24.08.2013 Views

dissertation global and local fracture properties of metal matrix ...

dissertation global and local fracture properties of metal matrix ...

dissertation global and local fracture properties of metal matrix ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

MONTANUNIVERSITÄT MONTANUNIVERSITÄT LE LEOBEN LE OBEN<br />

DISSERTATION<br />

GLOBAL AND LOCAL FRACTURE<br />

PROPERTIES OF METAL MATRIX<br />

COMPOSITES<br />

by Dipl. Ing. Ilchat Sabirov<br />

LEOBEN, 2004


Acknowledgements<br />

First <strong>of</strong> all, I would like to acknowledge gratefully Univ. Doz. Dr. Otmar Kolednik for being a<br />

patient supervisor <strong>and</strong> for supporting this work with new ideas. I could not have imagined<br />

having a better advisor <strong>and</strong> mentor for my PhD, <strong>and</strong>, without his common-sense, knowledge,<br />

kind participation in my work, <strong>and</strong> support <strong>of</strong> my research activity, this thesis would not<br />

appear.<br />

I would like to express my deep gratitude to Univ. Doz. Dr. Reinhard Pippan for his<br />

discussions <strong>and</strong> kind help during execution <strong>of</strong> this work. His helpful comments <strong>and</strong> notes are<br />

highly appreciated.<br />

I would like to thank O. Univ. Pr<strong>of</strong>. Dr. Robert Danzer from the Department <strong>of</strong> Structural <strong>and</strong><br />

Functional Ceramics for his kind assent to examine my PhD.<br />

I am indebted to all my colleagues at the Erich Schmid Institute <strong>of</strong> Materials Science <strong>of</strong><br />

Austrtian Academy <strong>of</strong> Sciences for a friendly atmosphere. My heartfelt thanks to D.I. Klaus<br />

Unterweger, Dr. Cristian Motz, Dr. Ernst Gach, Dr. Hans-Peter Brantner, D.I. Herbert<br />

Kreuzer, Dr. Rene Wadsack, D.I. Andreas Vorhauer, D.I Alois Maier, D.I. Gernot Trattnig,<br />

Mag. Günter Maier, etc., for their kind help in any matter. Special thanks to Franz Hubner,<br />

Günter Aschauer, Edeltraud Haberz, <strong>and</strong> Gabriele Moser for the excellent preparation <strong>of</strong><br />

specimens. I apologize to all colleagues whose names have been unintentionally omitted.<br />

I also wish to thank Univ. Doz. Dr. Heinz Pettermann <strong>and</strong> Dr. Dominik Duschlbauer from<br />

Vienna University <strong>of</strong> Technology for their cooperation during execution <strong>of</strong> this work.<br />

Finally, I would like to express my thanks to Pr<strong>of</strong>. Ruslan Valiev from the Ufa Aviation<br />

University for his encouragement in my research activity.<br />

1<br />

Leoben, April 2004.


TABLE OF CONTENTS<br />

1. Introduction 5<br />

2. Background 7<br />

2.1. Metal <strong>matrix</strong> composites as advanced materials 7<br />

2.2. The influence <strong>of</strong> the <strong>global</strong> microstructure <strong>and</strong> the <strong>matrix</strong> condition<br />

on the <strong>global</strong> mechanical <strong>properties</strong> in <strong>metal</strong> <strong>matrix</strong> composites 8<br />

2.3. The process <strong>of</strong> void initiation in materials with particles 9<br />

3. A method to estimate the <strong>local</strong> conditions for void initiation in materials with<br />

inclusions 15<br />

3.1. An automatic <strong>fracture</strong> surface analysis system 15<br />

3.2. The determination <strong>of</strong> the CODi- <strong>and</strong> CODvi–values 16<br />

3.3. Estimate <strong>of</strong> the maximum principal stresses in both phases 20<br />

3.3.1. Estimate <strong>of</strong> the HRR stress tensor at the moment <strong>of</strong> void initation 20<br />

3.3.2. The determination <strong>of</strong> the maximum stresses both in the particle <strong>and</strong> in the<br />

<strong>matrix</strong> by a Mori-Tanaka type mean-field correction 22<br />

3.3.2.1. The Eschelby approach 22<br />

3.3.2.2. Some general mean-field relations 24<br />

3.3.2.3. Solution procedure 25<br />

4. Local <strong>fracture</strong> <strong>properties</strong> <strong>of</strong> <strong>metal</strong> <strong>matrix</strong> composites 29<br />

4.1. Materials for investigation <strong>and</strong> their mechanical <strong>properties</strong> 29<br />

4.1.1. Cast MMCs 29<br />

4.1.1.1. Materials characterization 29<br />

4.1.1.2. Tensile tests 30<br />

4.1.1.3. Fracture mechanics tests 32<br />

4.1.2. Powder <strong>metal</strong>lurgy MMC 34<br />

4.1.2.1. Materials characterization 34<br />

4.1.2.2. Compression tests 35<br />

4.1.2.3. Fracture mechanics tests 37<br />

4.1.3. The relation between the mechanical <strong>properties</strong> 37<br />

2


4.2. The effect <strong>of</strong> the heat treatment on the <strong>fracture</strong> surface 40<br />

4.3. The effect <strong>of</strong> the particle location with respect to the crack tip on the <strong>local</strong><br />

<strong>fracture</strong> <strong>properties</strong> 42<br />

4.3.1. The relation between the particle location with respect to the crack tip<br />

<strong>and</strong> the CODi-values 42<br />

4.3.2. The relation between the particle location with respect to the crack tip<br />

<strong>and</strong> the CODvi-values 51<br />

4.4. The maximum particle stresses at the moment <strong>of</strong> void initiation 56<br />

4.5. The relation between the <strong>local</strong> conditions for void initiation <strong>and</strong> <strong>global</strong> mechanical<br />

<strong>properties</strong> 64<br />

4.6. Application <strong>of</strong> Weibull distribution for <strong>fracture</strong>d particles 66<br />

5. The relation between the inclusion size <strong>and</strong> the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> 69<br />

5.1. Material <strong>and</strong> mechanical <strong>properties</strong> 69<br />

5.1.1. Materials characterization 69<br />

5.1.2. Tensile tests 70<br />

5.1.3. Fracture mechanics tests 70<br />

5.2. The effect <strong>of</strong> the particle location <strong>and</strong> the inclusion size on the<br />

CODi- <strong>and</strong> CODvi-values 71<br />

6. Comparison <strong>of</strong> the <strong>local</strong> <strong>and</strong> <strong>global</strong> <strong>fracture</strong> <strong>properties</strong> 75<br />

7. Application <strong>of</strong> ECAP to improve the homogeneity <strong>of</strong> the particle distribution in<br />

MMCs 79<br />

7.1. Material 79<br />

7.2. A model by Tan <strong>and</strong> Zhang for the case <strong>of</strong> combination <strong>of</strong> extrusion<br />

<strong>and</strong> ECAP 80<br />

7.3. A quadrat method to estimate the homogeneity <strong>of</strong> the particle distribution<br />

in MMCs 83<br />

7.4. The effect <strong>of</strong> ECAP on the particle distribution, the <strong>global</strong>, <strong>and</strong> <strong>local</strong><br />

<strong>fracture</strong> <strong>properties</strong> in MMCs with severe clustered particle distribution 85<br />

7.4.1. The effect <strong>of</strong> ECAP on the microstructure 85<br />

7.4.2. The effect <strong>of</strong> ECAP on the <strong>global</strong> <strong>fracture</strong> <strong>properties</strong> 88<br />

7.4.3. The effect <strong>of</strong> ECAP on the <strong>fracture</strong> surface morphology 90<br />

3


7.4.4. The effect <strong>of</strong> ECAP on the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> 92<br />

7.5. About possible application <strong>of</strong> ECAP to improve the particle distribution<br />

in different MMCs 96<br />

8. Summary 98<br />

9. Appendix 101<br />

References 155<br />

4


1. Introduction<br />

Section 1<br />

In spite <strong>of</strong> the attractiveness <strong>of</strong> <strong>metal</strong> <strong>matrix</strong> composites (MMCs) because <strong>of</strong> their advantages,<br />

such as increased strength, stiffness, <strong>and</strong> modulus, their low <strong>fracture</strong> toughness in comparison<br />

with the <strong>fracture</strong> toughness <strong>of</strong> the <strong>matrix</strong> material is a primary obstacle against their<br />

widespread application in materials engineering. Many investigations have been already<br />

devoted to the mechanical behavior <strong>of</strong> MMCs <strong>and</strong> its relation to the microstructure, but both<br />

the microstructure <strong>and</strong> the <strong>properties</strong> have been considered at the <strong>global</strong> level: the variation <strong>of</strong><br />

strength <strong>and</strong> <strong>fracture</strong> toughness, measured by conventional tensile <strong>and</strong> <strong>fracture</strong> mechanics<br />

tests, have been studied as a function <strong>of</strong> parameters, such as the particle volume fraction, the<br />

average particle size, the particle shape, the mean particle orientation with respect to the crack<br />

plane, etc. [1-5].<br />

In comparison to previous investigations, the idea <strong>of</strong> this work is to study experimentally the<br />

