galop Mechanism: (a) a Mullins groove forms, (b) stress relaxation by<br />

local dislocation activity blunts the groove tip, (c) <strong>and</strong> (d) crack blunts again after its<br />

extension by a ∆L* increment following Mullins kinetics [150].<br />

Kinetics <strong>of</strong> grain boundary wetting, or grain boundary penetration by a liquid metal,<br />

was the subject <strong>of</strong> a round table chaired in 1998 by Glickman [157]. Many<br />

observations <strong>and</strong> experiences <strong>of</strong> the participants were put together, but no clear trend<br />

was exposed, nor was any general agreement found. An influence <strong>of</strong> stress on grain<br />

boundary wetting kinetics is mentioned, but not discussed.<br />

Stress does, however, play a central role in the intergranular penetration <strong>of</strong> liquid<br />

gallium along pure aluminium grain boudaries. This was demonstrated by<br />

synchrotron radiation X-ray microradiography [158-161]. A link between grain<br />

boundary wetting <strong>and</strong> liquid metal <strong>embrittlement</strong> is thus clearly expressed in [161].<br />

Since gallium intergranular penetration is an intrusive process (meaning that the<br />

neighbouring grains are mooved apart [162]), it is no surprise that stress, or more<br />

precisely elastic strain energy, controls the process; see [163]. Accordingly, a net<br />

driving force for Ga penetration along Al grain boundaries was proposed in [164]:<br />

−<br />

Eq.3-17<br />

with G is the Gibbs free energy, U the strain energy, <strong>and</strong> A the liquid penetration area<br />

∂<br />

∂ =−∂<br />

G U<br />

− 2γ SL + γ<br />

gb<br />

A ∂A in the grain boundary plane.<br />

Intergranular liquid penetration occurs by definition above the wetting transition<br />

[165]. Cahn showed that this transition can result from a change in temperature,<br />

pressure or composition [166]. The grain boundary wetting transition is observable in<br />

<strong>alloys</strong> within a two-phase region <strong>of</strong> the phase diagram; however grain boundary<br />

prewetting or premelting can also occur in a single-phase region, Fig. 3-38 [167].


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

Note that prewetting refers to a new liquid phase, chemically similar to the bulk, that<br />

appears at grain boundaries, whereas premelting characterizes the appearance <strong>of</strong> a<br />

chemically dissimilar phase that replaces the grain boundary [167].<br />

Figure 3-38: Schematic phase diagram representation with both bulk <strong>and</strong> grain<br />

boundary solvus lines. T min <strong>and</strong> T max are the minimum <strong>and</strong> maximum wetting transition<br />

temperature, corresponding to high-energy <strong>and</strong> low-energy grain boundaries<br />

respectively [167].<br />

It is expected that the occurrence <strong>of</strong> perfect grain boundary wetting has a strong<br />

influence on properties such as mobility, diffusivity, energy <strong>and</strong> composition [168].<br />

Nanometric grain boundary bismuth precursor films (equivalent to premelting films)<br />

were also detected at the grain boundaries <strong>of</strong> <strong>copper</strong> or nickel in contact with molten<br />

bismuth [169-171]. Stress-dependent nanometric film propagation was futhermore<br />

shown in [172-174]: under tensile stress, the film propagation rate was estimated to<br />

be increased by two orders <strong>of</strong> magnitude [172]. This was one motivation for<br />

proposing a new mechanism for liquid metal <strong>embrittlement</strong> [171].<br />

Liquid<br />

metal<br />

T<br />

A<br />

S<br />

Short<br />

micrometric<br />

film<br />

S+L<br />

Wt. % B<br />

Figure 3-39: Schematic representation <strong>of</strong> an intergranular liquid film consisting in a<br />

short micrometre-thick <strong>and</strong> a long nanometric film. According to the authors, LME must<br />

be controlled by phenomena occurring at the tip <strong>of</strong> the nanometric film.[171]<br />

σ<br />

σ<br />

L<br />

T m<br />

T wmax<br />

T wmin<br />

Very long<br />

nanometric<br />

film (2−4 nm)<br />

LME<br />

47


CHAPTER 3. LITERATURE REVIEW<br />

48<br />

Following the same idea, prewetting or premelting <strong>of</strong> grain boundaries allows a<br />

sufficiently high supply <strong>of</strong> embrittler atoms at the crack tip to induce grain boundary<br />

decohesion <strong>and</strong> high crack growth rates, as observed in LME. Below the perfect<br />

wetting transition temperature, it is suggested that the application <strong>of</strong> a tensile stress<br />

may induce prewetting or premelting <strong>and</strong> in turn provoke the rapid rupture [175].<br />

One may note the analogy with precursor films that also form ahead <strong>of</strong> a solid or<br />

liquid droplet deposited on a free surface [176]. This was also obseved in the Cu-Pb<br />

system [177-179]. High-resolution transmission electron microscope pictures, Fig. 3-<br />

40 clearly demonstrate the phenomenon, showing a Ti-27Ag-5Cu liquid precursor<br />

film on a (0006) TiC surface: the wetting liquid drop is namely found to be preceeded<br />

by a nanometre-thick liquid film [180].<br />

Figure 3-40: High resolution images <strong>of</strong> the tip <strong>of</strong> molten Ti-27.4Ag-4.9Cu alloy<br />

spreading on a SiC substrate at 800°C. Arrows denote the tip <strong>of</strong> the precursor film. (a)<br />

0 s, (b) 3.2 s, (c) 4.4 s, (d) 7.8 s; from [180].<br />

In summary, several theories have been proposed to explain liquid metal<br />

<strong>embrittlement</strong>. None <strong>of</strong> these provides a complete <strong>and</strong> predictive theory <strong>of</strong> the<br />

phenomenon. By <strong>and</strong> large, it nonetheless is recognized that, at the heart <strong>of</strong> LME, is


3.2 CU ALLOY EMBRITTLEMENT AT INTERMEDIATE TEMPERATURE<br />

the fact that the embrittling metal reduces the strength <strong>of</strong> atomic bonds in the base<br />

metal, just as do solid state embrittling interfacial segregants. This in turn leads to<br />

lowered capillary energy requirements for fracture.<br />

3.2.4 Copper intermediate temperature <strong>embrittlement</strong> - Conclusion<br />

Focusing back on the Cu-Pb system only, rationalization <strong>of</strong> the ductility trough is<br />

thus debated in the literature. It is recognized that ductility is recovered with the onset<br />

<strong>of</strong> dynamic recrystallization; however, there is no general agreement on a single<br />

cause for the lower temperature marking the appearance <strong>of</strong> the ductility trough.<br />

Some authors claim that the intermediate temperature <strong>embrittlement</strong> <strong>of</strong> <strong>leaded</strong> <strong>copper</strong><br />

is simply due to the presence <strong>of</strong> the finely dispersed molten <strong>and</strong> thus inherently weak<br />

liquid lead inclusions [47, 50]. While it cannot be disputed that this effect always<br />

exists, added mechanisms <strong>of</strong> grain boundary <strong>embrittlement</strong>, or alternatively <strong>of</strong> liquid<br />

metal <strong>embrittlement</strong>, have been advanced to explain mechanical or physical<br />

signatures <strong>of</strong> the observed <strong>embrittlement</strong> <strong>of</strong> <strong>copper</strong> by lead. Since, furthermore, (i)<br />

grain boundary sliding starts to occur at about 300 °C, (ii) thermally activated<br />

processes are known to operate around <strong>and</strong> above 0.5 <strong>of</strong> the homologous<br />

temperature, which is about 400 °C for <strong>copper</strong>, <strong>and</strong> (iii) the incipient melting<br />

temperature <strong>of</strong> lead is 327 °C, all three mechanisms start operating at around the<br />

same temperature, which complicates discrimination between these.<br />

It is, in conclusion <strong>and</strong> in view <strong>of</strong> what precedes, most likely that all three<br />

mechanisms in fact operate in the Cu-Pb system depending on circumstances under<br />

which the material is deformed. Most probably, it is not the absence <strong>of</strong> any <strong>of</strong> these<br />

mechanisms, but rather the boundary that separate it from others, i.e., the parametric<br />

location <strong>of</strong> dominance <strong>of</strong> each <strong>of</strong> these mechanisms over the others that remain to be<br />

determined.<br />

49


CHAPTER 3. LITERATURE REVIEW<br />

50


Chapter 4<br />

4 Equilibrium dihedral angle <strong>of</strong> Pb in Cu<br />

A new method is presented for the measurement <strong>of</strong> the dihedral angle <strong>of</strong><br />

intergranular, lenticular, lead inclusions. This method is based on the stereoscopic<br />

quantification <strong>of</strong> the three-dimensional (3D) surface <strong>of</strong> selected single inclusions.<br />

The dihedral angle is deduced when the shape <strong>of</strong> the intergranular inclusion is<br />

known, in particular when it consists in two spheres intersecting along the grain<br />

boundary plane.<br />

To this end, inclusions visible along a polished metallographic section are selectively<br />

dissolved. Scanning electron microscopy stereo image pairs are then taken <strong>and</strong><br />

processed so as to enable a 3D digital reconstruction <strong>of</strong> the inclusion/matrix interface<br />

along each inclusion. Spherical caps describing the Cu/Pb interface over non-facetted<br />

orientations are then fitted to the measured digital inclusion envelope<br />

reconstructions. Knowing the center <strong>and</strong> radius <strong>of</strong> these spheres, the true dihedral<br />

angle <strong>of</strong> each specific inclusion is then deduced.<br />

51


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

52<br />

4.1 Experimental procedure<br />

4.1.1 Material<br />

High purity Cu-1wt.% Pb was used in this study. The solubility <strong>of</strong> Cu in Pb is very<br />

low, <strong>and</strong> the solubility <strong>of</strong> Pb in Cu is almost nil, Fig. 4-1. Hence, the <strong>microstructure</strong><br />

<strong>of</strong> the alloy is <strong>of</strong> pure <strong>copper</strong> equiaxed grains with lead-rich inclusions. These are (i)<br />

spherical in shape when located within the grains, <strong>and</strong> (ii) lenticular in shape when at<br />

grain boundaries.<br />

Figure 4-1: Phase diagram <strong>of</strong> the Cu-Pb system [31].<br />

The material was prepared from 99.999% pure metals by induction melting in a<br />

quartz crucible under a high-purity (≥ 99.999 vol.%) argon atmosphere, <strong>and</strong> cast in a<br />

Ø 6 mm <strong>copper</strong> mold.<br />

4.1.2 Heat treatments<br />

Heat treatments were conducted in a primary vacuum (p ≈ 10-2 mbar) for all samples<br />

except for the 930 °C anneal, which was conducted in Formier-gas (95% N 2 , 5% H 2 ,<br />

<strong>and</strong> p ≈ 10 1 mbar). A two-step treatment was applied:<br />

(i) 1 h at 900 °C to induce microstructural coarsening, causing the inclusions to grow<br />

to a size <strong>of</strong> roughly 10 µm;


4.1 EXPERIMENTAL PROCEDURE<br />

(ii) a hold at fixed temperature above the melting point <strong>of</strong> lead, sufficiently long to<br />

reach shape equilibration <strong>of</strong> the lead inclusions. The specimens were either<br />

encapsulated in a glass ampoule (static vacuum), or positioned in a quartz tube in the<br />

presence <strong>of</strong> a titanium sponge oxygen scavenger under dynamic primary vacuum.<br />

The duration <strong>of</strong> the heat treatment was varied as a function <strong>of</strong> the temperature (it was<br />

60 h at 400 °C <strong>and</strong> 1 h at 900 °C). Following this hold, the samples were quenched in<br />

water. The quench cools the samples in water in a few seconds, as shown by simple<br />

estimation <strong>of</strong> the rate <strong>of</strong> Newtonian cooling with a heat-transfer coefficient typical<br />

for non-agitated water, h ≈ 10 3 W m -2 K -1 [181].<br />

It can be shown by diffusion rate analysis, see Section 4.2, that these time/<br />

temperature combinations are sufficient for shape equilibration <strong>of</strong> the 10 µm<br />

inclusions during the isothermal hold, while the inclusions do not have time to<br />

change shape during or after the quench.<br />

Additionnal thin (6 mm in diameter, <strong>and</strong> 1 mm thick) samples were heat-treated at<br />

600 °C for 1 week in flowing Ar - 10 H 2 in order to scavenge <strong>and</strong> remove oxygen<br />

<strong>and</strong> sulfur traces. Hydrogen diffuses troughout the sample <strong>and</strong> forms H 2 O <strong>and</strong> H 2 S,<br />

both <strong>of</strong> which evaporate for these long treatment times [182]. A further heat<br />

treatment at 800 °C for 24 h on one sample <strong>and</strong> at 400 °C for 1 week on a second<br />

sample was conducted in the same atmosphere. Samples were then transferred to the<br />

cold zone <strong>of</strong> the apparatus. In this way, the sample is brought to room temperature<br />

within less than 10 s. These treatments were kindly performed by Pr<strong>of</strong>. Dominique<br />

Chatain at CRMC2-CNRS in Marseille, France.<br />

4.1.3 Microstructure characterization<br />

St<strong>and</strong>ard metallographical sections were prepared by successive grinding on wet SiC<br />

papers down to 2500 grit, <strong>and</strong> then by polishing with 6 µm <strong>and</strong> 1 µm diamond<br />

suspension (Struers ® MD-Dur with Struers ® red lubricant). Final polishing<br />

comprises repeated etching <strong>and</strong> 1 µm diamond polishing so as to obtain a<br />

deformation-free surface. The etchant is Klemm III (20 g K 2 S 2 O 5 , 100 ml distilled<br />

water, <strong>and</strong> 11 ml saturated Na 2 S 2 O 5 ”stock solution”). The wet etching-time was<br />

90 s, following [183].<br />

53


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

54<br />

4.2 Time for equilibration<br />

Shape equilibration <strong>of</strong> a liquid inclusion occurs by the diffusion <strong>of</strong> atoms <strong>of</strong> the solid<br />

matrix through the liquid. Diffusion along the solid/liquid interface is usually not<br />

considered, since it is negligible as compared to the above mentionned diffusion path.<br />

It is <strong>of</strong>ten concluded that equilibrium shapes are very rapidly reached for nanometric<br />

inclusions based on a simple comparison <strong>of</strong> the inclusion diameter with diffusion<br />

distances estimated using the coefficient <strong>of</strong> diffusion in the liquid [184-186]. The<br />

physics <strong>of</strong> the situation are, however, a little more complicated than this: additional<br />

parameters, such as the (typically very low) solubility <strong>of</strong> the solid in the liquid, also<br />

intervene in the process. We therefore estimate the equilibration time in what follows.<br />

We assume here that the solubility <strong>of</strong> Pb in Cu is nil, <strong>and</strong> that the transport <strong>of</strong> matter<br />

occurs by diffusion through the liquid. Moreover, no kinetic limitation to the<br />

movement <strong>of</strong> the solid/liquid interface, nor any strain energy due to misfit, are<br />

considered.<br />

We consider a small portion <strong>of</strong> the curved interface between the liquid <strong>and</strong> solid<br />

phases. At local equilibrium, the concentration <strong>of</strong> Cu atoms in the liquid is X Cu :<br />

X = X + X<br />

⋅V<br />

RT<br />

0 0 γ SL Cu<br />

Cu Cu Cu<br />

X Cu<br />

Eq.4-1<br />

0<br />

where is the equilibrium atomic concentration <strong>of</strong> Cu in Pb, γSL is the solid/liquid<br />

V Cu<br />

interfacial energy, is the partial molar volume <strong>of</strong> Cu in Pb, κ is the curvature, <strong>and</strong><br />

RT has the usual meaning.<br />

⋅κ<br />

If the shape <strong>of</strong> the inclusion deviates from that which is dictated by capillary<br />

equilibrium, Cu atoms will diffuse through liquid Pb, bringing the inclusion<br />

geometry nearer that which is dictated by capillary forces, Fig. 4-2.


4.2 TIME FOR EQUILIBRATION<br />

Figure 4-2: Shape changes <strong>of</strong> a liquid inclusion due to capillary forces. Arrows<br />

describe the atomic flux <strong>of</strong> Cu atoms through the liquid Pb inclusion (shaded area).<br />

Suppose that, locally, the curvature κ deviates from its equilibrium value, κ e : a flux J<br />

<strong>of</strong> solute will then cause the interface to move, into the liquid by adjunction <strong>of</strong> more<br />

Cu atoms along the interface where κ (counted positively towards the solid) exceeds<br />

κ e , <strong>and</strong> towards the solid where κ < κ e . If we assimilate for simplicity the local<br />

curved inclusion surface to a spherical cap <strong>of</strong> radius r = 2/κ, a simple mass balance<br />

yields:<br />

2<br />

1 dr 1 r dκ<br />

J = ⋅ = − ⋅ ⋅<br />

Ω dt Ω 2 dt<br />

where Ω Cu is the atomic volume <strong>of</strong> <strong>copper</strong>.<br />

Eq.4-2<br />

The flux J depends on the precise geometry <strong>of</strong> the inclusions <strong>and</strong> on their capillary<br />

equilibrium shape. As we only aim here for an order-<strong>of</strong>-magnitude calculation, we<br />

simply write:<br />

Cu Cu<br />

J D<br />

Eq.4-3<br />

where ∆C is the concentration difference in volumetric solute concentration within<br />

C ∆<br />

≈<br />

L<br />

the inclusion between points <strong>of</strong> maximum curvature difference ∆κ along the<br />

interface, <strong>and</strong> L is an average diffusion distance between points <strong>of</strong> maximum<br />

difference in curvature. By analogy with Nabarro-Herring diffusional creep, for<br />

equiaxed inclusions this distance is [187]:<br />

equilibrium shape<br />

J Cu<br />

actual shape<br />

55


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

56<br />

Thus,<br />

r<br />

L = 2<br />

13. 3<br />

Eq.4-4<br />

J D<br />

Eq.4-5<br />

Neglecting variations in molar volume with composition in the liquid lead inclusions,<br />

C ∆<br />

≈ 66 .<br />

r<br />

from Eq. 4-1 we can write:<br />

0 γ SL ⋅ VCu<br />

∆κ<br />

J ≈66 . D⋅CCu ⋅<br />

RT r<br />

Eq.4-6<br />

Writing Eq. 4-2 at the point <strong>of</strong> maximum curvature κmax , from Eq. 4-6, <strong>and</strong><br />

assuming for simplicity that<br />

∆κ = κmax −κmin ≈2κmax −κe<br />

we obtain:<br />

Eq.4-7<br />

0<br />

1 dκ<br />

max 1 d∆κ<br />

13. 3⋅D⋅CCu<br />

⋅γ SL ⋅VCu ⋅ΩCu<br />

≈ ≈ −<br />

3<br />

∆κdt 2 ∆κdt<br />

RT ⋅ r<br />

Eq.4-8<br />

Since derivatives <strong>of</strong> κ <strong>and</strong> ∆κ are the same (the equilibrium shape being fixed). We<br />

thus have:<br />

therefore<br />

( )<br />

0<br />

1 d∆κ<br />

d ln( ∆κ) 26. 6⋅D⋅C<br />

⋅γ ⋅V ⋅Ω<br />

= ≈ −<br />

3<br />

∆κ<br />

dt dt<br />

RT ⋅ r<br />

0<br />

0 ⎛ 26. 6 ⋅D⋅C ⋅γ ⋅V ⋅Ω<br />

∆κ = ∆κ<br />

exp⎜−<br />

3<br />

⎝ RT ⋅ r<br />

Cu SL Cu Cu<br />

Cu SL Cu Cu<br />

Eq.4-9<br />

Eq.4-10<br />

where ∆κ ο is the initial maximum difference in curvature along the interface.<br />

Mathematically, the time to full equilibration is infinite; however, after time<br />

⎞<br />

⋅ t⎟<br />


4.2 TIME FOR EQUILIBRATION<br />

3<br />

4 RT ⋅ r<br />

∆t<br />

=<br />

0<br />

26. 6 ⋅D⋅CCu ⋅γSL ⋅VCu ⋅ΩCu<br />

Eq.4-11<br />

curvature differences have reduced to 2% <strong>of</strong> their initial value. ∆t given by Eq. 4-11<br />

can, thus, be taken as a simple estimation <strong>of</strong> the time for shape equilibration <strong>of</strong> the<br />

liquid inclusion.<br />

57


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

58<br />

4.3 Dihedral angle measurement<br />

The dihedral angle, φ, is dictated by capillary equilibrium at a three-phase junction.<br />

We consider here the case <strong>of</strong> a liquid lead inclusion located on a planar grain<br />

boundary, Fig. 4-3. The relevant interfacial energies are the grain boundary energy,<br />

γ gb , <strong>and</strong> the solid/liquid interfacial energy, γ SL . We have:<br />

γ = 2 γ<br />

gb SL<br />

⎛ φ<br />

cos<br />

⎞<br />

⎝ 2⎠<br />

grain boundary plane<br />

Figure 4-3: Schematic view <strong>of</strong> an equilibrated lenticular inclusion situated on a planar<br />

grain boundary: (a) 3D view, <strong>and</strong> (b) detail: the cut is perpendicular to both the grain<br />

boundary plane <strong>and</strong> the triple line.<br />

Eq.4-12<br />

A meaningful dihedral angle can be measured once the <strong>microstructure</strong> is at<br />

equilibrium. In this case, the grain boundaries are planar, <strong>and</strong> the solid/liquid<br />

interface displays a constant curvature.<br />

4.3.1 Classical 2D method<br />

φ<br />

detail<br />

Dihedral angles are classically measured along the plane <strong>of</strong> a polished<br />

metallographic section. Since this plane crosses the inclusions r<strong>and</strong>omly, a<br />

distribution <strong>of</strong> apparent angles is obtained, Fig. 4-4. It is <strong>of</strong>ten assumed that the (real<br />

3D) angle φ takes a unique value across all inclusions, i.e. the interfacial energies γ gb<br />

<strong>and</strong> γ SL are isotropic. φ is then taken as the median <strong>of</strong> the distribution [8, 9].<br />

(a)<br />

Pb<br />

φ<br />

γ SL<br />

γ SL<br />

Cu<br />

detail<br />

(b)<br />

γ gb


4.3 DIHEDRAL ANGLE MEASUREMENT<br />

Figure 4-4: Distribution <strong>of</strong> apparent dihedral angles on a 2D cut. Insets show three<br />

different scenarii where an intergranular inclusion intersects the polished section. The<br />

apparent angle (here: 53°, 60°, <strong>and</strong> 62°) varies between 0° <strong>and</strong> 180°.<br />

We determined the median <strong>of</strong> more than 100 apparent angles visible in the SEM on a<br />

2D longitudinal cut for samples treated at 400, 820, 900 <strong>and</strong> 930 °C. The angles were<br />

computed from measured width <strong>and</strong> thickness values <strong>of</strong> each 2D lens according to<br />

[6]:<br />

cumulative frequency [%]<br />

100<br />

75<br />

50<br />

25<br />

cos<br />

Eq.4-13<br />

with a <strong>and</strong> b the respective width <strong>and</strong> thickness <strong>of</strong> a regular lens-shaped inclusion.<br />

φ<br />

2 2<br />

⎛ ⎞<br />

⎝<br />

2 2<br />

2⎠<br />

=<br />

a − b<br />

a + b<br />

4.3.2 New 3D method<br />

0<br />

gb<br />

53°<br />

0 30 60 90 120 150 180<br />

The inclusions are first dissolved using a procedure designed by superposing the<br />

Pourbaix diagrams for lead <strong>and</strong> <strong>copper</strong>, Fig. 4-5. This reveals that, at a pH <strong>of</strong> 4 <strong>and</strong> at<br />

zero electrical potential, <strong>copper</strong> is immune while lead is corroded. Agitated <strong>and</strong><br />

desaerated pure acetic acid is thus used to selectively dissolve exposed lead<br />

inclusions along a polished metallographic cut through the samples, leaving the<br />

<strong>copper</strong> essentially intact. The acid is desaerated by bubbling N 2 in order to prevent<br />

oxidation <strong>of</strong> the metal surface.<br />

(i)<br />

(i)<br />

(iii)<br />

(ii)<br />

gb 60° gb 62°<br />

apparent dihedral angle φ [°]<br />

(ii) (iii)<br />

59


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

60<br />

Figure 4-5: Pourbaix diagram <strong>of</strong> pure <strong>copper</strong> <strong>and</strong> pure lead [188]. The domains <strong>of</strong><br />

corrosion, passivation, <strong>and</strong> immunity are specified. At zero potential, <strong>and</strong> at a pH <strong>of</strong> 4,<br />

<strong>copper</strong> is immune, whereas lead is dissolved.<br />

Stereo image pairs are then obtained from such samples in scanning electron<br />

microscopy using the “MeX” stereophotogrammetry s<strong>of</strong>tware [189]. The s<strong>of</strong>tware<br />

combines images captured after dual tilting <strong>of</strong> the stage to produce 3D<br />

reconstructions <strong>of</strong> the sample surfaces along the surface <strong>of</strong> dissolved inclusions.<br />

An important parameter with samples such as these is the surface roughness that<br />

exists along the metal surface where inclusions were dissolved. The s<strong>of</strong>tware bases<br />

its reconstruction <strong>of</strong> the dissolved inclusion surface using the two captured SEM<br />

images on identifying <strong>and</strong> matching selected points along this surface. This step <strong>of</strong><br />

the process is easily conducted on fractographs (for which the s<strong>of</strong>tware was<br />

originally designed); however, it becomes near-impossible if the <strong>copper</strong> surface is<br />

perfectly smooth after lead dissolution. A proper compromise must therefore be<br />

reached between (i) leaving, after etching, a surface that is sufficiently smooth so that<br />

the true overall geometry <strong>of</strong> the inclusion surface is preserved, <strong>and</strong> (ii) producing a<br />

small degree <strong>of</strong> surface roughening so as to aid the s<strong>of</strong>tware in identifying selected<br />

points that can be “matched” along the inclusion surface across the two SEM<br />

micrographs. In this respect, partial solidification along the interface <strong>of</strong> the (small)<br />

amount <strong>of</strong> <strong>copper</strong> initially dissolved into the liquid inclusion is perhaps somewhat<br />

helpful as it can produce small features along the interface. Practically, an adequate<br />

inclusion surface was obtained by fine-tuning the time <strong>of</strong> the etch (3 hours), <strong>and</strong> also<br />

by selecting the “best” inclusions in terms <strong>of</strong> relative surface roughness — in<br />

addition to other criteria such as their relative location with respect to the plane <strong>of</strong>


4.3 DIHEDRAL ANGLE MEASUREMENT<br />

polish <strong>and</strong> the degree <strong>of</strong> planarity <strong>of</strong> the grain boundary along which they are located,<br />

Fig. 4-6.<br />

Figure 4-6: Typical solid/liquid interface after selective dissolution <strong>of</strong> a lead<br />

intergranular inclusion. The sample was rotated <strong>and</strong> tilted in the SEM so that the triple<br />

line appears as a horizontal line on the image (i.e. the grain boundary plane is normal<br />

to the image, <strong>and</strong> dual tilting allows optimal 3D reconstruction).<br />

From these images, the 3D dataset is calculated <strong>and</strong> processed so as to enable a three-<br />

dimensional digital reconstruction <strong>of</strong> the visible <strong>copper</strong>/lead interface along each<br />

inclusion, Fig. 4-7a. Assuming that the solid-liquid interfacial energy is isotropic (see<br />

below), spherical caps describing the solid-liquid interface over non-facetted<br />

orientations can then be fitted with the aid <strong>of</strong> the least square method to the measured<br />

digital inclusion envelope reconstructions, Fig. 4-7b. This determines the radius <strong>and</strong><br />

center <strong>of</strong> the two spheres delineating the inclusion.<br />

Figure 4-7: 3D digital reconstruction <strong>of</strong> the solid-liquid interface <strong>of</strong> an intergranular<br />

inclusion, <strong>and</strong> (b) data points from each cap with the two fitting spheres. The Cu-1Pb<br />

sample was heat treated at 930 °C; φ = 78°.<br />

(b)<br />

61


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

62<br />

Provided the inclusion is located along a planar grain boundary, knowing the center<br />

<strong>and</strong> radius <strong>of</strong> these two spheres, the true dihedral angle φ along the circular line <strong>of</strong><br />

intersection <strong>of</strong> the two inclusions is then easily deduced, Fig. 4-8. We have:<br />

α = 2( π/ 2−φ)+<br />

φ → φ = π −α<br />

with α calculated from the scalar product:<br />

cos( α ) =<br />

CA 1 ⋅ CA 2<br />

CA ⋅ CA<br />

1 2<br />

A<br />

α<br />

π/2 − φ π/2 − φ<br />

φ<br />

C 1<br />

C 2<br />

Figure 4-8: Determination <strong>of</strong> φ from the center <strong>and</strong> radius <strong>of</strong> the two spheres<br />

Eq.4-14<br />

Eq.4-15


4.4 Results<br />

4.4.1 Equilibration time<br />

4.4 RESULTS<br />

An approximate value <strong>of</strong> the equilibration time is computed from Eq. 4-11,<strong>and</strong> is<br />

reported on Table 4.1 <strong>and</strong> Fig. 4-9. Typical values for r, the radius <strong>of</strong> curvature <strong>of</strong> the<br />

inclusion surface, are 2 µm <strong>and</strong> 5 µm for intragranular <strong>and</strong> intergranular inclusions<br />

respectively <strong>of</strong> our specimens. The diffusion coefficient, D, is computed from the<br />

litterature [190]:<br />

−9<br />

D = ⋅<br />

⎛ −4853,<br />

1<br />

419, 35 10 exp<br />

⎞<br />

⎝ T ⎠<br />

2<br />

⎡m<br />

⎤<br />

⎢<br />

⎣ s<br />

⎥<br />

⎦<br />

Eq.4-16<br />

The solid/liquid interfacial energy is computed according to the regular solution<br />

model, where [116]:<br />

liquid solid 2 liquid 2<br />

γ SL = A+ B( CCu −CPb<br />

) ≈ A+ B( CCu<br />

)<br />

Eq.4-17<br />

liquid<br />

solid<br />

where <strong>and</strong> are respectively the solidus <strong>and</strong> liquidus atomic<br />

C Cu<br />

concentrations in the Cu-Pb system.<br />

From the phase diagram [31], <strong>and</strong> literature data [26], we have A = 190 mJ/m 2 <strong>and</strong><br />

B = 248 mJ/m 2 .<br />

C Pb<br />

For the molar volume <strong>of</strong> <strong>copper</strong>, we use its value at room temperature, thus:<br />

− ⎡ m ⎤<br />

∀ T VCu<br />

= 712⋅10⎢⎥ ⎣mol<br />

⎦<br />

6<br />

3<br />

,<br />

Eq.4-18<br />

63


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

64<br />

Table4.1 : Equilibration time, ∆t as a function <strong>of</strong> temperature for typical intragranular inclusions<br />

(r = 2 µm) <strong>and</strong> intergranular inclusions (r = 5 µm) calculated from Eq. 4-11.<br />

T [°C]<br />

It is worthwhile noting that the equilibration time scales with the radius <strong>of</strong> curvature<br />

<strong>of</strong> the inclusion to the third power. Thus a large inclusion with r = 15 µm will reach<br />

its equilibrium shape in 30 min at 900 °C or after 10 days at 400 °C.<br />

Figure 4-9: Time to equilibration for lead inclusions <strong>of</strong> different radii embedded in a<br />

<strong>copper</strong> matrix as a function <strong>of</strong> the annealing temperature, according to Eq. 4-11.<br />

4.4.2 γ SL isotropy<br />

D<br />

10 -9 [m 2 /s]<br />

γ SL [J/m 2 ]<br />

C Cu . ΩCu<br />

[at.%]<br />

∆t<br />

r = 2 µm<br />

∆t<br />

r = 5 µm<br />

350 0.18 0.437 0.3 ~ 1 h ~ 16 hr<br />

400 0.31 0.437 0.4 ~ 30 min ~ 8 hr<br />

600 1.63 0.428 2.6 ~ 1 min ~ 20 min<br />

800 4.56 0.385 10.0 ~ 10 s ~ 2 min<br />

900 6.62 0.340 20.7 ~ 3 s ~ 1 min<br />

955 8.08 0.290 34.7 ~ 2 s ~ 30 s<br />

Equilibration time: ∆ t<br />

[s]<br />

10 6<br />

1000<br />

1<br />

0.001<br />

r=100 nm<br />

Cu-Pb<br />

r=0.2 µm<br />

r=2 µm<br />

r=1 µm<br />

r=5 µm<br />

r=10 µm<br />

400 500 600 700 800 900<br />

Temperature [°C]<br />

1 week<br />

1 day<br />

1 min<br />

Observation <strong>of</strong> intragranular inclusions confirms that the solid-liquid interfacial<br />

energy is isotropic in this system: intragranular inclusions are spherical, save for<br />

isolated facets that form within limited solid angles at low temperature. Faceted<br />

1 hr<br />

1 s


4.4 RESULTS<br />

inclusions were observed in samples heat-treated at 400 °C, the lowest temperature<br />

used, Fig. 4-10.<br />

Figure 4-10: SEM image <strong>of</strong> a dissolved intragranular inclusion after heat treatment at<br />

400 °C. Three visible facets are indicated by arrows on the figure.<br />

At 930 °C, a mean deviation to sphericity <strong>of</strong> 3.2% was measured out <strong>of</strong> 16’000 data<br />

points from a single spherical inclusion envelope, Fig. 4-11.<br />

Figure 4-11: (a) 3D digital reconstruction <strong>of</strong> the solid-liquid interface <strong>of</strong> an<br />

intragranular inclusion, <strong>and</strong> (b) data points with the fitting sphere. The Cu-1Pb sample<br />

was heat treated at 930 °C.<br />

4.4.3 Measured dihedral angle<br />

Using the present dissolution method, the dihedral angle φ <strong>of</strong> several inclusions was<br />

measured in specimens equilibrated during the second step <strong>of</strong> heat-treatment at<br />

temperatures ranging from 400 to 970 °C; individual results are plotted in Fig. 4-12.<br />

This plot also gives results for the measurement <strong>of</strong> φ using many inclusions<br />

according to the ”classical” statistical method, as well as data (also gathered using the<br />

(b)<br />

65


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

66<br />

”classical” statistical method) for this system from two references in the literature,<br />

namely the work <strong>of</strong> Ikeuye <strong>and</strong> Smith who performed about 250 measurements with<br />

an optical microscope along a longitudinal cut <strong>of</strong> a Cu-1Pb alloy annealed in a<br />

reducing atmosphere [22], <strong>and</strong> Eustathopoulos et al. who made 100 measurements<br />

with similar experimental conditions on a high-purity Cu-10Pb alloy [13].<br />

Dihedral angle [ ]<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Ikeuye & Smith 49<br />

Eustathopoulos 74<br />

3-D method, this work<br />

2-D "classical" method, this work<br />

0<br />

300 400 500 600 700 800 900 1000<br />

Temperature [ C]<br />

Cu - 1Pb (5N)<br />

Figure 4-12: Solid/liquid dihedral angle, φ(T), as a function <strong>of</strong> heat treatment<br />

temperature for the Cu-Pb system. Our measurements, circular symbols, are compared<br />

with literature data [13, 22]. Open dotted circles are measurements on the alloy <strong>of</strong> this<br />

work using the classical statistical method. Solid circles are measured dihedral angles<br />

<strong>of</strong> individual single inclusions using the method presented here.<br />

It is seen that the dihedral angle measured on different inclusions with the method<br />

presented here varies significantly from inclusion to inclusion at each temperature.<br />

Adjacent intergranular inclusions located on a single planar grain boundary are, on<br />

the other h<strong>and</strong>, characterized by the same dihedral angle, Fig. 4-13.


4.4 RESULTS<br />

Figure 4-13: SEM micrographs displaying two sets <strong>of</strong> neighbouring dissolved<br />

inclusions situated on a planar grain boundary. The dihedral angle measured by 3D<br />

reconstruction <strong>of</strong> each inclusion is indicated on the figure next to the inclusion.<br />

Fig. 4-14 shows dihedral angle values measured from samples treated in flowing Ar-<br />

10H 2 compared to φ values obtained after our vacuum heat treatments detailed in<br />

Section 4.1.2.<br />

The two samples heat treated under flowing hydrogenated argon display lower φ<br />

values with similar variations from inclusion to inclusion at each temperature.<br />

Average φ values are still larger than those reported in the literature [13, 22].<br />

Dihedral angle φ [°]<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

2-D classical method 3-D method<br />

.<br />

.<br />

Cu - 1Pb (5N)<br />

literature data<br />

vacuum treatment<br />

vacuum treatment Ar-10H2 treatment .<br />

0<br />

300 400 500 600 700 800 900 1000<br />

Temperature [°C]<br />

Figure 4-14: Solid/liquid dihedral angle, φ(T), as a function <strong>of</strong> heat treatment<br />

temperature for the Cu-Pb system. Measurements from the two samples heat treated<br />

under flowing Ar-10H 2 , open squares, are compared with those from vacuum treatments<br />

<strong>and</strong> literature data [13, 22] presented in Fig. 4-12.<br />

.<br />

.<br />

.<br />

67


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

68<br />

4.5 Discussion<br />

4.5.1 Inclusion equilibration kinetics<br />

The kinetics <strong>of</strong> morphological changes driven by capillarity were extensively studied<br />

by Mullins. Most geometries were treated comprising grain boundary grooves <strong>and</strong><br />

spherical cavities [14, 191]. Despite the simplicity <strong>of</strong> its derivation, Eq. 4-11 is in<br />

complete accordance with relaxation times reported in [191] for spherical inclusions.<br />

The present simple estimation is thus validated.<br />

• For nanometric particles <strong>of</strong> lead in Al or Cu slightly above their melting point, the<br />

corresponding equilibration times are relatively short: around one milisecond for<br />

particles 10 nm in radius, around one second for particles 100 nm in radius; at higher<br />

temperatures they are shorter still, Fig. 4-15. Shape equilibration <strong>of</strong> such particles in<br />

a hot-stage electron microscope is thus rapid [19, 185]. We note in passing that this<br />

reinforces the interpretation <strong>of</strong>fered by Gabrisch et al. for particle shape hysteresis<br />

observed with very small lead inclusions in Al upon thermal cycling as being due to<br />

interface kinetics (as opposed to diffusion) [19].<br />

Equilibration time: ∆ t [s]<br />

10 6<br />

1000<br />

1<br />

0.001<br />

r=1 µm<br />

r=100 nm<br />

r=10 nm<br />

r=0.2 µm<br />

r=5 µm<br />

r=2 µm<br />

350 400 450 500 550 600 650<br />

Temperature [°C]<br />

r=10 µm<br />

Al-Pb<br />

1 week<br />

1 day<br />

1 min<br />

Figure 4-15: Time to equilibration for lead inclusions <strong>of</strong> different radii embedded in an<br />

aluminum matrix as a function <strong>of</strong> the annealing temperature, according to Eq. 4-11.<br />

Thermodynamical data are from [192, 193].<br />

1 hr<br />

1 s


4.5 DISCUSSION<br />

• In as-cast melt-spun ribbons <strong>of</strong> <strong>alloys</strong> such as Al-Pb or Cu-Pb, cooling rates during<br />

solidification are typically on the order <strong>of</strong> 10 3 to 10 6 K/s, depending on ribbon size,<br />

<strong>and</strong> also on the moment <strong>of</strong> ribbon detachment from the wheel, something that is<br />

likely to occur well before final solidification <strong>of</strong> liquid inclusions in deep monotectic<br />

systems such as Al/Pb. Freezing ranges being at most a few hundred K, the time<br />

available for equilibration <strong>of</strong> the liquid inclusion shape is then on the order <strong>of</strong> 10 -5 to<br />

10 -1 s within the ribbons: depending on inclusion size <strong>and</strong> the precise value <strong>of</strong> the<br />

cooling rate, inclusions 10 to 100 nm wide may or may not have had time to reach<br />

their equilibrium shape right before they solidify.<br />

• In chill-cast samples, which typically solidify in a matter <strong>of</strong> several seconds, there is<br />

clearly no time for inclusions larger than 1 µm to equilibrate before they solidify,<br />

Fig. 4-15. This explains why lead particles in chill-cast samples Al/Pb <strong>of</strong> Ref. [185],<br />

which were above 1 µm in diameter, were only partially faceted in the as-cast alloy,<br />

<strong>and</strong> became more faceted after heat-treatment for 2’000 s at 350 °C (a time sufficient<br />

to equilibrate particles on the order <strong>of</strong> one micrometre wide).<br />

A size dependence has been reported for the shape <strong>of</strong> inclusions [194]. After<br />

prolonged anneals followed by a water quench, namely 100 h at 410 °C or 610 °C,<br />

inclusions below a critical diameter <strong>of</strong> about 0.4 µm were found to be faceted . On<br />

the other h<strong>and</strong>, larger inclusions were found to be spherical, as were Pb inclusions<br />

less than 200 nm in size observed in situ at 500 °C in a hot-stage TEM [19]. The<br />

cooling procedure <strong>of</strong> McCormick may thus not have been rapid enough to freeze the<br />

high temperature equilibrium shape <strong>of</strong> such micrometric (r = 0,2 µm) inclusions.<br />

Namely, considering the Al diffusion coefficient in liquid Pb, as well as the solid/<br />

liquid interfacial energy from [193], we estimate the time to equilibration for such<br />

inclusions as being in the order <strong>of</strong> 1 <strong>and</strong> 0.1 s, at 450 <strong>and</strong> 550 °C respectively, Fig. 4-<br />

15.<br />

The time <strong>of</strong> spheroidization, τ, <strong>of</strong> elongated Pb inclusions in swaged Al-5 wt.% Pb<br />

were found to follow an r 3 law, indicating a process controlled by volume diffusion.<br />

This was observed both by X-ray microphotography on micrometric inclusions<br />

69


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

70<br />

(Ø 25 - 75 µm) after interrupted anneals, Fig. 4-16, <strong>and</strong> in situ using hot-stage TEM<br />

on smaller inclusions, 0.1 to 1 µm in size [193, 195].<br />

0 min 855 min 1965 min<br />

Figure 4-16: Series <strong>of</strong> micrographs showing the change in shape <strong>of</strong> Pb inclusions in Al<br />

as a function <strong>of</strong> time at 625 °C [193].<br />

The time <strong>of</strong> spheroidization was found to depend on the equilibrium solubility, the<br />

diffusion coefficient, the solid/liquid interfacial energy, <strong>and</strong> the temperature. This is<br />

consistent with Eq. 4-11.<br />

3<br />

kBT⋅r τ ∝ 0 2<br />

D⋅CCu ⋅γSL ⋅ΩCu<br />

Eq.4-19<br />

Moreover the proportional constant that fits their experimental data is about 1/3. This<br />

value compares fairly well with the factor 4/26.6 in Eq. 4-11.<br />

The shape <strong>of</strong> micrometric inclusions is not influenced by misfit strains. This is unlike<br />

very small inclusions where, as shown in the Al/Pb system, “magic size” effects are<br />

observed below a diameter around 35 nm [17]. With inclusions larger than this, the<br />

relative importance <strong>of</strong> strain energy increases in the purely elastic regime, but so does<br />

the ease <strong>of</strong> misfit dislocation nucleation in the matrix [196]. Strain energy effects are<br />

thus expected to be minimized with larger inclusions.<br />

As an overall conclusion, the equilibrium shape <strong>of</strong> submicronic inclusions must be<br />

observed at elevated temperature during in-situ heating in the TEM. Larger micron-<br />

sized inclusions, on the other h<strong>and</strong>, can be observed at room temperature after rapid<br />

cooling from a higher annealing temperature.


4.5 DISCUSSION<br />

Focusing now on the Pb/Cu system we find that, for particles 5 µm in radius, the time<br />

for equilibration is around 8 hours at 400 °C <strong>and</strong> around one minute at 900°C,<br />

Table 4.1 <strong>and</strong> Fig. 4-9. This agrees well with our observations that, for reproducible<br />

shape equilibration <strong>of</strong> Pb inclusions roughly 10 µm wide in Cu, heat-treatment times<br />

<strong>of</strong> 60 h at 400 °C or 1 h at 900 °C are appropriate. After such a heat-treatment, if<br />

samples are quenched <strong>and</strong> reach room-temperature in a few seconds, there is no time<br />

for the shape <strong>of</strong> such large liquid inclusions to change significantly. As a<br />

consequence, interface geometries <strong>and</strong> dihedral angles measured using the technique<br />

described above correspond indeed to the capillary equilibrium shape <strong>of</strong> the inclusion<br />

at the heat-treatment temperature.<br />

4.5.2 Solid/liquid interfacial energy<br />

Apart from facets that appear around a few crystallographic directions at relatively<br />

low temperature (400 °C, Fig. 4-10), the solid Cu-liquid Pb interfacial energy is<br />

found to be isotropic in the temperature range explored. The small measured<br />

deviation <strong>of</strong> intragranular inclusions from a perfectly spherical shape can easily be<br />

attributed to roughness along the surface <strong>of</strong> the dissolved inclusions, <strong>and</strong>/or to error<br />

in the measurement <strong>and</strong> digital reconstruction <strong>of</strong> the interface.<br />

This agrees with the literature. It is known that Al/Pb <strong>and</strong> Cu/Pb systems are<br />

thermodynamically quite similar [185, 186, 197, 198]. In the former, more<br />

extensively characterized system, when the inclusions are liquid, partial faceting<br />

occurs along {111} planes below a temperature that increases with particle diameter<br />

in the range <strong>of</strong> 30 to 300 nm, approaching 500 °C for larger particles [19, 185].<br />

Above 500 °C, the interfacial energy is also documented to become isotropic [185].<br />

These observations for Al/Pb are fully consistent with present observations on larger<br />

lead inclusions in <strong>copper</strong>.<br />

71


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

72<br />

4.5.3 Dihedral angle measurement<br />

The method presented here yields measurements <strong>of</strong> true (3D) dihedral angles for<br />

individual inclusions that are significantly larger (10 µm) than those typically used in<br />

transmission electron microscopy (≤ 300 nm). Size effects, such as a size-dependence<br />

in particle shape or roughening temperature, which are characteristic <strong>of</strong> submicron<br />

inclusions <strong>and</strong> become exacerbated at very small sizes, are thus absent here. The<br />

triple line energy, which can intervene in nanometric inclusions, is also safely<br />

neglected.<br />

The dihedral angle value obtained with the present method results from an<br />

extrapolation <strong>of</strong> overall smooth inclusion interfaces to the triple line, essentially as it<br />

is done in the sessile drop technique for contact angles. As such, the measurement<br />

yields a ”macroscopic” angle φ that reflects overall energetic optimization <strong>of</strong> the<br />

inclusion shape – as implicit in the derivation <strong>of</strong> a capillary equilibrium equation<br />

such as Eq. 4-12. This implies for example that the effect <strong>of</strong> grain boundary facets,<br />

which are important at the atomic length scale <strong>and</strong> affect the local dihedral angle<br />

where they meet the inclusion apex, are neither important nor captured. To resolve<br />

details <strong>of</strong> the inclusions’ shape at the atomic scale, the use <strong>of</strong> scanning tunneling<br />

microscopy, STM, would be needed [199].<br />

The reproducibility <strong>of</strong> the method is apparent in analysis <strong>of</strong> several inclusions located<br />

along the same grain boundary, Fig. 4-13: φ is constant to within two degrees across<br />

several inclusions. Compared with ”statistical” methods, the present measurements<br />

are thus quite precise. We note in passing that φ values in Fig. 4-12 are near 90°,<br />

which as mentioned above is deemed the more difficult value for measurement <strong>of</strong> φ<br />

using ”classical” statistical methods. When φ is very low, on the other h<strong>and</strong>, the<br />

present method is less convenient because the stereoscopic reconstruction becomes<br />

challenging if the inclusion envelopes feature steep <strong>and</strong> narrow channel-like walls.<br />

4.5.4 Dihedral angles in Cu-Pb<br />

Variations in the value <strong>of</strong> φ that we find at each specific temperature using the present<br />

method do not reflect experimental error since adjacent intergranular inclusions


4.5 DISCUSSION<br />

display the same angle. The reason is rather that the grain boundary energy is far<br />

more anisotropic than the solid-liquid interfacial energy, in accordance with Clarke<br />

<strong>and</strong> Gees’ expectations [163]. If we consider the measurements made on the<br />

specimen annealed at 930 °C (which shows the largest degree <strong>of</strong> variation in<br />

measured φ), the angles vary between 64 <strong>and</strong> 103°. From Eq. 4-12, this implies a<br />

ratio <strong>of</strong> 4/3 between the corresponding extreme values <strong>of</strong> grain boundary energy, γ gb .<br />

Such a range <strong>of</strong> variation <strong>of</strong> high-angle grain boundary energy in <strong>copper</strong> is consistent<br />

with the literature, including both experimental data <strong>and</strong> results from atomistic<br />

computer simulations [149, 200].<br />

We note that such variations in φ as a result <strong>of</strong> high-angle grain boundary energy<br />

anisotropy have been noted before. Protsenko et al. observed a distribution in groove<br />

depths in polycrystalline nickel wetted by silver at 1040 °C <strong>and</strong> attributed this to the<br />

anisotropy <strong>of</strong> the grain boundary energy [201]. Disregarding any ”special” low<br />

energy grain boundary, a ~7/5 ratio between the higher <strong>and</strong> lower grain boundary<br />

energies in nickel (another fcc metal) can be estimated from their work.<br />

The disagreement between dihedral angle values we measure <strong>and</strong> the data available<br />

in the literature for Cu-Pb [13, 22] does not seem to arise from the method used: our<br />

”2D classical” method measurements yield values that are in good agreement with<br />

the average <strong>of</strong> our ”3D” measurements, Fig. 4-12, in accordance with De H<strong>of</strong>fs’<br />

mathematical treatment [12]. The disagreement therefore must arise from the<br />

presence <strong>of</strong> grain boundary contaminants in our alloy compared to the higher-purity<br />

samples <strong>of</strong> Refs. [13, 22].<br />

Results from the heat-treatments performed under flowing hydrogenated argon are in<br />

accordance with this conclusion. Indeed φ values for purified samples lie on average<br />

between data reported in the literature <strong>and</strong> values measured after vacuum treatment,<br />

Fig. 4-14. Contaminants such as sulphur or oxygen seem therefore to have been<br />

scavenged at least to some extent by hydrogen in the treatment.<br />

Since solute enrichment factors are reported to be more than one order <strong>of</strong> magnitude<br />

higher at grain boundaries than at solid-liquid interfaces [202], a segregated element<br />

will reduce γ gb more significantly than γ SL , <strong>and</strong> in turn increase the dihedral angle φ,<br />

Eq. 4-12. Waterhouse <strong>and</strong> Grubb also provide data documenting an increase in φ due<br />

73


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

74<br />

to grain boundary contamination in Cu-Pb annealed at 650 °C [34]: they report a<br />

value <strong>of</strong> 93°, which is close to values found here <strong>and</strong> well above the value <strong>of</strong> 60°<br />

documented in Refs. [13, 22] for high-purity Cu-Pb. According to Waterhouse, in<br />

reply to Stickels’ comments, the observed increase <strong>of</strong> φ is due to the presence <strong>of</strong><br />

0.08% P added as a desoxidizing agent, the phosphorous segregation thus reducing<br />

the γ gb /2γ SL ratio [35]. We note, however, that oxygen desegregation per se is also<br />

reported to cause a decrease <strong>of</strong> the dihedral angle <strong>of</strong> intergranular SiO 2 inclusions in<br />

internally oxidized <strong>copper</strong>, following a 1000 °C 24 h treatment in vacuum [6, 203].<br />

In one instance, though, it may be that disagreement between our results <strong>and</strong><br />

literature data is due to the method used. This is at 400 °C, where inclusions are<br />

facetted. Indeed, facets render the classical method fairly difficult to apply:<br />

”classically” measured values <strong>of</strong> φ in the literature may therefore be underestimated.<br />

Even in the case <strong>of</strong> uncontaminated Cu-Pb <strong>alloys</strong>, lead may segregate at grain<br />

boundaries at such an intermediate temperature, as when it spreads on a free surface<br />

<strong>of</strong> <strong>copper</strong> <strong>and</strong> forms a monolayer ahead <strong>of</strong> a liquid droplet <strong>of</strong> lead [177-179]. This<br />

would also produce a decrease <strong>of</strong> the apparent grain boundary energy, which may<br />

increase the dihedral angle, even in samples <strong>of</strong> very high purity.<br />

Segregation <strong>of</strong> dissolved lead at the <strong>copper</strong> grain boundaries was considered as<br />

negligible by Camel, from an analysis <strong>of</strong> the evolution <strong>of</strong> the dihedral angle with<br />

temperature [27]. This is surely true at high temperature. Indeed the evolution <strong>of</strong> φ<br />

with temperature can be back-calculated, Fig. 4-17, (i) from φ values at specific<br />

(relatively high) temperatures available in the literature [26], (ii) from an assumed<br />

linear decrease <strong>of</strong> γ gb with temperature, Eq. 4-20 [23], <strong>and</strong> (iii) from γ SL calculated<br />

according to the regular solution, Eq. 4-17.<br />

γ gb(<br />

T)<br />

γ gb( T) = γ gb(<br />

T ) + T T . ( T C) mJ<br />

T<br />

∂<br />

0 ⋅( − 0)= 620 −04⋅ − 947°<br />

[ ]<br />

∂<br />

Eq.4-20


Dihedral angle φ [°]<br />

4.5 DISCUSSION<br />

Figure 4-17: Evolution <strong>of</strong> φ back calculated from literature data [26] (squares), <strong>and</strong> the<br />

estimated evolution <strong>of</strong> γ SL <strong>and</strong> γ gb for the Cu-Pb system. Ikeuyes’ data [22] are added<br />

but were not considered for the calculations.<br />

Back-calculated φ values drop to zero at about 940 °C; this reflects the marked<br />

C Cu<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

back-calculated φ evolution<br />

literature data: Coudurier 77<br />

literature data: Ikeuye 49<br />

0<br />

300 400 500 600 700 800 900 1000<br />

Temperature [°C]<br />

liquid<br />

increase <strong>of</strong> at high temperature, Fig. 4-1. On the other h<strong>and</strong>, below 600 °C,<br />

back-calculated φ values are governed by the estimated evolution <strong>of</strong> γ gb , which<br />

causes φ to decrease. At these temperatures, the predicted evolution is thus in clear<br />

disagreement with both our measured φ values, <strong>and</strong> those <strong>of</strong> Ikeuye [22], Fig. 4-18.<br />

It is therefore possible that, below 600°C, Pb segregation to the Cu grain boundaries<br />

induces a γ gb reduction that increases φ above predictions <strong>of</strong> Eq. 4-20.<br />

75


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

76<br />

Dihedral angle φ [°]<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

literature data<br />

back-calculated φ<br />

Cu - 1Pb (5N)<br />

0<br />

300 400 500 600 700 800 900 1000<br />

Temperature [°C]<br />

our data<br />

vacuum treatment<br />

Ar-10H 2 treatment<br />

Figure 4-18: Solid/liquid dihedral angle, φ(T), as a function <strong>of</strong> heat treatment<br />

temperature for the Cu-Pb system. Back-calculated φ values (grain boundary<br />

segregation is not taken into account), solid line, are compared with literature data [13,<br />

22] <strong>and</strong> our data presented in Fig. 4-14.


4.6 Conclusion<br />

4.6 CONCLUSION<br />

A new method is presented for the determination <strong>of</strong> the dihedral angle <strong>of</strong> individual<br />

intergranular liquid inclusions; it is based on quantitative scanning electron<br />

microscopic analysis <strong>of</strong> metallographic surfaces along which quenched inclusions<br />

have been dissolved. The angle is derived from a mathematical fitting <strong>of</strong> the solid/<br />

liquid interface geometry around individual inclusions. It reflects the value dictated<br />

by global energy minimization <strong>of</strong> the inclusion shape under capillary forces, as<br />

expressed by Eq. 4-12. As such, this method parallels the sessile drop method for<br />

contact angle measurement, where the drop shape is used to deduce ”macroscopic”<br />

contact angles.<br />

Compared with classical methods <strong>of</strong> dihedral angle measurement, no statistical<br />

treatment <strong>of</strong> the data is needed. The method also overcomes some shortcomings <strong>of</strong><br />

other techniques, such as the direct measurement <strong>of</strong> dihedral angles at grain boundary<br />

grooves or in transmission electron microscopy.<br />

We show that the Cu/Pb solid-liquid interfacial energy is isotropic above 400 °C,<br />

while it exhibits a slight tendency for faceting at 400 °C. Even in the presence <strong>of</strong><br />

facets, the method presented here leads to reliable measurements <strong>of</strong> the true dihedral<br />

angle φ.<br />

For a specific temperature, we show that φ is not unique; this reflects the fact that<br />

high-angle grain boundary energies vary in <strong>copper</strong>.<br />

77


CHAPTER 4. EQUILIBRIUM DIHEDRAL ANGLE OF PB IN CU<br />

78


Chapter 5<br />

5 Influence <strong>of</strong> stress on the shape <strong>of</strong> an embedded liquid inclusion<br />

In this chapter, the combined effect <strong>of</strong> stress <strong>and</strong> capillarity on the equilibrium shape<br />

<strong>of</strong> a liquid inclusion is studied. This shape is determined by global minimization <strong>of</strong><br />

the elastic strain energy, quantified by means <strong>of</strong> the Eshelby method, <strong>and</strong> interfacial<br />

energies.<br />

Spheroidal intragranular inclusions are first considered. Predicted shapes are<br />

confronted with actual 3-D shapes imaged <strong>and</strong> quantified following the procedure<br />

described in Chapter 4. The shape <strong>of</strong> intergranular inclusions is also studied. It is<br />

found that the apparent global dihedral angle decreases under an applied remote<br />

stress. This is in qualitative accordance with our theoretical predictions; quantitative<br />

agreement is, however, more tenuous, which is not unexpected given simplifications<br />

made in analysis.<br />

79


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

80<br />

5.1 Theoretical predictions<br />

5.1.1 The Eshelby method<br />

The local elastic stress <strong>and</strong> strain <strong>of</strong> an inclusion embedded in a linear elastic<br />

stressed matrix can be approximated by the Eshelby calculation, which assumes that<br />

the inclusion is ellipsoidal in shape, Fig. 5-1. It is demonstrated that in such an<br />

inclusion, the stress <strong>and</strong> elastic strain are homogeneous [37]. The stress field outside<br />

the inclusion can also be rigourously calculated by adding the stress fields <strong>of</strong> two<br />

simplified situations [38]. The interaction strain energy, ∆W, due to the presence <strong>of</strong><br />

the inclusion can also be quantified. The procedure is briefly summarized in what<br />

follows.<br />

Figure 5-1: Oblate spheroidal inclusion with vertical rotation axis, the small <strong>and</strong> large<br />

semi-axis are b <strong>and</strong> a respectively. The direction <strong>of</strong> the applied uniaxial stress is parallel<br />

to the vertical axis.<br />

We use the Eshelby formalism, which is didactically described in [204]. The effect <strong>of</strong><br />

dissimilar elastic properties between the inclusion <strong>and</strong> the matrix can be accounted<br />

for by an equivalent transformation strain, ε* (or eigenstrain). The latter induces (in<br />

the virtual case <strong>of</strong> similar elastic properties between the ”ghost” homogeneous<br />

inclusion <strong>of</strong> same shape <strong>and</strong> the matrix) the same internal stresses in the matrix as ε,<br />

the elastic strain whithin the actual (inhomogeneous) inclusion. The eigenstrain is<br />

deduced from the elastic properties <strong>of</strong> the matrix <strong>and</strong> the inclusion, the applied<br />

remote stress, <strong>and</strong> the Eshelby tensor, S, defined by the inclusion shape <strong>and</strong> the<br />

Poisson ratio <strong>of</strong> the matrix:<br />

b<br />

a<br />

σapplied


5.1 THEORETICAL PREDICTIONS<br />

∗<br />

ε =<br />

−1<br />

∗ ( C−C )⋅S−C ∗ ( C<br />

−1<br />

C)⋅ C ⋅σapp<br />

Eq.5-1<br />

where C <strong>and</strong> C* are the rigidity tensors <strong>of</strong> the matrix <strong>and</strong> inclusion, respectively. The<br />

actual elastic stress <strong>and</strong> strain within the inclusion are respectively ε <strong>and</strong> σ :<br />

<strong>and</strong><br />

Eq.5-2<br />

σ = C ⋅ε<br />

Eq.5-3<br />

It is shown in [38] that, for an ellipsoidal inclusion, the interaction energy ∆W can<br />

*<br />

easily be computed from the eigenstrain, the inclusion volume, V, <strong>and</strong> the applied<br />

remote stress, σ app :<br />

1<br />

∗<br />

∆W =− V⋅σapp ⋅ε<br />

2<br />

Eq.5-4<br />

This interaction energy is counted as negative since it corresponds to work performed<br />

by an external load.<br />

5.1.2 Inclusion shape<br />

[ ] ⋅ −<br />

−1<br />

* 0 *<br />

ε = C ⋅ σ + S⋅ ε = ε + S⋅ε<br />

app<br />

( )<br />

In order to determine the equilibrium shape <strong>of</strong> an inclusion, one can consider the<br />

overall energy associated with the inclusion. Energy minimization as a function <strong>of</strong><br />

shape then allows to determine the equilibrium situation. Two contributions must be<br />

considered: (i) the strain energy ∆W, <strong>and</strong> (ii) the overall interfacial energy, Γ .<br />

The latter is only dependent on the interfacial energy, γ, <strong>and</strong> on the inclusion surface<br />

area, A. In the case <strong>of</strong> a spheroidal inclusion,<br />

A =<br />

4π<br />

c<br />

⋅ ⎛<br />

3 V 3 V 3 ⎛1+<br />

1−c<br />

⋅⎜⋅ ⋅ 1 − c + ⋅V⋅c⋅ln 2<br />

− c c<br />

⎜<br />

1 ⎝ 2<br />

4 ⎝1−<br />

1−c<br />

3 2<br />

where c is the aspect ratio <strong>of</strong> the inclusion defined as:<br />

2<br />

2<br />

⎞⎞<br />

⎟⎟<br />

⎠⎠<br />

Eq.5-5<br />

81


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

82<br />

b<br />

c =<br />

a<br />

Eq.5-6<br />

Thus we can express Γ, which is counted positive since an increase <strong>of</strong> the interfacial<br />

area consumes energy:<br />

Γ= γ ⋅A<br />

Eq.5-7<br />

The overall energy due to the presence <strong>of</strong> the inclusion is thus a function <strong>of</strong> its<br />

volume <strong>and</strong> shape, the applied remote stress, the elastic properties <strong>of</strong> matrix <strong>and</strong><br />

inclusion, <strong>and</strong> the interfacial energy.<br />

We make in what follows the assumption that the inclusion remains spheroidal, <strong>and</strong><br />

optimize its shape by minimizing the total excess energy associated with its presence<br />

in a stressed solid as a function <strong>of</strong> its aspect ratio at constant volume. This approach<br />

is classical for this type <strong>of</strong> problem [36, 39, 40] as it eases the problem considerably:<br />

the calculation is reduced to minimization <strong>of</strong> an analytical scalar function <strong>of</strong> a single<br />

scalar variable. It must, however, be borne in mind that this assumption, coupled with<br />

the underlying description <strong>of</strong> the surrounding matrix as a linear elastic solid (which is<br />

obviously not true for <strong>copper</strong> near an inclusion at elevated temperature), is a<br />

significant simplification <strong>of</strong> what is, in fact, a fully three-dimensional free boundary<br />

problem <strong>of</strong> the optimal shape <strong>of</strong> a liquid inclusion embedded in an elastoviscoplastic<br />

matrix.<br />

At equilibrium, (∆W + Γ) is at a minimum:<br />

with<br />

d( ∆W + Γ)<br />

= 0<br />

dc<br />

Eq.5-8<br />

σ app<br />

∆W = F() c ⋅V⋅ E<br />

Eq.5-9<br />

where F(c) is a function <strong>of</strong> the elastic properties <strong>of</strong> both the matrix <strong>and</strong> the inclusion,<br />

2<br />

<strong>and</strong> <strong>of</strong> the shape <strong>of</strong> the latter. Moreover,


Γ= Gc () ⋅V ⋅<br />

2 3 γ<br />

where G(c) is only a function <strong>of</strong> the geometry <strong>of</strong> the inclusion.<br />

5.1 THEORETICAL PREDICTIONS<br />

Thus equilibrium can be described by a single adimensional factor, Λ:<br />

Eq.5-10<br />

dG() c<br />

2 1/ 3<br />

Λ=− dc σ app ⋅V<br />

=<br />

dF() c γ ⋅ E<br />

dc<br />

Eq.5-11<br />

This last parameter describes a global equilibrium at the inclusion length scale. Note<br />

that this adimensional factor is defined here, Eq. 5-11, by taking into account the<br />

volume V <strong>of</strong> the inclusion to the power one third, whereas Λ defined in Eq. 3-4<br />

considers the initial pore diameter a 0 . Assuming a lenticular shape for the inclusion,<br />

the proportional factor is deduced from Eq. 5-12:<br />

V<br />

a 2<br />

= 2−3cos ⎛ ⎞<br />

cos<br />

2 3 ⎝ 2⎠ 2<br />

+<br />

⎡ π ⎛ φ ⎛ φ⎞⎞⎤<br />

⎢ ⎜<br />

⎝<br />

⎝ ⎠<br />

⎟<br />

⎣<br />

⎠⎥<br />

⎦<br />

1/ 3 0 3<br />

5.1.3 Intergranular inclusion<br />

Eq.5-12<br />

In the case <strong>of</strong> an intergranular inclusion, the grain boundary energy, γ gb must<br />

additionally be taken into account. Thus the overall interfacial energy due to the<br />

presence <strong>of</strong> the grain boundary inclusion is as follows:<br />

2<br />

Γ= γ ⋅A−γ gb ⋅π ⋅a<br />

Eq.5-13<br />

Note that the contribution <strong>of</strong> the grain boundary energy is counted as negative since a<br />

reduction <strong>of</strong> the grain boundary area restores energy, <strong>and</strong><br />

φ<br />

γ gb = 2γ<br />

⋅cos<br />

⎛ ⎞<br />

⎝ 2⎠<br />

Eq.5-14<br />

The actual equilibrium shape <strong>of</strong> an intergranular inclusion is not ellipsoidal. We<br />

consider two approximate shapes for the calculation <strong>of</strong> the interfacial area: (i) two<br />

spherical caps intersecting at the grain boundary plane, or (ii) the st<strong>and</strong>ard Eshelby<br />

1/ 3<br />

83


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

84<br />

ellipsoid. In each case the aspect ratio c is defined as the ratio between the small <strong>and</strong><br />

large axis, Fig. 5-2.<br />

Figure 5-2: An intergranular inclusion imbedded in a stressed material. The shape <strong>of</strong><br />

the inclusion is approximated by two symetric spherical caps (dashed contour) or by an<br />

oblate spheroid.<br />

The first approximation, which implies a constant curvature, matches the actual<br />

shape <strong>of</strong> the intergranular inclusion under stress-free conditions only. Under stress,<br />

local stresses are not constant along the inclusion/matrix interface; this implies in<br />

turn a non-uniform κ along the liquid/solid interface, Eq. 5-18.<br />

The surface area <strong>of</strong> the intergranular inclusion is calculated, respectively, as:<br />

or<br />

/<br />

A = V ⋅2π⋅ 1+<br />

c<br />

lenticular<br />

b<br />

( )<br />

2 3 2<br />

Eq.5-15<br />

Aellipsoidal<br />

c<br />

V<br />

c<br />

V<br />

c<br />

c V c<br />

c<br />

c<br />

Eq.5-16<br />

In both cases, (i) & (ii), the equilibrium shape <strong>of</strong> the intergranular inclusion can be<br />

=<br />

3<br />

4π<br />

⋅<br />

3<br />

2<br />

1 −<br />

⎛ 3<br />

⋅⎜⋅ ⋅<br />

⎝ 2<br />

⎛<br />

2 3 1+ 1 − + ⋅ ⋅ ⋅ln<br />

4<br />

⎜<br />

⎝1−<br />

2<br />

1−<br />

⎞⎞<br />

2 ⎟⎟<br />

1−<br />

⎠⎠<br />

described by the same adimensional factor Λ defined in Eq. 5-11.<br />

a<br />

⎡ 3<br />

⎢<br />

⎣⎢<br />

π ⋅c⋅ 3 + c<br />

2 ( )<br />

applied<br />

grain boundary<br />

The ideal case, where the remote stress direction is normal to one grain boundary<br />

plane along which an inclusion is located, is rarely met, Fig. 5-3 (a). Rotation <strong>and</strong> tilt<br />

⎤<br />

⎥<br />

⎦⎥<br />

2/ 3<br />

σ


5.1 THEORETICAL PREDICTIONS<br />

angles, respectively ξ <strong>and</strong> θ illustrated on Fig. 5-3 (b), must be defined in order to<br />

account for actual situations in a material subjected to uniaxial tensile stress.<br />

(a) (b)<br />

Figure 5-3: Schematic representation <strong>of</strong> an intergranular inclusion imbedded in an<br />

uniaxially stressed material. (a) The grain boundary plane is normal to the remote<br />

stress axis, whereas in (b), its orientation is r<strong>and</strong>om.<br />

Accordingly the stress tensor σσ kl considered in the determination <strong>of</strong> ∆W with the<br />

Eshelby formalism is transformed from the x, y, z system <strong>of</strong> axes to the x’’, y’’, z’’<br />

axes by successive rotation <strong>of</strong> ξ about the x axis (to x’, y’, z’ ), <strong>and</strong> θ about the y’<br />

axis (in the Einstein suffix notation):<br />

σσ = a a σσ<br />

kl ki lj ij<br />

a ij<br />

where is the direction cosine matrix.<br />

Eq.5-17<br />

The shape <strong>of</strong> the inclusion is again approximated by an oblate spheroid having the<br />

grain boundary plane as symmetric plane.<br />

5.1.4 Equilibration time<br />

σ app<br />

σ app<br />

The equilibrium shape <strong>of</strong> an inclusion is met once the chemical potential is<br />

homogeneous throughout the whole inclusion. At constant temperature <strong>and</strong><br />

σ app<br />

σ app<br />

ξ<br />

θ<br />

85


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

86<br />

homogeneous chemical composition at the interface, the chemical potential µ [36] is,<br />

quoting [205]:<br />

µκσ ( , )= µ 0 + γΩκ+ Ωw<br />

Eq.5-18<br />

where µ 0 is the equilibrium chemical potential <strong>of</strong> a planar interface under stress-free<br />

conditions, Ω is the atomic volume, γ the interfacial energy, κ the curvature; w is the<br />

local elastic strain energy. Note that κ is negative, since the <strong>copper</strong> interface <strong>of</strong><br />

interest is concave.<br />

The local driving force for shape equilibration is therefore dependent on both the<br />

local curvature <strong>and</strong> the local stress. The time for shape equilibration is therefore more<br />

difficult to estimate for a liquid inclusion embedded in a stressed matrix; however,<br />

since the driving force comprises the same capillary term as in the zero-stress<br />

problem, as a first rough estimate, we use Eq. 4-11.