<strong>local</strong> <strong>fracture</strong> <strong>properties</strong>: the conditions for void <strong>and</strong> <strong>fracture</strong> initiation. The motivation to<br />

study it is explained in Section 3.<br />

A method to estimate the <strong>local</strong> conditions for void initiation near the crack tip <strong>and</strong> the position<br />

<strong>of</strong> the particle location with respect to the crack tip is developed. The main tool <strong>of</strong> this method<br />

is a quantitative analysis <strong>of</strong> the <strong>fracture</strong> surfaces via automated stereophotogrammetry. By this<br />

tool, the <strong>local</strong> crack tip opening displacement at the moment <strong>of</strong> <strong>fracture</strong> initiation, CODi, (a<br />

measure <strong>of</strong> the <strong>local</strong> <strong>fracture</strong> initiation toughness) <strong>and</strong> the crack tip opening displacement at<br />

the moment <strong>of</strong> the void initiation, CODvi, (which can be considered as a <strong>local</strong> void initiation<br />

toughness) are determined. From the <strong>local</strong> void initiation toughness, the maximum normal<br />

stresses in the particles <strong>and</strong> at the interface at the moment <strong>of</strong> void initiation are evaluated by<br />

means <strong>of</strong> the HRR-field theory, a Mori-Tanaka type mean field analysis, <strong>and</strong> according to the<br />

model by Argon et al. This method is described in details in Section 4. The method is applied<br />

to Al6061 based MMCs. The results <strong>of</strong> investigation are presented <strong>and</strong> discussed in Sections 5<br />

<strong>and</strong> 6. The procedure is also applied to MnS-inclusions in the mild steel St37 (Section 6).<br />

Another topic <strong>of</strong> our investigations is an improvement <strong>of</strong> the homogeneity <strong>of</strong> particle<br />

distribution in MMCs reinforced by small (less than 10 µm) particles. As known, these<br />

materials are prone to cause an inhomogeneous particle distribution. A method for equal<br />

channel angular pressing (ECAP) is proposed to solve this problem. The effect <strong>of</strong> ECAP on<br />

the homogeneity <strong>of</strong> the particle distribution in an Al6061-20%Al2O3 MMC processed by<br />

powder <strong>metal</strong>lurgy (PM) technology on the <strong>global</strong> <strong>and</strong> <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> near the crack<br />

tip is investigated by automated stereophotogrammetry (Section 7).<br />

5


Section 1<br />

The main results <strong>of</strong> this investigation are following: the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> are<br />

influenced by the particle location with respect to the crack tip, the particle size, the material<br />

strength. A correlation between the <strong>local</strong> <strong>and</strong> <strong>global</strong> <strong>fracture</strong> <strong>properties</strong> is found.<br />

It should be noted that this work a part <strong>of</strong> the research devoted to the <strong>local</strong> deformation <strong>and</strong><br />

<strong>fracture</strong> <strong>properties</strong> <strong>of</strong> MMCs performed in frames <strong>of</strong> the FWF-project P14333-PHY.<br />

6


2. Background<br />

Section 2<br />

2.1. Metal <strong>matrix</strong> composites as advanced materials<br />

The development <strong>of</strong> MMCs for structural application is relatively new; it started<br />

approximately 20 years ago. The objectives <strong>of</strong> these developments are usually improvement<br />

<strong>of</strong> the performance <strong>of</strong> components, weight reduction, <strong>and</strong> replacement <strong>of</strong> more expensive<br />

conventional alloys. A vast amount <strong>of</strong> publications devoted to these problems prove the<br />

importance <strong>of</strong> composite materials in modern technology.<br />

Currently produced <strong>and</strong> practically applied are composite materials based on light <strong>metal</strong><br />

alloys, especially on aluminum, magnesium, <strong>and</strong> titanium, as well as on copper <strong>and</strong> high<br />

temperature superalloys reinforced by stable ceramic dispersion particles. Composite<br />

materials based on light <strong>metal</strong> alloys are reinforced with dispersion particles, platelets, short<br />

<strong>and</strong> continuous fibers <strong>and</strong> have very good mechanical <strong>properties</strong>. For instance, the MMCs on<br />

magnesium basis reinforced by SiC particles have mechanical <strong>properties</strong> which are by 30-<br />

40% better in comparison with the unreinforced magnesium alloys. Such MMCs are applied<br />

in the aircraft <strong>and</strong> car technology (space, satellite, <strong>and</strong> antenna structures) [1]. The advantages<br />

<strong>of</strong> aluminum composite materials, such as a high resistance to wear, a good strength, <strong>and</strong> a<br />

relatively good heat conduction, has resulted in their practical application in the car industry<br />

(discs <strong>of</strong> car brakes, Cardan shafts in passenger cars) [6], <strong>and</strong> in aerospace engineering<br />

(helicopter structures, jet engine fan blades) [1]. Titanium based composites have found a<br />

wide application as materials for high temperature structures [6]. Copper based composite<br />

materials, having a good heat conductivity, a good wear resistance, <strong>and</strong> an ability to join with<br />

<strong>metal</strong> alloys or special ceramics, are used in electronics as electrical contacts, resistance<br />

welding electrodes, elements <strong>of</strong> electronic system [6]. Nickel based superalloys have better<br />

high temperature mechanical <strong>properties</strong> than conventional high temperature alloys. Therefore,<br />

they are applied in jet engine technology <strong>and</strong> industrial turbines (as high temperature engine<br />

components, rotors for compressions, etc.) [1, 6].<br />

The industrial application <strong>of</strong> MMCs has been developing, <strong>and</strong> new composite materials have<br />

been <strong>of</strong>fered to industry by researchers. It is clear that detailed investigations devoted to the<br />

mechanical <strong>and</strong> physical <strong>properties</strong> <strong>of</strong> these materials must precede their industrial<br />

application. In the next section, a short review <strong>of</strong> relations between the composite architecture<br />

<strong>and</strong> mechanical <strong>properties</strong> is given.<br />

7


Section 2<br />

2.2. The influence <strong>of</strong> the <strong>global</strong> microstructure <strong>and</strong> the <strong>matrix</strong> condition on the <strong>global</strong><br />

mechanical <strong>properties</strong> in <strong>metal</strong> <strong>matrix</strong> composites<br />

The influence <strong>of</strong> the <strong>global</strong> microstructural parameters such as the particle volume fraction,<br />

the average particle size, etc., on the <strong>global</strong> mechanical <strong>properties</strong> has been already<br />

investigated in detail. The experimental results can be roughly generalized as follows [7]:<br />

With increasing particle volume fraction, the yield strength <strong>and</strong> the ultimate tensile strength<br />

increase, the ductility <strong>and</strong> the <strong>fracture</strong> toughness decrease, e.g., [8]. For a constant particle<br />

volume fraction, the tensile strength tends to increase with decreasing particle size, e.g.,<br />

[9,10]. No such simple pictures have been found for the more complex <strong>properties</strong> <strong>fracture</strong><br />

toughness <strong>and</strong> fatigue resistance; ambiguous results have been reported here [7, 11-12].<br />

Stronger <strong>matrix</strong> alloys produce stronger composites, e. g. [13,14]; the increase in strength due<br />

to the reinforcement decreases with increasing <strong>matrix</strong> strength; in the case <strong>of</strong> very high<br />

strength alloys, reinforcements may even lead to a reduction in strength [15]. Therefore, one<br />

<strong>of</strong> most important factors affecting the mechanical <strong>properties</strong> <strong>of</strong> MMCs is the heat treatment.<br />

In [16], the effect <strong>of</strong> the heat treatment on mechanical <strong>properties</strong> <strong>of</strong> the PM-MMC Al7093-<br />

15%SiC was investigated. An increase <strong>of</strong> the yield strength with increasing aging time up to<br />

the peak aged condition was found. A further increase <strong>of</strong> aging time led to a slight decrease <strong>of</strong><br />

the yield strength. The dependence <strong>of</strong> the <strong>fracture</strong> strain <strong>and</strong> the strain hardening coefficient<br />

on the aging time demonstrated an opposite character. A good correlation between the<br />