5.2 Experimental procedures<br />

5.2 EXPERIMENTAL PROCEDURES<br />

In order to measure the influence <strong>of</strong> stress on the shape <strong>of</strong> an inclusion, interrupted<br />

creep tests were performed on a Cu-Pb alloy at 400 °C (the inclusions are molten at<br />

this temperature). The 3D shape <strong>of</strong> the resolidified inclusions was observed on<br />

polished microsections <strong>and</strong> the ”global” dihedral angle was measured with the<br />

method described in Section 4.3.2.<br />

5.2.1 Material<br />

C99 is the industrial alloy used, a <strong>leaded</strong> <strong>copper</strong> containing 0.8 to 1.2 wt.% Pb <strong>and</strong><br />

0.01 to 0.04 wt.% P added as a desoxidizing agent. This material was provided by our<br />

industrial partner, Swissmetal Boillat, Reconvilier, Switzerl<strong>and</strong>. Processing<br />

comprised (i) billet casting, (ii) hot extrusion, <strong>and</strong> (iii) room temperature wire<br />

drawing.<br />

A heat treatment at 900 °C for 2 h under primary dynamic vacuum (p ≈ 10 -2 mbar)<br />

was given to the material in order to induce microstructural coarsening. The<br />

temperature was then lowered to 400 °C <strong>and</strong> kept constant for 60 h in order to allow<br />

the inclusion shape to equilibrate. The material was then water-quenched in order to<br />

keep the shape <strong>of</strong> the inclusions unaltered.<br />

5.2.2 Interrupted creep tests<br />

Ø 4 mm tensile test specimens, Fig. 5-4, are mounted in a modified hydraulic tensile<br />

test machine (MFL 100 kN), where up to 1000 N dead load (σ ≤ 80 MPa) can be<br />

suddenly applied after a desired hold at temperature under zero load. The heating is<br />

performed with a lamp furnace (<strong>Research</strong> Inc., Parabolic Clamshell Heater, Model<br />

4068-12-10), allowing rapid heating <strong>and</strong> temperature stabilization (< 120 s to<br />

400 °C). The temperature is monitored by a Ø 0.2 mm type K thermocouple attached<br />

to the center <strong>of</strong> the specimen. The deformation is measured by a high temperature<br />

extensometer (MTS 632.53F-30 equiped with Ø 3 mm, 190 mm long Al 2 O 3<br />

extension rods) attached to the specimen gage section.<br />

87


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

88<br />

M6<br />

Figure 5-4: Tensile specimen geometry. Dimensions are in mm.<br />

The load is applied after a 5 min hold at 400 °C <strong>and</strong> the temperature is kept constant<br />

for a further 15 min. The furnace is then switched <strong>of</strong>f <strong>and</strong> rapidly removed, while<br />

forced air is directed to the specimen surface. The specimen is unloaded once its<br />

temperature falls below 50 °C.<br />

Polished longitudinal cuts are then prepared. The visible lead inclusions are<br />

selectively dissolved <strong>and</strong> the dihedral angle <strong>of</strong> specific inclusions is measured with<br />

the method described in Section 4.3.2, recording their orientation with respect to the<br />

applied stress axis.<br />

R3<br />

110<br />

30<br />

4<br />

6<br />

15


5.3 Results<br />

5.3.1 Creep curves<br />

5.3 RESULTS<br />

Fig. 5-5 presents creep curves <strong>of</strong> C99 material subjected to constant loads for 15 min<br />

at 400 °C. Notice the thermal expansion <strong>of</strong> the specimen during heating <strong>and</strong> the<br />

successive holding time on the left-h<strong>and</strong> side <strong>of</strong> the viewgraph (t


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

90<br />

modulus <strong>of</strong> liquid lead, K Pb , was calculated from the tabulated velocity <strong>of</strong> sound<br />

[206]. Isotropic elastic properties <strong>of</strong> <strong>copper</strong> at 400 °C found in [207] were used.<br />

A numerical application for the typical case <strong>of</strong> a 5 µm diameter lead intragranular<br />

inclusion within uniaxially stressed <strong>copper</strong> at 400 °C is given in Fig. 5-6 (b). The<br />

interfacial energy is according to Eq. 4-17.<br />

Notice that no equilibrium configuration is predicted by the lower curve above a<br />

specific critical stress. Instability is predicted once the maximum <strong>of</strong> this curve is<br />

passed.<br />

Λ (c) [-]<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

2 1/3<br />

σapp<br />

⋅ V<br />

Λ(c) =<br />

γ ⋅ E<br />

0.9<br />

liquid<br />

inclusion<br />

0.7<br />

void<br />

(a) (b)<br />

Figure 5-6: Equilibrium shape <strong>of</strong> an intragranular inclusion imbedded in a material<br />

subjected to uniaxial remote stress: a) adimensionalized problem, <strong>and</strong> b) application to<br />

a typical ø5 µm Pb inclusion in our Cu-Pb alloy.<br />

b) Experiments<br />

0.5<br />

K Pb = 27,5 GPa<br />

"K " = 0 GPa<br />

Pb<br />

0.3<br />

aspect ratio: c [-]<br />

0.1<br />

sphere crack<br />

After interrupted creep at 400 °C, dissolved transgranular inclusions were imaged in<br />

the SEM. Extensive faceting was observed. An example is given in Fig. 5-7 with the<br />

corresponding datapoints obtained by the 3D reconstruction. Datapoints are best<br />

fitted by the sphere displayed; black datapoints lie outside the sphere <strong>and</strong> shaded<br />

σ [MPa]<br />

app<br />

200<br />

150<br />

100<br />

50<br />

0<br />

T = 400 °C<br />

γ SL = 0.432 J/m 2<br />

E Cu = 111 GPa<br />

V = 10 -16<br />

m 3<br />

K Pb = 27,5 GPa<br />

"K " = 0 GPa<br />

Pb<br />

0.9 0.7 0.5 0.3 0.1<br />

sphere crack<br />

aspect ratio: c [-]


5.3 RESULTS<br />

ones lie inside. Note that Fig. 5-7 (b) adjusts to Fig. 5-7 (a) after a 180° rotation about<br />

the horizontal axis.<br />

One can see that the actual shape resembles a prolate spheroid, instead <strong>of</strong> the<br />

predicted oblate one. Namely the inclusion is elongated in the stress direction, Fig. 5-<br />

7 (a).<br />

(a) (b)<br />

Figure 5-7: Dissolved intragranular inclusion in C99 crept at 400 °C with σ=59 MPa,<br />

the stress axis is vertical: a) SEM image, <strong>and</strong> b) corresponding datapoints from the<br />

reconstruction <strong>of</strong> the unfaceted interface. The datapoints are best fitted by the sphere<br />

displayed (stress axis is vertical).<br />

5.3.3 Intergranular inclusion<br />

The predicted shape <strong>of</strong> an intergranular inclusion is presented in Fig. 5-8. A<br />

comparison is given between a first set <strong>of</strong> predictions where the inclusion is made <strong>of</strong><br />

two spherical caps intersecting at the grain boundary (solid line, the surface area <strong>of</strong><br />

the inclusion is calculated according to Eq. 5-15) <strong>and</strong> a second set <strong>of</strong> predictions<br />

where the shape <strong>of</strong> the inclusion is approximated by an oblate ellipsoid (dashed line,<br />

Eq. 5-16).<br />

In both cases, two calculations were conducted:<br />

91


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

92<br />

(i) one in which the compressibility <strong>of</strong> the liquid inclusion equalled that <strong>of</strong> liquid lead<br />

(Bulk modulus K Pb = 27.5 GPa);<br />

(ii) one in which the liquid inclusion bulk modulus is set to zero, i.e., in which the<br />

inclusion behaves, from the mechanical point <strong>of</strong> view, like a void (K Pb = 0).<br />

Results from the calculation are given in adimensional form in Fig. 5-8 (a), <strong>and</strong> in<br />

Fig. 5-8 (b) in dimensional form for the case <strong>of</strong> a 10 µm diameter lead inclusion<br />

located at a grain boundary that is normal to applied uniaxial stress. The horizontal<br />

axis is given in terms <strong>of</strong> the “apparent macroscopic” dihedral angle <strong>of</strong> the inclusion,<br />

related to its aspect ratio according to:<br />

2<br />

⎛1<br />

− c ⎞<br />

φ = 2 arccos⎜<br />

2 ⎟<br />

⎝1<br />

+ c<br />

⎠<br />

Eq.5-19<br />

As seen, regardless <strong>of</strong> the geometrical assumption made in estimating the capillary<br />

energy, there is a clear difference in inclusion behaviour between a liquid inclusion <strong>of</strong><br />

finite compressibility <strong>and</strong> a void:<br />

(i) with zero bulk modulus, there is a maximum in the curve, past which the inclusion<br />

shape is unstable. Past this point, the inclusion is predicted to collapse into a crack;<br />

this result was derived previously [36, 39, 40];<br />

(ii) with a finite bulk modulus, on the other h<strong>and</strong>, the curve shows an inflexion point<br />

but increases monotonously. The shape instability predicted for a void thus<br />

disappears. It was ascertained by conducting additional calculations that, even with a<br />

very small bulk modulus (K Pb = 1 GPa), the bifurcation point disappears. The<br />

analysis therefore indicates that, under increasing stress, liquid inclusions remain<br />

stable if there is no void nucleation in their midst.


Λ (c) [-]<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0.6<br />

2 1/3<br />

σapp<br />

⋅ V<br />

Λ(c) =<br />

γ ⋅ E<br />

lenticular<br />

shape<br />

ellipsoidal<br />

shape<br />

0.5<br />

0.4<br />

5.3 RESULTS<br />

(a) (b)<br />

Figure 5-8: Equilibrium shape <strong>of</strong> an intergranular inclusion imbedded in a material<br />

subjected to an uniaxial remote stress, which is normal to the grain boundary: a)<br />

adimensionalized problem, <strong>and</strong> b) application to a typical gb inclusion in our Cu-Pb<br />

alloy. Two sets <strong>of</strong> data are given, (i) the area <strong>of</strong> the interface is either approximated by<br />

two spherical caps intersecting at the grain boundary (Eq. 5-15), or (ii) it is<br />

approximated by an oblate ellipsoid (Eq. 5-16). For each set <strong>of</strong> data, the compressibility<br />

<strong>of</strong> the liquid lead is taken into account (KPb =27.5 GPa) or the inclusion properties are<br />

that <strong>of</strong> a void (”KPb ”=0 GPa).<br />

a) Experiment<br />

0.3<br />

Apparent (macroscopic) dihedral angles were measured in interrupted (after 15 min)<br />

creep tests at 43, 59, <strong>and</strong> 75 MPa. An additional creep test performed at 400 °C at a<br />

stress <strong>of</strong> 83 MPa was interrupted after 10 min since creep had entered its third stage.<br />

After longitudinal cut preparation <strong>of</strong> this specimen <strong>and</strong> inclusion dissolution, no<br />

single inclusion displayed a sufficiently regular shape to allow the measurement <strong>of</strong><br />

any relevant dihedral angle.<br />

The angle determination was the more difficult to perform, the higher was the remote<br />

stress. This is due to the plastic flow <strong>of</strong> the material, <strong>and</strong> to the migration <strong>of</strong> the grain<br />

boundaries, Fig. 5-9 (a).<br />

K Pb = 27,5 GPa<br />

"K " = 0 GPa<br />

Pb<br />

0.2<br />

aspect ratio: c [-]<br />

0.1<br />

crack<br />

σ [MPa]<br />

app<br />

200<br />

150<br />

100<br />

50<br />

φ 0<br />

0<br />

100<br />

=120°<br />

T = 400 °C<br />

γ SL = 0.432 J/m 2<br />

E Cu = 111 GPa<br />

V = 10 -16 m 3<br />

80<br />

K Pb = 27,5 GPa<br />

60<br />

"K " = 0 GPa<br />

Pb<br />

40<br />

20<br />

apparent dihedral angle: φ [°]<br />

crack<br />

93


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

94<br />

(a) (b)<br />

Figure 5-9: Intergranular inclusion dissolved after an interrupted creep test<br />

performed at 400 °C under 59 MPa. Material is C99. (a) The grain boundary is<br />

serrated, <strong>and</strong> the former solid/liquid interface is extensively distorted. This did not<br />

allow the measurement <strong>of</strong> φ following the method described in Chapter 4. On the other<br />

h<strong>and</strong> (b) a relevant dihedral angle could be measured (φ = 79°). The dashed contours<br />

correspond to the intersection <strong>of</strong> the polished plane with the two calculated spheres<br />

allowing the measurement <strong>of</strong> φ.<br />

Nevertheless some inclusions displayed a sufficiently regular shape to allow the<br />

measurement <strong>of</strong> a relevant apparent dihedral angle, Fig. 5-9 (b). Resulting data<br />

values show a significant spread, as do data from stress-free samples, Fig. 4-14;<br />

however it nevertheless emerges from the data that φ is sensitive to the applied load.<br />

Decreasing values are generally measured for increasing load, Fig. 5-10 <strong>and</strong><br />

Appendix 10.4.<br />

measured dihedral angle: φ [°]<br />

130<br />

120<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

C99<br />

400 °C<br />

0 20 40 60 80 100<br />

applied remote stress: σapp [MPa]<br />

Figure 5-10: Measured φ values after interrupted creep test as a function <strong>of</strong> the applied<br />

remote stress. Each datapoint corresponds to a measurement on a single inclusion<br />

according to the procedure described in Section 4.3.2.


5.3 RESULTS<br />

Let us first assume that there is a single zero-stress dihedral angle φ o . To account for<br />

the influence <strong>of</strong> inclusion orientation with respect to stress, the theoretical Λ(φ)<br />

curves are drawn for the corresponding specific grain boundary ξ <strong>and</strong> θ values,<br />

Fig. 5-3 for each inclusion from Appendix 10.4. An illustration <strong>of</strong> this is given in<br />

Fig. 5-11 for the specific inclusion marked by an asterisk in Appendix 10.4. As<br />

shown in Fig. 5-11, the predicted φ values are deduced from the two theoretical<br />

curves <strong>and</strong> Λ, for the case where the inclusion properties are those <strong>of</strong> a void<br />

(K Pb = 0). Predictions for a finite K Pb show less agreement.<br />

[-]<br />

Λ<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

φ 0<br />

measured φ<br />

Λ = 0.587<br />

0<br />

100 80 60 40 20<br />

=120°<br />

apparent dihedral angle: φ [°]<br />

crack<br />

Figure 5-11: Determination <strong>of</strong> the predicted φ values for an inclusion located on a<br />

grain boundary characterized by ξ = 37° <strong>and</strong> θ = 25° (i.e. changing the stress tensor<br />

accordingly, Eq. 5-17). Here, the experimentally observed inclusion <strong>of</strong> interest is<br />

characterized by Λ = 0.587 <strong>and</strong> φ = 69°.<br />

This way, actual measured dihedral angles, Appendix 10.4, can be compared to<br />

predicted φ values for each inclusion, Fig. 5-12.<br />

"K " = 0 GPa<br />

Pb<br />

lenticular<br />

shape<br />

ellipsoidal<br />

shape<br />

95


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

96<br />

Figure 5-12: Predicted φ values compared to measured values. The inclusion properties<br />

are taken as those <strong>of</strong> a void (”K Pb ”=0 GPa). Both lenticular <strong>and</strong> ellipsoidal shape are<br />

considered.<br />

As is clear from Fig. 5-12, experimental data show much greater spread than<br />

predicted values.<br />

predicted dihedral angle: φ [°]<br />

130<br />

120<br />

110<br />

100<br />

90<br />

σ app = 0<br />

80<br />

"K " = 0 GPa<br />

Pb<br />

70<br />

lenticular<br />

shape<br />

60<br />

50<br />

slope = 1<br />

ellipsoidal<br />

shape<br />

50 60 70 80 90 100 110 120 130<br />

measured dihedral angle: φ [°]


5.4 Discussion<br />

5.4.1 Theoretical predictions<br />

5.4 DISCUSSION<br />

Our theoretical predictions can be compared to those <strong>of</strong> the literature for<br />

intragranular [36], <strong>and</strong> for intergranular voids [39, 40]. This can be done by<br />

comparing the respective predicted critical adimensional parameter recalculated in<br />

Table 1 according to our definition, Eq. 5-11.<br />

Eq. 5-11<br />

Table 1 : Predicted critical parameter Λ c according to Eq. 5-11 for a void imbedded in a<br />

uniaxially stressed material.<br />

intragranular inclusion intergranular inclusion φ 0 =120°<br />

Sun et al.<br />

[36]<br />

ellipsoidal<br />

shape<br />

2 1/ 3<br />

σ app ⋅<br />

Λ=<br />

γ ⋅<br />

V<br />

E<br />

our analysis<br />

Fig. 5-6 (a)<br />

ellipsoidal<br />

shape<br />

Raj [40]<br />

Wang et al.<br />

[39]<br />

penny shape ellipsoidal<br />

shape<br />

lenticular<br />

shape<br />

our analysis<br />

Fig. 5-8 (a)<br />

ellipsoidal<br />

shape<br />

0.73 1.07 0.62 0.38 0.61 0.44<br />

Our theoretical predictions are in good accordance with the literature. Differences in<br />

Λ c may be due to different assumptions considered in the analysis. As an illustration,<br />

for the Poisson’s ratio, we used ν = 0.355 (tabulated value for pure <strong>copper</strong> at 400 °C<br />

[207]), whereas ν = 1/3 was taken in [36] for their numerical calculations. It must<br />

also be emphasized that Λ is numerically calculated by taking the ratio between the<br />

first derivative <strong>of</strong> two complicated functions, Eq. 5-11. It is moreover clear from<br />

Figs. 5-13 <strong>and</strong> 5-14 below, that the shape approximations defined in Section 5.1.3<br />

induce some degree <strong>of</strong> error. In particular, it is seen that, when Γ is estimated using<br />

Eq. 5-16, i.e., by assimilating the inclusion to the same spheroid used in calculating<br />

the mechanical energy, the predicted shape at low stress is inconsistent with the<br />

assumed stress-free dihedral angle φ o . This is clearly because Eq. 5-15, (lenticular<br />

shape) appropriate for such an inclusion at zero stress, is required to obtain a global<br />

97


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

98<br />

inclusion shape that satisfies local capillary equilibrium at the triple line. This<br />

advantage <strong>of</strong> Eq. 5-15 instead <strong>of</strong> Eq. 5-16 used by Wang et al. [39] motivated the<br />

present modification <strong>of</strong> the analysis.<br />

5.4.2 Measured inclusion shape<br />

Theoretical predictions discussed above describe equilibrium situations. Fig. 5-7<br />

suggests, however, that the rate <strong>of</strong> shape equilibration is lower than the rate <strong>of</strong><br />

viscous flow <strong>of</strong> our creeping material. Indeed, the intragranular inclusion in Fig. 5-7<br />

is found to be slightly elongated in the applied tensile stress direction, whereas it<br />

should be a prolate ellipsoid shape oriented normal to the stress. Since shape<br />

equilibration is governed by diffusion, it is clear that equilibration requires lower<br />

strain rates <strong>and</strong> smaller inclusions, as observed in [208]. Notice however that<br />

relatively coarse inclusions, more than 5 µm wide, are needed for accurate dihedral<br />

angle measurement with the method presented in Section 4.3.2.<br />

Estimated times for equilibration in the absence <strong>of</strong> applied stress Eq. 4-11 are, at<br />

400°C, 8 hours for an inclusion 10 µm in diameter, <strong>and</strong> one hour for an inclusion 5<br />

µm in diameter. Although (as mentioned in Section 5.1.4), equilibration times thus<br />

estimated are not strictly valid in the presence <strong>of</strong> stress (because stress changes<br />

concentration gradients across the inclusion), the indication that shape equilibration<br />

is probably incomplete in the present samples is consistent with Fig. 5-7. Apparent<br />

dihedral angles measured here may, therefore, be somewhat superior than the<br />

equilibration value, φ(σ), corresponding to the applied stress, for two reasons:<br />

(i) tensile stress lowers φ, Figs. 5-6 <strong>and</strong> 5-8, thus measured φ values probably lie<br />

between φ o <strong>and</strong> φ(σ)<br />

(ii) creep will open up pores that lie more or less perpendicular to the stress axis.<br />

Another difficulty is that the frozen <strong>microstructure</strong> <strong>of</strong> the creeping C99 material is<br />

not at equilibrium. Indeed grain boundaries are serrated, indicating grain boundary<br />

movement, Fig. 5-9 (a). Our method for dihedral angle measurement is based on the<br />

assumption that the solid/liquid interfaces are spherical. This is not the case in the<br />

presence <strong>of</strong> stress, since the inclusion equilibrium shape can not display a constant


5.4 DISCUSSION<br />

curvature κ when w is not constant along the whole interface, Eq. 5-18. Nevertheless,<br />

relevant ”global” dihedral angle measurements could be performed on selected<br />

inclusions, Fig. 5-9 (b).<br />

5.4.3 Comparison <strong>of</strong> theory with experiment<br />

It must first be recalled that our theoretical predictions are based on several<br />

assumptions.<br />

(i) The interaction energy ∆W is calculated with the Eshelby formalism. It is<br />

therefore assumed that the behaviour <strong>of</strong> both the matrix <strong>and</strong> the inclusion are linear<br />

elastic. Moreover, a spheroidal shape is assumed for the inclusion. It is clear that our<br />

lenticular lead inclusions are not spheroidal; nor does the creeping <strong>copper</strong> matrix<br />

behave as a linear elastic solid.<br />

(ii) The shape approximation for the inclusion is also shown to have indeed a<br />

noticable influence on the predicted curves, see e.g. the solid <strong>and</strong> dashed lines on<br />

Fig. 5-8.<br />

(iii) It was concluded in Chapter 4 that, at a specific temperature, the dihedral angle is<br />

not unique, reflecting mainly the anisotropy <strong>of</strong> γ gb . In unstressed C99 equilibrated at<br />

400 °C, φ is not unique. It was indeed found to vary between 102 <strong>and</strong> 125°,<br />

Appendix 10.4. In our theoretical predictions however, we considered a unique value<br />

for the ratio <strong>of</strong> γ gb /γ SL at 400 °C.<br />

Experimental data <strong>of</strong> the present study are not the only that document a decrease <strong>of</strong><br />

dihedral angle φ <strong>of</strong> liquid inclusions under applied stress. In stressed <strong>leaded</strong> nickel,<br />

Stickels et al. reported a similar decrease <strong>of</strong> φ [32, 33]. In this study, reported<br />

dihedral angles are averages measured as indicated in Section 4.3.1 from two-<br />

dimensional metallographic cuts, taking the average value <strong>of</strong> 500 measurements for<br />

each data point. The data <strong>of</strong> Stickels are plotted in Fig. 5-13. Also plotted in that<br />

figure are curves predicted by the present analysis, considering the elastic properties<br />

<strong>of</strong> nickel at 370 °C [207], the interfacial energy <strong>of</strong> the Cu-Pb system, <strong>and</strong> V = 10- 16<br />

m3 corresponding to a 10 µm wide inclusion, similarly as for Fig. 5-8 (b). This<br />

latter value was guessed, since no indication <strong>of</strong> the lead particle size was given<br />

99


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

100<br />

(smaller inclusions render the apparent dihedral angle measurement very difficult<br />

under an optical microscope with a 2000 x magnification, as performed in the sixties<br />

by Stickels) [32, 33].<br />

σ [MPa]<br />

app<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

φ 0<br />

0<br />

=51°<br />

40<br />

T = 370 °C<br />

γ SL = 0.432 J/m 2<br />

E Ni = 195 GPa<br />

V = 10 -16 m 3<br />

Figure 5-13: Apparent dihedral angle data from the literature ploted against applied<br />

hydrostatic or uniaxial stress on Ni-2Pb samples at 370 °C; data are from [32, 33].<br />

Lines correspond to our theoretical predictions considering (upper curves) an<br />

intergranular Pb liquid inclusion embedded in hydrostatically stressed Ni, <strong>and</strong> (lower<br />

curves) a grain boundary void in an uniaxially stressed matrix.<br />

It is seen that the shape <strong>of</strong> curves traced by the two sets <strong>of</strong> data points reproduce<br />

relatively well the overall shape <strong>of</strong> the two curve sets: data for hydrostatic<br />

compression tests reproduce curves derived assuming that the inclusion has the bulk<br />

modulus <strong>of</strong> liquid lead at that temperature, while uniaxial (tensile <strong>and</strong> compressive)<br />

test data points delineate a curve that is closer, in shape <strong>and</strong> value, to the curve<br />

derived under the assumption that the liquid inclusion behaves as a void. All else<br />

being equal, a far higher isostatic compressive stress (e.g., 344 MPa) is needed to<br />

lower the dihedral angle than is the case for uniaxial stress (the corresponding<br />

measured – tensile – stress is 14 MPa, see Fig. 5-13).<br />

30<br />

Ni-Pb<br />

20<br />

K = 27,5 GPa<br />

Pb<br />

σ<br />

hydrostatic app<br />

"K " = 0 GPa<br />

Pb<br />

σ<br />

uniaxial app<br />

10<br />

apparent dihedral angle: φ [°]<br />

crack<br />

theoretical predictions<br />

lenticular shape<br />

ellipsoidal shape<br />

experimental data<br />

hydrostatic stress<br />

compressive<br />

uniaxial stress<br />

tensile & compressive


5.4 DISCUSSION<br />

This makes good sense: hydrostatic compression does not cause global deformation<br />

<strong>of</strong> the material. Under uniaxial tension or compression, on the other h<strong>and</strong>, the<br />

material deforms (by creep <strong>and</strong>/or plasticity), such that the various void nucleation<br />

mechanisms that have been invoked to explain creep fracture can operate to nucleate<br />

a void in the inclusion [209].<br />

Such pore nucleation was indeed also reported in the literature: (i) pores were<br />

namely found to grow to a micrometre size in a T91 steel after prolonged annealing<br />

under stress in liquid lead [210]. (ii) Pore nucleation <strong>and</strong> growth at liquid Bi<br />

inclusions was also reported for an Al-0.23Ti-1.5Bi alloy strained at elevated<br />

temperature [211].<br />

The influence <strong>of</strong> an applied tensile stress on φ was also documented in the literature<br />

for a Cu-1Pb-0.08P alloy, close to our C99 material (which contains 0.8 to<br />

1.2 wt.% Pb <strong>and</strong> 0.01 to 0.04 wt.% P) [34, 35], Fig. 3-12. Again experimental φ<br />

values compare relatively well with our predictions assuming the presence <strong>of</strong><br />

voiding in the inclusion (”K Pb ” = 0 GPa), Fig. 5-14. Elastic properties were<br />

extrapolated from [207], γ SL is according to Eq. 4-17, <strong>and</strong> the inclusion volume V<br />

was estimated from a metallographic plate in [34].<br />

101


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

102<br />

σ [MPa]<br />

app<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

φ 0<br />

=93°<br />

Cu-Pb<br />

T = 650 °C<br />

γ SL = 0.422 J/m 2<br />

E Cu = 100 GPa<br />

V = 10 -16 m 3<br />

80<br />

Figure 5-14: Apparent dihedral angle data from the literature ploted against applied<br />

uniaxial stress on Cu-1Pb-0.08P samples at 650 °C, data are from [34, 35]. Lines<br />

correspond to our theoretical predictions considering a grain boundary void in a<br />

stressed matrix.<br />

In both comparisons <strong>of</strong> theoretical predictions with data from the literature, the<br />

quantitative agreement is far from perfect, but nonetheless acceptable given the many<br />

assumptions made in the derivation (that the matrix remains elastic, for example).<br />

Turning now to experimental data from the present study, comparison is somewhat<br />

more difficult because our data are more specific: each inclusion that was<br />

characterized has a known size, <strong>and</strong> also a specific (variable but measured)<br />

orientation with respect to the applied stress axis.<br />

This second factor creates the complication that a single parameter (Λ) no longer<br />

describes entirely the mechanical <strong>and</strong> capillary state <strong>of</strong> every inclusion: the stress<br />

state assumed in the derivation is no longer a simple tensile stress normal to the grain<br />

boundary; rather, it is multiaxial, given by Eq. 5-17.<br />

70<br />

60<br />

50<br />

To account for this, theoretical Λ(φ) curves were drawn for each inclusion orientation<br />

(given in Appendix 10.4) <strong>and</strong> compared with the data; an example is given in Fig. 5-<br />

11. As seen the curve reaches higher Λ values than that <strong>of</strong> Fig. 5-8 (a): this obeys<br />

intuition, since a higher stress will be required to lower the dihedral angle <strong>of</strong> an<br />

40<br />

theoretical predictions<br />

30<br />

lenticular shape<br />

ellipsoidal shape<br />

experimental data<br />

apparent dihedral angle: φ [°]<br />

"K " = 0 GPa<br />

Pb<br />

uniaxial stress<br />

20<br />

10<br />

crack


5.4 DISCUSSION<br />

inclusion that is oriented along a grain boundary that is inclined with respect to the<br />

stress axis. Again, we stress the key simplification made in the present derivation,<br />

namely that the inclusion shape evolves along a single family <strong>of</strong> (spheroidal/<br />

lenticular) shapes: under shear, as experienced along grain boundaries inclined with<br />

respect to the stress axis, asymmetric (sheared) inclusion geometries may be closer to<br />

equilibrium.<br />

These theoretical curves were drawn under a second assumption, namely that the<br />

stress-free dihedral angle for the inclusion is known <strong>and</strong> constant : we have taken<br />

φ ο = 120°. As seen in the previous chapter, dihedral angles in fact vary strongly<br />

depending on the specifics <strong>of</strong> the grain boundary along which they lie.<br />

Comparison <strong>of</strong> predicted <strong>and</strong> measured dihedral angle values is given in Fig. 5-12:<br />

perfect agreement would cause the data points to lie along the diagonal <strong>of</strong> the plot. As<br />

seen, for both models (which differ by details <strong>of</strong> the assumed inclusion shape while<br />

both assimilating the inclusion to a void, as exposed above), the predicted influence<br />

<strong>of</strong> applied stress is less than the observed spread in values: most predicted values (all<br />

but two) lie, for the stress state <strong>and</strong> inclusion size characteristic <strong>of</strong> each inclusion,<br />

relatively close to the zero stress equilibrium value (φ ο = 120° for the model<br />

assuming lenticular inclusions).<br />

This discrepancy is also apparent in the data <strong>of</strong> Waterhouse, Fig. 5-14: the observed<br />

range <strong>of</strong> variation <strong>of</strong> measured average dihedral angles is from 93 to 75° as stress<br />

increases from 0 to 21 MPa (note that no account can be taken here <strong>of</strong> the effect <strong>of</strong><br />

inclusion orientation, not reported for data from that study). Theory on the other h<strong>and</strong><br />

predicts a range <strong>of</strong> reduction <strong>of</strong> apparent dihedral angle <strong>of</strong> six degrees at most, Fig. 5-<br />

14.<br />

This discrepancy cannot be explained by the following factors:<br />

- lack <strong>of</strong> time for equilibration would cause the observed dihedral angles to decrease<br />

less than is predicted; experimental variations should therefore be less than is<br />

predicted by theory;<br />

103


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

104<br />

- similarly, opening <strong>of</strong> lenticular voids by tensile creep deformation normal to the<br />

grain boundary will increase φ, not lower it below theoretical predictions;<br />

- lack <strong>of</strong> voiding in the inclusion (implying a finite liquid phase bulk modulus) would<br />

also lower the rate <strong>of</strong> change <strong>of</strong> inclusion aspect ratio (or apparent dihedral angle)<br />

with stress Fig. 5-8, whereas experimental data vary more than is predicted.<br />

The somewhat greater apparent dihedral angle reduction seen experimentally to<br />

result from increasing stress may be related to the many assumptions made in the<br />

derivation. In particular, the assumption that the inclusions are smooth spheroids may<br />

underestimate significantly the elastic energy associated with the presence <strong>of</strong> the<br />

inclusion, <strong>and</strong> hence with variations <strong>of</strong> its shape. Alternatively, the following<br />

explanations could also be proposed:<br />

- shear deformation along a sliding grain boundary may elongate the inclusions, in<br />

turn decreasing their aspect ratio c <strong>and</strong> hence their apparent “macroscopic” dihedral<br />

angle φ,<br />

- grain boundaries that were observed to become mobile in the stressed Cu-Pb<br />

samples <strong>of</strong> this work, Fig. 5-9 (a), may orient themselves under the action <strong>of</strong> stress<br />

away from orientations <strong>of</strong> lowest energy reached during annealing. This, in turn,<br />

would bias the intrinsic, stress-free dihedral angle φ ο <strong>of</strong> the inclusions towards lower<br />

values, creating an effect that is over <strong>and</strong> above the effect expected from stress.


5.5 Conclusion<br />

5.5 CONCLUSION<br />

Analysis predicts a change with stress in the shape <strong>of</strong> a liquid inclusion embedded in<br />

a stressed matrix. Two cases are treated; (i) the inclusion is considered as a void, or<br />

alternatively (ii) the finite compressibility <strong>of</strong> the inclusion content is taken into<br />

account. The ”apparent dihedral angle”, defined at the inclusion length scale <strong>and</strong><br />

influenced by stress is predicted to decrease while the adimensional parameter Λ<br />

defined in Eq. 5-11 increases. This occurs at a far greater rate if the inclusion<br />

compressibility is nil.<br />

In case (i), a critical value <strong>of</strong> Λ is found in accordance to literature [36, 39-42], above<br />

which the inclusion is unstable <strong>and</strong> is predicted to collapse to a crack. On the other<br />

h<strong>and</strong>, no instability is predicted to occur in case (ii).<br />

A finite decrease <strong>of</strong> the average dihedral angle φ with increasing tensile stress was<br />

measured on C99 samples subjected to interrupted creep uniaxial tension tests at<br />

400 °C, as compared to φ measured on unstressed specimen. Our experimental data<br />

as well as data from the literature follow the same trends as predictions, namely that<br />

φ decreases under increased tensile stress. φ values are best described by predictions<br />

accounting for the shape evolution <strong>of</strong> a void – case (i) above – within a uniaxially<br />

stressed solid, suggesting that voids nucleate within the liquid inclusions or along the<br />

matrix/inclusion interface.<br />

Experimental data show a greater rate <strong>of</strong> decrease <strong>of</strong> the inclusion apparent dihedral<br />

angle with increasing stress. More work would be needed to elucidate whether the<br />

dispcrepancy is due to limited data, assumptions in the derivation, or an effect not<br />

accounted for in the analysis, such as the possible consequences <strong>of</strong> grain boundary<br />

sliding or mobility.<br />

105


CHAPTER 5. INFLUENCE OF STRESS ON THE SHAPE OF AN EMBEDDED LIQUID INCLUSION<br />

106


Chapter 6<br />

6 Cu-<strong>embrittlement</strong>: the role <strong>of</strong> Pb<br />

Small additions <strong>of</strong> lead, on the order <strong>of</strong> 1 wt.%, are frequently used to improve the<br />

machinability <strong>of</strong> <strong>copper</strong> <strong>alloys</strong>. At intermediate temperature, namely above<br />

approximately 300 °C, the reduction in ductility that is classically observed in <strong>copper</strong><br />

<strong>and</strong> its <strong>alloys</strong> is markedly accentuated by the presence <strong>of</strong> such lead additions.<br />

There has been some debate in the literature regarding whether the underlying<br />

mechanism is liquid metal <strong>embrittlement</strong> (LME) or an accentuation <strong>of</strong> grain<br />

boundary <strong>embrittlement</strong> (GBE) caused in the solid state by segregated lead, see<br />

Section 3.2.4. This chapter is a contribution to the question, which shows that both<br />

mechanisms operate: GBE dominates at low strain rate, whereas LME is dominant at<br />

high strain rates.<br />

Tensile tests were conducted on both pure <strong>and</strong> <strong>leaded</strong> <strong>copper</strong> at different strain rates<br />

below <strong>and</strong> above 327 °C, the melting point <strong>of</strong> lead. In un<strong>leaded</strong> samples, damage was<br />

identified as evolving from ductile cavitation to grain boundary decohesion as the<br />

strain rate is reduced; this is thus a clear symptom <strong>of</strong> GBE. The addition <strong>of</strong> 1 wt. %<br />

Pb induces a sharp decrease in the reduction <strong>of</strong> area at all investigated temperatures.<br />

At a strain rate <strong>of</strong> 10 -4 s -1 , damage is recognized as being intergranular decohesion<br />

throughout the whole sample in both <strong>leaded</strong> <strong>and</strong> un<strong>leaded</strong> samples. If the strain rate is<br />

increased to 10 s -1 , stress-strain curves <strong>of</strong> <strong>leaded</strong> <strong>and</strong> un<strong>leaded</strong> <strong>copper</strong> superimpose<br />

very well until fracture occurs in the former well below the fracture strain <strong>of</strong> the<br />

latter; this is a signature <strong>of</strong> LME. We confirm this conclusion with additional<br />

experiments, in which sample surface contact with liquid lead is shown to cause<br />

cracking <strong>of</strong> the samples.<br />

107


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

108<br />

6.1 Experimental procedures<br />

6.1.1 Materials<br />

Copper, 99.997% pure (hereafter Cu4N) as well as 99,95% pure (hereafter CuOFHC)<br />

were provided by Swissmetal ® . These were in the form <strong>of</strong> Ø 18 mm cast bars <strong>and</strong><br />

Ø 6 mm strain-hardened bars, respectively.<br />

Cu4N material was hot-extruded at 900 °C to Ø 8 mm, <strong>and</strong> successively cold-drawn<br />

to Ø 7 <strong>and</strong> 6 mm. Both Cu4N <strong>and</strong> CuOFHC were heat-treated at 500 °C for 15 min in<br />

air to allow complete recrystallization. The resulting grains are equiaxed, 25 µm in<br />

size.<br />

Cu-1 wt.% Pb <strong>alloys</strong> were either prepared from high-purity metals in the laboratory,<br />

(hereafter Cu1Pb) or provided by Swissmetal ® (industrial alloy C99, see<br />

Section 5.2.1). The Cu1Pb material was cast into small billets, 20 mm in diameter<br />

<strong>and</strong> 60 mm in height following the procedure described in Section 4.1.1. Similarly as<br />

for material Cu4N, Ø 6 mm bars <strong>of</strong> recrystallized Cu1Pb material were prepared by<br />

hot-extrusion, wire drawing, <strong>and</strong> heat treatment at 500 °C for 15 min.<br />

Lead coating on CuOFHC samples was either obtained<br />

(i) by hot dipping, where samples were immersed for ≥ 15 min at 600 °C in a lead<br />

bath saturated in <strong>copper</strong>, removed from the bath <strong>and</strong> cooled with compressed air, or<br />

(ii) by electrodeposition, as follows. 500 ml electrolyte was prepared by the<br />

dissolution in distilled water <strong>of</strong> 21 g HBF 4 , 15 g H 3 BO 3 , 0.1 g peptone, <strong>and</strong> 25 g<br />

PbO according to [212]. The cathode current was set to 200 A/m 2 (i.e. 0.1 A for a<br />

Ø 4 mm, 30 mm-long cylinder) according to [213]. Optimal coating was obtained<br />

after 20 min deposition.<br />

Industrial Cu-15 wt.% Ni-8 wt.% Sn <strong>alloys</strong> with <strong>and</strong> without a 1 wt.% Pb addition<br />

were also provided by Swissmetal ® . These <strong>alloys</strong> are respectively denominated NP8<br />

(<strong>leaded</strong>) <strong>and</strong> CN8 (lead-free). These <strong>alloys</strong> were processed on the industrial site by<br />

successive Osprey spray deposition, hot extrusion <strong>and</strong> room temperature rolling<br />

<strong>and</strong> drawing. Strain-hardened CN8 material was provided in the form <strong>of</strong> 6 mm


6.1 EXPERIMENTAL PROCEDURES<br />

diameter bars. NP8 was supplied as air-cooled hot-extruded Ø 20 mm bars (NP8, the<br />

<strong>leaded</strong> grade <strong>of</strong> CN8 is not produced industrially). Single-phased material was<br />

obtained afterwards by solutionization at 850 °C for 1 h under flowing Ar (200 L/h),<br />

followed by rapid cooling (within 20 s to 200 °C) in a fluidized bath. For both CN8<br />

<strong>and</strong> NP8, the <strong>microstructure</strong> is equiaxed, the average grain size is 100 µm, <strong>and</strong> lead<br />

inclusions in NP8 are below 1 µm in size. A pseudo-binary phase diagram as well as<br />

a TTT diagram <strong>of</strong> the Cu-15Ni-8Sn alloy are available in the literature, [214], Fig. 6-<br />

1.<br />

(a) (b)<br />

Figure 6-1: (a) Isopleth at 15 wt.% Ni <strong>of</strong> the ternary Cu-Ni-Sn phase diagram, <strong>and</strong> (b)<br />

TTT diagram <strong>of</strong> the Cu15Ni8Sn alloy obtained by TEM characterizations <strong>and</strong> electrical<br />

resistivity measurements [214].<br />

6.1.2 Mechanical testing<br />

a) Tensile testing<br />

Elevated temperature tensile testing was performed on a servohydraulic MFL 100 kN<br />

machine (modified for the interrupted creep tests, see Section 5.2.2). The maximum<br />

displacement rate <strong>of</strong> the piston is 400 mm/s; thus, strain rates up to 10 1 s -1 can be<br />

reached for samples having a 30 mm long gage section.<br />

The lamp furnace described in Section 5.2.2 was used. A Type K, Ø 0.2 mm<br />

thermocouple was spot-welded on the center <strong>of</strong> the CuNiSn <strong>alloys</strong>. For low-alloy<br />

109


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

110<br />

<strong>copper</strong> specimens, this was not possible because <strong>of</strong> their high conductivity. In this<br />

case, the thermocouple wires were attached mechanically, using Ni wires, Fig. 6-2.<br />

Figure 6-2: CuOFHC (top) <strong>and</strong> NP8 (bottom) tensile test specimen with attached<br />

thermocouple. The reduced section was grinded before thermocouple attachement.<br />

Tarnishing <strong>of</strong> the samples is due to the heat treatment, 15 min at 500 °C in air for the<br />

CuOFHC material, <strong>and</strong> 1 h at 850 °C under flowing Ar for the NP8 material.<br />

In addition to the high-temperature extensometer described in Section 5.2.2, a clip-on<br />

extensometer was used for the room-temperature tests. The compliance <strong>of</strong> the<br />

machine was thus determined, allowing precise determination <strong>of</strong> the sample<br />

deformation from the cross-head position.<br />

Threaded-end round tensile specimen were machined according to ASTM St<strong>and</strong>ard<br />

E8.A: a subsize specimen proportional to the st<strong>and</strong>ard was chosen, Fig. 5-4. A 4 mm-<br />

diameter sample allows sufficiently rapid heating (about 50 °C/s with a graphite<br />

coating) in the lamp furnace, <strong>and</strong> contains a sufficient number (more than 20 [215])<br />

<strong>of</strong> grains across its diameter.<br />

A peripherical notch (semi-elliptical in section; depth: a = 250 µm, width:<br />

2b = 250 µm) was machined on several specimens in order to reach higher local<br />

stresses.<br />

The reduction in area, RA was measured post mortem with the aid <strong>of</strong> a pr<strong>of</strong>ilometer.<br />

Four measurement <strong>of</strong> the diameter (sample is step-rotated about its axis) were done<br />

on each <strong>of</strong> the two parts <strong>of</strong> the broken specimen at the fracture surface, Ø f , <strong>and</strong> away<br />

from it, Ø hom , in order to measure the homogeneous lateral contraction, RA hom .


RA =<br />

RA<br />

hom<br />

ø − ø<br />

2<br />

ø<br />

100<br />

f 0<br />

2 2<br />

=<br />

b) Impact testing<br />

6.1 EXPERIMENTAL PROCEDURES<br />

Eq.6-1<br />

Eq.6-2<br />

Elevated temperature instrumented Charpy impact testing was performed on CN8<br />

<strong>and</strong> NP8 material with a 100 J apparatus (MFL G.m.b.H. / Model SPH-III / impact<br />

speed: 5000 mm/s).<br />

0<br />

ø − ø<br />

2<br />

ø<br />

[ ] %<br />

100 hom %<br />

0<br />

2 2<br />

[ ]<br />

Rapid heating to the desired temperature was achieved by dipping the samples in a<br />

salt bath. Temperature hold within the latter varied from a few seconds to several<br />

minutes. After that hold, the V-notched 10x10x55 specimen, Fig. 6-3, was rapidly<br />

positioned <strong>and</strong> fractured. A type K thermocouple spot-welded on the specimen<br />

surface allowed the aquisition <strong>of</strong> the temperature until specimen rupture.<br />

45°<br />

detail<br />

R0.25<br />

0<br />

detail<br />

55<br />

8<br />

10<br />

Figure 6-3: V-notched specimen for Charpy impact testing. Dimensions are in mm.<br />

10<br />

111


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

112<br />

6.2 Results<br />

6.2.1 Ductility trough<br />

a) Leaded <strong>and</strong> un<strong>leaded</strong> pure <strong>copper</strong><br />

The measured ductility trough for pure <strong>copper</strong> is displayed in Fig. 6-4. The reduction<br />

in area is shown to decrease sharply with increasing temperature only at low strain<br />

rate, <strong>and</strong> particularly for the less pure material. Results obtained at 400 °C are<br />

striking: extensive ductility is obtained for both CuOFHC <strong>and</strong> Cu4N material at a<br />

strain rate <strong>of</strong> 10 1 s -1 , whereas at 10 -4 s -1 the ductility loss is more manifest for the<br />

CuOFHC material (open symbols) than for the Cu4N material (crossed symbols).<br />

Reduction in area: RA<br />

[%]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Cu annealed<br />

15 min at 500°C<br />

CuOFHC<br />

Cu4N<br />

CuOFHC<br />

Cu4N<br />

ε = 10 s -1 .<br />

ε = 10 -4 s -1 .<br />

ε = 10 -4 s -1<br />

0 100 200 300 400 500 600<br />

Temperature [°C]<br />

ε = 10 s -1<br />

Figure 6-4: Evolution <strong>of</strong> the reduction <strong>of</strong> area measured after elevated temperature<br />

tensile tests at two different strain rates. Square symbols st<strong>and</strong> for 10-4 s-1 , <strong>and</strong> round<br />

ones st<strong>and</strong> for 10 s-1 . Curve fits underline data gathered on the CuOFHC material.<br />

Some data obtained at 400 °C with the Cu4N material are added for comparison<br />

(crossed symbols).<br />

The influence <strong>of</strong> an addition <strong>of</strong> 1 wt.% Pb on the intermediate temperature ductility<br />

<strong>of</strong> OFHC <strong>copper</strong> is shown on Fig. 6-5. Arrows are added to guide the eye while<br />

comparing the ductility <strong>of</strong> <strong>leaded</strong> <strong>and</strong> un<strong>leaded</strong> material for the same test conditions.<br />

.<br />

.


Reduction in area: RA [%]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

anneal<br />

15 min at 500°C<br />

CuOFHC<br />

Cu1Pb<br />

CuOFHC<br />

Cu1Pb<br />

ε = 10 s -1 .<br />

ε = 10 -4 s -1 .<br />

(i)<br />

ε = 10 -4 s -1<br />

0 100 200 300 400 500 600<br />

Temperature [°C]<br />

ε = 10 s -1<br />

6.2 RESULTS<br />

Figure 6-5: Reduction in area <strong>of</strong> <strong>leaded</strong> <strong>and</strong> un<strong>leaded</strong> pure OFHC <strong>copper</strong> after<br />

elevated temperature tensile test for two different strain rates. Square symbols st<strong>and</strong> for<br />

10-4 s-1 , <strong>and</strong> round ones st<strong>and</strong> for 10 s-1 . Curve fits underline data gathered on the<br />

CuOFHC material. Filled symbols are for Cu1Pb material. Arrows are added in order<br />

to help direct comparison between un<strong>leaded</strong> <strong>and</strong> <strong>leaded</strong> material under the same test<br />

conditions.<br />

The flow curves <strong>of</strong> the CuOFHC, Cu1Pb, <strong>and</strong> C99 materials strained at 10 s -1 at<br />

400 °C are displayed in Fig. 6-6. The three flow curves superimpose very well, but<br />

differ in the moment <strong>of</strong> rupture. The Cu1Pb material is characterized by a far lower<br />

ductility (ε r ≈ 10%). The reduction in area at fracture <strong>of</strong> CuOFHC, Cu1Pb, <strong>and</strong> C99<br />

are respectively 97, 15, <strong>and</strong> 33%. Necking <strong>of</strong> the CuOFHC is thus almost to a point.<br />

.<br />

(ii)<br />

.<br />

(iii)<br />

113


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

114<br />

eng. stress [MPa]<br />

200<br />

150<br />

100<br />

50<br />

0<br />

ε .<br />

= 10 s -1<br />

400°C<br />

CuOFHC<br />

RA = 97%<br />

Cu1Pb<br />

RA = 15%<br />

C99<br />

RA = 33%<br />

0 10 20 30 40 50 60<br />

eng. strain [%]<br />

Figure 6-6: Flow curves <strong>of</strong> the CuOFHC (symbol is an open circle), Cu1Pb (open<br />

square), <strong>and</strong> C99 (open diamond) materials at 400 °C, <strong>and</strong> at a strain rate <strong>of</strong> 10 s-1 . For<br />

each material, fracture is emphasized by a solid symbol.<br />

The tensile behaviour <strong>of</strong> these materials at 400 °C, for a strain rate <strong>of</strong> 10 -4 s -1 , is<br />

shown in Fig. 6-7. The behaviour <strong>of</strong> the Cu4N material is also shown. Dynamic<br />

recrystallization <strong>of</strong> this material is clearly operative since waves are observable on<br />

the σ(ε) curve. The yield stress <strong>and</strong> ultimate tensile strength σ max <strong>of</strong> these materials<br />

are also much reduced at lower strain rate, compare Fig. 6-6 with Fig. 6-7. Indeed for<br />

CuOFHC: σ max ( = 10 s -1 ) ≈ 150 MPa <strong>and</strong> σ max ( = 10 -4 s -1 ) ≈ 70 MPa. Moreover,<br />

ε •<br />

the flow curves <strong>of</strong> the different materials do not superpose at low strain rate. Note<br />

also the high elongation <strong>of</strong> the C99 industrial material.<br />

ε •


eng. stress [MPa]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

ε = 10 -4 s -1<br />

.<br />

400°C<br />

CuOFHC<br />

RA = 28%<br />

Cu1Pb<br />

RA = 16%<br />

C99<br />

RA = 38%<br />

Cu4N<br />

RA<br />

= 51%<br />

0 10 20 30 40 50<br />

eng. strain [%]<br />

6.2 RESULTS<br />

Figure 6-7: Flow curves <strong>of</strong> the CuOFHC (symbol is an open circle), Cu1Pb (open<br />

square), C99 (open diamond), <strong>and</strong> Cu4N (open triangle) materials at 400 °C, <strong>and</strong> at a<br />

strain rate <strong>of</strong> 10-4 s-1 . For each material, fracture is emphasized by a solid symbol.<br />

The influence <strong>of</strong> lead as an alloying element was shown above. Liquid lead supplied<br />

from the external surface <strong>of</strong> <strong>copper</strong> specimen also influences the tensile behaviour <strong>of</strong><br />

pure Cu at 400 °C, as shown on Fig. 6-8. All the curves superimpose well except for<br />

that <strong>of</strong> the dipped CuOFHC sample, which is somehow lower (probably due to<br />

coarser grains): ductility loss is obvious after such a treatment. Concerning Pb<br />

electroplated samples, only the notched specimen exhibits <strong>embrittlement</strong>. The<br />

reduction in area <strong>of</strong> the unnotched specimen is in accordance with the elongation at<br />

fracture. RA is namely 96 <strong>and</strong> 49% for OFHC <strong>copper</strong> electroplated with lead, <strong>and</strong><br />

dipped in lead, respectively. By comparison, RA=97% for CuOFHC material.<br />

It can also be noted that the <strong>copper</strong>-<strong>embrittlement</strong> induced by liquid lead is far<br />

stronger if the lead is pre-alloyed with <strong>copper</strong>. This is obvious if one compares the<br />

behaviour <strong>of</strong> the Cu1Pb, Fig. 6-6, <strong>and</strong> that <strong>of</strong> Pb-coated CuOFHC, Fig. 6-8.<br />

115


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

116<br />

Figure 6-8: 400 °C flow curves <strong>of</strong> CuOFHC material at a rapid strain rate <strong>of</strong> 10 s -1 :<br />

the influence <strong>of</strong> a Pb coating, <strong>and</strong> the effect <strong>of</strong> a notch are shown. Square symbols st<strong>and</strong><br />

for electroplated material, <strong>and</strong> diamond are for dipped material. Grey filled symbols<br />

represent notched specimen. For each material, fracture is emphasized by a solid<br />

symbol.<br />

b) Leaded <strong>and</strong> un<strong>leaded</strong> Cu-Ni-Sn <strong>alloys</strong><br />

The intermediate temperature tensile behaviour <strong>of</strong> CN8 <strong>and</strong> NP8 materials was also<br />

quantified, Fig. 6-9. In the whole temperature range studied, the NP8 material is<br />

more brittle than CN8. For both materials, the ductility decreases above 200 °C. In<br />

addition, above 327 °C (the Pb melting point), the ductility drops sharply for the<br />

lead-containing NP8 material, <strong>and</strong> then remains roughly constant as the temperature<br />

increases.<br />

eng. stress [MPa]<br />

200<br />

150<br />

100<br />

50<br />

0<br />

= 10 s -1<br />

ε .<br />

400°C<br />

(pn)<br />

(n)<br />

CuOFHC<br />

CuOFHC notched (n)<br />

Cu + Pb plated (p)<br />

Cu + Pb plated & notched (pn)<br />

Cu + Pb dipped (d)<br />

0 10 20 30 40 50 60<br />

eng. strain [%]<br />

(d)<br />

(p)


Reduction in area: RA<br />

[%]<br />

6.2 RESULTS<br />

Figure 6-9: Reduction in area <strong>of</strong> CN8 <strong>and</strong> NP8 material strained at 10 -2 s -1 . The<br />

melting temperature <strong>of</strong> lead, 327 °C is indicated.<br />

The influence <strong>of</strong> room temperature strain hardening on the intermediate temperature<br />

tensile properties <strong>of</strong> solutionized <strong>and</strong> quenched CuNiSn <strong>alloys</strong> is shown in Fig. 6-10.<br />

Notice the comparatively high flow stress (~200 MPa) <strong>of</strong> the CN8 alloy in the<br />

annealed monophase state. In the same conditions, the flow curves <strong>of</strong> CN8 <strong>and</strong> NP8<br />

superimpose quite well. The ductility <strong>of</strong> the lead-containing NP8 is about 1% at<br />

400 °C, <strong>and</strong> drops to essentially zero for specimens strain-hardened at room<br />

temperature.<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

CN8<br />

NP8<br />

0 100 200 300 327 400 500 600<br />

Temperature [°C]<br />

CN8 & NP8<br />

annealed<br />

60 min at 850°C<br />

& quenched<br />

.<br />

ε = 10 -2 s -1<br />

117


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

118<br />

eng. stress [MPa]<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

Figure 6-10: 400 °C flow curves <strong>of</strong> CN8 & NP8 material at a strain rate <strong>of</strong> 10 -2 s -1 .<br />

Room temperature strain hardened specimen are differenciated with grey filled symbols.<br />

The ductility <strong>of</strong> NP8 is much lower than that <strong>of</strong> CN8. For each material, fracture is<br />

emphasized by a solid symbol.<br />

6.2.2 Impact energy<br />

0<br />

400 °C<br />

CN8<br />

CN8 (RT strain hardened)<br />

NP8<br />

NP8 (RT strain hardened)<br />

0 3 6 9 12 15<br />

eng. strain [%]<br />

Charpy impact tests were performed on both CN8, <strong>and</strong> NP8 material at room- <strong>and</strong><br />

intermediate temperatures. Results are summarized on Fig. 6-11. The impact energy<br />

<strong>of</strong> the un<strong>leaded</strong> CN8 material is much higher than that <strong>of</strong> the NP8 material over the<br />

whole temperature range explored. Notice that between 400 <strong>and</strong> 500 °C, phase<br />

transformation occurs before impact, due to high rate <strong>of</strong> the phase transformation,<br />

Fig. 6-1 (b), <strong>and</strong> finite heating time within the salt bath (about 5 s to 400 °C, plus 5 s<br />

to 500 °C, <strong>and</strong> an additionnal 5 s to position the sample <strong>and</strong> run the test).<br />

Comparing the Pb-containing NP8 material with CN8, it is clear that lead inclusions<br />

induce a sharp decrease in impact energy at all temperatures.<br />

CN8 & NP8<br />

annealed<br />

60 min at 850°C<br />

& quenched<br />

ε = 10 -2 s -1<br />

.


6.2.3 Damage<br />

Impact energy: CVN [J]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

CN8<br />

NP8<br />

6.2 RESULTS<br />

Figure 6-11: Impact energy <strong>of</strong> monophased CN8 <strong>and</strong> NP8 material. Grey symbols<br />

underline the occurence <strong>of</strong> rapid phase transformation just before impact. The melting<br />

temperature <strong>of</strong> lead, 327 °C is indicated.<br />

Damage in Cu4N, Cu1Pb, CuOFHC, <strong>and</strong> C99 materials strained at 400 °C at both<br />

10 s -1 , <strong>and</strong> 10 -4 s -1 was observed on polished longitudinal sections. SEM images are<br />

shown on Fig. 6-12. The lead-containing materials were imaged in the Back<br />

Scattered Electron mode, BSE, in order to reveal the chemical contrast. Left-h<strong>and</strong><br />

images were taken from a specimen strained at 10 s -1 , whereas right-h<strong>and</strong> images are<br />

from samples tested at 10 -4 s -1 .<br />

Intergranular decohesion is most marked in the CuOFHC <strong>and</strong> Cu1Pb materials<br />

strained at 10 -4 s -1 . Under the same testing conditions, the Cu4N material is less<br />

damaged, <strong>and</strong> the C99 is almost sound. These observations remain valid if one<br />

considers the damage remote from the fracture surface. Note that for all materials<br />

tested at low strain rate, intergranular damage is more pronounced near the surface <strong>of</strong><br />

the specimen than in its interior.<br />

0 100 200 300 327 400 500 600<br />

Temperature [°C]<br />

CN8 & NP8<br />

annealed<br />

60 min at 850°C<br />

At the higher strain rate <strong>of</strong> 10 s -1 , the CuOFHC, <strong>and</strong> Cu4N materials fracture in a<br />

ductile manner with almost perfect necking. There is therefore no intergranular<br />

damage near the fracture surface nor in the specimen interior. On the other h<strong>and</strong>, the<br />

119


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

120<br />

lead-containing materials fracture intergranularly, <strong>and</strong> damage is only concentrated in<br />

the vicinity <strong>of</strong> the fracture surface. No damage was detected away from the fracture<br />

surface in the four materials tested. There is also no evidence for the transient<br />

existence <strong>of</strong> a Pb-rich liquid film in grain boundaries subjected to tensile stress: the<br />

inclusions appear rather as being mostly elongated in the tensile direction, Fig. 6-<br />

12 (g).


121<br />

(a) Cu4N dε/dt=10 s -1 RA=96% (b) Cu4N dε/dt=10 -4 s -1 RA=51%<br />

(c) Cu1Pb dε/dt=10 s -1 RA=15% (d) Cu1Pb dε/dt=10 -4 s -1 RA=16%<br />

(e) CuOFHC dε/dt=10 s -1 RA=97% (f) CuOFHC dε/dt=10 -4 s -1 RA=28%<br />

(g) C99 dε/dt=10 s -1 RA=33% (h) C99 dε/dt=10 -4 s -1 RA=38%<br />

Figure 6-12: Longitudinal cuts at the fracture surface <strong>of</strong> Cu4N, Cu1Pb, CuOFHC, <strong>and</strong><br />

C99 materials after 400 °C tensile test in air. Strain rate is 10 s -1 (left images) <strong>and</strong> 10 -<br />

4 s -1 (right images). The tensile axis is vertical. Pb-containing materials were imaged in<br />

the BSE mode, <strong>and</strong> Pb-free materials in the SE mode.


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

122<br />

Liquid lead surface-coatings are on the other h<strong>and</strong> found to migrate very rapidly with<br />

a propagating crack. This is illustrated in Fig. 6-13 comparing a CuOFHC tensile<br />

specimen dipped in liquid lead for 18 h at 600 °C before (a), <strong>and</strong> after (b) tensile<br />

testing at 400°C <strong>and</strong> = 10 s -1 .<br />

(a) (b)<br />

Figure 6-13: SEM images <strong>of</strong> longitudinal cuts in the BSE mode: lead appears brighter<br />

than <strong>copper</strong>. (a) Mullins-type groove <strong>of</strong> a CuOFHC sample dipped in a Pb bath<br />

saturated in Cu for 18 h at 600 °C, the depth <strong>of</strong> the groove is 12 µm. (b) Similar sample<br />

subsequently strained at 400 °C, <strong>and</strong> dε/dt = 10 s-1 : a subcritical crack nucleated from<br />

a grain boundary groove, <strong>and</strong> propagated to a depth d=560 µm, see Fig. 6-8.<br />

6.2.4 Fractography<br />

ε •<br />

It is clear from Fig. 6-14 that the fracture <strong>of</strong> Cu1Pb strained at intermediate<br />

temperature <strong>and</strong> elevated strain rate is intergranular. At 400 °C, above the Pb melting<br />

point, the lead-rich inclusions are molten. Inclusions indeed did not melt in the<br />

Cu1Pb material tested at 300 °C; nevertheless, the material fractured intergranularly.


(a) Cu1Pb 300 °C dε/dt=10 s -1 RA=28% (b)<br />

(c) Cu1Pb 400 °C dε/dt=10 s -1 RA=15% (d)<br />

6.2 RESULTS<br />

(e) C99 400 °C dε/dt=10 s-1 RA=33% (f)<br />

Figure 6-14: SEM fractographs in the BSE mode. The Cu1Pb material was tensile<br />

tested in air at dε/dt=10 s-1 at (a) & (b) 300 °C, below the Pb melting point, <strong>and</strong> (c) &<br />

(d) at 400 °C, where the lead is molten. Fractographs <strong>of</strong> the C99 material tested at<br />

400 °C are added for comparison (e) & (f).<br />

123


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

124<br />

6.3 Discussion<br />

6.3.1 Embrittlement Mechanisms<br />

The effect <strong>of</strong> lead in <strong>copper</strong> may be three-fold: (i) lead inclusions act as stress<br />

concentration sites <strong>and</strong> pin the grain boundaries; this renders cavitation easier <strong>and</strong><br />

reduces the grain boundary mobility retarding dynamic recrystallization, (ii)<br />

segregated lead may decrease the grain boundary cohesion; (iii) once liquid, lead<br />

may promote liquid metal <strong>embrittlement</strong>.<br />

a) Grain Boundary Embrittlement<br />

GBE is a time-dependent process, which therefore becomes exacerbated with<br />

decreasing strain rate. This results in a widening <strong>and</strong> deepening <strong>of</strong> the ductility<br />

trough. A clear illustration <strong>of</strong> GBE for pure <strong>copper</strong> is given on Fig. 6-4, where our<br />

CuOFHC <strong>and</strong> Cu4N materials display a much reduced RA when strained at 400 °C<br />

<strong>and</strong> = 10 -4 s -1 as compared to the same materials at = 10 s -1 . These results are<br />

ε •<br />

replotted on Fig. 6-15 (a) <strong>and</strong> are shown to compare very well with literature, Fig. 6-<br />

15 (b) from Ref. [49].<br />

We note that GBE operates in all materials we studied at low strain rate, since none<br />

demonstrates full ductility, Fig. 6-7. On the other h<strong>and</strong>, at elevated strain rate, there is<br />

no time allowed for grain boundary <strong>embrittlement</strong>, <strong>and</strong> pure <strong>copper</strong> samples exhibit<br />

necking to a point.<br />

ε •


Reduction in area: R [%]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 200 400 600<br />

6.3 DISCUSSION<br />

(a) (b)<br />

Figure 6-15: (a) Same Figure as Fig. 6-4 displaying the ductility trough <strong>of</strong> CuOFHC &<br />

Cu4N material tested in air. (b) At the same scale, results from the literature [49]<br />

(tensile test under vacuum: p=10-2 Pa): pure <strong>copper</strong> is Cu 99,99% polycrystal, <strong>and</strong><br />

internally oxydized <strong>copper</strong> contains dispersed Ø 0.05 µm Al2O3 particles.<br />

Our results are also plotted on the Ashby fracture-mechanism map [52] for CuOFHC,<br />

Fig. 6-16. The ultimate tensile strength is plotted as a function <strong>of</strong> temperature for our<br />

CuOFHC samples strained at 10 s -1 <strong>and</strong> 10 -4 s -1 . Solid <strong>and</strong> open symbols st<strong>and</strong><br />

respectively for intergranular <strong>and</strong> transgranular fracture. The agreement with the<br />

fracture-mechanisms map is excellent. We note that complete necking due to DRX is<br />

predicted for a temperature above 600 °C, in agreement with the results presented on<br />

Fig. 6-15 (b). Ultra-high purity material can, however, display dynamic<br />

recrystallization at a much lower temperature [68]. In this case, the boundary <strong>of</strong> the<br />

”Rupture” field on Fig. 6-16 is shifted to the left. The occurence <strong>of</strong> dynamic<br />

recrystallization is manifest on Fig. 6-7 for the Cu4N material. This is recognized by<br />

the wavy evolution <strong>of</strong> the σ−ε curve. Cu4N is the purest material we tested; at 400 °C<br />

<strong>and</strong> = 10 -4 s -1 , it demonstrates indeed the highest ductility in term <strong>of</strong> both strain<br />

ε •<br />

CuOFHC<br />

Cu4N<br />

CuOFHC<br />

Cu4N<br />

e = 10 s -1 .<br />

e = 10 -4 s -1 .<br />

Temperature [ C]<br />

<strong>and</strong> reduction in area at fracture, Fig. 6-7.<br />

125


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

126<br />

ultimate tensile stress UTS [MPa]<br />

1000<br />

100<br />

10<br />

Figure 6-16: Ashby fracture mechanism map for CuOFHC material, redrawn from<br />

Fig. 3-18, with our datapoints. For these, the y-axis value is the UTS. Circle symbols<br />

st<strong>and</strong> for a strain rate <strong>of</strong> 10 s-1 , <strong>and</strong> squares for 10-4 s-1 . Solid, <strong>and</strong> open symbols<br />

indicate an intergranular, <strong>and</strong> intragranular fracture respectively.<br />

Influence <strong>of</strong> segregated lead<br />

Ductile Fracture<br />

Transgranular<br />

Creep Fracture<br />

Considering Fig. 6-12 (b) & (d), it is clear that the addition <strong>of</strong> lead to <strong>copper</strong><br />

exacerbates grain boundary <strong>embrittlement</strong>. Indeed Cu1Pb is more damaged than<br />

Cu4N. Moreover the former is much less ductile than the latter, Fig. 6-7. These<br />

findings are in accordance with the literature. It was namely reported that the 400 °C<br />

ductility <strong>of</strong> Cu OFHC, tested in air, is reduced by a prior treatment at 800 °C under a<br />

Pb vapor atmosphere [216]. It can therefore be concluded that lead, as an impurity in<br />

<strong>copper</strong>, induces GBE even at very small concentrations.<br />

Influence <strong>of</strong> segregated oxygen<br />

Dynamic Fracture<br />

Intergranular<br />

Creep Fracture<br />

Rupture<br />

(DRX)<br />

Cu OFHC<br />

ε = 10<br />

1<br />

0 200 400<br />

Temperature [°C]<br />

600 800<br />

-4 s -1<br />

ε = 10 s<br />

.<br />

-1 .<br />

Impurities such as oxygen, segregated to the grain boundaries <strong>of</strong> <strong>copper</strong>, are also<br />

expected to induce GBE. Within the materials we tested at = 10 -4 s -1 , intergranular<br />

damage is found throughout the whole specimen <strong>and</strong> is more pronounced near the<br />

specimen surface than in its interior. This suggests that stress-driven oxygen diffusion<br />

occurs from the surrounding air during the test, <strong>and</strong> weakens the grain boundaries.<br />

This is in agreement with the literature. Oxygen diffusion along the grain boundaries<br />

<strong>of</strong> <strong>copper</strong> was explicitly reported to occur at intermediate temperature only if a<br />

ε •


6.3 DISCUSSION<br />

tensile stress is applied [217]. Moreover, intergranular damage revealed by<br />

synchrotron microradiography in oxygen-saturated <strong>copper</strong> subjected to creep at<br />

500 °C was also found to be on one h<strong>and</strong> governed by a diffusion mechanism, <strong>and</strong>,<br />

on the other h<strong>and</strong>, to be controlled primarily by the maximum principal stress [109].<br />

Intergranular damage was furthermore documented by Bleakney in pure <strong>copper</strong><br />

subjected to creep in air. It is reported that the creep rupture <strong>embrittlement</strong> <strong>of</strong> high-<br />

purity <strong>copper</strong> is caused by a combination <strong>of</strong> stress-corrosion <strong>and</strong> internal oxidation <strong>of</strong><br />

the grain boundaries [218].<br />

From the above-mentioned considerations, we expect therefore that the Cu4N<br />

material subjected to low strain rate tensile testing at intermediate temperature would<br />

display a higher ductility if tested under vacuum.<br />

Phosphorus is added to the industrial C99 material as a desoxidizing element <strong>of</strong> the<br />

melt. Therefore the important reduction <strong>of</strong> intergranular damage <strong>of</strong> this material as<br />

compared to the others, see Fig. 6-12 (left column), can be attributed to the presence<br />

<strong>of</strong> phosphorus acting as an oxygen scavenger. This parallels findings on bronze:<br />

oxygen is indeed scavenged by boron or magnesium added to a Cu-8Sn bronze,<br />

suppressing the penetration <strong>of</strong> oxygen along the grain boundaries, <strong>and</strong> restoring in<br />

turn the high temperature ductility <strong>of</strong> the bronze [71].<br />

Note that the embrittling effect <strong>of</strong> oxygen is reversible. A heat treatment at 950 °C<br />

under a 10 -5 Torr vacuum was namely found to remove the oxygen <strong>and</strong> restore both<br />

tensile <strong>and</strong> creep ductility <strong>of</strong> coarse-grained oxygen saturated 99,995% Cu that was<br />

treated for 2 h at 800 °C under an O 2 atmosphere [108]. Miura et al. accordingly<br />

report that a 950 °C 6 h degassing treatment in vacuum alleviate the GBE <strong>of</strong><br />

internally oxidized Cu-SiO 2 polycrystals [57].<br />

This suggests that, for <strong>copper</strong> <strong>alloys</strong> <strong>of</strong> st<strong>and</strong>ard purity, intermediate temperature<br />

ductility is enhanced if desegregation occurs or if the embrittling segregated species<br />

are scavenged. On the other h<strong>and</strong> it is clear from Fig. 6-4 that st<strong>and</strong>ard purity <strong>copper</strong><br />

is not embrittled at intermediate temperature if rapidly strained, as one would expect<br />

127


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

128<br />

for such a time-dependent damage phenomenon in the absence <strong>of</strong> other<br />

<strong>embrittlement</strong> mechanisms.<br />

b) Liquid Metal Embrittlement<br />

Above the lead melting point, the impact energy <strong>of</strong> pure <strong>copper</strong> is found to be much<br />

reduced due to the addition <strong>of</strong> 1 wt.% Pb to Cu, Fig. 6-17 (b). This is a clear<br />

manifestation <strong>of</strong> LME, since pure <strong>copper</strong> does not exhibit any <strong>embrittlement</strong> at such<br />

high strain rates above 327 °C [130]. It was also reported by Roth that mixed<br />

intergranular <strong>and</strong> transgranular fracture is obtained for <strong>leaded</strong> <strong>copper</strong> samples tested<br />

just below 327 °C. This is probably due to remaining undercooled liquid inclusions,<br />

since the sample was impacted while cooling from a temperature higher than 350 °C<br />

[130].<br />

impact energy [J]<br />

40<br />

30<br />

20<br />

10<br />

0<br />

monophase NP8<br />

850 °C annealed<br />

& quenched<br />

(a) (b)<br />

Figure 6-17: Intermediate temperature impact energy <strong>of</strong> (a) the monophased NP8<br />

material, see Fig. 6-11 <strong>and</strong> (b) pure <strong>copper</strong>, <strong>and</strong> Cu-1Pb alloy [130]. V-notched samples<br />

were tested (a) 10x10 mm 2 in section, <strong>and</strong> (b) 8 mm in diameter.<br />

The resemblance <strong>of</strong> these literature results [130] with our results obtained on the<br />

lead-containing NP8 material is clear, Fig. 6-17. This suggests therefore that this<br />

latter material is prone to LME.<br />

327 °C<br />

100 200 300<br />

Temperature [°C]<br />

Comparing our <strong>leaded</strong> <strong>and</strong> un<strong>leaded</strong> materials tensile tested at 400 °C <strong>and</strong> 10 s-1 leads to the same conclusion when considering the relative reduction in area at<br />

fracture, Fig. 6-6. At this high strain rate, the flow curves <strong>of</strong> the different materials<br />

tested superpose very well until the sudden rupture <strong>of</strong> the Pb-containing materials,<br />

Fig. 6-6. This is a classical signature <strong>of</strong> LME [132].


6.3 DISCUSSION<br />

The onset <strong>of</strong> a ductility trough due to LME is associated with incipient melting <strong>of</strong> the<br />

embrittler. A drastic change in fracture surface occurs, from transgranular to fully<br />

intergranular, see e.g. [219]. At 400 °C <strong>and</strong> = 10 s -1 , fracture is indeed found to be<br />

fully intergranular in Cu1Pb <strong>and</strong> C99 materials, Fig. 6-14, <strong>and</strong> damage is<br />

concentrated at the fracture surface. Un<strong>leaded</strong> <strong>copper</strong> materials tested in the same<br />

conditions exhibit fully ductile transgranular fracture, Fig. 6-12 (left column).<br />

The sharp ductility loss at 400 °C measured on our Cu1Pb material subjected to high<br />

strain rate tensile testing, Fig. 6-5, also compares well with results obtained on a<br />

phosphorus bronze containing 2 wt.% lead, where it was observed that above 327 °C,<br />

secondary cracks were filled by liquid lead [65]. Intergranular lead inclusion also act<br />

as liquid metal reservoir at grain boundaries: in the present samples: an emptied<br />

lenticular inclusion is illustrated on Fig. 6-18.<br />

(a) (b)<br />

Figure 6-18: SEM images (BSE mode) <strong>of</strong> a longitudinal cut <strong>of</strong> a prolonged annealed<br />

(900 °C 1 h <strong>and</strong> 400 °C 24 h) Cu1Pb material tensile tested at 400 °C <strong>and</strong> 10 s-1 , the<br />

stress axis is horizontal: (a) intergranular microcrack at low magnification. (b) detailed<br />

view <strong>of</strong> the tip <strong>of</strong> the same microcrack.<br />

Pb-coated samples<br />

Turning to Pb-coated samples, we also find a behaviour characteristic <strong>of</strong> LME. Pb-<br />

coating caused varying effects on the ductility <strong>of</strong> our CuOFHC specimen. Namely no<br />

influence was noted for the electroplated smooth sample, Fig. 6-8, whereas<br />

spectacular <strong>embrittlement</strong> was observed in the dipped samples, Fig. 6-13. Such<br />

erratic manifestation <strong>of</strong> LME is well known: notched specimen can be embrittled in<br />

conditions under which smooth samples are immune [132 , 142]. To rationalize the<br />

ε •<br />

129


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

130<br />

different behaviour <strong>of</strong> the dipped <strong>and</strong> the electroplated samples, it must be<br />

emphasized that the former exhibited slight grain boundary grooves, in intimate<br />

contact with lead. In adddition, electroplated notched specimen displayed<br />

<strong>embrittlement</strong>, Fig. 6-8. This suggests that the migration <strong>of</strong> liquid lead is associated<br />

with crack propagation; however, crack initiation requires that additional conditions<br />

be met. According to Roth, the presence <strong>of</strong> a liquid metal wetting the grain boundary<br />

decreases the grain boundary cohesion <strong>and</strong> favours the nucleation <strong>of</strong> a brittle crack<br />

once a slip b<strong>and</strong> impinges, on a grain boundary ahead <strong>of</strong> an intergranular liquid<br />

inclusion [66].<br />

Our results on Pb-coated specimen also agree with results reported on the LME<br />

susceptibility <strong>of</strong> martensitic steel in contact with liquid Pb-Bi. Namely the presence<br />

<strong>of</strong> an oxide layer on the T91 steel renders it immune to LME attacks [144].<br />

It is moreover worthwhile noting that Sanchez Medina claimed that Cu-Pb is not<br />

prone to LME since OFHC Cu samples surrounded by a ring <strong>of</strong> liquid Pb behave like<br />

uncoated specimen at temperature ranging from RT to 1000 °C (similarly to our<br />

electroplated smooth samples, Fig. 6-8) [47].<br />

Summary <strong>of</strong> observations<br />

We conclude, in summary, that at high strain rates, Cu-Pb <strong>alloys</strong> fracture by liquid<br />

metal <strong>embrittlement</strong>. Furthermore, the data provide indications that a critical stress<br />

must be exceeded in order to cause such LME fracture:<br />

• in Fig. 6-10, it is seen that at 400 °C, strain-hardened NP8 fails at 220 MPa before<br />

any macroscopic plastic deformation, whereas with annealed NP8, the initial flow<br />

stress <strong>of</strong> which is lower, failure occurs after some plastic deformation when roughly<br />

the same stress is reached (200 MPa);<br />

• in lead-coated specimens, failure by LME was only triggered when there was a site<br />

<strong>of</strong> stress concentration along the surface <strong>of</strong> the sample, taking the form <strong>of</strong> a grain<br />

boundary groove or a pre-machined notch, Fig. 6-8.