<strong>fracture</strong> strain <strong>and</strong> the strain hardening coefficient was found. It was shown that the <strong>fracture</strong><br />

toughness has an inverse dependence on strength <strong>and</strong> correlates well to the strain hardening<br />

coefficient. Fracture surface inspection revealed the dominance <strong>of</strong> the particle <strong>fracture</strong> for the<br />

solution treated, under-aged, peak-aged, <strong>and</strong> slightly over-aged conditions <strong>of</strong> the <strong>matrix</strong><br />

whereas, in the highly overaged condition, (near-)interface debonding failure dominated.<br />

In [17], a model was proposed to predict the <strong>fracture</strong> toughness in this material. This model is<br />

based on a prediction proposed by Hahn <strong>and</strong> Rosenfield [18],<br />

8<br />

1<br />

2<br />

1 ⎡<br />

⎤ 1<br />

3 −<br />

⎢ ⎛π<br />

⎞<br />

6<br />

= 2σ<br />

E⎜<br />

⎟ d⎥<br />

f . (2.1)<br />

⎢ ⎝ 6 ⎠ ⎥<br />

⎣<br />

⎦<br />

K IC<br />

y<br />

In Eq. 2.1 σy is the yield strength <strong>of</strong> material, E the Young modulus, d the particle size, <strong>and</strong> f<br />

the particle volume fraction. It was shown in [17] that Eq. 2.1 significantly overestimates the


Section 2<br />

<strong>fracture</strong> toughness data. Using the analyses <strong>of</strong> Rice <strong>and</strong> Johnson [19] <strong>and</strong> Shih [20], P<strong>and</strong>ey et<br />

al. proposed a modified interpretation <strong>of</strong> the Hahn <strong>and</strong> Rosenfield model<br />

K<br />

c<br />

9<br />

1<br />

2<br />

⎡ βσ y Ed ⎤<br />

= 0.<br />

77 ⋅ ⎢<br />

2 ⎥ f<br />

⎣d<br />

n ( 1−<br />

v ) ⎦<br />

1<br />

−<br />

6<br />

, (2.2)<br />

where β = 0.5, dn is a dimensionless constant depending on the strain hardening coefficient<br />

[20], <strong>and</strong> v is the Poisson ratio. This prediction has provided a better agreement with <strong>fracture</strong><br />

toughness data for this material.<br />

In [21], the effect <strong>of</strong> the aging process on the <strong>properties</strong> <strong>of</strong> an Al6061 alloy reinforced by 20%<br />

<strong>of</strong> Al2O3 particles was studied. A similar dependence <strong>of</strong> mechanical <strong>properties</strong> on the heat<br />

treatment condition was revealed as in [16]. A shift from particle <strong>fracture</strong> to an interface (or<br />

near-interface) decohesion when moving from the underaged to the overaged regimes, <strong>and</strong><br />

higher <strong>local</strong> volume fraction <strong>of</strong> spinel products at the particle/<strong>matrix</strong> debonded interface <strong>of</strong><br />

overaged specimens were observed.<br />

The plastic deformation behavior <strong>of</strong> MMCs was investigated in [22]. It was shown that, in an<br />

Al6061-SiC MMC in under-aged condition, plastic strain is distributed homogeneously<br />

throughout the whole length <strong>of</strong> the specimen, whereas strain <strong>local</strong>ization is noted for the<br />

MMC in the peak-aged condition [22]. This difference is also reflected in the damage<br />

mechanism. In the under-aged condition, the particle cracking increases with strain over the<br />

whole length <strong>of</strong> the specimen, <strong>and</strong> the density <strong>of</strong> cracked particles near the <strong>fracture</strong> surface is<br />

comparable to that in the rest <strong>of</strong> the specimen. In the peak-aged condition, the extent <strong>of</strong> the<br />

particle cracking is less than in the under-aged condition <strong>and</strong> tends to be highest near the<br />

<strong>fracture</strong> surface. The percentage <strong>of</strong> cracked particles is less than about 5% <strong>of</strong> the total particle<br />

content, <strong>and</strong> experiments show that the <strong>fracture</strong> strain can be recovered after pre-strain levels<br />

which produce extensive particle <strong>fracture</strong>. The <strong>fracture</strong> strain is very sensitive to heat<br />

treatment, decreasing with increasing strength <strong>and</strong> decreasing strain hardening coefficient, so<br />

the <strong>matrix</strong> condition is very important in the <strong>fracture</strong> process [22].<br />

2.3. The problem <strong>of</strong> the homogeneity <strong>of</strong> the particle distribution <strong>and</strong> its influence on the<br />

mechanical <strong>properties</strong><br />

A poor reinforcement distribution restricts the application <strong>of</strong> MMCs in industry due to the<br />

degraded mechanical <strong>properties</strong> <strong>and</strong> limited formability. As well known, in order to gain a


Section 2<br />

high strength, fine reinforcements <strong>and</strong> a relatively large particle volume fraction are preferred<br />

[2]. However, it is difficult to take advantage <strong>of</strong> both <strong>of</strong> these requirements because they are<br />

prone to cause an inhomogeneous particle distribution [23].<br />

A number <strong>of</strong> studies have shown that during the deformation <strong>of</strong> particulate reinforced MMCs<br />

there is a progressive development <strong>of</strong> damage within the material. It has been observed that<br />

the damage tends to originate preferentially in particle clustered regions [24, 25]. Thus, one <strong>of</strong><br />

the most important microstructural features <strong>of</strong> the MMCs is the spatial distribution <strong>of</strong> the<br />

particles. In [26, 27], an Eshelby based approach was used to create a model <strong>of</strong> a clustered<br />

composite. A composite with non-uniform particle distribution was considered as a two phase<br />

material comprising regions <strong>of</strong> higher than average <strong>local</strong> volume fraction as stiffer regions<br />

within a more compliant dilutely reinforced <strong>matrix</strong>. This analysis suggests somewhat<br />

surprisingly, that a composite containing clustered particles should have a higher flow stress.<br />

It was also found that larger stresses may be expected in both the <strong>matrix</strong> <strong>and</strong> particles in<br />

clustered regions compared with a composite containing a homogeneous particle distribution.<br />

On the other h<strong>and</strong>, it is suggested in [28] that reinforcement clustering tends to reduce the<br />

composite flow stress, relative to an equi-spaced array. Zhiru et al. [29] have proposed that<br />

the <strong>matrix</strong> triaxial stresses are generally higher in clustered regions <strong>of</strong> four or more particles,<br />

thus shielding the center <strong>of</strong> the cluster from plastic flow due to the higher <strong>local</strong> levels <strong>of</strong><br />

constraint imposed on the <strong>matrix</strong>. Plastic flow is, thus, only expected in the center <strong>of</strong> the<br />

clusters once the more dilutely reinforced regions <strong>of</strong> the <strong>matrix</strong> have become sufficiently<br />

work hardened. In [30], a decrease <strong>of</strong> experimentally determined <strong>global</strong> <strong>fracture</strong> toughness<br />

with increasing degree <strong>of</strong> clustering was observed.<br />

2.4. The process <strong>of</strong> void initiation in materials with particles<br />

As well known, micro-ductile <strong>fracture</strong> consists <strong>of</strong> three stages: void nucleation, void growth,<br />

<strong>and</strong> void coalescence [31]. In ductile materials reinforced with hard second phase particles,<br />

void nucleation occurs either by particle/<strong>matrix</strong> decohesion or the particle <strong>fracture</strong><br />

mechanism. Void nucleation can determine the ductility, since the subsequent stages <strong>of</strong> void<br />

growth <strong>and</strong> coalescence can be extremely rapid (especially in the case <strong>of</strong> large ceramic<br />

particles) once the void initiation has occurred. Therefore, the full underst<strong>and</strong>ing <strong>of</strong> void<br />

initiation process is very important to predict the <strong>fracture</strong> <strong>properties</strong> <strong>of</strong> MMCs.<br />

10


Section 2<br />

An energy criterion <strong>and</strong> a stress criterion have been suggested to characterize the<br />

particle/<strong>matrix</strong> interface separation [32]. A detailed analysis by Tanaka et al. [32] has shown<br />

that for particles having a size smaller than 250 Å, the particle/<strong>matrix</strong> separation is primarily<br />

controlled by the energy criterion. In this case, the particle/<strong>matrix</strong> separation will occur when<br />

the elastic strain energy which is released upon decohesion becomes equal to the newly<br />

created surface energy. When an inclusion is larger than 250 Å, the energy criterion is always<br />

fulfilled <strong>and</strong> the particle/<strong>matrix</strong> separation is primarily controlled by the stress criterion [32].<br />