Is LME gouverned by a critical value <strong>of</strong> dimensionless parameter Λ ?<br />

6.3 DISCUSSION<br />

The observation that a critical stress must be exceeded for LME to cause fracture <strong>of</strong><br />

Cu-Pb specimens above the melting point <strong>of</strong> lead suggests the possibility that<br />

fracture is dominated by the dimensionless parameter Λ defined in Eq. 5-11 . This<br />

parameter is shown to trigger unstable flattening <strong>of</strong> liquid inclusions under tensile<br />

stress, i.e., crack formation <strong>and</strong> catastrophic growth when Λ exceeds a critical value<br />

Λ c , defined at the maximum <strong>of</strong> curves for K Pb = 0 in Fig. 5-8 (a). This catastrophic<br />

crack propagation is in particular suggested by numerical simulations <strong>of</strong> Wang,<br />

which indicate much accelerated shape changes for pores (changing shape by surface<br />

<strong>and</strong> grain boundary diffusion) if Λ>Λ c , Fig. 3-15 [41]. That very rapid crack<br />

propagation can occur at the tip <strong>of</strong> a crack wetted by the liquid (i.e., by LME) is<br />

shown quantitatively by Robertson [152].<br />

The suggestion that fracture by LME is triggered by attainment <strong>of</strong> a critical value, Λ c ,<br />

<strong>of</strong> parameter Λ is supported by the following observations.<br />

1 - As mentioned above, a critical stress must be exceeded;<br />

2 - this critical stress is lowered when the inclusion diameter increases.<br />

This is seen in Fig. 6-19, which compares tensile data for fine-grained <strong>copper</strong> <strong>and</strong><br />

Cu-Pb <strong>alloys</strong>, <strong>and</strong> their coarse-grained equivalents obtained by annealings at 900°C<br />

for 1 h prior to testing. During grain growth the lead inclusions coalesce, causing<br />

their average size to increase (see Section 4.1.2). As seen, solid lines drawn on<br />

Fig. 6-19 for the fine-grained materials CuOFHC, C99 <strong>and</strong> Cu1Pb superimpose quite<br />

well, the <strong>leaded</strong> samples fracturing prematurely at 212 MPa (C99) <strong>and</strong> 118 MPa (Cu-<br />

1Pb). This is also true for the dashed lines corresponding to coarse-grained materials<br />

C99 <strong>and</strong> Cu1Pb; however, in these, the critical stress at which LME causes premature<br />

failure is significantly lower, namely 129 MPa for C99, <strong>and</strong> 80 MPa for Cu-1Pb.<br />

131


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

132<br />

Figure 6-19: Flow curves <strong>of</strong> the CuOFHC (symbol is an open circle), Cu1Pb (open<br />

square), <strong>and</strong> C99 (open diamond) materials at 400 °C, <strong>and</strong> at a strain rate <strong>of</strong> 10 s-1 .<br />

Crossed symbol correspond to the fracture. The flow curves <strong>of</strong> the fine grained material<br />

are evidenced by solid lines. Tensile behaviour <strong>of</strong> coarse grained materials (no data for<br />

CuOFHC) follow the dashed lines. Lead inclusions size in the fine grained Cu1Pb &<br />

C99 samples is about 1 µm, whereas it is around 10 µm in the coarse grained samples.<br />

3 - The critical fracture stress is decreased by a lowering <strong>of</strong> the stress-free dihedral<br />

angle φ o .<br />

true stress [MPa]<br />

250<br />

200<br />

150<br />

100<br />

50<br />

Liquid Ga is known to perfectly wet the general grain boundaries <strong>of</strong> Al (φ o = 0°). In<br />

this case, as is intuitively obvious (<strong>and</strong> also borne out by the analysis <strong>of</strong> Chapter 5<br />

with φ o = 0°), the critical value <strong>of</strong> Λ c , <strong>and</strong> hence the critical fracture stress σ c , are nil.<br />

Liquid lead, on the other h<strong>and</strong>, does not wet perfectly the <strong>copper</strong> grain boundaries at<br />

intermediate temperature, since φ o > 0°, see e.g. Fig. 4-14. In the Cu-Pb system, Λ c is<br />

thus finite positive.<br />

0<br />

ε .<br />

= 10 s -1<br />

400°C<br />

coarse<br />

grained<br />

fine<br />

grained<br />

0 10 20 30 40 50<br />

true strain [%]<br />

CuOFHC Fracture<br />

C99 Fracture<br />

Cu1Pb Fracture<br />

The influence <strong>of</strong> φ o is also illustrated in Fig. 6-19, where it is seen that, for both<br />

inclusion sizes, Cu1Pb is more brittle than C99. This is probably due to the slight<br />

difference in interfacial chemical composition between the two <strong>alloys</strong>, reflected by<br />

an average lower φ o in the purer Cu1Pb material. Indeed at 400 °C, the average value<br />

<strong>of</strong> φ o in Cu-1Pb is 100°, Fig. 4-14; whereas φ is about 120° for the C99 material; this<br />

is shown in Appendix 10.4, where the results <strong>of</strong> four dihedral angle measurements in<br />

C99 at 400°C are given. Now, Λ c depends on the stress-free initial dihedral angle φ o


6.3 DISCUSSION<br />

(see Chapter 5): a decrease <strong>of</strong> φ o results in a decrease in Λ c , <strong>and</strong> hence in σ c , all else<br />

being constant, Fig. 6-20.<br />

We note in passing that such a dependence on the dihedral angle, whereby increasing<br />

the dihedral angle increases the resistance to LME, is in broad agreement with the<br />

grain boundary engineering concept <strong>of</strong> Watanabe illustrated on Al bicrystals wetted<br />

by Ga [220], <strong>and</strong> with Kargols’ findings that a lower grain boundary energy (higher<br />

φ) ensures a higher resistance to LME in Al bicrystals in contact with liquid Hg-<br />

3 at.% Ga [148].<br />

Λ [-]<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

120<br />

Figure 6-20: Predicted Λ(φ) curves for the C99 <strong>and</strong> Cu1Pb materials at 400 °C. Inputs<br />

for the theoretical calculations are identical save for φ 0 , the equilibrium dihedral angle<br />

under stress-free conditions. Predictions are for a grain boundary void, lenticular in<br />

shape.<br />

Finally, we note that quantitative comparison <strong>of</strong> the predictions <strong>of</strong> the analysis in<br />

Chapter 5 <strong>and</strong> data <strong>of</strong> Fig. 6-19 is satisfying given the assumptions made in the<br />

model. At 400 °C, we have γ SL = 0.432 J/m 2 , <strong>and</strong> E = 111 GPa. Inclusions 1 µm <strong>and</strong><br />

10 µm in diameter have a volume <strong>of</strong> about 10 -19 <strong>and</strong> 10 -16 m 3 respectively. From the<br />

Λ c values shown on Fig. 6-20, the predicted critical stress, σ c is determined from<br />

Eq. 6-3, <strong>and</strong> given in Table 6-1.<br />

1<br />

φ 0 C99<br />

400 °C<br />

100<br />

φ 0 Cu1Pb<br />

C99<br />

80<br />

Cu1Pb<br />

60<br />

40<br />

apparent dihedral angle: φ [°]<br />

Λ c C99 = 0.61<br />

Λ c Cu1Pb = 0.41<br />

20<br />

crack<br />

133


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

134<br />

σ<br />

Eq.6-3<br />

Table 6-1: Predicted critical stress σ c for the Cu1Pb <strong>and</strong> C99 materials subjected to uniaxial tension at<br />

400 °C according to Eq. <strong>and</strong> Fig. 6-20. Experimental values are from Fig. 6-19.<br />

400 °C<br />

As seen, values do not match with precision, which is not surprising given<br />

assumptions <strong>of</strong> the calculation <strong>and</strong> the absence <strong>of</strong> any adjustable parameter; however,<br />

variations caused by changes in alloy composition or grain size, as well as the order<br />

<strong>of</strong> magnitude <strong>of</strong> the fracture stress <strong>of</strong> the C99 <strong>and</strong> Cu1Pb materials tensile tested at<br />

400 °C, Fig. 6-19, are quite well predicted.<br />

Similarly, a fracture stress can be predicted for the lead-containing NP8 material if<br />

the dimension <strong>and</strong> geometry <strong>of</strong> the lead inclusions are known. φ can be measured<br />

from fractographs, Fig. 6-21 (b), using the method described in Section 4.3.2,<br />

slightly modified since we use the 3D reconstruction <strong>of</strong> the grain boundary plane <strong>and</strong><br />

<strong>of</strong> only one (but whole) former liquid inclusion/solid metal interface, Fig. 6-22.<br />

Optimal measurement is achieved with an emptied inclusion on an intergranular<br />

fracture surface, which has to be unoxidized <strong>and</strong> free <strong>of</strong> distortion due to plastic<br />

deformation.<br />

γ<br />

V<br />

SL ⋅ E<br />

c = Λc 1/ 3<br />

theoretical predictions experimental values<br />

σ c Cu1Pb σ c C99 σ max Cu1Pb σ max C99<br />

fine grained 205 MPa 251 MPa 118 MPa 212 MPa<br />

coarse grained 64 MPa 79 MPa 80 MPa 129 MPa


(a) (b)<br />

6.3 DISCUSSION<br />

Figure 6-21: Fracture surface <strong>of</strong> a NP8 cracked during quenching in oil from 900 °C.<br />

Figure 6-22: Data points (x’s) from the 3D reconstruction <strong>of</strong> the former solid/liquid<br />

interface <strong>of</strong> the inclusion illustrated on Fig. 6-21 (b). The sphere fitting these<br />

datapoints, as well as the reconstructed grain boundary plane are shown. φ = 64° <strong>and</strong><br />

V = 3.5 10-17 m3 .<br />

The inclusion volume V, <strong>and</strong> dihedral angle φ are calculated from the determined<br />

equations <strong>of</strong> the grain boundary plane <strong>and</strong> fitting sphere. For the inclusion (the only<br />

characterized here) displayed in Fig. 6-21 (b): φ = 64° <strong>and</strong> V = 3.5 10 -17 m 3 .<br />

A rough estimate <strong>of</strong> Λ c is calculated taking into account φ o = 64°, the elastic<br />

properties <strong>of</strong> the NP8 at 800 °C (E = 80 GPa, ν = 0.364), γ SL = 0.387 J/m 2 at 800 °C<br />

estimated for the Cu-Pb system according to the regular solution model, Eq. 4-17,<br />

<strong>and</strong> taking K Pb = 0 GPa (i.e. considering the inclusion as a void). The calculations<br />

give Λ c = 0.16 <strong>and</strong> 0.10 for lenticular <strong>and</strong> ellipsoidal inclusion geometries,<br />

respectively.<br />

Grain Boundary plane<br />

135


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

136<br />

From Eq. 6-3, this gives σ c = 39 <strong>and</strong> 30 MPa respectively. These values are in good<br />

agreement with the measured resistance <strong>of</strong> NP8 at 800 °C, namely R m = 44 MPa,<br />

Fig. 7-6, if one remembers that there is no adjustable parameter in the estimation.<br />

Finally, we note that, according to this analysis;<br />

(i) single crystals are also expected to be prone to LME. In that case, a dihedral angle<br />

<strong>of</strong> 180° (hence γ gb = 0) must be assumed in the calculation <strong>of</strong> Λ c ;<br />

(ii) <strong>embrittlement</strong> resulting from an external contact with liquid metal can also be<br />

rationalized according to [42]: an estimation <strong>of</strong> σ c is given in Eq. 3-11 [42];<br />

(iii) this approach also predicts intuitively an increasing susceptibility to LME with<br />

increasing hardness <strong>of</strong> the alloy, since higher elastic strain energy can be stored;<br />

however, the present analysis cannot describe this effect, since the mechanical<br />

analysis is limited to linearly elastic materials.<br />

Implications<br />

If, on the basis <strong>of</strong> what precedes, we now consider Λ c as a parameter quantifying the<br />

specific resistance <strong>of</strong> a given liquid inclusion-containing alloy to LME, what are the<br />

practical implications ?<br />

Since the critical fracture stress is then given by Eq. 6-3, increasing the resistance <strong>of</strong><br />

a given system to LME is tantamount to increasing the right-h<strong>and</strong> side <strong>of</strong> Eq. 6-3.<br />

Increasing E is a rather difficult task; however, controling V, the inclusion size (in all<br />

rigor the largest inclusions’ size) can be achieved relatively easily, by deformation<br />

processing for example.<br />

A second possibility is to increase the equilibrium dihedral angle φ o , as this results in<br />

an increase <strong>of</strong> Λ c . Since:<br />

cos<br />

Eq. 3-2<br />

this can be achieved either by increasing the liquid/solid interfacial energy, or by<br />

φ ⎛ γ o ⎞ gb<br />

⎝ 2 ⎠ 2γSL<br />

=<br />

decreasing the highest grain boundary energy. An increase <strong>of</strong> γ SL can be obtained by


6.3 DISCUSSION<br />

changing the composition <strong>of</strong> the liquid metal; however, gross composition changes<br />

would be required since segregants to the solid/liquid interface do not increase γ SL .<br />

On the other h<strong>and</strong>, a decrease <strong>of</strong> γ gb (e.g. resulting from grain boundary segregation,<br />

<strong>of</strong>ten detrimental to GBE) should enhance resistance to LME. So should ”grain<br />

boundary engineering” if it can be used to decrease all grain boundary energies in the<br />

alloy. Finally, as mentionned above, increasing γ SL should not be envisaged unless<br />

the inclusion composition can be changed radically.<br />

The upper bound <strong>of</strong> the ductility trough.<br />

In LME, the onset <strong>of</strong> <strong>embrittlement</strong> occurs at the melting point <strong>of</strong> the liquid metal;<br />

this sets the lower bound <strong>of</strong> the ductility trough. On the other h<strong>and</strong>, a higher limit <strong>of</strong><br />

the ductility trough corresponding to the recovery <strong>of</strong> ductility may be explained by<br />

the fact that, at this temperature – T R – the resistance <strong>of</strong> the ”normally ductile metal”<br />

(i.e. R m <strong>of</strong> pure Cu) becomes lower than the critical stress σ c , Fig. 6-23. In this high-<br />

temperature region, the difference between the liquid inclusion-containing <strong>and</strong> the<br />

embrittler-free alloy should disappear: in short, signs <strong>of</strong> LME may disappear at<br />

elevated temperature.<br />

stress [MPa]<br />

250<br />

200<br />

150<br />

100<br />

50<br />

R m<br />

CuOFHC<br />

ε = 10-3 s-1 .<br />

24 h 500°C<br />

anneal<br />

15 min 500°C<br />

anneal<br />

σ c<br />

T E<br />

0<br />

0 100 200 300 400 500 600 700 800<br />

Temperature [°C]<br />

Figure 6-23: Evolution <strong>of</strong> the tensile strength <strong>of</strong> CuOFHC material as compared to a<br />

schematic curve <strong>of</strong> σ c . Both limits <strong>of</strong> the ductility trough T E <strong>and</strong> T R are added.<br />

Also, if the strain rate is sufficiently lowered, in the creep-deformation regime the<br />

flow stress <strong>and</strong> R m will decrease significantly. This could also cause the<br />

T R<br />

137


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

138<br />

disappearance <strong>of</strong> LME; this was observed in our <strong>leaded</strong> <strong>copper</strong> strained at 400 °C <strong>and</strong><br />

10 -4 s -1 .<br />

On the other h<strong>and</strong>, a larger inclusion size would result in a shift <strong>of</strong> the dashed line to<br />

lower stress values, extending the range <strong>of</strong> strain rates <strong>and</strong> temperatures at which<br />

LME is observed.<br />

c) stress concentration sites<br />

It is clear from the above subsections that <strong>leaded</strong> <strong>copper</strong> is subjected at intermediate<br />

temperature both to Grain Boundary, <strong>and</strong> Liquid Metal Embrittlement. The former<br />

mechanism dominates at lower strain rates, whereas the latter operates mainly at<br />

higher strain rate. In some instances, however, <strong>embrittlement</strong> may be aided<br />

mechanically by the presence <strong>of</strong> the solid micrometric intergranular lead inclusions.<br />

These act as stress concentration sites, <strong>and</strong> in turn favour the nucleation <strong>of</strong> cavities,<br />

as suggested by [47, 50]. The sharp ductility loss at 300 °C measured on the Cu1Pb<br />

material subjected to high strain rate tensile testing, Fig. 6-5, must thus be, at least in<br />

part, due to this purely mechanical influence <strong>of</strong> the finely dispersed lead inclusions.


6.4 Conclusion<br />

6.4 CONCLUSION<br />

Three different mechanisms have been proposed <strong>and</strong> opposed in the litterature to<br />

account for the occurence <strong>of</strong> the ductility trough in <strong>leaded</strong> <strong>copper</strong>. We show here that<br />

all three phenomena operate. Their respective importance depend mainly on the<br />

temperature <strong>and</strong> on the strain rate. This is illustrated on Fig. 6-5, (see arrows):<br />

(i) Below 327 °C <strong>and</strong> at high strain rate, cavitation at intergranular solid lead<br />

inclusions seems to trigger fracture, predominantly along the grain boundaries.<br />

(ii) Segregation-induced grain boundary <strong>embrittlement</strong> dominates at low strain rates<br />

in the whole temperature range <strong>of</strong> the ductility trough.<br />

(iii) Above 327 °C, liquid metal <strong>embrittlement</strong> is manifest at high strain rates. This<br />

mechanism requires that a critical stress be reached, which decreases as the<br />

inclusions volume increases <strong>and</strong> the dihedral angle decreases.<br />

This last observation, coupled with the observed dependence <strong>of</strong> the high strain rate<br />

fracture stress in these <strong>alloys</strong>, suggests that fracture by LME occurs once the<br />

adimensional parameter Λ, defined in Eq. 5-11, reaches a critical value Λ c , reflecting<br />

the influence <strong>of</strong> the stress applied, the elastic modulus <strong>of</strong> the solid, the interfacial<br />

energies, <strong>and</strong> the volume <strong>of</strong> the inclusions.<br />

139


CHAPTER 6. CU-EMBRITTLEMENT: THE ROLE OF PB<br />

140


Chapter 7<br />

7 Quench cracking in <strong>leaded</strong> <strong>copper</strong> <strong>alloys</strong>: a case study<br />

Ternary <strong>copper</strong>-nickel-tin <strong>alloys</strong> are <strong>of</strong>ten called ”spinodal <strong>alloys</strong>” since they exhibit<br />

spinodal decomposition at intermediate temperature. These <strong>alloys</strong> are high-strength<br />

<strong>copper</strong> <strong>alloys</strong> showing substantial age hardening [221]. A yield stress as high as<br />

1300 MPa with 46% reduction area at fracture can be reached for a Cu12Ni8Sn<br />

<strong>alloys</strong> after appropriate thermomechanical treatments [222]. This makes them<br />

attractive for applications in electronics requiring high spring strength, good<br />

formability, <strong>and</strong> the other advantages <strong>of</strong> a high-<strong>copper</strong> alloy, such as electrical<br />

conductivity <strong>and</strong> corrosion resistance. In addition, these <strong>alloys</strong> are demonstrated to<br />

have excellent wear properties.<br />

The major drawback <strong>of</strong> these <strong>alloys</strong> is that they exhibit substantial solute segregation<br />

during conventional solidification. Powder metallurgy allows the production <strong>of</strong> small<br />

parts, homogeneous in composition. Large fine-grained parts up to 65 cm in<br />

diameter, are also processed with the patented EquaCast process at Brush Wellman<br />

Inc. [223]. Osprey spray deposition is an alternative process that allows the<br />

production <strong>of</strong> large billets, typically 160 mm in diameter [224]. Rapid solidification<br />

<strong>of</strong> the sprayed liquid metal droplets ensures chemical homogeneity at the 100 µm-<br />

scale. Billets are subsequently hot-extruded. The extruded bars, typically 15 to<br />

25 mm in diameter, are immediately water quenched to avoid any hardening phase<br />

transformation.<br />

NP8, the <strong>leaded</strong> version <strong>of</strong> CN8, cracks systematically during the quench. Our goal<br />

was to identify the underlying <strong>embrittlement</strong> mechanisms, <strong>and</strong> devise strategies to<br />

allow the production <strong>of</strong> this alloy grade.<br />

141


CHAPTER 7. QUENCH CRACKING IN LEADED COPPER ALLOYS: A CASE STUDY<br />

142<br />

7.1 experimental procedures<br />

7.1.1 Materials<br />

The materials used in this case study were <strong>leaded</strong> <strong>and</strong> un<strong>leaded</strong> CuNiSn <strong>alloys</strong> NP8<br />

<strong>and</strong> CN8, already described under Section 6.1.1. Annealed monophase material was<br />

studied, since this corresponds to the state <strong>of</strong> the matter when quenched. The heat-<br />

treatment procedure is also described in Section 6.1.1.<br />

7.1.2 Thermal properties<br />

The thermal properties <strong>of</strong> CN8 <strong>and</strong> NP8 were considered to be the same: the thermal<br />

conductivity, λ, as well as the thermal expansion coeficient, α, were determined<br />

experimentally with CN8 material only.<br />

The thermal expansion coefficient was measured between room temperature <strong>and</strong><br />

850 °C under argon in a differential dilatometer (Bähr Thermoanalyse, model DIL<br />

802 V).<br />

The evolution <strong>of</strong> λ with temperature was estimated up to 900 °C from experimental<br />

values between RT <strong>and</strong> 100 °C by the ”stationary method” (constant heat flux<br />

through a reference brass material <strong>and</strong> a Ø 18 mm, 50 mm long CN8 cylinder) <strong>and</strong><br />

from Refs. [225, 226]. The specific heat, C p , as well as the density, ρ, were also<br />

estimated for the same temperature range from Refs. [225, 227].<br />

The temperature evolution <strong>of</strong> an instrumented quenched bar was measured in order to<br />

back-calculate the heat transfer coeficient, h(T) <strong>of</strong> varying quenching media (from<br />

agitated water to ambient air). A CN8 Ø 18 mm bar was instrumented with three<br />

Ø 1 mm Thermocoax ® thermocouples. One was located at the center <strong>of</strong> the bar, <strong>and</strong><br />

the tips ot the two others were located just below the bar surface, Fig. 7-1.


18<br />

7.1 EXPERIMENTAL PROCEDURES<br />

Figure 7-1: Section <strong>of</strong> the instrumented CN8 bar prolonged by a 2 m long stainless<br />

steel tube. Thermocouple extension wires contained within the tube are not shown.<br />

The CN8 bar attached to a 2 m long stainless steel tube was heated whithin a<br />

resistance furnace to a homogeneous temperature <strong>of</strong> 900 °C, <strong>and</strong> subsequently<br />

immersed in the quenching medium. The temperature acquisition rate was 10 Hz.<br />

Such measurements were conducted in the cooling unit <strong>of</strong> the extrusion press on the<br />

industrial site, as well as in agitated water, oil, fluidized bath or ambient air at the<br />

laboratory.<br />

CN8 bar<br />

From material parameters <strong>and</strong> with the adequate thermal boundary conditions, the<br />

exact temperature evolution, T(t,r) at any point whithin a quenched cylindrical bar<br />

can be determined by FEM. Since it is assumed that only radial temperature gradients<br />

occur, the thermal problem at h<strong>and</strong> is one-dimensional, Fig. 7-2.<br />

The boundary conditions are:<br />

Q( 0, l ) = 0<br />

( )<br />

150<br />

TC tips<br />

Eq.7-1<br />

QR ( , l ) = h⋅2πRL⋅ Tinit −T∞<br />

Eq.7-2<br />

with Q, the heat flux, R <strong>and</strong> L, the bar radius <strong>and</strong> length, Tinit , the temperature <strong>of</strong> the<br />

bar just before quenching, <strong>and</strong> T ∞ the temperature <strong>of</strong> the quenching medium.<br />

60<br />

1<br />

Steel tube<br />

143


CHAPTER 7. QUENCH CRACKING IN LEADED COPPER ALLOYS: A CASE STUDY<br />

144<br />

Figure 7-2: Parameters involved for the resolution <strong>of</strong> the thermal problem <strong>of</strong> a<br />

quenched round bar by FEM. Axial symetry is assumed.<br />

7.1.3 Mechanical analysis<br />

l<br />

r<br />

R = D/2<br />

C p, ρ, λ, T init<br />

h, T<br />

Mechanical properties <strong>of</strong> both CN8 <strong>and</strong> NP8 materials were measured between RT<br />

<strong>and</strong> 800 °C, at 100 °C intervals. Elevated temperature tensile tests were performed<br />

with the aid <strong>of</strong> the apparatus described in Section 6.1.2. A strain rate <strong>of</strong> 10 -2 s -1 was<br />

chosen, since it corresponds roughly to the estimated strain rate during the quench.<br />

The deformation <strong>of</strong> the sample was measured in all tests using a high-temperature<br />

extensometer, in order to gather precise strain data. The dependence <strong>of</strong> the Poissons’<br />

ratio, ν, on temperature was estimated up to 900 °C from the literature [228].<br />

The elastoplastic mechanical properties <strong>of</strong> CN8 <strong>and</strong> NP8 materials were introduced<br />

in a second FEM code to calculate the thermal displacement field from the T(t,r) <strong>and</strong><br />

α(T) values, Fig. 7-3. From this <strong>and</strong> the conditions for mechanical equilibrium, with<br />

the (known) temperature-dependent alloy flow stress, the local constrained σ−ε<br />

conditions were calculated. The temperature distribution <strong>and</strong> evolution is given by<br />

the above-mentioned thermal problem, whereas the boundary conditions are:<br />

ur( 0,l)= 0<br />

ul ( r,0)=<br />

0<br />

ur ( R,l)= const1<br />

ul ( r, L)= const2<br />

Eq.7-3<br />

Eq.7-4<br />

Eq.7-5<br />

Eq.7-6


7.1 EXPERIMENTAL PROCEDURES<br />

Figure 7-3: Parameters involved for the resolution <strong>of</strong> the mechanical problem <strong>of</strong> a<br />

quenched round bar by FEM. Axial symetry is assumed.<br />

Finally, the transient thermal stresses are determined from the finite element<br />

elastoplastic simulation knowing the thermal strain distribution, Fig. 7-4.<br />

Figure 7-4: Schematic representation <strong>of</strong> a quenched round infinite bar. The<br />

temperature <strong>and</strong> stress pr<strong>of</strong>iles are schematically added.<br />

FEM calculations presented in what follows were conducted by Dr. Andreas Rossoll<br />

at EPFL.<br />

l<br />

r<br />

T(t, r), α, E, ν, σ − ε<br />

R = D/2<br />

145


CHAPTER 7. QUENCH CRACKING IN LEADED COPPER ALLOYS: A CASE STUDY<br />

146<br />

7.2 Results<br />

7.2.1 Mechanical properties <strong>of</strong> CN8 & NP8 between RT <strong>and</strong> 800 °C<br />

The evolution <strong>of</strong> the reduction in area at fracture, as well as the 400 °C tensile<br />

properties <strong>of</strong> both CN8 <strong>and</strong> NP8 materials are shown on Fig. 6-9 <strong>and</strong> Fig. 6-10. At<br />

400 °C, the ductility is significantly reduced by the addition <strong>of</strong> 1 wt.% lead. Fig. 7-5<br />

shows the resulting fracture surface <strong>of</strong> CN8 <strong>and</strong> NP8 materials. Pb-containing NP8<br />

fractures intergranularly, whereas the CN8 fracture is mixed transgranular <strong>and</strong><br />

intergranular. On the intergranular fracture surface <strong>of</strong> the NP8 material, traces <strong>of</strong><br />

intergranular lead inclusions are apparent. These hemispherical cups are empty;<br />

suggesting that liquid lead migration occurred.<br />

(a) CN8 dε/dt=10-2 s-1 RA=18% (b) NP8 dε/dt=10-2 s-1 RA=6%<br />

Figure 7-5: Fracture surface after 400 °C tensile testing in air at dε/dt=10-2 s-1 for (a)<br />

CN8, <strong>and</strong> (b) NP8 material in the monophased state (solutionized <strong>and</strong> quenched).<br />

The <strong>embrittlement</strong> is much more marked at 800 °C: NP8 breaks before the onset <strong>of</strong><br />

global plasticity, whereas CN8 displays dynamic recrystallization, Fig. 7-6.


eng. stress [MPa]<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

7.2 RESULTS<br />

Figure 7-6: 800 °C flow curves <strong>of</strong> CN8 & NP8 material at a strain rate <strong>of</strong> 10 -2 s -1 . For<br />

each material, fracture is emphasized by a solid symbol.<br />

7.2.2 Heat transfer coefficient <strong>of</strong> different quenching media<br />

The measured temperature evolution at the center <strong>of</strong> a Ø 18 mm CN8 bar is shown on<br />

Fig. 7-7. ”Noses” <strong>of</strong> the CN8 TTT-diagram are reproduced from [214]. Strictly<br />

speaking, since the TTT-diagram is developped using small samples isothermally<br />

transformed, the ”nose” <strong>of</strong> the curves should be moved somewhat to the right for<br />

continuous cooling applications [181]. Hence can be concluded that a Ø 18 mm CN8<br />

bar does not experience any phase transformation while quenched into water. It is<br />

also not surprising that air-cooled material is not single-phased.<br />

Notice that the cooling curves measured with our instrumented bar were very similar<br />

for a water quench in the bath at our laboratory or in the industrial cooling unit <strong>of</strong> the<br />

extrusion press.<br />

800 °C<br />

CN8<br />

NP8<br />

0 1 2 3 4 5 6<br />

eng. strain [%]<br />

CN8 & NP8<br />

annealed<br />

60 min at 850°C<br />

ε = 10 -2 s -1<br />

& quenched<br />

.<br />

147


CHAPTER 7. QUENCH CRACKING IN LEADED COPPER ALLOYS: A CASE STUDY<br />

148<br />

Temperature [°C]<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

water<br />

oil<br />

fluidized bath<br />

Figure 7-7: Cooling curves <strong>of</strong> a CN8 Ø 18 mm bar immersed into different quenching<br />

media. Temperature was measured at the center <strong>of</strong> the bar. The TTT-diagram (not CCT)<br />

<strong>of</strong> the alloy is added.<br />

Transient thermal stresses arise due to temperature gradients whithin a material<br />

characterized by a non-zero expansion coefficient. Newtonian cooling is defined as<br />

the case were temperature gradients are negligible. It is characterized by Bi ≤ 0.1,<br />

where the Biot number is:<br />

spinodal<br />

0<br />

0,1 1 10 100 1000 10 4<br />

air<br />

time [s]<br />

hR<br />

Bi =<br />

λ<br />

Eq.7-7<br />

Charts displaying the temperature response <strong>of</strong> an infinite cylinder initially at a<br />

uniform temperature, T init , <strong>and</strong> then subjected to a convective environnement at T ∞<br />

are available in e.g. [181]. Further adimensional parameters, the Fourier number, Fo ,<br />

<strong>and</strong> the equivalent temperature, T equ , are used:<br />

λ t<br />

Fo = 2<br />

R Cpρ Eq.7-8<br />

T − T∞<br />

Tequ<br />

=<br />

Tinit − T∞<br />

Eq.7-9<br />

Relevant curves from the reference charts [181] are compared to present FEM results<br />

in Fig. 7-8. Adimensionalization <strong>of</strong> our results was done by considering room-<br />

DO 22<br />

Grain Boundary +<br />

intragranular γ<br />

Discontinuous γ<br />

L1 2 + DO 22


7.2 RESULTS<br />

temperature values for the parameters λ, C p , <strong>and</strong> ρ. Values taken were 28 W/m K,<br />

380 J/kg K, <strong>and</strong> 8.94 10 3 kg/m 3 respectively [225].<br />

As seen, for our Ø 18 mm CN8 bar, air cooling can safely be considered as<br />

Newtonian. Therefore negligible transient thermal stresses are expected. Water<br />

quenching induces severe temperature gradients, while the fluidized bath, on the<br />

other h<strong>and</strong>, seems to be a suitable cooling medium (thermal stresses are negligible)<br />

since Bi is about 0.1.<br />

T equ centre [-]<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

water<br />

oil<br />

fluidized bath<br />

air<br />

Bi = 20<br />

0.1 1<br />

Fo [-]<br />

10<br />

Figure 7-8: Adimensionalized cooling curves from Fig. 7-7 in comparison with<br />

cooling curves from the analytical approach for variing Biot numbers from [181].<br />

The heat transfer coeficient, h(T) <strong>of</strong> various quenching media was back-calculated by<br />

FEM from experimental cooling data, Fig. 7-9. h describes the efficiency <strong>of</strong> the<br />

quench. In the case <strong>of</strong> a water-quench, this parameter varies by more than an order <strong>of</strong><br />

magnitude in the temperature range studied. This is mainly due to phase<br />

transformations in the quench-medium: boiling for water, cracking for oil. Due to<br />

differences in chemical composition <strong>and</strong> molecular weight, each oil is characterized<br />

by a different cooling capability. The latter can also be tailored by small additions <strong>of</strong><br />

water. Concerning the fluidized bath, no phase transformation occurs. We<br />

accordingly find a smooth variation <strong>of</strong> h. Estimation <strong>of</strong> the heat transfer coeficient<br />

according to the analytical equations describing cooling at constant h [181], Fig. 7-8<br />

lead to values for h ranging between 6 <strong>and</strong> 60 kW/m 2 K (Bi=2 to 20) for water<br />

quenching. These values agree well with those determined by FEM, Fig. 7-9.<br />

2<br />

4<br />

0.4<br />

0.8<br />

0.2<br />

Bi = 0.1<br />

149


CHAPTER 7. QUENCH CRACKING IN LEADED COPPER ALLOYS: A CASE STUDY<br />

150<br />

[kWm -2 K -1 ]<br />

h<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

0.001<br />

water<br />

0 100 200 300 400 500 600 700 800 900<br />

Temperature [°C]<br />

Figure 7-9: Heat transfer coeficient for different quenching media calculated by FEM<br />

[Andreas Rossoll, EPFL].<br />

FEM simulation-derived transient thermal stresses are plotted as a function <strong>of</strong><br />

temperature, <strong>and</strong> compared to the ultimate tensile strengh <strong>of</strong> both CN8 <strong>and</strong> NP8<br />

materials, Fig. 7-10. Transient stresses plotted are the principal stresses acting at the<br />

surface <strong>of</strong> a Ø 18 mm bar quenched into water or fluidized bath. It is interesting to<br />

note the presence <strong>of</strong> compressive residual stresses at the surface <strong>of</strong> the quenched bar.<br />

These are due to the occurence <strong>of</strong> plastic deformation during the quench.<br />

Fig. 7-10 shows that the mechanical resistance <strong>of</strong> the CN8 material always exceeds<br />

the transient thermal stresses in the whole temperature range, for both quench<br />

conditions. On the other h<strong>and</strong>, this is not the case for the lead-containing NP8<br />

material. In conclusion, NP8 Ø 18 mm bars should crack during water-quenching,<br />

but not during a quench in the fluidized bath.<br />

oil<br />

fluidized bath<br />

air


σ surf or R m [MPa]<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

-100<br />

-200<br />

-300<br />

σsurf water quench<br />

σsurf cooling in fluidized bath<br />

0 100 200 300 400 500 600 700 800 900<br />

Temperature [°C]<br />

CN8<br />

NP8<br />

7.2 RESULTS<br />

Figure 7-10: FEM calculated surface transient thermal stresses for a Ø 18 mm bar<br />

water-quenched or cooled in fluidized bath are compared to CN8 <strong>and</strong> NP8 tensile<br />

strength (each symbol is the result <strong>of</strong> a single tensile test).<br />

R m<br />

R m<br />

151


CHAPTER 7. QUENCH CRACKING IN LEADED COPPER ALLOYS: A CASE STUDY<br />

152<br />

7.3 Implications<br />

It is clear from Fig. 7-10 that water-quenched NP8 bars, 18 mm in diameter, crack<br />

because <strong>of</strong> transient thermal stresses. This is a manifestation <strong>of</strong> LME. It is recalled<br />

that, in production, this material is rapidly cooled from the extrusion temperature to<br />

avoid the occurence <strong>of</strong> spinodal decomposition; since the material in the monophased<br />

state is most suitable for subsequent cold working. We show on Fig. 7-7 that water is<br />

the only quenching medium that ensures sufficiently rapid cooling for Ø 18 mm bars.<br />

Therefore a trade-<strong>of</strong>f must be found between rapid cooling to avoid phase<br />

transformation, <strong>and</strong> moderate cooling to ensure low thermal stresses.<br />

We propose three main strategies to overcome this problem: (i) lower the diameter <strong>of</strong><br />

the extruded bar, (ii) change the cooling fluid, or (iii) change the alloy composition.<br />

7.3.1 Bar diameter<br />

Cooling curves can be extrapolated for any bar dimension, provided that the bar<br />

geometry <strong>and</strong> the cooling medium are kept constant. Namely the T equ -Fo curve is<br />

intrinsic for any cooling medium, Fig. 7-8. We show thus on Fig. 7-11 that 6 mm is<br />

the critical radius for a bar to experience no phase transformation during quenching<br />

in a fluidized bath.