Void nucleation occurs at the attainment <strong>of</strong> a critical normal stress at the particle/<strong>matrix</strong><br />

interface.<br />

Some models based on a dislocation theory were developed. Ashby proposed a model where<br />

primary deformation incompatibilities do not produce cavities directly, but initiate highly<br />

organized secondary dislocation loops from the interface <strong>of</strong> the inclusion with the <strong>matrix</strong> to<br />

reduce the <strong>local</strong> shear stresses [33]. These loops can form reverse pile-ups <strong>and</strong> can build up<br />

increasing interfacial stresses until they reach the interfacial strength. In [31], the <strong>local</strong> stress<br />

on the particle interface has been estimated as a function <strong>of</strong> the <strong>local</strong> dislocation density<br />

around a particle. It was demonstrated that the rate <strong>of</strong> dislocation storage around a particle, in<br />

a plastically deforming <strong>matrix</strong>, decreases with an increase in particle size <strong>and</strong>, above a critical<br />

particle size <strong>of</strong> 1 µm, a continuum plasticity model can be applied to describe the microvoid<br />

nucleation process, since the <strong>local</strong> strain hardening effect around the particle is <strong>of</strong> the same<br />

magnitude as that in the <strong>matrix</strong>.<br />

Argon et al. [34] used a continuum mechanics theory to estimate the maximum interfacial<br />

stresses. They considered a rigid cylindrical inclusion embedded into an incompressible<br />

power-law hardening material subjected to a pure shear deformation. The actual behavior <strong>of</strong><br />

the material is bounded by two limiting idealizations: a rigid ideally plastic behavior <strong>and</strong> a<br />

linear behavior. The stress concentrations obtained for these two extreme idealizations were<br />

compared to estimate the actual interfacial stresses. It was shown that the maximum<br />

interfacial stresses can be calculated as a sum <strong>of</strong> the von Mises equivalent stress, σeq, <strong>and</strong> the<br />

hydrostatic stress, σm,<br />

σ interf max = σeq + σm. (2.3)<br />

In [35], Argon <strong>and</strong> Im investigated the behavior <strong>of</strong> equiaxed Fe3C particles in a 1045 Steel,<br />

Cu-Cr particles in a Cu-0.6%Cr alloy, <strong>and</strong> TiC particles in a maraging steel. Cylindrical<br />

tensile specimens with <strong>and</strong> without machined notches were tested up to failure. The broken<br />

specimen halves were sectioned along the axis, <strong>metal</strong>lographically polished, <strong>and</strong> lightly<br />

11


Section 2<br />

etched to mark the inclusions. The inclusions which have initiated voids were detected, <strong>and</strong><br />

the number <strong>of</strong> voids was plotted as a function <strong>of</strong> the distance from the <strong>fracture</strong> surface. The<br />

point along the axis where the void density drops to zero was taken as characterizing the<br />

critical conditions for interface separation. A finite element analysis was performed, [36], to<br />

compute the stress conditions at this point. The critical interfacial stresses were then<br />

calculated according to Eq. (2.3). The maximum interfacial stresses were 1670 MPa, 990<br />

MPa, <strong>and</strong> 1820 MPa for Fe3C, Cu-Cr, <strong>and</strong> TiC particles, respectively.<br />

Arsenault <strong>and</strong> Flom evaluated the interfacial strength between particles <strong>and</strong> the <strong>matrix</strong> in an<br />

Al6061 based MMC with 1% SiC particles [37]. Specimens similar to that used in [35] were<br />

machined <strong>and</strong> procedure used in [35] was applied to the specimens. The interfacial strength<br />

was calculated by Eq (2.3), taking into account the stress triaxiality for the tip <strong>of</strong> the notch,<br />

where the triaxiality reaches its maximum. The interfacial strength was determined as 1690<br />

MPa. This result does not seem to be correct, since the number <strong>of</strong> voids associated with the<br />

debonding <strong>of</strong> SiC particles was much smaller than the total number <strong>of</strong> voids related to the<br />

<strong>fracture</strong>. A few examples <strong>of</strong> the areas where debonded SiC particles are located were<br />

observed, but they were distant from the tip <strong>of</strong> the notch.<br />

Beremin [38] studied the conditions for void initiation for elongated MnS inclusions in a low<br />

alloy steel A508. Notched tensile specimens were loaded to different <strong>global</strong> deformations <strong>and</strong><br />

subsequently sectioned <strong>and</strong> polished to determine the outer boundary <strong>of</strong> the region within<br />

which voids have been initiated. The <strong>local</strong> stresses at this boundary were computed with a<br />

finite element analysis, <strong>and</strong> the maximum principal stress in the inclusion at the moment <strong>of</strong><br />

void initiation evaluated by the equation<br />

σ p max =σm + (2/3+λ)σeq. (2.4)<br />

This equation was deduced from a non-linear Eshelby-type approach, following [39]. The<br />

parameter λ is a function <strong>of</strong> the particle shape; λ = 1 for a spherical particle. Beremin found<br />

that the σ p max-values for the MnS-inclusions in the steel A508 are temperature independent,<br />

but depend on the specimen orientation: σ p max = 1120±60 MPa in the longitudinal direction,<br />

where MnS-inclusions are <strong>fracture</strong>d, <strong>and</strong> σ p max = 810±50 MPa in the short transverse<br />

direction, where MnS-inclusion/<strong>matrix</strong> decohesion prevails.<br />

Toda et al. [40] determined the <strong>fracture</strong> strength <strong>of</strong> second-phase particles in a notched 3-<br />

point bend specimen made <strong>of</strong> a wrought Al-Li alloy. A side-surface <strong>of</strong> the specimen was<br />

polished <strong>and</strong> etched before the testing. An interrupted in-situ loading experiment was<br />

12


Section 2<br />

conducted in the scanning electron microscope (SEM). The notch radius was 150 µm. The<br />

position <strong>and</strong> size <strong>of</strong> each particle was measured within a region <strong>of</strong> 500 µm around the notch<br />

<strong>and</strong> the extending crack tip. The chemical composition <strong>of</strong> the inclusions was determined by a<br />

SEM-EDX analysis. After each loading step, a micrograph was taken to detect the broken<br />

particles. From the value <strong>of</strong> the J-integral, J, the particle <strong>fracture</strong> stress was evaluated using<br />

the stress field approach proposed by Hutchinson [41], Rice, <strong>and</strong> Rosengren [42] (HRR-<br />

theory) <strong>and</strong> a mean field correction following Tanaka et al. [32]. The process zone where<br />

large deformations appear directly at the growing crack tip was excluded from the analysis.<br />

Toda et al. found values <strong>of</strong> the maximum normal stress <strong>of</strong> σ p max = 710 MPa for CuAl2 <strong>and</strong><br />

Al2CuMg particles with diameters between 5 <strong>and</strong> 8 µm. Whereas the <strong>fracture</strong> strength <strong>of</strong> the<br />

CuAl2 particles did not change much with increasing particle size, larger Al2CuMg particles<br />

exhibited considerably smaller values, e.g., σ p max = 540 MPa for diameters between 14 <strong>and</strong><br />

18 µm. Al3Zr <strong>and</strong> Al3Ti particles did not <strong>fracture</strong>, thus a lower limit <strong>of</strong> their <strong>fracture</strong> strength<br />

could be determined: σ p max ≥ 1000 MPa <strong>and</strong> 910 MPa, respectively.<br />

The study <strong>of</strong> Toda et al. is remarkable, however, a few disadvantages has prevented the direct<br />

application <strong>of</strong> the procedure for our materials:<br />

(1.) The HRR-field is, strictly speaking, not applicable for a region around a notch, nor is<br />

it for a growing crack. Regarding the growing cracks, Toda et al. claim that the HRR-<br />

field should be still valid if a parameter Ω = Eεfr /σy is smaller than 34.5 (E is the<br />

Young modulus, εfr denotes the <strong>fracture</strong> strain in a uniaxial tensile test, <strong>and</strong> σy the<br />

yield stress). Hereby, Toda et al. refer to a paper by Chan [43] who used the<br />

parameter Ω to distinguish between stable <strong>and</strong> unstable crack growth regimes. As has<br />

been shown by Hutchinson <strong>and</strong> Paris [44], the conditions <strong>of</strong> J-controlled crack<br />

growth will be lost for any material, if the crack extension becomes too large.<br />

Moreover, it is demonstrated in [44] that a tough material, where dJ/da <strong>and</strong> the<br />

parameter Ω are large, will exhibit J-control for a longer regime <strong>of</strong> crack extension<br />

than a less tough material with a small Ω.<br />

(2.) The mean-field analysis was performed assuming an inclusion in a <strong>matrix</strong> under a<br />

uniaxial tensile stress. A particle near a crack tip will, however, experience a highly<br />

triaxial stress state.<br />

(3.) The measurement at the side surface <strong>of</strong> the specimen might be not representative for<br />

the bulk behavior. It is well known that, because <strong>of</strong> the lower constraint, void<br />

initiation <strong>and</strong> void growth are retarded near the side surface compared to the<br />

midsection region.<br />

13


Section 2<br />

(4.) The specimen preparation might pre-damage some <strong>of</strong> the particles so that they fail at<br />

a lower stress.<br />

As one can see, despite a large number <strong>of</strong> investigations dealing with the <strong>fracture</strong> behavior <strong>of</strong><br />

MMCs, the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> in MMCs have not been investigated so far, probably,<br />

due to the absence <strong>of</strong> appropriate tools. But this topic is very important for inhomogeneous<br />

materials. As the void initiation is the first stage <strong>of</strong> the ductile <strong>fracture</strong>, the conditions for void<br />

nucleation call great interest, especially, near the crack front in the interior <strong>of</strong> the specimen.<br />

As shown in next section, an application <strong>of</strong> recently developed tools for <strong>fracture</strong> surface<br />

analysis could solve such a problem as the impossibility to investigate the <strong>fracture</strong> processes<br />

in the interior <strong>of</strong> the specimen <strong>and</strong> give a possibility to determine the <strong>local</strong> conditions for void<br />

initiation for individual particles near the crack tip.<br />

14


Section 3<br />

3. A method to estimate the <strong>local</strong> conditions for void initiation in materials<br />

with inclusions<br />

This section describes the proposed procedure to estimate the <strong>local</strong> conditions for void<br />

initiation near the crack tip. The procedure consists <strong>of</strong>:<br />

1) For individual particles lying in the midsection region near the crack tip, the values <strong>of</strong> the<br />

crack tip opening displacement at the moment <strong>of</strong> void initiation, CODvi, are determined.<br />

This is done by using stereophotogrammetry <strong>and</strong> a digital image analysis system.<br />

2) The mesoscopic stress tensor at the position <strong>of</strong> each considered particle is estimated by the<br />

HRR-theory from the determined CODvi-value.<br />

3) The stress tensors in both the particle <strong>and</strong> in the <strong>matrix</strong> are calculated by a non-linear<br />

Mori-Tanaka secant approach.<br />

Below, each step is outlined in detail.<br />

3.1. An automatic <strong>fracture</strong> surface analysis system<br />

To determine the values <strong>of</strong> the crack tip opening displacement at the moment <strong>of</strong> void<br />

initiation, CODvi, <strong>and</strong> at the moment <strong>of</strong> <strong>fracture</strong> initiation, CODi, as well as the polar<br />

coordinates <strong>of</strong> the particle location with respect to the crack plane, an automatic <strong>fracture</strong><br />

surface analysis system is used [45, 46]. The key part <strong>of</strong> this system is a matching algorithm<br />

which is able to find automatically homologue points in a stereo image pair which is taken in<br />

a scanning electron microscope (SEM). Homologue points are points in the two images which<br />

belong to the same physical point on the specimen. The stereo image pair is produced by<br />

tilting the specimen in the SEM. Each image consists <strong>of</strong> 1024 x 768 pixels at 256 gray levels.<br />

On average, between 20000 to 30000 homologue points are found. It should be noted that<br />

images with a resolution <strong>of</strong> up to 4000x3200 pixels can be taken in new SEMs. For such<br />

pictures, the number <strong>of</strong> homologue points may increase by a factor 10 or more.<br />

From the homologue points <strong>and</strong> the three image parameters, magnification, working distance,<br />

<strong>and</strong> tilt angle, a three-dimensional model <strong>of</strong> the depicted <strong>fracture</strong> surface region is generated,<br />

called “digital elevation model” (DEM). In Figure 3.1, a DEM for a cast Al6061-10%Al2O3<br />

MMC is given as an example. The DEM can be cut arbitrarily to extract <strong>fracture</strong> surface<br />

15


Section 3<br />

Fig. 3.1. The digital elevation model <strong>of</strong> a cast Al6061-10%Al2O3 MMC.<br />

pr<strong>of</strong>iles. Further important capabilities <strong>of</strong> the automatic <strong>fracture</strong> surface analysis system are<br />

the automatic evaluation <strong>of</strong> pr<strong>of</strong>ile <strong>and</strong> surface roughness parameters <strong>and</strong> the determination <strong>of</strong><br />

fractal dimensions [47].<br />

The first version <strong>of</strong> the automatic <strong>fracture</strong> surface analysis system used a simple matching<br />

algorithm, based on the evaluation <strong>of</strong> a cross-correlation coefficient [48]. Later, the system<br />

has been completely modified with respect to the matching algorithm <strong>and</strong> the user interface<br />

[46]. A comprehensive description <strong>of</strong> the principles <strong>of</strong> the new matching algorithm is given in<br />

[49]. Both systems have been applied successfully to analyze ductile <strong>and</strong> cleavage <strong>fracture</strong><br />

surfaces <strong>of</strong> <strong>metal</strong>s [50, 51], <strong>metal</strong>lic glasses [52], <strong>and</strong> inter<strong>metal</strong>lic alloys [53], but it has been<br />

also proven valuable for the inspection <strong>of</strong> the surface structure <strong>of</strong> many classes <strong>of</strong> materials,<br />

from <strong>metal</strong>s to biological materials.<br />

In the following section, it is described how the automatic <strong>fracture</strong> surface analysis system is<br />

applied to determine the <strong>local</strong> CODi- <strong>and</strong> CODvi-values in materials with inclusions.<br />

3.2. The determination <strong>of</strong> the CODi- <strong>and</strong> CODvi –values<br />

The SEM-micrographs <strong>of</strong> Figure 3.2 show corresponding regions near the midsection on the<br />

two halves <strong>of</strong> a broken compact tension (CT) specimen made <strong>of</strong> a cast Al6061-10%Al2O3<br />

16


Section 3<br />

Fig. 3.2. Corresponding stereo image pairs with pr<strong>of</strong>iles passed through a particle<br />

(the particle location is marked).<br />

MMC. The pre-fatigue regions (in the bottom part <strong>of</strong> pictures) <strong>and</strong> the regions <strong>of</strong> micro-<br />

ductile <strong>fracture</strong> can be clearly distinguished. From the DEMs <strong>of</strong> the corresponding regions,<br />

crack pr<strong>of</strong>iles are extracted perpendicularly to the pre-fatigue crack front. The pr<strong>of</strong>iles are<br />

drawn so that they cross an alumina particle in front <strong>of</strong> the crack tip. Sometimes, the paths<br />

have a zigzag shape to find easier the corresponding path on the second specimen half.<br />

In Figure 3.3a, the two corresponding pr<strong>of</strong>iles through an alumina particle <strong>of</strong> the specimen are<br />

arranged so that the moment <strong>of</strong> <strong>fracture</strong> initiation is depicted, i.e., the moment when the<br />

ligament between the first void in front <strong>of</strong> the tip <strong>and</strong> the blunted pre-crack fails. The upper<br />

pr<strong>of</strong>ile (drawn as the red line) corresponds to the left-h<strong>and</strong> side <strong>of</strong> Figure 3.2, the lower<br />

pr<strong>of</strong>ile (blue line) corresponds to the right-h<strong>and</strong> side. The location <strong>of</strong> the particle in the center<br />

<strong>of</strong> the void is marked. The critical crack tip opening displacement, CODi, which is a measure<br />

<strong>of</strong> the <strong>local</strong> <strong>fracture</strong> initiation toughness [50], can be determined: CODi = 17 µm.<br />

If the lower pr<strong>of</strong>ile <strong>of</strong> Figure 3.3a is shifted vertically so that two pr<strong>of</strong>iles touch each other at<br />

the location <strong>of</strong> the marked particle, we get a sketch <strong>of</strong> the crack at the moment <strong>of</strong> void<br />

initiation (Fig. 3.3b) <strong>and</strong> the crack tip opening displacement at the moment <strong>of</strong> void initiation,<br />

CODvi, can be measured. For the considered particle, we get CODvi = 10 µm. Analogously to<br />

17


height [µm]<br />

height [µm]<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

COD i =17µm<br />

Section 3<br />

-60 -20 20 60 100<br />

COD vi=10µm<br />

distance [µm]<br />

a)<br />

-60 -20 20 60 100<br />

distance [µm]<br />

b)<br />

18<br />

r<br />

Θ<br />

particle<br />

Fig. 3.3. Crack pr<strong>of</strong>ile through a particle in the cast Al6061-10%Al2O3 MMC: a) at the<br />

moment <strong>of</strong> <strong>fracture</strong> initiation; b) at the moment <strong>of</strong> void initiation at the particle location.<br />

the <strong>local</strong> <strong>fracture</strong> initiation toughness CODi (which can be transferred to critical values <strong>of</strong> the<br />

stress intensity or the J-integral), CODvi gives the resistance <strong>of</strong> the material against void<br />

initiation <strong>and</strong>, thus, it can be considered as a <strong>local</strong> “void initiation toughness” [54].