Temperature [°C]<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

7.3 IMPLICATIONS<br />

Figure 7-11: Calculated cooling curves for a CN8 bar <strong>of</strong> different diameters immersed<br />

in fluidized bath. TTT-diagram is added [214].<br />

7.3.2 Cooling medium<br />

ø 8<br />

ø 6<br />

ø 4<br />

ø 18<br />

ø 12<br />

0<br />

0,1 1 10 100 1000 10 4<br />

time [s]<br />

CuNiSn <strong>alloys</strong> are intended to compete with the Cu-Be alloy family. The very high<br />

strength <strong>of</strong> both <strong>of</strong> these alloy classes is obtained by strain hardening <strong>and</strong> subsequent<br />

precipitation hardening. In Cu15Ni8Sn <strong>alloys</strong>, nucleation <strong>of</strong> discontinuous γ-<br />

precipitates corresponds to a loss in hardening [224]. The matrix is in any case<br />

depleted in the alloying elements that form ordered precipitates (spinodal, DO 22 or<br />

L1 2 ). Therefore the ”nose” corresponding to the appearance <strong>of</strong> the discontinuous γ-<br />

precipitates must definitely be avoided during alloy processing. Fig. 7-7 suggests that<br />

air-cooling is not sufficiently rapid to avoid discontinuous precipitation <strong>of</strong> the γ-<br />

phase; however, this is not the case for a quench in a fluidized bath.<br />

The industrial production <strong>of</strong> NP8 could therefore become possible with an<br />

appropriate cooling fluid, the severity <strong>of</strong> which is lower than that <strong>of</strong> an oil but higher<br />

than that <strong>of</strong> agitated air. It is recognized in [229] that fluidized-beds are used as<br />

cooling media only to a limited extend. Alternate media according to [229] are slow<br />

DO 22<br />

spinodal<br />

Grain Boundary +<br />

intragranular γ<br />

Discontinuous γ<br />

L1 2 + DO 22<br />

oils, aqueous polymer solutions, forced gases, fogs, sprays or dry dies.<br />

153


CHAPTER 7. QUENCH CRACKING IN LEADED COPPER ALLOYS: A CASE STUDY<br />

154<br />

7.3.3 Alloy composition<br />

Cu15Ni8Sn <strong>alloys</strong> are characterized by very rapid kinetics <strong>of</strong> phase transformation,<br />

since spinodal decomposition starts after only 4 s at 400 °C, Fig. 7-12 (a) [214]. It is<br />

also known from the literature that Cu-7.5Ni-5Sn <strong>alloys</strong> are characterized by much<br />

slower kinetics, Fig. 7-12 (b). Notice that earlier resistivity measurements indicated<br />

the occurence <strong>of</strong> a phase transformation after a 100 s isothermal stay at 400 °C for a<br />

Cu-9Ni-6Sn alloy [230]. Reducing the alloying element contents renders the water<br />

quenching-step unnecessary; however, this lowers the peak hardness obtainable<br />

[223].<br />

(a) (b)<br />

Figure 7-12: TTT-diagrams (a) <strong>of</strong> the Cu15Ni8Sn alloy [214], <strong>and</strong> (b) <strong>of</strong> the<br />

Cu7,5Ni5Sn alloy [231] obtained by TEM characterizations <strong>and</strong> electrical resistivity<br />

measurements.


7.4 Conclusion<br />

7.4 CONCLUSION<br />

We have shown that water-quenched NP8 bars, 18 mm in diameter, crack because <strong>of</strong><br />

transient thermal stresses during the quench. Failure is under these circumstances can<br />

be attributed to liquid metal <strong>embrittlement</strong>.<br />

We propose three main strategies to overcome this problem:<br />

(i) lower the diameter <strong>of</strong> the extruded bar,<br />

(ii) change the cooling fluid, or<br />

(iii) change the alloy composition.<br />

155


CHAPTER 7. QUENCH CRACKING IN LEADED COPPER ALLOYS: A CASE STUDY<br />

156


Chapter 8<br />

8 General conclusions<br />

• A new method is presented for the determination <strong>of</strong> the dihedral angle <strong>of</strong> individual<br />

intergranular formerly liquid inclusions. The angle is derived from a mathematical<br />

fitting <strong>of</strong> the solid/liquid interface 3-D geometry around individual inclusions, <strong>and</strong><br />

therefore reflects the dihedral angle dictated by global energy minimization <strong>of</strong> the<br />

inclusion shape under capillary forces.<br />

• In a high-purity Cu-1 wt.% Pb alloy, <strong>and</strong> for a specific temperature, we show that φ<br />

is not unique; this reflects the fact that high-angle grain boundary energies vary in the<br />

metal.<br />

• A decrease <strong>of</strong> φ was measured on Cu-1Pb-0.04P samples subjected to interrupted<br />

uniaxial tensile creep tests at 400 °C, as compared to φ measured on unstressed<br />

specimens. φ values are best described by predictions accounting for the shape<br />

evolution <strong>of</strong> a zero bulk modulus inclusion within a uniaxially stressed solid,<br />

suggesting that voids nucleate within the liquid inclusions or along the matrix/<br />

inclusion interface.<br />

• We demonstrate that LME as well as GBE operate in <strong>leaded</strong>-<strong>copper</strong> at intermediate<br />

temperature. Their respective importance depends mainly on the strain rate.<br />

• According to analysis accounting for inclusion shape stability, we calculate a<br />

critical parameter Λ c , above which inclusions are predicted to collapse to a crack.<br />

157


CHAPTER 8. GENERAL CONCLUSIONS<br />

158<br />

2 1/ 3<br />

σ ⋅ V<br />

Λ =<br />

γ SL ⋅ E<br />

<strong>and</strong><br />

Λc = max( Λ) ⇒ σ c =<br />

γ SL ⋅ E<br />

Λc<br />

1/ 3<br />

V<br />

Eq.8-1<br />

The calculated respective critical stress σc is found to be in good agreement with the<br />

respective fracture stress <strong>of</strong> both <strong>leaded</strong> pure <strong>copper</strong> <strong>and</strong> <strong>leaded</strong> <strong>copper</strong>-nickel-tin<br />

<strong>alloys</strong> tested here, found to fail by LME at high strain-rate.


Chapter 9<br />

9 Future work<br />

In a continuation <strong>of</strong> this project, it would be <strong>of</strong> interest to:<br />

(i) Couple the present individual grain boundary inclusion dihedral angle<br />

measurement method with EBSD methods for characterization <strong>of</strong> grain orientation in<br />

the scanning electron microscope. This would in particular extend the present method<br />

for dihedral angle measurement to provide a new indirect SEM-based method for the<br />

measurement <strong>of</strong> grain boundary energies.<br />

(ii) Further heat treatments under reducing atmosphere should be performed in order<br />

to cover the whole temperature range between 350 <strong>and</strong> 1000 °C. Dihedral angle<br />

measurements should be performed in parallel with EBSD analysis in order to check<br />

whether the φ variations observed at a specific temperature reflects indeed the γ gb<br />

anisotropy.<br />

(iii) Make more measurements <strong>of</strong> φ from fractographs <strong>of</strong> the quenched NP8, so as to<br />

determine a representative distribution <strong>of</strong> inclusion sizes.<br />

Applying also that technique to tensile tested coarse-grained C99 <strong>and</strong> Cu1Pb <strong>and</strong><br />

comparing those measurements with our st<strong>and</strong>ard procedure (i.e. after metallographic<br />

preparation <strong>and</strong> selective dissolution) would also be useful to ascertain the validity <strong>of</strong><br />

using fracture surfaces.<br />

159


CHAPTER 9. FUTURE WORK<br />

160<br />

(iv) From the known evolution with temperature <strong>of</strong> both the elastic properties <strong>and</strong><br />

interfacial energies, the critical parameter Λ c (T) could be calculated as a function <strong>of</strong><br />

temperature. For a given inclusion size, the critical stress associated with LME<br />

fracture σ c (T) could then be estimated. A new line could be accordingly drawn on the<br />

Ashby fracture-mechanisms map, Fig. 3-18.<br />

(v) In order to further verify the influence <strong>of</strong> the liquid inclusion size on σ c , tensile<br />

tests should be performed at 400 °C, varying both V <strong>and</strong> , all else being kept<br />

ε •<br />

constant. With variations between 10 -4 to 10 s -1 , the material would therefore be<br />

subjected to maximum stresses ranging from about 70 to 150 MPa for <strong>leaded</strong> pure<br />

<strong>copper</strong>, exhibiting LME failure or not, depending on the (largest) inclusion size.<br />

Hardening <strong>of</strong> the <strong>copper</strong> matrix (e.g. by oxide dispersion strenghtening) would also<br />

be opportune, since a linear elastic solid is considered in the Eshelby analysis.<br />

(vi) Since an increase <strong>of</strong> φ is predicted to enhance the resistance to LME, small<br />

additions should be made to <strong>leaded</strong> <strong>alloys</strong> in order to decrease γ gb or increase γ SL ,<br />

resulting respectively from beneficial segregation or liquid metal alloying.<br />

(vii) Dihedral angle measurements should also be performed on samples subjected to<br />

elevated temperature hydrostatic compression. This mainly for two reasons: (a) it<br />

could be checked whether the dihedral angle is correctly predicted making the<br />

assumption that the inclusion has a finite positive compressibility (no voiding); <strong>and</strong><br />

(b) the <strong>microstructure</strong> is expected to be at equilibrium, since neither global plasticity,<br />

nor grain boundary motion, should occur to a significant degree.<br />

(viii) Finally, the mechanical analysis should be extended, to address the influence <strong>of</strong><br />

plastic <strong>and</strong> viscous deformation <strong>of</strong> the surrounding metal, coupled perhaps with grain<br />

boundary sliding, as all are indeed observed in the experimental data. This is a<br />

formidable task, requiring finite-element simulation to be conducted in realistic<br />

fashion. The link <strong>of</strong> the present analysis with fracture mechanics, which is relatively<br />

obvious given the nature <strong>of</strong> the critical parameter Λ c , may perhaps provide pathways<br />

easing this exploration.<br />

ε •


Chapter 10<br />

10 Appendices<br />

10.1 Least square method <strong>and</strong> dihedral angle measurement<br />

The MeX program allows to get the coordinates <strong>of</strong> points belonging to the former<br />

solid/liquid interface. Assuming that these points belong to a spherical cap, their<br />

coordinates (x i ;y i ;z i ) obey:<br />

( ) + ( − ) + ( − ) − =<br />

f( a, b, c, R)≡ x −a<br />

y b z c R<br />

2 2 2 2<br />

i i i i<br />

where (a;b;c) are the coordinates <strong>of</strong> the center <strong>of</strong> the sphere <strong>of</strong> radius R.<br />

With more than four measuring points:<br />

2<br />

2<br />

2 2<br />

( x1−a) + ( y1−b) + ( z1−c) − R = f1( a, b, c, R)+ r1<br />

2<br />

2<br />

2 2<br />

( x2−a) + ( y2−b) + ( z2−c) − R = f2( a, b, c, R)+ r2<br />

( ) + ( − ) + ( − ) − =<br />

2 2 2 2<br />

xn − a yn b zn c R fn( a, b, c,R)+<br />

rn the system is over-determined, with no solution for a, b, c <strong>and</strong> R in general.<br />

Eq.10-1<br />

Eq.10-2<br />

a, b, c <strong>and</strong> R are therefore calculated as the values that give the lowest mean-square<br />

deviation <strong>of</strong> f from zero over all points <strong>of</strong> measurement (x i ;y i ;z i ). Since function f is<br />

<strong>of</strong> second order, it is approximated by its Taylor series about an approximated<br />

solution a o , b o , c o <strong>and</strong> R o calculated analytically from four measured points:<br />

0<br />

161


CHAPTER 10. APPENDICES<br />

162<br />

with<br />

this yields a system <strong>of</strong> linear equations:<br />

with<br />

4<br />

∂f<br />

fi( a, b, c, R)≈ fi( a0, b0, c0, R0)+<br />

∑ ( a0, b0, c0, R0)⋅ xk+ ri<br />

∂ x ∆<br />

x<br />

k = 1<br />

⎛ a⎞<br />

⎛ a0<br />

⎞ ⎛ ∆a⎞<br />

⎜ b⎟<br />

0<br />

= ⎜ ⎟ = x + ∆x<br />

⎜ c⎟<br />

⎜ ⎟<br />

⎝ R⎠<br />

⎜ b ⎟ ⎜ b⎟<br />

0 ∆<br />

= ⎜ ⎟ + ⎜ ⎟<br />

⎜ c0<br />

⎟ ⎜ ∆c<br />

⎟<br />

⎜ ⎟ ⎜ ⎟<br />

⎝ R ⎠ ⎝∆<br />

R⎠<br />

k k k<br />

y = A ∆ x + r<br />

[ n× 1] [ n× m] [ m× 1] [ n×<br />

1]<br />

⎡ ∂f<br />

⎢ ∂a<br />

⎢∂f<br />

A=<br />

⎢<br />

⎢ ∂a<br />

⎢<br />

⎢∂f<br />

⎣⎢<br />

∂a<br />

⎛ ∆a⎞<br />

⎜ ∆b⎟<br />

∆ x = ⎜ ⎟<br />

⎜ ∆c<br />

⎟<br />

⎜ ⎟<br />

⎝∆<br />

R⎠<br />

∂f<br />

∂b<br />

∂f<br />

∂b<br />

∂f<br />

∂b<br />

∂f<br />

∂c<br />

∂f<br />

∂c<br />

∂f<br />

∂c<br />

0<br />

∂f<br />

⎤<br />

∂R<br />

⎥<br />

∂f<br />

⎥<br />

⎥<br />

∂R<br />

⎥<br />

⎥<br />

∂f<br />

⎥<br />

∂R<br />

⎦<br />

1 1 1 1<br />

2 2 2 2<br />

n n n n<br />

⎥<br />

( a , b , c , R )<br />

0 0 0 0<br />

( )− ( )<br />

( )− ( )<br />

⎡ f1 a, b, c, R f1 a0, b0, c0, R0<br />

⎤<br />

⎢<br />

⎥<br />

y = ⎢<br />

f2 a, b, c, R f2 a0, b0, c0, R0<br />

⎥<br />

⎢<br />

⎥<br />

⎢<br />

⎥<br />

⎣⎢<br />

fn( a, b, c, R)− fn( a0, b0, c0, R0)<br />

⎦⎥<br />

⎡r1<br />

⎤<br />

⎢r<br />

⎥<br />

2<br />

r = ⎢ ⎥<br />

⎢ ⎥<br />

⎢ ⎥<br />

⎣rn<br />

⎦<br />

k<br />

Eq.10-3<br />

Eq.10-4<br />

Eq.10-5<br />

Eq.10-6<br />

Eq.10-7<br />

Eq.10-8<br />

Eq.10-9


Its solution is ∆ ˆx :<br />

∆ =( ) − T 1 T<br />

ˆx A A A y<br />

10.1 LEAST SQUARE METHOD AND DIHEDRAL ANGLE MEASUREMENT<br />

Eq.10-10<br />

The coordinates <strong>of</strong> the center <strong>of</strong> the sphere <strong>and</strong> its radius, respectively (a;b;c;R) are<br />

finally obtained by additing ∆ ˆx<br />

to the approximate solution (ao ;bo ;co ;Ro ).<br />

The Mathcad spreadsheet developed to this end is given below for the specific case<br />

<strong>of</strong> a grain boundary inclusion after shape equilibration at 930 °C. Measured φ was<br />

78°, Fig. 4-7.<br />

163


CHAPTER 10. APPENDICES<br />

164<br />

Determination <strong>of</strong> the equation <strong>of</strong> spheres knowing >4 points from its<br />

surface with the least square method<br />

1) Equation <strong>of</strong> the spheres from an approached solution<br />

sphere A:<br />

A1,B1,C1,D1: 4 points from the sphere surface with center S1 <strong>and</strong> radius r1<br />

A1 :=<br />

Given<br />

�<br />

�<br />

�<br />

�<br />

vec1 := Find S1 , S1 , S1 , r1<br />

0 1 2<br />

vec1 =<br />

−0.67<br />

8.68<br />

6.16<br />

Approximations:<br />

( A1 − S1<br />

0 0)<br />

2<br />

( A1 − S1<br />

1 1)<br />

2<br />

+ ( A1 − S1<br />

2 2)<br />

2<br />

+ r1 2<br />

− = 0<br />

( B1 − S1<br />

0 0)<br />

2<br />

( B1 − S1<br />

1 1)<br />

2<br />

+ ( B1 − S1<br />

2 2)<br />

2<br />

+ r1 2<br />

− = 0<br />

( C1 − S1<br />

0 0)<br />

2<br />

( C1 − S1<br />

1 1)<br />

2<br />

+ ( C1 − S1<br />

2 2)<br />

2<br />

+ r1 2<br />

− = 0<br />

( D1 − S1<br />

0 0)<br />

2<br />

( D1 − S1<br />

1 1)<br />

2<br />

+ ( D1 − S1<br />

2 2)<br />

2<br />

+ r1 2<br />

− = 0<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

10.715<br />

−0.13<br />

9.83<br />

14.856<br />

( )<br />

�<br />

�<br />

�<br />

�<br />

B1 :=<br />

�<br />

�<br />

�<br />

�<br />

−0.51<br />

−8.96<br />

5.74<br />

�<br />

�<br />

S1 :=<br />

�<br />

�<br />

�<br />

�<br />

C1 :=<br />

−5<br />

−5<br />

5<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

Center & radius <strong>of</strong> the sphere A<br />

centreA :=<br />

−0.26<br />

1.97<br />

0.04<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

r1 := 10<br />

vec1 0<br />

vec1 1<br />

vec1 2<br />

�<br />

�<br />

�<br />

�<br />

D1 :=<br />

�<br />

�<br />

�<br />

�<br />

−3.23<br />

−1.59<br />

4.92<br />

�<br />

�<br />

rayonA := vec1<br />

3


centreA =<br />

Sphere A<br />

�<br />

�<br />

�<br />

�<br />

10.715<br />

−0.13<br />

9.83<br />

XA0 := centreA<br />

0<br />

YA0 := centreA<br />

1<br />

ZA0 := centreA<br />

2<br />

RA0 := rayonA<br />

�<br />

�<br />

10.1 LEAST SQUARE METHOD AND DIHEDRAL ANGLE MEASUREMENT<br />

sphere B:<br />

A2,B2,C2,D2: 4 points from the sphere surface with center S2 <strong>and</strong> radius r2<br />

�<br />

0.12<br />

0.10<br />

0.22<br />

A2 := � 8.62 B2 := � −7.80<br />

C2 := � 1.35 D2 :=<br />

�<br />

�<br />

6.59<br />

6.10<br />

0.11<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

Approximations:<br />

Given<br />

S2 := 5<br />

�<br />

� 5�<br />

r2 := 10<br />

( A2 − S2<br />

0 0)<br />

2<br />

( A2 − S2<br />

1 1)<br />

2<br />

+ ( A2 − S2<br />

2 2)<br />

2<br />

+ r2 2<br />

− = 0<br />

( B2 − S2<br />

0 0)<br />

2<br />

( B2 − S2<br />

1 1)<br />

2<br />

+ ( B2 − S2<br />

2 2)<br />

2<br />

+ r2 2<br />

− = 0<br />

( C2 − S2<br />

0 0)<br />

2<br />

( C2 − S2<br />

1 1)<br />

2<br />

+ ( C2 − S2<br />

2 2)<br />

2<br />

+ r2 2<br />

− = 0<br />

( D2 − S2<br />

0 0)<br />

2<br />

( D2 − S2<br />

1 1)<br />

2<br />

+ ( D2 − S2<br />

2 2)<br />

2<br />

+ r2 2<br />

− = 0<br />

vec2 := Find S2 , S2 , S2 , r2<br />

0 1 2<br />

vec2 =<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

−13.049<br />

0.367<br />

8.327<br />

15.639<br />

( )<br />

�<br />

�<br />

�<br />

�<br />

rayonA = 14.856<br />

�<br />

�<br />

�<br />

�<br />

5<br />

�<br />

centreB :=<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

Sphere B<br />

centreB =<br />

vec2 0<br />

vec2 1<br />

vec2 2<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

−13.049<br />

0.367<br />

8.327<br />

XB0 := centreB<br />

0<br />

YB0 := centreB<br />

1<br />

ZB0 := centreB<br />

2<br />

RB0 := rayonB<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

2.38<br />

−1.48<br />

6.57<br />

Center & radius <strong>of</strong> the sphere B<br />

rayonB := vec2<br />

3<br />

�<br />

�<br />

rayonB = 15.639<br />

165


CHAPTER 10. APPENDICES<br />

166<br />

2) Import the files containing m>4 points x i , y i et z i<br />

XYZA :=<br />

mA := rows( XYZA)<br />

nA := mA − 1<br />

nA 6.462 10 3<br />

= ×<br />

AA<br />

iA, 0<br />

AA<br />

iA, 1<br />

AA<br />

iA, 2<br />

AA<br />

iA, 3<br />

iA := 0.. nA<br />

XA :=<br />

iA<br />

YA :=<br />

iA<br />

ZA :=<br />

iA<br />

3) Determine the matrices A (mx4)<br />

:= 2⋅ XA0 − XA<br />

iA<br />

:= 2⋅ YA0 − YA<br />

iA<br />

:= 2⋅ ZA0 − ZA<br />

iA<br />

:=<br />

\\Imxlmmpc21\..\pr<strong>of</strong>iles_exp15_sphere A.xls<br />

Number <strong>of</strong> measuring points: n<br />

XYZA<br />

iA, 0<br />

XYZA<br />

iA, 1<br />

XYZA<br />

iA, 2<br />

( )<br />

( )<br />

( )<br />

−2RA0 XYZB :=<br />

mB := rows( XYZB)<br />

nB := mB − 1<br />

nB 5.091 10 3<br />

= ×<br />

iB := 0.. nB<br />

XB :=<br />

iB<br />

YB :=<br />

iB<br />

ZB :=<br />

iB<br />

AB<br />

iB, 0<br />

AB<br />

iB, 1<br />

AB<br />

iB, 2<br />

AB<br />

iB, 3<br />

\\Imxlmmpc21\..\pr<strong>of</strong>iles_exp15_sphere B.xls<br />

XYZB<br />

iB, 0<br />

XYZB<br />

iB, 1<br />

XYZB<br />

iB, 2<br />

:= 2⋅ XB0 − XB<br />

iB<br />

:= 2⋅ YB0 − YB<br />

iB<br />

:= 2⋅ ZB0 − ZB<br />

iB<br />

:=<br />

( )<br />

( )<br />

( )<br />

−2RB0


Sphere A:<br />

yA<br />

iA, 0<br />

Sphere B:<br />

yB<br />

iB, 0<br />

4) Determine the matrices y (mx1)<br />

∆xA =<br />

( ) 1<br />

∆xA AA T AA<br />

−<br />

:=<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

centerA :=<br />

−0.664<br />

0.237<br />

−0.688<br />

−0.959<br />

�<br />

�<br />

�<br />

�<br />

�<br />

radiusA := RA0 + ∆xA3 centerA =<br />

�<br />

�<br />

�<br />

�<br />

0<br />

1<br />

2<br />

radiusA = 13.897<br />

10.1 LEAST SQUARE METHOD AND DIHEDRAL ANGLE MEASUREMENT<br />

( ) 2<br />

( ) 2<br />

RB0 2<br />

:= − XB − XB0 − YB − YB0 − ZB − ZB0<br />

iB iB iB<br />

5) We can determine ∆x = (A T A) -1 A T y<br />

AA T yA<br />

6) The precise solution is: (X0+∆x 0 ;Y0+∆x 1 ;Z0+∆x 2 ) for the center<br />

XA0 + ∆xA0<br />

YA0 + ∆xA1<br />

ZA0 + ∆xA2<br />

0<br />

10.051<br />

0.107<br />

9.141<br />

�<br />

�<br />

�<br />

( ) 2<br />

( ) 2<br />

∆xB AB T AB<br />

−<br />

:=<br />

∆xB =<br />

( ) 2<br />

( ) 1<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

0.538<br />

−0.028<br />

−0.151<br />

−0.503<br />

R0+∆x 3 for the radius<br />

( ) 2<br />

RA0 2<br />

:= − XA − XA0 − YA − YA0 − ZA − ZA0<br />

iA iA iA<br />

centerB :=<br />

�<br />

�<br />

�<br />

�<br />

�<br />

radiusB := RB0 + ∆xB3 centerB =<br />

�<br />

�<br />

�<br />

�<br />

XB0 + ∆xB0<br />

YB0 + ∆xB1<br />

ZB0 + ∆xB2<br />

0<br />

1<br />

2<br />

AB T yB<br />

0<br />

-12.511<br />

0.339<br />

8.176<br />

radiusB = 15.136<br />

�<br />

�<br />

�<br />

167


CHAPTER 10. APPENDICES<br />

168<br />

x:= 6<br />

Given<br />

7) Equations <strong>of</strong> the spheres:<br />

sphereB( x, y,<br />

z)<br />

:= x − centerB<br />

0<br />

8) Determination <strong>of</strong> a point at the intersection <strong>of</strong> the two spheres<br />

sphereA( x, y,<br />

z)<br />

= 0<br />

sphereB( x, y,<br />

z)<br />

= 0<br />

inter1 := Find( x, y,<br />

z)<br />

inter1 =<br />

9) Vectors between the center <strong>of</strong> the spheres <strong>and</strong> the intersection point: V1 et V2<br />

V1 := centerA − inter1<br />

V1 =<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

( ) 2<br />

sphereA( x, y,<br />

z)<br />

:= x − centerA<br />

0<br />

( ) 2<br />

y := 5<br />

−0.093<br />

6.656<br />

2.262<br />

10.144<br />

−6.549<br />

6.879<br />

�<br />

�<br />

10) Dihedral angle: φ = 180° - angle V1V2<br />

φ := 180 −<br />

�<br />

�<br />

�<br />

�<br />

�<br />

180<br />

π acos ⋅<br />

φ = 77.95<br />

( y − centerA<br />

1)<br />

2<br />

+ ( z − centerA<br />

2)<br />

2<br />

+ radiusA 2<br />

−<br />

( y − centerB<br />

1)<br />

2<br />

+ ( z − centerB<br />

2)<br />

2<br />

+ radiusB 2<br />

−<br />

z:= 3<br />

V2 := centerB − inter1<br />

−12.418<br />

V2 = −6.317<br />

5.914<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

V1⋅V2 V1 ⋅ V2<br />

�<br />

�<br />

��<br />

��


xa<br />

i, j<br />

ya<br />

i, j<br />

za<br />

i, j<br />

10.1 LEAST SQUARE METHOD AND DIHEDRAL ANGLE MEASUREMENT<br />

11) Representation <strong>of</strong> the spheres <strong>and</strong> the measuring points<br />

i := 0.. 50 j := 0.. 50<br />

θi i 2π<br />

:= ⋅ φj j<br />

50<br />

π<br />

:= ⋅<br />

50<br />

sphere A sphere B<br />

( )<br />

( )<br />

:= centerA + radiusA⋅sin φ<br />

0<br />

j ⋅cos<br />

θi xb := centerB + radiusB⋅sin φ<br />

i, j 0<br />

j ⋅cos<br />

θi ( )<br />

( )<br />

:= centerA + radiusA⋅sin φ<br />

1<br />

j ⋅sin<br />

θi yb := centerB + radiusB⋅sin φ<br />

i, j 1<br />

j ⋅sin<br />

θi ( )<br />

:= centerA + radiusA⋅cos φ<br />

2<br />

j<br />

zb := centerB +<br />

i, j 2<br />

( xa, ya,<br />

za)<br />

, ( xb, yb,<br />

zb)<br />

, ( XA, YA,<br />

ZA)<br />

,<br />

( XB, YB,<br />

ZB)<br />

( )<br />

( )<br />

radiusB⋅cos( φj) ( )<br />

( )<br />

169


CHAPTER 10. APPENDICES<br />

170<br />

10.2 Bulk modulus <strong>of</strong> liquid lead<br />

According to Iida <strong>and</strong> Guthrie [206] p.93,The isentropic compressibility <strong>of</strong> a liquid<br />

κS is dependent on its density ρ, <strong>and</strong> on the velocity <strong>of</strong> sound in this medium U:<br />

1 ∂V<br />

⎞<br />

κ S =− ⎟<br />

V ∂p<br />

⎠<br />

1 1<br />

U = → κ S = 2<br />

ρκ ρ U<br />

S<br />

Moreover the isothermal compressibility κ T is:<br />

1 ∂V<br />

⎞<br />

κ T =− ⎟<br />

V ∂p<br />

⎠<br />

Eq.10-11<br />

Eq.10-12<br />

Eq.10-13<br />

2<br />

Cp<br />

α MT<br />

κT = κS = κS+<br />

CV<br />

ρ Cp<br />

Eq.10-14<br />

where Cp <strong>and</strong> Cv are the heat capacity at constant pressure <strong>and</strong> volume respectively,<br />

M is the atomic weight, <strong>and</strong> α is the isobaric thermal expansivity:<br />

α =<br />

1<br />

V<br />

∂V<br />

⎞<br />

⎟<br />

∂ ⎠<br />

T p<br />

Combining Eq. 10-12 <strong>and</strong> Eq. 10-14, we have:<br />

κ<br />

T<br />

2<br />

1 α MT<br />

= + 2<br />

ρ U ρ C<br />

The bulk modulus K is therefore:<br />

S<br />

T<br />

p<br />

Eq.10-15<br />

Eq.10-16<br />

1<br />

K =<br />

κ T<br />

1<br />

=<br />

2<br />

1 α MT<br />

+ 2<br />

ρ U ρ<br />

Cp<br />

Eq.10-17<br />

The velocity <strong>of</strong> sound as a function <strong>of</strong> temperature is tabulated. A linear decrease is<br />

assumed from the incipient melting temperature T m [206]:


A linear decrease with temperature is assumed for ρ [206]:<br />

10.2 BULK MODULUS OF LIQUID LEAD<br />

Eq.10-18<br />

3 3 kg<br />

ρ( T)= 10. 67 ⋅10 −1. 32( T −Tm)→<br />

ρ(<br />

400° C)=<br />

10. 57 ⋅10<br />

⎡ ⎤<br />

3<br />

⎣⎢ m ⎦⎥<br />

Eq.10-19<br />

The heat capacity Cp is also assumed to decrease linearly with temperature [206]:<br />

With<br />

UT ( )= 1810 − 0. 38( T−Tm)→ U( 400° C<br />

⎡<br />

)= 1782<br />

⎣⎢<br />

2<br />

−3<br />

⎡ kg m<br />

Cp( T)= 32. 43 −3. 10 ⋅ 10 ( T)→ Cp( 400° C)=<br />

30. 34 ⎢<br />

⎣ K mol s<br />

T K 600<br />

m<br />

we have:<br />

= [ ]<br />

− kg<br />

M = 207 5⋅10 ⎡ ⎤<br />

⎣⎢ mol ⎦⎥<br />

3<br />

.<br />

−4 −1<br />

α = 123 . ⋅ 10 [ K ]<br />

K 400 C 27 5 GPa<br />

° ( )= [ ] .<br />

Pb<br />

m<br />

s<br />

⎤<br />

⎦⎥<br />

2<br />

⎤<br />

⎥<br />

⎦<br />

Eq.10-20<br />

Eq.10-21<br />

Eq.10-22<br />

Eq.10-23<br />

Eq.10-24<br />

171


CHAPTER 10. APPENDICES<br />

172<br />

10.3 Eshelby analysis<br />

The Mathcad spreadsheet is given below for the specific case <strong>of</strong> a grain boundary<br />

liquid inclusion, lenticular in shape. The Λ(c) curve obtained is Fig. 5-8 (a).<br />

φo := 120deg<br />

Stable intergranular lenticular inclusion:<br />

Aspect ratio : b/a = c < 1<br />

Volume <strong>of</strong> the inclusion: Vol<br />

γSL 0.432 J<br />

m 2<br />

:= ⋅<br />

�<br />

�<br />

γgb 2γSL cos φo<br />

:= ⋅ �<br />

2<br />

�<br />

�<br />

Area <strong>of</strong> the lenticular inclusion is :<br />

2<br />

3<br />

3 2<br />

Area( Vol, c)<br />

:= Vol ⋅2⋅π⋅1 + c ⋅<br />

( )<br />

�<br />

�<br />

�<br />

3 1<br />

π<br />

c 3 c 2<br />

⋅<br />

⋅ +<br />

( )<br />

Interfacial energy <strong>of</strong> the inclusion is Γ=ΓSL - Γgb:<br />

ΓSL( Vol, c)<br />

:= γSL⋅Area( Vol, c)<br />

2<br />

3<br />

Γgb( Vol, c)<br />

γSL⋅Vol ⋅2 cos φo<br />

3<br />

:=<br />

⋅ � ⋅π⋅<br />

2<br />

�<br />

�<br />

�<br />

�<br />

Γ( Vol, c)<br />

:=<br />

ΓSL( Vol, c)<br />

− Γgb( Vol, c)<br />

�<br />

�<br />

�<br />

2<br />

�<br />

�<br />

�<br />

3 1<br />

π<br />

c 3 c 2<br />

⋅<br />

⋅ +<br />

( )<br />

�<br />

�<br />

�<br />

2


ECu 111 10 9<br />

:= ⋅ ⋅Pa<br />

ν := 0.355<br />

C :=<br />

C =<br />

C11 :=<br />

C12 :=<br />

Eshelby analysis for an oblate spheroid<br />

Cu elastic modulus at 400°C<br />

Cu Poissons' ratio at 400°C<br />

ECu⋅( 1 − ν)<br />

( 1 + ν)<br />

⋅(<br />

1 − 2⋅ν) ECu⋅ν 1 + ν ⋅ −<br />

ECu<br />

C44 :=<br />

21+ ν<br />

( ) ( 1 2⋅ν) ( )<br />

Elastic constants <strong>of</strong> Cu, taken as isotropic<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

C11<br />

C12<br />

C12<br />

0<br />

0<br />

0<br />

C12<br />

C11<br />

C12<br />

0<br />

0<br />

0<br />

1.822 10 11<br />

×<br />

1.003 10 11<br />

×<br />

1.003 10 11<br />

×<br />

0<br />

0<br />

0<br />

C12<br />

C12<br />

C11<br />

0<br />

0<br />

0<br />

KPb 27.5 10 9<br />

:= ⋅ ⋅Pa<br />

0<br />

0<br />

0<br />

C44<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

C44<br />

0<br />

1.003 10 11<br />

×<br />

1.822 10 11<br />

×<br />

1.003 10 11<br />

×<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

C44<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

1.003 10 11<br />

×<br />

1.003 10 11<br />

×<br />

1.822 10 11<br />

×<br />

Pb bulk modulus at 400°C<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

4.096 10 10<br />

×<br />

Elastic constant <strong>of</strong> liquid Pb: G=0 (no resistance to shear)<br />