Section 3<br />

The location <strong>of</strong> the particle center with respect to the crack tip, given in polar coordinates (r,<br />

θ), is also determined from the crack pr<strong>of</strong>ile. The polar coordinates <strong>of</strong> the particle center shall<br />

be measured from the initial position <strong>of</strong> the tip <strong>of</strong> the fatigue pre-crack before the blunting<br />

process starts. This can be performed more accurately from the crack pr<strong>of</strong>iles depicting the<br />

moment <strong>of</strong> void initiation, Figure 3.3b. For many materials, the tip <strong>of</strong> the fatigue pre-crack<br />

can be found on the SEM micrographs as the beginning <strong>of</strong> the stretched zone.<br />

As follows from the procedure (parallel shifting <strong>of</strong> pr<strong>of</strong>iles), the method <strong>of</strong> determination <strong>of</strong><br />

CODvi is based on the assumption that the void growth rate in a direction perpendicular to the<br />

crack plane,<br />

.<br />

R , is equal to the COD growth rate,<br />

y<br />

19<br />

.<br />

COD .<br />

An early numerical analysis <strong>of</strong> void growth near a crack tip, [55], suggests that the COD<br />

growth rate should be several times higher than the void growth rate. If this would be valid,<br />

we would get negative CODvi-values for most <strong>of</strong> the considered below particles, so that we<br />

conclude that this cannot be true. In [56], the growth <strong>of</strong> a single void near the crack tip was<br />

investigated for different specimen geometries by the finite element method. In Figure 3.4, the<br />

Fig. 3.4. Plots <strong>of</strong> the crack tip opening displacement, B, the dimension <strong>of</strong> the void size in two<br />

directions, Rx <strong>and</strong> Ry, <strong>and</strong> the size <strong>of</strong> the ligament between the crack tip <strong>and</strong> the void , D, in<br />

plane strain case (taken from [56]).<br />

i


Section 3<br />

Fig. 3.5. The observations <strong>of</strong> the growth <strong>of</strong> an assemble <strong>of</strong> voids near the crack tip for plane<br />

strain conditions (taken from [57]).<br />

void size in two directions, Rx <strong>and</strong> Ry, the size <strong>of</strong> the ligament, D, the value <strong>of</strong> the crack tip<br />

opening displacement, B, are plotted as a function <strong>of</strong> the dimensionless loading parameter <strong>of</strong><br />

the crack,<br />

J<br />

σ<br />

o o R<br />

, for different specimen geometries: the double-edge cracked (DEC), the<br />

center-cracked plain (CCP), three point bending (TPB), <strong>and</strong> the single-edge cracked (SEC).<br />

From Figure 3.4, the void <strong>and</strong> COD growth rates can be estimated for a TPB specimen (which<br />

has very similar constraint conditions as a compact tension specimen) <strong>and</strong> for plane strain<br />

conditions. The result is<br />

.<br />

.<br />

COD ≈ 1. 1⋅<br />

D . It should be noted that results obtained for other<br />

specimen types differ: for instance, for the SEC specimen,<br />

20<br />

.<br />

CODi .<br />

≈ 3⋅ D . In a recent study,<br />

[57], the growth <strong>of</strong> an assemble <strong>of</strong> voids in a polycrystalline microstructure near a crack tip<br />

has been modeled. An analysis <strong>of</strong> Figure 3.5 shows again that the void <strong>and</strong> COD growth rates<br />

are approximately equal. Thus, it can be concluded that the inaccuracy <strong>of</strong> determination <strong>of</strong> the<br />

CODvi-values related with the simple shifting <strong>of</strong> pr<strong>of</strong>iles is not high.<br />

3.3. Estimate <strong>of</strong> the maximum principal stresses in both phases<br />

3.3.1. Estimate <strong>of</strong> the HRR-stress tensor in the moment <strong>of</strong> void initiation<br />

The stress tensor at the moment <strong>of</strong> void initiation is estimated from the HRR-field solution<br />

(Eq. 3.1) which gives the stress field around a crack tip as a function <strong>of</strong> the <strong>global</strong> material<br />

<strong>properties</strong> <strong>and</strong> the polar coordinates with respect to the crack tip (Fig. 3.6) [41, 42],


Section 3<br />

Fig. 3.6. Schematic view <strong>of</strong> the HRR-theory.<br />

21<br />

(3.1)<br />

In Eq. 3.1, E is the Young modulus, J is the <strong>fracture</strong> toughness <strong>of</strong> material, σ0 is the yield<br />

strength, r <strong>and</strong> θ are the polar coordinates, IN <strong>and</strong> dN are dimensionless constants both<br />

depending on the strain hardening coefficient, N, <strong>and</strong> on σ0 /E, ~ σ ( N,<br />

θ ) is a dimensionless<br />

function listed in [59].<br />

The J-integral in Eq. 3.1 can be substituted by COD using the relation proposed by Shih in<br />

[58]<br />

, (3.2)<br />

where dN is dimensionless constant depending on the strain hardening coefficient, N, [58].<br />

The material is assumed to follow a st<strong>and</strong>ard power-law work hardening behavior<br />

, (3.3)<br />

where according to [59], the σ0 is set to the yield strength, σy, <strong>and</strong> α is determined from the<br />

tensile true stress-strain curve.<br />

⎡ E J ⎤<br />

σ σ<br />

~<br />

ij = 0 ⎢<br />

σ ,<br />

2 ⎥<br />

ij<br />

⎢⎣<br />

ασ 0 I N r ⎥⎦<br />

J =<br />

d<br />

1<br />

σ 0<br />

N<br />

To evaluate the stress tensor at the moment <strong>of</strong> void initiation, σ HRR vi, the measured r, θ, <strong>and</strong><br />

the CODvi-values are inserted into Eqs. (3.2) <strong>and</strong> (3.1). From the stress tensor σ HRR vi, the<br />

maximum principal stress, σ HRR max, can be calculated.<br />

1/(<br />

N + 1)<br />

COD<br />

ε<br />

⎛ σ ⎞<br />

= α ⎜<br />

⎟<br />

ε 0 ⎝σ<br />

0 ⎠<br />

N<br />

r<br />

θ<br />

( N θ )<br />

ij<br />

σij


Section 3<br />

3.3.2. The determination <strong>of</strong> the maximum stresses both in the particle <strong>and</strong> in the <strong>matrix</strong> by<br />

a Mori-Tanaka type mean-field correction<br />

So far, the material has been considered as a material with homogenized <strong>properties</strong>. The<br />

considered length scale is called the “mesoscopic scale” <strong>and</strong> the stresses are denoted as<br />

“mesoscopic stresses”, i.e., the HRR-stresses are meso-stresses. However, for the evaluation<br />

<strong>of</strong> <strong>local</strong> phenomena, such as particle failure, the “microscopic scale” has to be considered. To<br />

do this, a non-linear Mori-Tanaka type mean field approach is applied which is based on the<br />

Eshelby theory. In the next section, the Eshelby approach is presented in detail.<br />

3.3.2.1. The Eshelby approach<br />

Internal stresses are present in any inhomogeneous material consisting <strong>of</strong> several phases.<br />

They arise as a result <strong>of</strong> some kind <strong>of</strong> misfit between the constituents (<strong>matrix</strong> <strong>and</strong> the<br />

reinforcement). Such a misfit could arise from a temperature change when the constituents<br />

have different thermal expansion coefficients, but a closely related situation is created during<br />

mechanical loading – when a stiff inclusion tends to deform less than the surrounding <strong>matrix</strong>.<br />

Analysis <strong>of</strong> the internal stresses in inclusions allows the prediction <strong>of</strong> <strong>global</strong> composite<br />

<strong>properties</strong> such as thermal expansion <strong>and</strong> stiffness. For the case <strong>of</strong> an ellipsoid, an analytical<br />

technique proposed by J.D.Eshelby [60] can be employed. Eshelby could prove that the<br />

ellipsoid has a uniform stress at all points within it. The technique is based on<br />

representing the actual inclusion by<br />

Fig. 3.7. Eshelby’s cutting <strong>and</strong> welding exercises for the uniform stress-free transformation <strong>of</strong><br />

an ellipsoid region.<br />

22


Section 3<br />

one made <strong>of</strong> the <strong>matrix</strong> material which has an appropriate misfit strain, such that the stress<br />

field is the same as for the actual inclusion. In Figure 3.7, a region (the inclusion) is cut from<br />

the unstressed elastically homogeneous material, <strong>and</strong> is then imagined to undergo a shape<br />

change (the transformation strain, εεεε T ) free from the constraining material. The inclusion<br />

cannot now be replaced directly back into the hole from it came. Instead surface tractions are<br />

first applied in order to return it to its original shape. Once back in position, the two regions<br />

are then welded together, i.e. there is no movement or sliding along the interface, <strong>and</strong> the<br />

surface tractions are then removed. Equilibrium is then reached between the <strong>matrix</strong> <strong>and</strong> the<br />

inclusion at a constrained strain, ε c , <strong>of</strong> the inclusion relative to its initial shape before<br />

removal.<br />

Since the inclusion is strained uniformly throughout, the stress within it can be calculated<br />

using Hooke’s Law in terms <strong>of</strong> the elastic strain (ε c - ε T ) <strong>and</strong> the stiffness tensor <strong>of</strong> the<br />

material, Cm.<br />

σσσσI = Cm(εεεε c - εεεε T ) (3.4)<br />

Eshelby found that ε c can be obtained from ε T by means <strong>of</strong> a tensor termed the Eshelby tensor<br />

“S”, which can be calculated in terms <strong>of</strong> the shape (aspect ratio) <strong>of</strong> the inclusion <strong>and</strong> <strong>of</strong> the<br />

Poisson’s ratio <strong>of</strong> the <strong>matrix</strong>, see e.g. [2, 60]<br />

εεεε c = Sεεεε T (3.5)<br />

The tensor S thus expresses the relationship between the final constrained inclusion shape <strong>and</strong><br />

the shape <strong>of</strong> the inclusion after its cutting out <strong>and</strong> deformation. For the case <strong>of</strong> the spherical<br />

inclusions (aspect ratio is equal to 1), the only non-zero components <strong>of</strong> the Eshelby tensor are<br />

[61]<br />

m<br />

7 − 5ν<br />

S(<br />

1,<br />

1)<br />

=<br />

S(<br />

2,<br />

2)<br />

= S(<br />

3,<br />

3)<br />

=<br />

m<br />

15(<br />

1−ν<br />

)<br />

m<br />

5ν<br />

−1<br />

S(<br />

1,<br />

2)<br />

= S(<br />

2,<br />

1)<br />

= S(<br />

1,<br />

3)<br />

= S(<br />

3,<br />

1)<br />

= S(<br />

2,<br />

3)<br />

= S(<br />

3,<br />

2)<br />

=<br />

m<br />

15(<br />

1−ν<br />

)<br />

m<br />

2 ⋅ ( 4 − 5ν<br />

)<br />

S(<br />

4,<br />

4)<br />

= S(<br />

5,<br />

5)<br />

= S(<br />

6,<br />

6)<br />

=<br />

m<br />

15(<br />

1−ν<br />

)<br />

23


Section 3<br />

The Eshelby tensor does not depend on the size <strong>of</strong> the inclusion. Therefore, micromechanical<br />

methods based on the Eshelby tensor do not have an intrinsic length scale, i.e. the results do<br />

not depend on the size <strong>of</strong> inclusions.<br />

3.3.2.2. Some general mean-field relations<br />

It should be noted that the Eshelby theory is developed for the case when both constituents the<br />

<strong>matrix</strong> <strong>and</strong> the reinforcements are elastic. However, the MMCs usually have elastic<br />

reinforcements <strong>and</strong> elastic-plastic <strong>matrix</strong>. A mean field approach based on the Eshelby theory<br />

is used to estimate stresses in the particles <strong>and</strong> in the <strong>matrix</strong> for such materials.<br />

Mean field approaches operate on the basis <strong>of</strong> averaged stress <strong>and</strong> strain fields in the<br />

constituent phases. Some general relations apply for linking the homogeneous meso-fields to<br />

the micro-fields [62] by employing so called ‘concentration tensors’ as<br />

24<br />

(3.6a)<br />

(3.6b)<br />

Here, B p <strong>and</strong> B m denote the stress concentration tensors, <strong>and</strong> σ p <strong>and</strong> σ m are the (averaged)<br />

micro-stress tensors for the particle <strong>and</strong> <strong>matrix</strong> phases, respectively; σ is the meso-stress<br />

tensor (Fig. 3.6). Equivalent relations hold for the strains. Note that Eqs. (3.6a,b) are tensorial<br />

relations taking into account the full 3D stress state. In our case, the meso-stress tensor σ is<br />

determined by the HRR-theory (Eq. 3.1).<br />

Following Benveniste’s [63] interpretation <strong>of</strong> the Mori-Tanaka approach, the phase<br />

concentration tensors are evaluated as<br />

ξ is the particle volume fraction, I the unit tensor, <strong>and</strong><br />

concentration tensor, which can be written according to [62] as<br />

B<br />

p<br />

dil<br />

σ<br />

σ<br />

p =<br />

m =<br />

B<br />

B<br />

p<br />

m<br />

σ<br />

σ<br />

[ ] 1 p −<br />

( 1 − ξ ) I +<br />

p p<br />

B B dil ξ<br />

= B<br />

dil<br />

[ ] 1 p −<br />

( 1−<br />

ξ ) I +<br />

m<br />

B ξ<br />

= B<br />

dil<br />

[ ] 1<br />

-1<br />

-1 −<br />

m<br />

p m<br />

I + Es<br />

( I − S)(<br />

E − s )<br />

= E<br />

, (3.7a)<br />

. (3.7b)<br />

p<br />

B dil the dilute particle stress<br />

, (3.8)


Section 3<br />

m<br />

E s is the secant tensor <strong>of</strong> the <strong>matrix</strong> material,<br />

Eshelby tensor.<br />

25<br />

p<br />

E the particle elasticity tensor, <strong>and</strong> S the<br />

It should be noted that Eq. 3.8 reflects the original approach by Eshelby [60], where a single<br />

inclusion embedded in an infinite <strong>matrix</strong> is assumed; thus, the dilute concentration tensors are<br />

independent <strong>of</strong> the volume fraction. In the Mori-Tanaka method (or similar approaches) [64-<br />

66], the dilute-case assumption is removed <strong>and</strong> Eqs. (3.7a,b) are functions <strong>of</strong> the particle<br />

volume fraction.<br />

3.2.2.3. Solution procedure<br />

In the solution procedure, the <strong>matrix</strong> is assumed to follow the st<strong>and</strong>ard power-law work<br />

hardening behavior (Eq. 3.3). In Eq. 3.3, the uniaxial stress <strong>and</strong> strain, σ <strong>and</strong> ε, are replaced<br />

by the equivalent stress <strong>and</strong> equivalent strain, σ m eq <strong>and</strong> ε m eq, for multiaxial load cases. As our<br />

“region <strong>of</strong> interest” is located close to the crack tip, well within the plastic zone, the Poisson’s<br />

ratio is assumed to be ν m = 0.5. It is noted that with specified E <strong>and</strong> ν all components <strong>of</strong> the<br />

elasticity tensor are determined for a homogeneous isotropic material. Similarly, the<br />

components <strong>of</strong> the loading dependent secant tensor are determined from the secant modulus<br />

Es m = σ m eq/ε m eq.<br />

As for a given σ HRR , the <strong>matrix</strong> stress <strong>and</strong> strain <strong>and</strong>, thus, the secant modulus are initially<br />

unknown, an implicit system <strong>of</strong> equations is set up which is solved by an iterative procedure.<br />

The aim <strong>of</strong> the iterative procedure is to determine the stress concentration tensors, B p <strong>and</strong> B m .<br />

The particle behavior is easy to h<strong>and</strong>le, because the secant modulus is independent <strong>of</strong> the<br />

(HRR)<br />

σij,vi θ<br />

(HRR)<br />

σij,vi Fig. 3.8. A schematic view <strong>of</strong> the mean-field approach.<br />

r<br />

particle<br />

σij,vi <strong>matrix</strong><br />

σij,vi


equivalent stress<br />

Section 3<br />

Fig. 3.9. A schematic illustration <strong>of</strong> the Mori-Tanaka approach.<br />

stress or strain level <strong>and</strong> equal to Young modulus. Therefore, the elasticity tensor,<br />

26<br />

p<br />

E , that is<br />

inserted into the Mori-Tanaka calculations, is always the same. But the nonlinear <strong>matrix</strong><br />

behavior causes problems because the magnitude <strong>of</strong> the secant modulus depends on the stress<br />

<strong>and</strong> the strain level <strong>and</strong> so the <strong>matrix</strong> elasticity tensors,<br />

m<br />

E s , are a function <strong>of</strong> the equivalent<br />

stress <strong>and</strong> strain. To explain it more in detail, in Figure 3.9, a schematic illustration <strong>of</strong> the<br />