�<br />

KPb KPb KPb 0 0 0<br />

� KPb<br />

�<br />

KPb<br />

Cstar :=<br />

�<br />

0<br />

KPb<br />

KPb<br />

0<br />

KPb<br />

KPb<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0 0 0 0 0 0<br />

0 0 0 0 0 0<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

4.096 10 10<br />

×<br />

0<br />

10.3 ESHELBY ANALYSIS<br />

0<br />

0<br />

0<br />

0<br />

0<br />

4.096 10 10<br />

×<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

Pa<br />

173


CHAPTER 10. APPENDICES<br />

174<br />

Sc ( )<br />

aspect ratio <strong>of</strong> the oblate spheroid: c=a3/a1


x1 :=<br />

x1b ξ<br />

( ) := Q( θ)<br />

⋅x1b(<br />

ξ)<br />

x1c ξ, θ<br />

�<br />

�<br />

�<br />

�<br />

(ii) rotation de θ autour de x2b -> x1c x2c x3c<br />

Transformation matrix: Q<br />

Cosine matrix: Aij<br />

( ) := x1c( ξ, θ)<br />

⋅x1<br />

A11 ξ, θ<br />

( ) := x2c( ξ, θ)<br />

⋅x1<br />

A21 ξ, θ<br />

( ) := x3c( ξ, θ)<br />

⋅x1<br />

A31 ξ, θ<br />

1<br />

0<br />

0<br />

�<br />

�<br />

Transformation matrix : P<br />

( ) := P( ξ)<br />

⋅x1<br />

Stress tensor<br />

-> after successive ξ <strong>and</strong> θ rotation<br />

A( ξ, θ)<br />

x2 :=<br />

(i) rotation de ξ autour de x1 -> x1b x2b x3b<br />

x2b ξ<br />

:=<br />

x2c ξ, θ<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

0<br />

1<br />

0<br />

�<br />

�<br />

P( ξ)<br />

:=<br />

( ) := P( ξ)<br />

⋅x2<br />

Q( θ)<br />

( ) := Q( θ)<br />

⋅x2b(<br />

ξ)<br />

A12 ξ, θ<br />

A22 ξ, θ<br />

A32 ξ, θ<br />

( )<br />

( )<br />

( )<br />

A11 ξ, θ<br />

A21 ξ, θ<br />

A31 ξ, θ<br />

�<br />

�<br />

�<br />

�<br />

:=<br />

1 0<br />

0 cos ξ<br />

0 −sin<br />

ξ<br />

�<br />

�<br />

�<br />

�<br />

cos( θ)<br />

0<br />

sin θ<br />

( )<br />

( )<br />

( )<br />

( ) := x1c( ξ, θ)<br />

⋅x2<br />

( ) := x2c( ξ, θ)<br />

⋅x2<br />

( ) := x3c( ξ, θ)<br />

⋅x2<br />

( )<br />

( )<br />

( )<br />

A12 ξ, θ<br />

A22 ξ, θ<br />

A32 ξ, θ<br />

�<br />

( )<br />

( ) �<br />

0<br />

sin ξ<br />

cos ξ<br />

x3 :=<br />

( ) �<br />

0 −sin<br />

θ<br />

1 0<br />

0 cos( θ)<br />

( ) �<br />

( )<br />

( ) �<br />

A13 ξ, θ<br />

A23 ξ, θ<br />

A33 ξ, θ<br />

x3b ξ<br />

�<br />

�<br />

�<br />

�<br />

�<br />

0<br />

0<br />

1<br />

�<br />

�<br />

x3c ξ, θ<br />

A13 ξ, θ<br />

A23 ξ, θ<br />

A33 ξ, θ<br />

10.3 ESHELBY ANALYSIS<br />

( ) := P( ξ)<br />

⋅x3<br />

( ) := Q( θ)<br />

⋅x3b(<br />

ξ)<br />

( ) := x1c( ξ, θ)<br />

⋅x3<br />

( ) := x2c( ξ, θ)<br />

⋅x3<br />

( ) := x3c( ξ, θ)<br />

⋅x3<br />

175


CHAPTER 10. APPENDICES<br />

176<br />

Stress tensor before transformation: σij<br />

Stress tensor after transformation: σkl<br />

σ1( ξ, θ)<br />

σ2( ξ, θ)<br />

σ3( ξ, θ)<br />

σ4( ξ, θ)<br />

σ5( ξ, θ)<br />

σ6( ξ, θ)<br />

stress:<br />

strain:<br />

eigenstrain ε∗<br />

2 2<br />

:=<br />

A( ξ, θ)<br />

0i , ⋅A( ξ, θ)<br />

0, j⋅σij<br />

� � ( i, j)<br />

i= 0 j = 0<br />

2 2<br />

:=<br />

A( ξ, θ)<br />

1i , ⋅A( ξ, θ)<br />

1, j⋅σij<br />

� � ( i, j)<br />

i= 0 j = 0<br />

2 2<br />

:=<br />

A( ξ, θ)<br />

2i , ⋅A( ξ, θ)<br />

2, j⋅σij<br />

� � ( i, j)<br />

i= 0 j = 0<br />

2 2<br />

:=<br />

A( ξ, θ)<br />

1i , ⋅A( ξ, θ)<br />

2, j⋅σij<br />

� � ( i, j)<br />

i= 0 j = 0<br />

2<br />

:= �<br />

i= 0<br />

2<br />

A( ξ, θ)<br />

0i , ⋅A( ξ, θ)<br />

2, j⋅σij<br />

� ( i, j)<br />

j = 0<br />

2<br />

:= �<br />

i= 0<br />

2<br />

( A( ξ, θ)<br />

0i , ⋅A( ξ, θ)<br />

1, j⋅σij<br />

� i, j)<br />

j = 0<br />

( ) := C S c<br />

εstar c, σ , ξ,<br />

θ<br />

( )<br />

σ° σ, ξ,<br />

θ<br />

ε° σ, ξ,<br />

θ<br />

Interaction strain energy: ∆W=-1/2 Vol σ° ε∗<br />

[ ⋅( ( ) − I)<br />

− Cstar⋅S( c)<br />

] 1 −<br />

⋅ ( Cstar − C)<br />

⋅ε°<br />

σ, ξ,<br />

θ<br />

−1<br />

( )<br />

∆W Vol, c,<br />

σ,<br />

ξ,<br />

θ<br />

( ) C 1<br />

:=<br />

:=<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

:=<br />

( ) �<br />

(<br />

(<br />

(<br />

(<br />

) �<br />

) �<br />

)<br />

�<br />

�<br />

) �<br />

( ) �<br />

σ1 ξ, θ<br />

σ2 ξ, θ<br />

σ3 ξ, θ<br />

σ4 ξ, θ<br />

σ5 ξ, θ<br />

σ6 ξ, θ<br />

�<br />

0 0 0<br />

σij := � 0 0 0<br />

�<br />

0 0 1<br />

�<br />

10 6<br />

⋅ ⋅σ⋅Pa ( )<br />

− ⋅σ°<br />

σ, ξ,<br />

θ<br />

�� ( ) ��<br />

( )<br />

( )<br />

2 Vol ⋅ σ° σ, ξ θ , ⋅ ⋅εstar<br />

c, σ , ξ,<br />

θ<br />

�<br />


Fcξ , , θ<br />

Adimensionalisation:<br />

Gc ( ) 2⋅π 1 c 2<br />

3<br />

3 1<br />

⋅(<br />

+ )<br />

π<br />

c 3 c 2<br />

2<br />

⋅ � ⋅ � 2 cos<br />

� �<br />

⋅ +<br />

φo<br />

3<br />

� �<br />

:=<br />

− ⋅ � ⋅π⋅<br />

� 2 �<br />

�<br />

�<br />

�<br />

−1<br />

�<br />

( ) := �<br />

2<br />

�<br />

�<br />

�<br />

�<br />

Λ( c, 0deg,<br />

0deg)<br />

σ1 ξ, θ<br />

�<br />

(<br />

(<br />

(<br />

(<br />

(<br />

,<br />

,<br />

) � �<br />

�<br />

) � �<br />

) � �<br />

)<br />

�⋅�<br />

� �<br />

) � �<br />

( ) �<br />

� �<br />

σ2 ξ, θ<br />

σ5 ξ, θ<br />

σ6 ξ, θ<br />

( )<br />

�<br />

�<br />

�<br />

�<br />

3 1<br />

π<br />

c 3 c 2<br />

⋅<br />

⋅ +<br />

( )<br />

σ3 ξ θ<br />

⋅ [ C + ( Cstar − C)<br />

⋅S(<br />

c)<br />

]<br />

σ4 ξ θ<br />

1 −<br />

( C − Cstar)<br />

ECu C 1 −<br />

⋅�<br />

⋅�<br />

⋅ ⋅<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0.6<br />

Λ c, ξ,<br />

θ<br />

Λ=σ 2 V 1/3 /γ SL E<br />

0.53<br />

d<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

( )<br />

c ( ) Gc<br />

d<br />

0.47<br />

c Fcξ , θ ,<br />

:= −<br />

d<br />

d<br />

0.4<br />

( )<br />

0.34<br />

c<br />

0.27<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

0.21<br />

10.3 ESHELBY ANALYSIS<br />

2<br />

�<br />

��<br />

0.14<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

�<br />

( ) ����<br />

���<br />

( ) ����<br />

( ) ����<br />

( )<br />

����<br />

����<br />

( ) ����<br />

( ) ���<br />

����<br />

σ1 ξ, θ<br />

σ2 ξ, θ<br />

σ3 ξ, θ<br />

σ4 ξ, θ<br />

σ5 ξ, θ<br />

σ6 ξ, θ<br />

0.076<br />

0.01<br />

177


CHAPTER 10. APPENDICES<br />

178<br />

10.4 Measured values after interrupted creep test<br />

The measured dihedral angles in 400 °C crept C99 samples at different stresses are<br />

listed. For each single inclusion treated, its specific volume <strong>and</strong> the orientation <strong>of</strong> the<br />

grain boundary are indicated. The adimensional parameter Λ is also added:<br />

2 1/ 3<br />

σ app ⋅<br />

Λ=<br />

γ ⋅<br />

V<br />

E<br />

σ app [MPa] φ [°] V [10 -16 m 3 ] ξ [°] θ [°] Λ [-]<br />

0 102<br />

117<br />

125<br />

117<br />

43 125<br />

88<br />

59 79<br />

96<br />

74<br />

69 *<br />

65<br />

84<br />

0.54<br />

0.26<br />

1.06<br />

7.38<br />

2.93<br />

6.32<br />

1.20<br />

1.64<br />

3.32<br />

5.28<br />

14.11<br />

2.08<br />

-<br />

-<br />

-<br />

-<br />

3<br />

64<br />

38<br />

86<br />

54<br />

37<br />

71<br />

10<br />

-<br />

-<br />

-<br />

-<br />

3<br />

8<br />

0<br />

0<br />

7<br />

25<br />

10<br />

0<br />

Eq. 5-11<br />

0<br />

0<br />

0<br />

0<br />

0.256<br />

0.331<br />

0.358<br />

0.398<br />

0.503<br />

0.587<br />

0.814<br />

0.430<br />

75 88 1.93 62 7 0.678


Chapter 11<br />

11 List <strong>of</strong> References<br />

1.Kuyucak, S. <strong>and</strong> M. Sahoo, A review <strong>of</strong> the machinability <strong>of</strong> <strong>copper</strong>-base <strong>alloys</strong>.<br />

Canadian Metallurgical Quarterly, 1996. 35(1): p. 1-15.<br />

2.Smith, C.S., Grains, Phases, <strong>and</strong> Interfaces — an Interpretation <strong>of</strong> Microstructure.<br />

Transactions <strong>of</strong> the American Institute <strong>of</strong> Mining <strong>and</strong> Metallurgical Engineers, 1948.<br />

175: p. 15-51.<br />

3.Lupis, C.H.P., Chemical Thermodynamics <strong>of</strong> Materials. 1983, New York,<br />

Amsterdam: North-Holl<strong>and</strong>, Elsevier. 375-380.<br />

4.Wray, P.J., The geometry <strong>of</strong> two-phase aggregates in which the shape <strong>of</strong> the second<br />

phase is determined by its dihedral angle. Acta Metallurgica, 1976. 24(2): p. 125-<br />

135.<br />

5.Eustathopoulos, N., Energetics <strong>of</strong> solid/liquid interfaces <strong>of</strong> metals <strong>and</strong> <strong>alloys</strong>.<br />

International Metals Reviews, 1983. 28(4): p. 189-210.<br />

6.Mori, T., et al., Determination <strong>of</strong> the Energies <strong>of</strong> 001 Twist Boundaries in Cu with<br />

the Shape <strong>of</strong> Boundary SiO2 Particles. Philosophical Magazine Letters, 1988. 58(1):<br />

p. 11-15.<br />

7.Harker, D. <strong>and</strong> E.R. Parker, Grain Shape <strong>and</strong> Grain Growth. Transactions <strong>of</strong> the<br />

American Society for Metals, 1945. 34: p. 156-201.<br />

8.Riegger, O.K. <strong>and</strong> L.H. Van Vlack, Dihedral Angle Measurement. Transactions <strong>of</strong><br />

the American Institute <strong>of</strong> Mining <strong>and</strong> Metallurgical Engineers, 1960. 218(5): p. 933-<br />

935.<br />

9.Stickels, C.A. <strong>and</strong> E.E. Hucke, Measurement <strong>of</strong> Dihedral Angles. Transactions <strong>of</strong><br />

the Metallurgical Society <strong>of</strong> AIME, 1964. 230(4): p. 795-801.<br />

10.Jurewicz, S.R. <strong>and</strong> A.J.G. Jurewicz, Distribution <strong>of</strong> Apparent Angles on R<strong>and</strong>om<br />

Sections with Emphasis on Dihedral Angle Measurements. Journal <strong>of</strong> Geophysical<br />

<strong>Research</strong>-Solid Earth <strong>and</strong> Planets, 1986. 91(B9): p. 9277-9282.<br />

179


CHAPTER 11. LIST OF REFERENCES<br />

180<br />

11.Fortes, M.A. <strong>and</strong> A.C. Ferro, Determination <strong>of</strong> Dihedral Angle Distributions in<br />

Polycrystals - Application to Alpha+Beta-Brass. Metallurgical Transactions A-<br />

Physical Metallurgy <strong>and</strong> Materials Science, 1988. 19(5): p. 1147-1151.<br />

12.De H<strong>of</strong>f, R.T., Estimation <strong>of</strong> Dihedral Angles from Stereological Counting<br />

Measurements. Metallography, 1986. 19(2): p. 209-217.<br />

13.Eustathopoulos, N., et al., Etude thermodynamique de la tension interfaciale<br />

solide/liquide pour un système métallique binaire. III. Journal de chimie physique,<br />

1974. 71(11-12): p. 1465-1471.<br />

14.Mullins, W.W., Grain Boundary Grooving by Volume Diffusion. Transactions <strong>of</strong><br />

the Metallurgical Society <strong>of</strong> AIME, 1960. 218: p. 354-361.<br />

15.Rabkin, E. <strong>and</strong> L. Klinger, The fascination <strong>of</strong> grain boundary grooves. Advanced<br />

Engineering Materials, 2001. 3(5): p. 277-282.<br />

16.Gabrisch, H., U. Dahmen, <strong>and</strong> E. Johnson, In situ measurement <strong>of</strong> dihedral angles<br />

at liquid grain boundary inclusions. Microscopy <strong>Research</strong> <strong>and</strong> Technique, 1998.<br />

42(4): p. 241-247.<br />

17.Dahmen, U., et al., Magic-size equilibrium shapes <strong>of</strong> nanoscale Pb inclusions in<br />

Al. Physical Review Letters, 1997. 78(3): p. 471-474.<br />

18.Johnson, E., et al., Shapes <strong>and</strong> sizes <strong>of</strong> nanoscale Pb inclusions in Al. Materials<br />

Science <strong>and</strong> Engineering A, 2001. 304-306: p. 187-193.<br />

19.Gabrisch, H., et al., Equilibrium shape <strong>and</strong> interface roughening <strong>of</strong> small liquid<br />

Pb inclusions in solid Al. Acta Materialia, 2001. 49(20): p. 4259-4269.<br />

20.Robertson, W.M., Grain Boundary Grooving <strong>and</strong> Scratch Decay on Copper in<br />

Liquid Lead. Transactions <strong>of</strong> the Metallurgical Society <strong>of</strong> AIME, 1965. 233(July): p.<br />

1232-1236.<br />

21.Ratke, L. <strong>and</strong> H.J. Vogel, Theory <strong>of</strong> Grain-Boundary Grooving in the Convective-<br />

Diffusive Regime. Acta Materialia, 1991. 39(5): p. 915-923.<br />

22.Ikeuye, K.K. <strong>and</strong> C.S. Smith, Studies <strong>of</strong> Interface Energies in Some Aluminum <strong>and</strong><br />

Copper Alloys. Metals Transactions, 1949. 185: p. 762-768.<br />

23.Eustathopoulos, N., Contribution à l'étude de l'interface solide-liquide de métaux<br />

et alliages, in Ecole Nationale Supérieure d'Electrochimie et d'Electrométallurgie.<br />

1974, Université Scientifique et médicale de Grenoble: Grenoble. p. 118.<br />

24.Joud, J.C., et al., Détermination de la Tension Superficielle des Alliages Ag-Pb et<br />

Cu-Pb par la Méthode de la Goutte Posée. Journal de chimie physique, 1973. 70(9):<br />

p. 1290-1294.<br />

25.Eustathopoulos, N., J.C. Joud, <strong>and</strong> P.J. Desre, Etude thermodynamique de la<br />

tension interfaciale solide/liquide pour un système métallique binaire. II.-<br />

Applications. Journal de chimie physique, 1974. 71(5): p. 777-787.


26.Coudurier, L., et al., Corrosion Intergranullaire du Cuivre par le Plomb Liquide<br />

sous l'effet des Forces Capillaires. Journal de chimie physique, 1977. 74(3): p. 289-<br />

294.<br />

27.Camel, D., N. Eustathopoulos, <strong>and</strong> P. Desré, Chemical adsorption <strong>and</strong><br />

temperature dependence <strong>of</strong> the solid-liquid interfacial tension <strong>of</strong> metallic binary<br />

<strong>alloys</strong>. Acta Metallurgica, 1980. 28(3): p. 239-247.<br />

28.Sanchez Medina, M., L. Coudurier, <strong>and</strong> N. Eustathopoulos, Microstructure<br />

d'équilibre et tensions interfaciales dans les systèmes biphasés solide-liquide CuPbM<br />

(M : Ag, Ni, Sn, Sb). Mémoires et Etudes Scientifiques Revue de Métallurgie, 1981.<br />

78(11): p. 607-617.<br />

29.Straumal, B., D. Molodov, <strong>and</strong> W. Gust, Grain boundary wetting phase<br />

transitions in the Al-Sn <strong>and</strong> Al-Sn-Pb systems. Materials Science Forum, 1996. 207-<br />

209: p. 437-440.<br />

30.Straumal, B., et al., The wetting transition in high <strong>and</strong> low energy grain<br />

boundaries in the Cu(In) system. Acta Metallurgica et Materialia, 1992. 40(5): p.<br />

939-945.<br />

31.Baker, H., et al., eds. ASM H<strong>and</strong>book Volume 03: Alloy Phase Diagrams. 1992,<br />

ASM International, Materials Park, OH.<br />

32.Stickels, C.A. <strong>and</strong> E.E. Hucke, Some Effects <strong>of</strong> Temperature + Hydrostatic<br />

Pressure on Interfacial Tensions in Nickel-Lead System. Transactions <strong>of</strong> the<br />

Metallurgical Society <strong>of</strong> Aime, 1964. 230(1): p. 20-27.<br />

33.Stickels, C.A. <strong>and</strong> E.E. Hucke, Effect <strong>of</strong> Stress on Dihedral Angle in Leaded<br />

Nickel. Journal <strong>of</strong> the Institute <strong>of</strong> Metals, 1964. 92(8): p. 234-237.<br />

34.Waterhouse, R.B. <strong>and</strong> D. Grubb, Embrittlememt <strong>of</strong> Alpha-Brasses by Liquid<br />

Metals. Journal <strong>of</strong> the Institute <strong>of</strong> Metals, 1963. 91(6): p. 216-219.<br />

35.Stickels, C.A., Discussion <strong>of</strong> "Embrittlememt <strong>of</strong> Alpha-Brasses by Liquid Metals".<br />

Journal <strong>of</strong> the Institute <strong>of</strong> Metals, 1963. 91(12): p. 422-423.<br />

36.Sun, B., Z. Suo, <strong>and</strong> A.G. Evans, Emergence <strong>of</strong> cracks by mass transport in elastic<br />

crystals stressed at high temperatures. Journal <strong>of</strong> the Mechanics <strong>and</strong> Physics <strong>of</strong><br />

Solids, 1994. 42(11): p. 1653-1677.<br />

37.Eshelby, J.D., The Determination <strong>of</strong> the Elastic Field <strong>of</strong> an Ellipsoidal Inclusion,<br />

<strong>and</strong> Related Problems. Proceedings <strong>of</strong> the Royal Society <strong>of</strong> London Series A-<br />

Mathematical <strong>and</strong> Physical Sciences, 1957. 241(1226): p. 376-396.<br />

38.Mura, T., Micromechanics <strong>of</strong> Defects in Solids. second edition ed. Mechanics <strong>of</strong><br />

Elastic <strong>and</strong> Inelastic Solids, ed. S. Nemat-Nasser <strong>and</strong> G.A. Oravas. 1987, Dordrecht/<br />

Boston /London: Kluver Academic Publishers. 587.<br />

181


CHAPTER 11. LIST OF REFERENCES<br />

182<br />

39.Wang, H. <strong>and</strong> Z.H. Li, The three-dimensional analysis for diffusive shrinkage <strong>of</strong> a<br />

grain-boundary void in stressed solid. Journal <strong>of</strong> Materials Science, 2004. 39(10): p.<br />

3425-3432.<br />

40.Raj, R., Unstable Spreading <strong>of</strong> a Fluid Inclusion in a Grain Boundary under<br />

Normal Stress. Journal <strong>of</strong> the American Ceramic Society, 1986. 69(9): p. 708-712.<br />

41.Wang, W. <strong>and</strong> Z. Suo, Shape change <strong>of</strong> a pore in a stressed solid via surface<br />

diffusion motivated by surface <strong>and</strong> elastic energy variation. Journal <strong>of</strong> the Mechanics<br />

<strong>and</strong> Physics <strong>of</strong> Solids, 1997. 45(5): p. 709-729.<br />

42.Suo, Z. <strong>and</strong> H. Yu, Crack nucleation on an elastic polycrystal surface in a<br />

corrosive environment: Low dimensional dynamical models. Acta Materialia, 1997.<br />

45(6): p. 2235-2245.<br />

43.Rice, J.R. <strong>and</strong> T.J. Chuang, Energy Variations in Diffusive Cavity Growth. Journal<br />

<strong>of</strong> the American Ceramic Society, 1981. 64(1): p. 46-53.<br />

44.Srolovitz, D.J. <strong>and</strong> S.H. Davis, Do stresses modify wetting angles? Acta<br />

Materialia, 2001. 49(6): p. 1005-1007.<br />

45.Chuang, T.-J. <strong>and</strong> J.R. Rice, The shape <strong>of</strong> intergranular creep cracks growing by<br />

surface diffusion. Acta Metallurgica, 1973. 21(12): p. 1625-1628.<br />

46.Takahashi, Y., K. Takahashi, <strong>and</strong> K. Nishiguchi, A numerical analysis <strong>of</strong> void<br />

shrinkage processes controlled by coupled surface <strong>and</strong> interface diffusion. Acta<br />

Metallurgica et Materialia, 1991. 39(12): p. 3199-3216.<br />

47.Sanchez Medina, M., et al., Influence des additions de Pb et de Sn sur la ductilité<br />

du cuivre OFHC entre la température ambiante et 1000 °C. Mémoires et Etudes<br />

Scientifiques Revue de Métallurgie, 1981. 78(2): p. 101-109.<br />

48.Foulger, R.V. <strong>and</strong> E. Nicholls, Influence <strong>of</strong> composition <strong>and</strong> <strong>microstructure</strong> on<br />

mechanical working properties <strong>of</strong> <strong>copper</strong>-base <strong>alloys</strong>. Metals Technology, 1976. 3: p.<br />

366-369.<br />

49.Davies, P.W., et al., Fracture behaviour <strong>of</strong> Copper during Hot Deformation.<br />

Journal <strong>of</strong> the Institute <strong>of</strong> Metals, 1971. 99: p. 195-197.<br />

50.Sanchez Medina, M., R. Sangiorgi, <strong>and</strong> N. Eustathopoulos, On the <strong>embrittlement</strong><br />

<strong>of</strong> Cu by Liquid Pb Inclusions. Scripta Materialia, 1981. 15(7): p. 737-738.<br />

51.Reid, b.J. <strong>and</strong> J.N. Greenwood, Intergranular Cavitation in Stressed Copper-<br />

Nickel Alloys. Transactions <strong>of</strong> the Metallurgical Society <strong>of</strong> AIME, 1958. 212: p. 503-<br />

507.<br />

52.Ashby, M.F., C. G<strong>and</strong>hi, <strong>and</strong> D.M.R. Taplin, Fracture-Mechanism Maps <strong>and</strong> their<br />

Construction for F.C.C. Metals <strong>and</strong> Alloys. Acta Materialia, 1979. 27(5): p. 699-729.<br />

53.Hondros, E.D. <strong>and</strong> D. McLean, Cohesion Margin <strong>of</strong> Copper. Philosophical<br />

Magazine, 1974. 29(4): p. 771-795.


54.Watanabe, T., Grain Boundary Sliding <strong>and</strong> Stress Concentration during Creep.<br />

Metallurgical Transactions A, 1983. 14A: p. 531-545.<br />

55.Onaka, S., M. Kato, <strong>and</strong> R. Tanaka, Intermediate Temperature Embrittlement <strong>of</strong><br />

Dispersion Hardened Cu-GeO2 Polycrystals. Transactions <strong>of</strong> the Japan Institute <strong>of</strong><br />

Metals, 1987. 28(1): p. 32-40.<br />

56.Ayensu, A. <strong>and</strong> T.G. Langdon, The Inter-Relationship between Grain Boundary<br />

Sliding <strong>and</strong> Cavitation during Creep <strong>of</strong> Polycrystalline Copper. Metallurgical<br />

Transactions A, 1996. 27A: p. 901-907.<br />

57.Miura, H., T. Sakai, <strong>and</strong> M. Kato, Ductility <strong>and</strong> Fracture <strong>of</strong> Copper Bicrystals<br />

with Boundary SiO2 Particles. Materials Science <strong>and</strong> Technology, 1994. 10(12): p.<br />

1044-1049.<br />

58.Afshar, S. <strong>and</strong> M. Biscondi, Intergranular cavitation observed during OFHC<br />

<strong>copper</strong> bicrystal creep. Scripta Metallurgica, 1986. 20(5): p. 707-711.<br />

59.Watanabe, T. <strong>and</strong> S. Tsurekawa, The control <strong>of</strong> brittleness <strong>and</strong> development <strong>of</strong><br />

desirable mechanical properties in polycrystalline systems by grain boundary<br />

engineering. Acta Materialia, 1999. 47(15-16): p. 4171-4185.<br />

60.Miura, H., et al., Correlation between Intermediate-Temperature-Embrittlement<br />

<strong>and</strong> Grain-Boundary Energy. Materials Science Forum, 1996. 207-209: p. 649-652.<br />

61.Monzen, R., et al., Bismuth <strong>embrittlement</strong> <strong>of</strong> [011] twist boundaries in <strong>copper</strong><br />

bicrystals. Metallurgical <strong>and</strong> Materials Transactions a-Physical Metallurgy <strong>and</strong><br />

Materials Science, 1999. 30(2): p. 483-485.<br />

62.Monzen, R. <strong>and</strong> T. Okamoto, Grain-boundary diffusion <strong>of</strong> Bi in Cu bicrystals.<br />

Materials Science Forum, 1999. 294-296: p. 561-564.<br />

63.Kamdar, M.H., Solid Metal Induced Embrittlement, in Corrosion. 2001, ASM<br />

International: Materials Park, OH, USA. p. 184-189.<br />

64.Gavin, S.A., et al., Effect <strong>of</strong> trace impurities on hot ductility <strong>of</strong> as-cast cupronickel<br />

<strong>alloys</strong>. Metals Technology, 1978: p. 397-401.<br />

65.Eborall, R. <strong>and</strong> P. Gregory, The Mechanism <strong>of</strong> Embrittlement by a Liquid Phase.<br />

Journal <strong>of</strong> the Institute <strong>of</strong> Metals, 1955. 84: p. 88-90.<br />

66.Roth, M.C., G.C. Weatherly, <strong>and</strong> W.A. Miller, The temperature dependence <strong>of</strong> the<br />

mechanical properties <strong>of</strong> aluminum <strong>alloys</strong> containing low-melting-point inclusions.<br />

Acta Metallurgica, 1980. 28(7): p. 841-853.<br />

67.Wolley, D.J. <strong>and</strong> A.G. Fox, The <strong>embrittlement</strong> <strong>of</strong> <strong>leaded</strong> <strong>and</strong> un<strong>leaded</strong> a+b 60/40<br />

brasses in the temperature range 300 to 500°C. Journal <strong>of</strong> Materials Science Letters,<br />

1988. 7: p. 763-765.<br />

68.Fujiwara, S. <strong>and</strong> K. Abiko, Ductility <strong>of</strong> Ultra High Purity Copper. Journal de<br />

Physique IV, 1995. 5(C7): p. 295-300.<br />

183


CHAPTER 11. LIST OF REFERENCES<br />

184<br />

69.Misra, R.D.K., V.S. Prasad, <strong>and</strong> P.R. Rao, Dynamic <strong>embrittlement</strong> in an agehardenable<br />

<strong>copper</strong>-chromium alloy. Scripta Materialia, 1996. 35(1): p. 129-133.<br />

70.Kanno, M., Effect <strong>of</strong> a Small Addition <strong>of</strong> Zirconium on Hot Ductility <strong>of</strong> a Cu-Cr<br />

Alloy. Zeitschrift für Metallkunde, 1988. 79: p. 684-688.<br />

71.Kanno, M., et al., Effects <strong>of</strong> a Small Amount <strong>of</strong> Added Elements on the Ductility <strong>of</strong><br />

a Cu-8 mass%Sn Alloy at Elevated Temperature. Transactions <strong>of</strong> the Japan Institute<br />

<strong>of</strong> Metals, 1987. 28(1): p. 73-80.<br />

72.Kanno, M. <strong>and</strong> N. Shimodaira, Improvement in Hot Ductility in a Cu-4.4mol.% Sn<br />

Alloy by Adding a Small Amount <strong>of</strong> Third Element. Transactions <strong>of</strong> the Japan Institute<br />

<strong>of</strong> Metals, 1987. 28(9): p. 742-748.<br />

73.Kanno, M. <strong>and</strong> N. Shimodaira, Improvement in Ductility at Intermediate<br />

Temperatures by a small Addition <strong>of</strong> Yttrium in a Cu-30Zn Alloy. Scripta Materialia,<br />

1987. 21: p. 1487-1492.<br />

74.Kubosono, K., et al., Copper-Nickel Based Alloy. 1995, Mitsubishi Denki<br />

Kabushiki Kaisha: US Patent.<br />

75.Ohmori, M., K. Wakasa, <strong>and</strong> Y. Yoshinaga, Effect <strong>of</strong> strain Rate on Tensile<br />

Ductility <strong>of</strong> Polycrystalline Copper <strong>and</strong> alfa-Brasses at Intermediate Temperatures.<br />

Journal <strong>of</strong> the Japan Institute <strong>of</strong> Metals, 1973. 37(11): p. 1188-1194.<br />

76.Ohmori, M., Y. Yoshinaga, <strong>and</strong> H. Maniwa, The Effect <strong>of</strong> Strain Rate on<br />

Intermediate Temperature Embrittlement <strong>of</strong> Copper. Journal <strong>of</strong> the Japan Institute <strong>of</strong><br />

Metals, 1970. 8: p. 791-797.<br />

77.Ohmori, M., et al., High Temperature Tensile Ductility <strong>of</strong> Copper in Vacuum.<br />

Journal <strong>of</strong> the Japan Institute <strong>of</strong> Metals, 1976. 40(8): p. 802-807.<br />

78.Perovic, D.D., G.C. Weatherly, <strong>and</strong> W.A. Miller, On the strain rate dependence <strong>of</strong><br />

ductility troughs observed under liquid metal <strong>embrittlement</strong> <strong>and</strong> creep conditions.<br />

Scripta Materialia, 1987. 21: p. 345-347.<br />

79.Nowosielski, R., Ductility Minimum Temperature in Selected Monophase Binary<br />

Brasses. Journal <strong>of</strong> Materials Processing Technology, 2001. 109: p. 142-153.<br />

80.Sato, S., T. Otsu, <strong>and</strong> E. Hata, Embrittlement <strong>of</strong> Copper Alloys during Annealing.<br />

Journal <strong>of</strong> the Institute <strong>of</strong> Metals, 1970. 98: p. 135-141.<br />

81.Wang, G.X., Effect <strong>of</strong> Microstructural Modification on Tensile <strong>and</strong> Fatigue<br />

Properties <strong>of</strong> Cu-35%Ni-3.5%Cr Alloy. Acta Metallurgica et Materialia, 1994. 42(7):<br />

p. 2547-2554.<br />

82.Shibayanagi, T., S. Saji, <strong>and</strong> S. Hori, Improvement in low ductility failure in<br />

Copper. Supplement to Transactions <strong>of</strong> the Japan Institute <strong>of</strong> Metals, 1986. 27<br />

Supplement(Supplement): p. 943-950.


83.McQueen, H.J., Dynamic Recovery <strong>and</strong> Recrystallization, in Encyclopedia <strong>of</strong><br />

Materials: Science <strong>and</strong> Technology, K.H.J. Buschow, et al., Editors. 2001, Elsevier<br />

Science Ltd.: Oxford. p. 2375-2381.<br />

84.Gottstein, G., D.A. Molodov, <strong>and</strong> L.S. Shvindlerman, Grain Boundary Migration<br />

in Metals: Recent Developments. Interface Science, 1998. 6: p. 7-22.<br />

85.Barrera, E.V., et al., Quasi-Static Intergranular Cracking in a Cu-Sn Alloy; an<br />

Analog <strong>of</strong> Stress Relief Cracking <strong>of</strong> Steels. Scripta Metallurgica et Materialia, 1992.<br />

27: p. 205-210.<br />

86.Muthiah, R.C., et al., Tin-Induced Dynamic Embrittlement <strong>of</strong> Cu-7%Sn Bicrystals<br />

with a Σ5 Boundary. Acta Materialia, 1999. 47(9): p. 2797-2808.<br />

87.Pfaendtner, J.A., et al., Time-dependent Interfacial Failure in Metallic Alloys.<br />

Materials Science <strong>and</strong> Engineering, 1999. A260: p. 1-11.<br />

88.Hondros, E.D. <strong>and</strong> M.P. Seah, Segregation to Interfaces. International Metals<br />

Reviews, 1977: p. 262-300.<br />

89.Misra, R.D.K., C.J. McMahon Jr., <strong>and</strong> A. Guha, Brittle Behavior <strong>of</strong> a Dilute<br />

Copper-Beryllium Alloy at 200°C in Air. Scripta Materialia, 1994. 31(11): p. 1471-<br />

1474.<br />

90.Hondros, E.D., et al., Interfacial <strong>and</strong> Surface Microchemistry, in Physical<br />

Metallurgy: Fourth, Revised <strong>and</strong> Enhanced Edition, R.W. Cahn <strong>and</strong> P. Haasen,<br />

Editors. 1996, Elsevier Science Publishers, BV: Amsterdam. p. 1201-1289.<br />

91.Raynolds, J.E., et al., Impurity effects on adhesion at an interface between NiAl<br />

<strong>and</strong> Mo. Acta Materialia, 1999. 47(11): p. 3281-3289.<br />

92.Van der Ven, A. <strong>and</strong> G. Ceder, The thermodynamics <strong>of</strong> decohesion. Acta<br />

Materialia, 2004. 52(5): p. 1223-1235.<br />

93.Messmer, R.P. <strong>and</strong> C.L. Briant, The role <strong>of</strong> chemical bonding in grain boundary<br />

<strong>embrittlement</strong>. Acta Metallurgica, 1982. 30(2): p. 457-467.<br />

94.Keast, V.J., et al., Chemistry <strong>and</strong> bonding changes associated with the segregation<br />

<strong>of</strong> Bi to grain boundaries in Cu. Acta Materialia, 1998. 46(2): p. 481-490.<br />

95.Bruley, J., V.J. Keast, <strong>and</strong> D.B. Williams, An EELS study <strong>of</strong> segregation-induced<br />

grain-boundary <strong>embrittlement</strong> <strong>of</strong> <strong>copper</strong>. Acta Materialia, 1999. 47(15-16): p. 4009-<br />

4017.<br />

96.Pearson, D.H., C.C. Ahn, <strong>and</strong> B. Fultz, Measurements <strong>of</strong> 3d Occupancy from Cu<br />

L(2,3) Electron-Energy- Loss Spectra <strong>of</strong> Rapidly Quenched CuZr, CuTi, CuPd, CuPt,<br />

<strong>and</strong> CuAu. Physical Review B, 1994. 50(17): p. 12969-12972.<br />

97.Alber, U., H. Mullejans, <strong>and</strong> M. Ruhle, Bismuth segregation at <strong>copper</strong> grain<br />

boundaries. Acta Materialia, 1999. 47(15-16): p. 4047-4060.<br />

185


CHAPTER 11. LIST OF REFERENCES<br />

186<br />

98.Rez, P. <strong>and</strong> J.R. Alvarez, Calculation <strong>of</strong> cohesion <strong>and</strong> changes in electronic<br />

structure due to impurity segregation at boundaries in iron. Acta Materialia, 1999.<br />

47(15): p. 4069-4075.<br />

99.Keast, V.J. <strong>and</strong> D.B. Williams, Quantitative compositional mapping <strong>of</strong> Bi<br />

segregation to grain boundaries in Cu. Acta Materialia, 1999. 47(15-16): p. 3999-<br />

4008.<br />

100.Roy, A., U. Erb, <strong>and</strong> H. Gleiter, Grain boundary <strong>embrittlement</strong> as a function <strong>of</strong><br />

boundary structure (energy). Acta Metallurgica, 1982. 30(10): p. 1847-1850.<br />

101.Watanabe, T., Grain boundary Design for the Control <strong>of</strong> Intergranular Fracture.<br />

Materials Science Forum, 1989. 46: p. 25-48.<br />

102.Joshi, A. <strong>and</strong> D.F. Stein, An Auger Spectroscopic Analysis <strong>of</strong> Bismuth Segregated<br />

to Grain Boundaries in Copper. Journal <strong>of</strong> the Institute <strong>of</strong> Metals, 1971. 99(JUN): p.<br />

178-181.<br />

103.Powell, B.D. <strong>and</strong> H. Mykura, The segregation <strong>of</strong> bismuth to grain boundaries in<br />

<strong>copper</strong>-bismuth <strong>alloys</strong>. Acta Metallurgica, 1973. 21(8): p. 1151-1156.<br />

104.Marcus, H.L. <strong>and</strong> N.E. Paton, Metastable Grain-Boundary Segregation in<br />

Copper Containing Tellurium. Metallurgical Transactions, 1974. 5(10): p. 2135-<br />

2138.<br />

105.Menyhard, M., B. Blum, <strong>and</strong> C.J. McMahon Jr., Grain Boundary Segregation<br />

<strong>and</strong> Transformations in Bi-doped polycristalline Copper. Acta Metallurgica<br />

Materialia, 1989. 37(2): p. 549-557.<br />

106.Pineau, A., et al., Grain Boundary Cosegregation <strong>and</strong> Diffusion in Cu(Fe,S)<br />

Solid Solution. Acta Metallurgica, 1983. 31(7): p. 1047-1052.<br />

107.Westbrook, J.H., Segregation at Interfaces. Metallurgical Reviews, 1964. 9(36):<br />

p. 415-471.<br />

108.Nieh, T.G. <strong>and</strong> W.D. Nix, Embrittlement <strong>of</strong> Copper Due to Segregation <strong>of</strong><br />

Oxygen to Grain- Boundaries. Metallurgical Transactions a-Physical Metallurgy <strong>and</strong><br />

Materials Science, 1981. 12(5): p. 893-901.<br />

109.Benci, J.E., D.P. Pope, <strong>and</strong> J.L. Bassani, A combined experimental <strong>and</strong> analytical<br />

investigation <strong>of</strong> creep damage development in <strong>copper</strong>. Acta Metallurgica et<br />

Materialia, 1994. 42(1): p. 225-238.<br />

110.Tipler, H.R. <strong>and</strong> D. McLean, Influence <strong>of</strong> Interface Energy on Creep Rupture.<br />

Metal Science Journal, 1970. 4: p. 103-107.<br />

111.Briant, C.L. <strong>and</strong> R.P. Messmer. Chemical Bonding <strong>and</strong> Grain Boundary<br />

Embrittlement. in Embrittlement by Liquid Metals. 1982. Watervliet NY: TMS<br />

Warrendale, PA, publ. 1983.


112.Seah, M.P., Adsorption-induced interface decohesion. Acta Metallurgica, 1980.<br />

28(7): p. 955-962.<br />

113.Rice, J.R. <strong>and</strong> J.-C. Wang, Embrittlement <strong>of</strong> Interfaces by Solute Segregation.<br />

Materials Science <strong>and</strong> Engineering, 1989. A107: p. 23-40.<br />

114.Chatain, D., P. Wynblatt, <strong>and</strong> G.S. Rohrer, Anisotropic phenomena at interfaces<br />

in bismuth-saturated <strong>copper</strong>. Scripta Materialia, 2004. 50(5): p. 565-569.<br />

115.Chang, L.-S., et al., Thermodynamic aspects <strong>of</strong> the grain boundary segregation<br />

in Cu(Bi) <strong>alloys</strong>. Acta Materialia, 1999. 47(15-16): p. 4041-4046.<br />

116.Chang, L.-S., et al., Kinetic aspects <strong>of</strong> the grain boundary segregation in Cu(Bi)<br />

<strong>alloys</strong>. Acta Materialia, 1999. 47(10): p. 2951-2959.<br />

117.Saindrenan, G., R. Le Gall, <strong>and</strong> F. Christien, Endommagement interfacial des<br />

métaux. TECHNOSUP. 2002, Paris: Ellipses édition. 256.<br />

118.Hirth, J.P. <strong>and</strong> J.R. Rice, On the Thermodynamics <strong>of</strong> Adsorption at Interfaces as<br />

It Influences Decohesion. Metallurgical Transactions a-Physical Metallurgy <strong>and</strong><br />

Materials Science, 1980. 11(9): p. 1501-1511.<br />

119.Mishin, Y., P. S<strong>of</strong>ronis, <strong>and</strong> J.L. Bassani, Thermodynamic <strong>and</strong> kinetic aspects <strong>of</strong><br />

interfacial decohesion. Acta Materialia, 2002. 50(14): p. 3609-3622.<br />

120.Bika, D. <strong>and</strong> C.J. McMahon Jr., A Model for Dynamic Embrittlement. Acta<br />

Metallurgica et Materialia, 1995. 43(5): p. 1909-1916.<br />

121.McMahon, C.J., Brittle fracture <strong>of</strong> grain boundaries. Interface Science, 2004.<br />

12(2-3): p. 141-146.<br />

122.Pfaendtner, J.A., R. Muthiah, <strong>and</strong> C.J. McMahon Jr. Quasi-Static Brittle Failure.<br />

in Materials <strong>Research</strong> Society Symposium Procedings. 1999.<br />

123.Xu, Y. <strong>and</strong> J.L. Bassani, A steady-state model for diffusion-controlled fracture.<br />

Materials Science <strong>and</strong> Engineering A, 1999. 260(1-2): p. 48-54.<br />

124.Christien, F., R. Le Gall, <strong>and</strong> G. Saindrenan, Influence <strong>of</strong> stress on intergranular<br />

brittleness <strong>of</strong> a martensitic stainless steel, in Diffusion, Segregation <strong>and</strong> Stresses in<br />

Materials. 2003. p. 275-283.<br />

125.Misra, R.D.K., Issues concerning the effects <strong>of</strong> applied tensile stress on<br />

intergranular segregation in a low alloy steel. Acta Materialia, 1996. 44(3): p. 885-<br />

890.<br />

126.Kaur, I. <strong>and</strong> W. Gust, H<strong>and</strong>book <strong>of</strong> Grain <strong>and</strong> Interphase Boundary Diffusion<br />

Data. Vol. 1. 1989: Ziegler Press, Stuttgart.<br />

127.Misra, R.D.K. <strong>and</strong> B.P. Kashyap, An analysis <strong>of</strong> intergranular segregation <strong>of</strong><br />

sulphur in a low alloy steel by stress-induced diffusion process. Scripta Materialia,<br />

1996. 35(6): p. 755-760.<br />

187


CHAPTER 11. LIST OF REFERENCES<br />

188<br />

128.Nicholas, M.G. A Survey <strong>of</strong> Litterature on Liquid Metal Embrittlement <strong>of</strong> metals<br />

<strong>and</strong> Alloys. in Embrittlement by Liquid <strong>and</strong> Solid Metals. 1982. St.-Louis, Missouri:<br />

The Metallurgical Society <strong>of</strong> AIME.<br />

129.Rostoker, W., Liquid Metal Embrittlement, in Metals H<strong>and</strong>book, Desk Edition,<br />

H.E. Boyer <strong>and</strong> T.L. Gall, Editors. 1985, American Society for Metals: Metals Park,<br />

OH. p. 32.26-32.27.<br />

130.Roth, M.C., G.C. Weatherly, <strong>and</strong> W.A. Miller, The Role <strong>of</strong> Liquid Metal<br />

Embrittlement in Free-machining Alloys containing low-melting-point Inclusions.<br />

Canadian Metallurgical Quaterly, 1979. 18: p. 341-348.<br />

131.Westwood, A.R.C., C.M. Preece, <strong>and</strong> M.H. Kamdar, Adsorption-Induced Brittle<br />

Fracture in Liquid-Metal Environments, in Fracture An Advanced Treatise, H.<br />

Liebowitz, Editor. 1971, Academic Press: New York <strong>and</strong> London. p. 589-644.<br />

132.Nicholas, M.G. <strong>and</strong> C.F. Old, Review Liquid Metal Embrittlement. Journal <strong>of</strong><br />

Materials Science, 1979. 14: p. 1-18.<br />

133.Stol<strong>of</strong>f, N.S. Metal Induced Embrittlement - A Historical Perspective. in<br />

Embrittlement by Liquid <strong>and</strong> Solid Metals. 1982. St.-Louis, Missouri: The<br />

Metallurgical Society <strong>of</strong> AIME.<br />

134.Lynch, S.P. Mechanisms <strong>of</strong> Fracture in Liquid Metal Environments. in<br />

Embrittlement by Liquid <strong>and</strong> Solid Metals. 1982. St.-Louis, Missouri: The<br />

Metallurgical Society <strong>of</strong> AIME.<br />

135.Lynch, S.P., Environmentally Assisted Cracking: Overview <strong>of</strong> Evidence for an<br />

Adsorption-Induced Localised-Slip Process. Acta Metallurgica, 1988. 36(10): p.<br />

2639-2661.<br />

136.Fern<strong>and</strong>es, P.J.L., R.E. Clegg, <strong>and</strong> D.R.H. Jones, Failure by Liquid Metal<br />

Induced Embrittlement. Engineering Failure Analysis, 1994. 1(1): p. 51-63.<br />

137.Fern<strong>and</strong>es, P.J.L. <strong>and</strong> D.R.H. Jones, Mechanisms <strong>of</strong> liquid metal induced<br />

<strong>embrittlement</strong>. International Materials Reviews, 1997. 42(6): p. 251-261.<br />

138.Joseph, B., M. Picat, <strong>and</strong> F. Barbier, Liquid metal <strong>embrittlement</strong>: A state-<strong>of</strong>-theart<br />

appraisal. The European Physical Journal Applied physics, 1999. 5: p. 19-31.<br />

139.Kamdar, M.H., Liquid-Metal Embrittlement, in Corrosion. 2001, ASM<br />

International: Materials Park, OH, USA. p. 171-184.<br />

140.Kamdar, M.H., ed. Embrittlement by Liquid <strong>and</strong> Solid Metals. 1982, The<br />

Metallurgical Society <strong>of</strong> AIME: St.-Louis, Missouri.<br />

141.Shunk, F.A. <strong>and</strong> W.R. Warke, Specificity as an aspect <strong>of</strong> liquid metal<br />

<strong>embrittlement</strong>. Scripta Metallurgica, 1974. 8(5): p. 519-526.<br />

142.Fern<strong>and</strong>es, P.J.L. <strong>and</strong> D.R.H. Jones, Specificity in liquid metal induced<br />

<strong>embrittlement</strong>. Engineering Failure Analysis, 1996. 3(4): p. 299-302.


143.Fern<strong>and</strong>es, P.J.L. <strong>and</strong> D.R.H. Jones, The effects <strong>of</strong> <strong>microstructure</strong> on crack<br />

initiation in liquid-metal environments. Engineering Failure Analysis, 1997. 4(3): p.<br />

195-204.<br />

144.Auger, T., et al., Effect <strong>of</strong> contact conditions on <strong>embrittlement</strong> <strong>of</strong> T91 steel by<br />

lead-bismuth. Journal <strong>of</strong> Nuclear Materials, 2004. 335(2): p. 227-231.<br />

145.Perovic, D.D., G.C. Weatherly, <strong>and</strong> W.A. Miller, Influence <strong>of</strong> Temperature <strong>and</strong><br />

Strain-rate on Liquid Metal Embrittlement <strong>of</strong> an Aluminum Alloy containing Bismuth<br />

Inclusions. Acta Materialia, 1988. 36(8): p. 2249-2257.<br />

146.Glickman, E.E. Mechanism <strong>of</strong> Liquid Metal Embrittlement by Simple<br />

Experiments: from Atomics to Life-Time. in Multiscale Phenomena in Plasticity.<br />

2000. Ouranopolis, Greece: Kluwer Academic Publisher.<br />

147.Gordon, P., Metal-Induced-Embrittlement <strong>of</strong> Metals - An Evaluation <strong>of</strong><br />

Embrittler Transport Mechanisms. Metallurgical Transactions A, 1978. 9A: p. 267-<br />

273.<br />

148.Kargol, J.A. <strong>and</strong> D.L. Albright, The Effect <strong>of</strong> Relative Crystal Orientation on The<br />

Liquid Metal Induced Grain Boundary Fracture <strong>of</strong> Aluminum Bicrystals.<br />

Metallurgical Transactions A, 1977. 8A: p. 27-34.<br />

149.Hasson, G., et al., Theoretical <strong>and</strong> experimental determinations <strong>of</strong> grain<br />

boundary structures <strong>and</strong> energies: Correlation with various experimental results.<br />

Surface Science, 1972. 31: p. 115-137.<br />

150.Glickman, E.E., Grain boundary grooving accelerated by local plasticity as a<br />

possible mechanism <strong>of</strong> liquid metal <strong>embrittlement</strong>. Interface Science, 2003. 11(4): p.<br />

451-459.<br />

151.Wolski, K., et al., Liquid metal <strong>embrittlement</strong> studies on model systems with<br />

respect to the spallation target technology: The importance <strong>of</strong> nanometre-thick films.<br />

Journal de Physique IV, 2002. 12: p. Pr8-249-261.<br />

152.Robertson, W.M., Propagation <strong>of</strong> a Crack Filled with Liquid Metal. Transactions<br />

<strong>of</strong> the Metallurgical Society <strong>of</strong> AIME, 1966. 236: p. 1478-1482.<br />

153.Stol<strong>of</strong>f, N.S. <strong>and</strong> T.L. Johnston, Crack Propagation in a Liquid Metal<br />

Environment. Acta Materialia, 1963. 11(4): p. 251-256.<br />

154.Westwood, A.R. <strong>and</strong> M.H. Kamdar, Concerning Liquid Metal Embrittlement,<br />

Particularly <strong>of</strong> Zinc Monocrystals by Mercury. Philosophical Magazine, 1963. 8(89):<br />

p. 787-804.<br />

155.Lynch, S.P., Solid-metal-induced <strong>embrittlement</strong> <strong>of</strong> aluminium <strong>alloys</strong> <strong>and</strong> other<br />

materials. Materials Science <strong>and</strong> Engineering A, 1989. 108: p. 203-212.<br />

156.Glickman, E.E., Stress, surface energy <strong>and</strong> segregation effects in liquid metal<br />

<strong>embrittlement</strong>: Role <strong>of</strong> grain boundary grooving accelerated by local plasticity.<br />

Defect <strong>and</strong> Diffusion Forum, 2003. 216-217: p. 207-216.<br />

189


CHAPTER 11. LIST OF REFERENCES<br />

190<br />

157.Glickman, E.E., et al., Round table discussion II: Grain Boundary Wetting <strong>and</strong><br />

Liquid Grooving. Defect <strong>and</strong> Diffusion Forum, 1998. 156: p. 265-272.<br />

158.Pereiro-Lopez, E., et al., Grain boundary liquid metal wetting: A synchrotron<br />

micro-radiographic investigation. Nuclear Instruments <strong>and</strong> Methods in Physics<br />

<strong>Research</strong> Section B: Beam Interactions with Materials <strong>and</strong> Atoms, 2003. 200: p. 333-<br />

338.<br />

159.Pereiro-Lopez, E., et al., Liquid-metal wetting at grain boundaries: a<br />

synchrotron-micro- radiographic investigation, in Diffusion, Segregation <strong>and</strong><br />

Stresses in Materials. 2003. p. 241-248.<br />

160.Pereiro-Lopez, E., W. Ludwig, <strong>and</strong> D. Bellet, Discontinuous penetration <strong>of</strong><br />

liquid Ga into grain boundaries <strong>of</strong> Al polycrystals. Acta Materialia, 2004. 52(2): p.<br />

321-332.<br />

161.Ludwig, W., E. Pereiro-Lopez, <strong>and</strong> D. Bellet, In situ investigation <strong>of</strong> liquid Ga<br />

penetration in Al bicrystal grain boundaries: grain boundary wetting or liquid metal<br />

<strong>embrittlement</strong>? Acta Materialia, 2005. 53(1): p. 151-162.<br />

162.Ludwig, W. <strong>and</strong> D. Bellet, Penetration <strong>of</strong> liquid gallium into the grain<br />

boundaries <strong>of</strong> aluminium: a synchrotron radiation microtomographic investigation.<br />

Materials Science <strong>and</strong> Engineering A, 2000. 281(1-2): p. 198-203.<br />

163.Clarke, D.R. <strong>and</strong> M.L. Gee, Wetting <strong>of</strong> surfaces <strong>and</strong> grain boundaries, in<br />

Materials Interfaces Atomic-level structure <strong>and</strong> properties, D.Y. Wolf, S., Editor.<br />

1992, Chapman & Hall: London. p. 255-272.<br />

164.Hugo, R.C. <strong>and</strong> R.G. Hoagl<strong>and</strong>, In-Situ TEM Observation <strong>of</strong> Aluminum<br />

Embrittlement by Liquid Gallium. Scripta Materialia, 1998. 38(3): p. 523-529.<br />

165.Laporte, V. <strong>and</strong> K. Wolski. Liquid Bismuth Induced Intergranular Embrittlement<br />

<strong>of</strong> Polycrystalline Copper. in Environmental Degradation <strong>of</strong> Engineering Materials.<br />

2003. Bordeaux.<br />

166.Cahn, J.W., Critical-Point Wetting. Journal <strong>of</strong> Chemical Physics, 1977. 66(8): p.<br />

3667-3672.<br />

167.Straumal, B. <strong>and</strong> B. Baretzky, Grain boundary phase transitions <strong>and</strong> their<br />

influence on properties <strong>of</strong> polycrystals. Interface Science, 2004. 12(2-3): p. 147-155.<br />

168.Straumal, B.B., P. Zieba, <strong>and</strong> W. Gust, Grain boundary phase transitions <strong>and</strong><br />

phase diagrams. International Journal <strong>of</strong> Inorganic Materials, 2001. 3(8): p. 1113-<br />

1115.<br />

169.Marié, N., K. Wolski, <strong>and</strong> M. Biscondi, Grain boundary penetration <strong>of</strong> nickel by<br />

liquid bismuth as a film <strong>of</strong> nanometric thickness. Scripta Materialia, 2000. 43(10): p.<br />

943-949.


170.Marié, N., K. Wolski, <strong>and</strong> M. Biscondi, Intergranular penetration <strong>and</strong><br />

<strong>embrittlement</strong> <strong>of</strong> solid nickel through bismuth vapour condensation at 700°C. Journal<br />

<strong>of</strong> Nuclear Materials, 2001. 296(1-3): p. 282-288.<br />

171.Wolski, K., N. Marié, <strong>and</strong> M. Biscondi, AES quantification <strong>of</strong> intergranular film<br />

thickness in the Ni-Bi system with respect to the liquid metal <strong>embrittlement</strong><br />

phenomenon. Surface <strong>and</strong> Interface Analysis, 2001. 31: p. 280-286.<br />

172.Wolski, K., et al., About the Importance <strong>of</strong> Nanometer-Thick Intergranular<br />

Penetration in the Analysis ol Liquid Metal Embrittlement. Interface Science, 2001.<br />

9: p. 183-189.<br />

173.Joseph, B., et al., Rapid penetration <strong>of</strong> liquid Bi along Cu grain boundaries.<br />

Scripta Materialia, 1998. 39(6): p. 775-781.<br />

174.Joseph, B., F. Barbier, <strong>and</strong> M. Aucouturier, Embrittlement <strong>of</strong> <strong>copper</strong> by liquid<br />

bismuth. Scripta Materialia, 1999. 40(8): p. 893-897.<br />

175.Rabkin, E. Grain Boundary Embrittlement by Liquid Metals. in Multiscale<br />

Phenomena in Plasticity. 2000. Ouranopolis, Greece: Kluwer Academic Publisher.<br />

176.de Gennes, P.G., Wetting - Statics <strong>and</strong> Dynamics. Reviews <strong>of</strong> Modern Physics,<br />

1985. 57(3): p. 827-863.<br />

177.Rao, G., D.B. Zhang, <strong>and</strong> P. Wynblatt, A determination <strong>of</strong> interfacial energy <strong>and</strong><br />

interfacial composition in Cu-Pb <strong>and</strong> Cu-Pb-X <strong>alloys</strong> by solid state wetting<br />

measurements. Acta Metallurgica et Materialia, 1993. 41(11): p. 3331-3340.<br />

178.Moon, J., et al., Effects <strong>of</strong> concentration dependent diffusivity on the growth <strong>of</strong><br />

precursing films <strong>of</strong> Pb on Cu(111). Surface Science, 2001. 488(1-2): p. 73-82.<br />

179.Prevot, G., et al., Macroscopic <strong>and</strong> mesoscopic surface diffusion from a deposit<br />

formed by a Stranski-Krastanov type <strong>of</strong> growth: Pb on Cu(100) at above one layer <strong>of</strong><br />

coverage. Physical Review B, 2000. 61(15): p. 10393-10403.<br />

180.Iwamoto, C. <strong>and</strong> S.-I. Tanaka, Atomic morphology <strong>and</strong> chemical reactions <strong>of</strong> the<br />

reactive wetting front. Acta Materialia, 2002. 50(4): p. 749-755.<br />

181.Poirier, D.R. <strong>and</strong> G.H. Geiger, Transport Phenomena in Materials Processing.<br />

Second Edition ed. 1994, Warrendale, PA: TMS. 267.<br />

182.Chatain, D., V. Ghetta, <strong>and</strong> P. Wynblatt, Equilibrium shape <strong>of</strong> <strong>copper</strong> crystals<br />

grown on sapphire. Interface Science, 2004. 12(1): p. 7-18.<br />

183.Weck, E. <strong>and</strong> E. Leistner, Metallographic instructions for colour etching by<br />

immersion. Fachbuchserie Schweisstechnick. Vol. 77. 1995, Düsseldorf: Deutscher<br />

Verlag für Schweisstechnick. 60.<br />

184.Moore, K.I., K. Chattopadhyay, <strong>and</strong> B. Cantor, In situ Transmission Electron-<br />

Microscope Measurements <strong>of</strong> Solid Al-Solid Pb <strong>and</strong> Solid Al-Liquid Pb Surface-<br />

Energy Anisotropy in Rapidly Solidified Al-5-Percent-Pb (by Mass). Proceedings <strong>of</strong><br />

191


CHAPTER 11. LIST OF REFERENCES<br />

192<br />

the Royal Society <strong>of</strong> London Series a- Mathematical Physical <strong>and</strong> Engineering<br />

Sciences, 1987. 414(1847): p. 499-&.<br />

185.Moore, K.I., D.L. Zhang, <strong>and</strong> B. Cantor, Solidification <strong>of</strong> Pb particles embedded<br />

in Al. Acta Metallurgica et Materialia, 1990. 38(7): p. 1327-1342.<br />

186.Goswami, R. <strong>and</strong> K. Chattopadhyay, The superheating <strong>and</strong> the crystallography<br />

<strong>of</strong> embedded Pb particles in f.c.c. Al, Cu <strong>and</strong> Ni matrices. Acta Metallurgica et<br />

Materialia, 1995. 43(7): p. 2837-2847.<br />

187.Gibbs, G.B., Fluage Par Diffusion Dans Les Solides Polycristallins. Memoires<br />

Scientifiques De La Revue De Metallurgie, 1965. 62(10): p. 781-786.<br />

188.Pourbaix, M., Atlas <strong>of</strong> Electrochemical Equilibria in Aqueous Solutions. 2nd<br />

english edition ed. 1974, Houston, Texas, USA: NACE.<br />

189.MeX s<strong>of</strong>tware, Alicona Imaging GmbH. 2004: Graz.<br />

190.Gorman, J.W. <strong>and</strong> G.W. Preckshot, Molecular Diffusion <strong>and</strong> Interphase Transfer<br />

in the Solid Copper-Molten Lead System. Transactions <strong>of</strong> the American Institute <strong>of</strong><br />

Mining <strong>and</strong> Metallurgical Engineers, 1958. 212: p. 367-373.<br />

191.Nichols, F.A. <strong>and</strong> W.W. Mullins, Surface- (Interface-) <strong>and</strong> Volume-Diffusion<br />

Contributions to Morphological Changes Driven by Capillarity. Transactions <strong>of</strong> the<br />

Metallurgical Society <strong>of</strong> AIME, 1965. 233: p. 1840-1847.<br />

192.Predel, B., Al-Pb (Aluminum-Lead), in L<strong>and</strong>olt-Börnstein - Group IV Physical<br />

Chemistry. 1991, Springer-Verlag: Heidelberg.<br />

193.McLean, M., Kinetics <strong>of</strong> Spheroidization <strong>of</strong> Lead Inclusions in Aluminum.<br />

Philosophical Magazine, 1973. 27(6): p. 1253-1266.<br />

194.McCormick, M.A., E.B. Evans, <strong>and</strong> U. Erb, On the Shapes <strong>of</strong> Liquid-Lead<br />

Inclusions in Aluminum. Philosophical Magazine a-Physics <strong>of</strong> Condensed Matter<br />

Structure Defects <strong>and</strong> Mechanical Properties, 1986. 53(2): p. L27-L30.<br />

195.McLean, M. <strong>and</strong> M.S. Loveday, In-situ observations <strong>of</strong> the annealing <strong>of</strong> liquid<br />

lead inclusions entrained in an aluminium matrix. Journal <strong>of</strong> Materials Science,<br />

1974. 9: p. 1104-1114.<br />

196.Matthews, J.W., Misfit Dislocations, in Dislocations in Solids, F.R.N. Nabarro,<br />

Editor. 1979, North-Holl<strong>and</strong> Publ.: New York. p. 461-545.<br />

197.Kim, W.T. <strong>and</strong> B. Cantor, Solidification behaviour <strong>of</strong> Pb droplets embedded in a<br />

Cu matrix. Acta Metallurgica et Materialia, 1992. 40(12): p. 3339-3347.<br />

198.Goswami, R., K. Chattopadhyay, <strong>and</strong> P.L. Ryder, Melting <strong>of</strong> Pb particles<br />

embedded in Cu-Zn matrices. Acta Materialia, 1998. 46(12): p. 4257-4271.<br />

199.Bonzel, H.P., Surface morphologies: Transient <strong>and</strong> equilibrium shapes. Interface<br />

Science, 2001. 9(1-2): p. 21-34.


200.Wolf, D., Correlation between Structure, Energy, <strong>and</strong> Ideal Cleavage Fracture<br />

for Symmetrical Grain-Boundaries in Fcc Metals. Journal <strong>of</strong> Materials <strong>Research</strong>,<br />

1990. 5(8): p. 1708-1730.<br />

201.Protsenko, P., et al., Misorientation effects on grain boundary grooving <strong>of</strong> Ni by<br />

liquid Ag, in Diffusion, Segregation <strong>and</strong> Stresses in Materials. 2003. p. 225-230.<br />

202.Chatain, D., C. Vahlas, <strong>and</strong> N. Eustathopoulos, Etude des tensions interfaciales<br />

liquide-liquide et solide-liquide dans les systemes à monotectique Zn-Pb et Zn-Pb-<br />

Sn. Acta Metallurgica, 1984. 32(2): p. 227-234.<br />

203.Miura, H., M. Kato, <strong>and</strong> T. Mori, Temperature-Dependence <strong>of</strong> the Energy <strong>of</strong> Cu<br />

[110] Symmetrical Tilt Grain-Boundaries. Journal <strong>of</strong> Materials Science Letters,<br />

1994. 13(1): p. 46-48.<br />

204.Clyne, T.W. <strong>and</strong> P.J. Withers, An Introduction to Metal Matrix Composites.<br />

Cambridge Solid State Science Series, ed. E.A. Davis <strong>and</strong> L.M. Ward. 1993,<br />

Cambridge, U.K.: Cambridge University Press. 509 pp.<br />

205.Herring, C., Surface Tension as a Motivation for Sintering, in The Physics <strong>of</strong><br />

Powder Metallurgy, W.E. Kingston, Editor. 1951, McGraw-Hill: New-York.<br />

206.Iida, T. <strong>and</strong> R.I.L. Guthrie, The Physical Properties <strong>of</strong> Liquid Metals. 1993,<br />

Oxford, U.K.: Oxford Science Publications, Clarendon Press. 288.<br />

207.Simmons, G. <strong>and</strong> H. Wang, Single Crystal Elastic Constants <strong>and</strong> Calculated<br />

Aggregate Properties: A HANDBOOK. 1971, Cambridge, MA: MIT Press.<br />

208.Miura, H. <strong>and</strong> T. Sakai, Shape change <strong>of</strong> liquid B2O3 particles dispersed in<br />

<strong>copper</strong> crystals during high-temperature deformation. Materials Science <strong>and</strong><br />

Engineering A, 2001. 319-321: p. 756-759.<br />

209.Courtney, T.H., Mechanical Behavior <strong>of</strong> Materials. Mc Graw-Hill Series in<br />

Materials Science <strong>and</strong> Engineering, ed. M.B. Bever <strong>and</strong> C.A. Wert. 1990, New York,<br />

NY: Mc Graw-Hill. 502-561.<br />

210.Gamaoun, F., et al., Cavity formation <strong>and</strong> accelerated plastic strain in T91 steel<br />

in contact with liquid lead. Scripta Materialia, 2004. 50(5): p. 619-623.<br />

211.Perovic, D.D., W.A. Miller, <strong>and</strong> G.C. Weatherly, On the solidification <strong>of</strong> bismuth<br />

inclusions in stressed aluminum. Scripta Metallurgica, 1987. 21(5): p. 701-703.<br />

212.Agullo, E., et al., Influence <strong>of</strong> an ultrasonic field on lead electrodeposition on<br />

<strong>copper</strong> using a fluoroboric bath. New Journal <strong>of</strong> Chemistry, 1999. 23(1): p. 95-101.<br />

213.Lowenheim, F.A., Electroplating, ed. J. Robison <strong>and</strong> R. Weine. 1978, New York:<br />

McGraw-Hill. 594.<br />

214.Zhao, J.-C. <strong>and</strong> M.R. Notis, Spinodal decomposition, ordering transformation,<br />

<strong>and</strong> discontinuous precipitation in a Cu-15Ni-8Sn alloy. Acta Materialia, 1998.<br />

46(12): p. 4203-4218.<br />

193


CHAPTER 11. LIST OF REFERENCES<br />

194<br />

215.Kocks, U.F., The Relation Between Polycrystal Deformation <strong>and</strong> Single-Crystal<br />

Deformation. Metallurgical Transactions, 1970. 1(5): p. 1121.<br />

216.Beckman, J.P. <strong>and</strong> D.A. Woodford, Gas Phase Embrittlement by Metal Vapors.<br />

Metallurgical Transactions A, 1989. 20A: p. 184-187.<br />

217.Lagarde, P. <strong>and</strong> M. Biscondi, Fluage intergranulaire de bicristaux orientés de<br />

cuivre. Memoires Scientifiques De La Revue De Metallurgie, 1974. 71(2): p. 121-<br />

131.<br />

218.Bleakney, H.H., The Creep-Rupture Embrittlement <strong>of</strong> Copper. Canadian<br />

Metallurgical Quarterly, 1965. 4(1): p. 13-29.<br />

219.Benson, B.A. <strong>and</strong> R.G. Hoagl<strong>and</strong>, Crack growth behavior <strong>of</strong> a high strength<br />

aluminum alloy during LME by gallium. Scripta Metallurgica, 1989. 23(11): p. 1943-<br />

1948.<br />

220.Watanabe, T., M. Tanaka, <strong>and</strong> S. Karashima. Liquid Metal Gallium-Induced<br />

Intergranular Frcture <strong>of</strong> Aluminium Bicrystals. in Embrittlement by Liquid <strong>and</strong> Solid<br />

Metals. 1982. St.-Louis, Missouri: The Metallurgical Society <strong>of</strong> AIME.<br />

221.Biselli, C. <strong>and</strong> P. Isler, Neue Kupferlegierungen für Mechanik, Elektrik und<br />

Elektronik. Metall, 2000. 54(4): p. 180-183.<br />

222.Plewes, J.T., High-Strength Cu-Ni-Sn Alloys by Thermomechanical Processing.<br />

Metallurgical Transactions, 1975. A 6(3): p. 537-544.<br />

223.Cribb, W.R. <strong>and</strong> J.O. Ratka, Copper Spinodal Alloys. Advanced Materials <strong>and</strong><br />

Processes, 2002. 160(11): p. 1-4.<br />

224.Hermann, P. <strong>and</strong> D.G. Morris, Relationship between Microstructure <strong>and</strong><br />

Mechanical Properties <strong>of</strong> a Spinodally Decomposing Cu-15Ni-8Sn Alloy Prepared by<br />

Spray Deposition. Metallurgical Transactions A, 1994. 25: p. 1403-1412.<br />

225.Boillat, S., List <strong>of</strong> products. 2002, UMS Usines Métallurgiques Suisse SA, Usine<br />

Boillat: Reconvilier.<br />

226.Touloukian, Y.S., et al., Thermal Conductivity: Metallic Elements <strong>and</strong> Alloys.<br />

Thermophysical Properties <strong>of</strong> Matter, ed. Y.S. Touloukian. Vol. 1. 1970, New York -<br />

Washington: IFI/Plenum.<br />

227.Zinov'ev, V.E., H<strong>and</strong>book <strong>of</strong> thermophysical properties <strong>of</strong> metals at high<br />

temperatures. 1996, New York: Nova Science Publishers. 581.<br />

228.Ledbetter, H.M. <strong>and</strong> E.R. Naimon, Elastic Properties <strong>of</strong> Metals <strong>and</strong> Alloys.<br />

II. Copper. Journal <strong>of</strong> Physical <strong>and</strong> Chemical Reference Data, 1974. 3(4): p. 897-929.<br />

229.Davis, J.R., et al., eds. ASM H<strong>and</strong>book Volume 04: Heat Treating. 1991, ASM<br />

International, Materials Park, OH.


230.Schwartz, L.H., S. Mahajan, <strong>and</strong> J.T. Plewes, Spinodal decomposition in a Cu-9<br />

wt% Ni-6 wt% Sn alloy. Acta Metallurgica, 1974. 22(5): p. 601-609.<br />

231.Zhao, J.-C. <strong>and</strong> M.R. Notis, Microstructure <strong>and</strong> precipitation kinetics in a Cu-<br />

7.5Ni-5Sn alloy. Scripta Materialia, 1998. 39(11): p. 1509-1516.<br />

195


Laurent FELBERBAUM<br />

Av. des Tilleuls 24 né le 5 juin 1975<br />

1203 Genève suisse<br />

Tél. privé 022 344 24 57 marié, un enfant<br />

Tél. pr<strong>of</strong>. 021 693 68 02<br />

laurent.felberbaum@epfl.ch<br />

EXPÉRIENCE<br />

PROFESSIONNELLE<br />

C<strong>and</strong>idat au Doctorat<br />

2000 à ce jour<br />

Laboratoire de Métallurgie Mécanique,<br />

Ecole Polytechnique Fédérale de Lausanne,<br />

EPFL<br />

Projet en collaboration avec Swissmetal<br />

Mécanismes de fragilisation et <strong>microstructure</strong>s d’alliages de cuivre au plomb<br />

Ingénieur en Matériaux Laboratoires de Métallurgie Physique et Métallurgie<br />

1998-1999 Mécanique, EPFL<br />

Projet en collaboration avec ABB,<br />

Calcom,<br />

EMPA,<br />

SR Technics et Sulzer<br />

Essais mécaniques à haute température sur des superalliages réparés par<br />

soudage laser ou brasage.<br />

Enseignement<br />

Section des Matériaux, EPFL<br />

1998-2004 Animation de nombreux travaux pratiques en Métallurgie.<br />

Conduite de deux projets de diplôme (4 mois).<br />

Direction de séances d’exercices ”Déformation et Rupture” (niveau master).<br />

1998<br />

FORMATION<br />

Diplôme d’Ingénieur en<br />

Département des matériaux, EPFL<br />

Sciences des Matériaux 3ème<br />

année en échange à l’ETH Zurich<br />

1993 Maturité type A Gymnase Cantonal du Bugnon, VD<br />

Baccalauréat Latin-Grec-Mathématiques<br />

LOISIRS<br />

Montagne VTT, r<strong>and</strong>onnées en cabane, ski et snowboard<br />

Sports d’équipe Volleyball et surtout Football

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!