Mori-Tanaka approach is given. The dashed lines represents the effective composite behavior.<br />

The <strong>matrix</strong> stress field is linked to the <strong>global</strong> composite stress field by the <strong>matrix</strong> stress<br />

concentration tensor, B m , according to Eq. 3.7b. On the one h<strong>and</strong>, B m depends on the <strong>matrix</strong><br />

secant modulus, m<br />

E s , on the other h<strong>and</strong> this <strong>matrix</strong> secant modulus is needed to calculate the<br />

B m . The problem is to find a secant modulus. The flow chart <strong>of</strong> the iterative procedure which<br />

is employed to solve this problem is shown in Figure 3.10. The solution yields the particle <strong>and</strong><br />

<strong>matrix</strong> stress <strong>and</strong> strain tensors. From the stress tensors σ p <strong>and</strong> σ m (Eq. 3.6), the maximum<br />

normal stresses in the particle, σ p max, <strong>and</strong> in the <strong>matrix</strong>, σ m max, are evaluated.<br />

It should be noted, that on our solutions, the equivalent <strong>matrix</strong> stress is evaluated according to<br />

equation proposed by Hu in [67]<br />

?<br />

particle<br />

from HRR-theory<br />

?<br />

different secant moduli<br />

composite<br />

<strong>matrix</strong><br />

equivalent strain


σ<br />

2<br />

eq<br />

Section 3<br />

⎛ m G<br />

σ ⎜<br />

3<br />

= −<br />

⎜<br />

⎝ ξ<br />

27<br />

2<br />

<strong>matrix</strong><br />

composite<br />

. (3.9)<br />

Here, equivalent <strong>matrix</strong> stress is defined from the elastic distortional energy <strong>of</strong> the <strong>matrix</strong>,<br />

which can be evaluated from the variation <strong>of</strong> the compliance tensor <strong>of</strong> the<br />

composite,⎺Ccomposite, with respect to the variation <strong>of</strong> the shear modulus <strong>of</strong> the <strong>matrix</strong>, G<strong>matrix</strong>.<br />

σ m is a macroscopic load, ξ<strong>matrix</strong> is a <strong>matrix</strong> volume fraction. This approach is known to<br />

predict a more realistic non-linear behavior <strong>of</strong> the porous materials or materials with stiff<br />

reinforcements when the high values <strong>of</strong> stress triaxiality are present [67, 68].<br />

<strong>matrix</strong><br />

δC<br />

δG<br />

<strong>matrix</strong><br />

⎞<br />

⎟σ<br />

⎟<br />

⎠<br />

m


Assume σ m eq<br />

Calculate <strong>matrix</strong> secant modulus,<br />

E m sec, input<br />

Section 4<br />

Set up the elasticity tensor <strong>and</strong> employ<br />

MORI-TANAKA procedure ⇒ calculate<br />

stress-consentration tensor, B m<br />

Calculate the <strong>matrix</strong> stress<br />

tensor, σ m = B m σ HRR<br />

Calculate equivalent<br />

<strong>matrix</strong> stress, σ m eq<br />

Calculate new secant modulus,<br />

E m sec, output<br />

(E m sec, input - E m sec, output) ≥ accuracy<br />

accuracy<br />

(E m sec, input - E m sec, output) < accuracy<br />

Fig. 3.10. An iterative procedure solution: flow chart.<br />

28<br />

Calculate new <strong>matrix</strong><br />

secant modulus,<br />

E m sec, input = f (E m sec, input ; E m sec, output)<br />

converged solution


Section 4<br />

4. Local <strong>fracture</strong> <strong>properties</strong> <strong>of</strong> <strong>metal</strong> <strong>matrix</strong> composites<br />

4.1. Materials <strong>and</strong> their <strong>global</strong> mechanical <strong>properties</strong><br />

To study the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> <strong>of</strong> the MMCs, two different materials cast MMCs <strong>and</strong> a<br />

powder <strong>metal</strong>lurgy MMC are chosen as an object for this investigation. Below, more detailed<br />

information about investigated materials is given.<br />

4.1.1. Cast MMCs<br />

4.1.1.1. Materials characterization<br />

Cast <strong>and</strong> extruded MMC with an Al-6061 <strong>matrix</strong> reinforced by Al2O3 particles (three<br />

different particle volume fractions, ξ = 10, 15, <strong>and</strong> 20%, are considered) is investigated. The<br />

100 µm<br />

a) b)<br />

c)<br />

Fig. 4.1. a) Longitudinal sections <strong>of</strong> Al6061-based MMC with: a) 10% Al2O3 particles, b)<br />

15% Al2O3 particles; c) 20% Al2O3 particles.<br />

29<br />

100 µm<br />

100 µm


Section 4<br />

Table 4.1. Chemical composition <strong>of</strong> the Al-6061 alloy<br />

Si Fe Cu Mn Mg Zn Cr Ti<br />

0.4÷0.8 0.7 0.15÷0.4 0.15 0.8÷1.2 0.25 0.04÷0.35 0.15<br />

chemical composition <strong>of</strong> the <strong>matrix</strong> is given in Table 4.1. The mean alumina particle size is<br />

about 10 µm. Metallographic sectioning shows that the particles are distributed quite<br />

homogeneously in all composites, but a few particle clusters are observed, as well (Fig. 4.1).<br />

The particles have a shape <strong>of</strong> spheroids. The MMCs were supplied in the shape <strong>of</strong> bars with a<br />

section <strong>of</strong> 40x12.5 mm by AMAG (Austria).<br />

The materials were annealed at 560°C for 30 minutes, quenched in water, <strong>and</strong> kept at room<br />

temperature for 1 week [69]. To study the effect <strong>of</strong> the <strong>matrix</strong> <strong>properties</strong> on the composite<br />

behavior, the materials were subjected to different heat treatments:<br />

(1.) aging at room temperature;<br />

(2.) aging at 160°C for 8h;<br />

(3.) aging at 160°C for 24h;<br />

(4.) aging at 160°C for 200h.<br />

In the following, the investigated specimens are referred to by their volume percentage <strong>and</strong><br />

the heat treatment, e.g., Specimen Al2O3-10-RT or Specimen Al2O3-15-8h. The term<br />

“Increasing aging condition” will be used when specimens with the conditions RT, 160°C/8h,<br />

160°C/24h, 160°C/200h are compared.<br />

4.1.1.2. Tensile tests<br />

To determine the <strong>global</strong> material parameters, conventional tensile mechanical tests are<br />

performed. The tensile specimens have a cylindrical shape with a diameter <strong>of</strong> 3 mm <strong>and</strong> a<br />

gage length <strong>of</strong> 15 mm (Fig. 4.2). The tensile tests are conducted on a mechanical testing<br />

machine “ZWICK” at a loading rate <strong>of</strong> 5.6·10 -4 s -1 . The materials are assumed to follow a<br />

st<strong>and</strong>ard power-law work hardening behavior (Eq. 3.3). The strain hardening coefficient, N,<br />

<strong>and</strong> the coefficient, α, are determined from the log(ε/ε0) vs. log(σ/σ0) curves. The fit was<br />

taken so that it covers the better part <strong>of</strong> the stress-strain behavior (Fig. 4.3a). The results <strong>of</strong> the<br />

tensile tests are collected in Table 4.2. The mechanical <strong>properties</strong> <strong>of</strong> the <strong>matrix</strong> material in<br />

each considered aging condition are given in Table 3.3, as well.<br />

30


Section 4<br />

Fig 4.2. A view <strong>of</strong> a tensile test specimen <strong>of</strong> the cast Al6061 MMC fixed in the holders for<br />

Material E<br />

tensile test.<br />

Table 4.2. Mechanical <strong>properties</strong> <strong>of</strong> the cast Al6061 MMCs.<br />

[GPa]<br />

σy<br />

[MPa]<br />

σUTS<br />

[MPa]<br />

εfr<br />

[%]<br />

N α dn Ji<br />

31<br />

[kN/m]<br />

J0.2 / Bl<br />

[kN/m]<br />

Al2O3-0-RT 71 167 290 16.0 0.26 4.34 0.36 42.4 150.4<br />

Al2O3-0-8h 71 218 300 13.3 0.19 4.39 0.47 28.2 63.5<br />

Al2O3-0-24h 71 260 288 8.0 0.07 1.43 0.61 14.0 28.8<br />

Al2O3-0-200h 71 303 309 4.0 0.03 1.76 0.73 9.8 25.2<br />

Jic (HR)<br />

[kN/m]<br />

Al2O3-10-RT 86 185 318 14.8 0.20 2.21 0.37 5.0 10.4 10.5 3.2<br />

Al2O3-10-8h 86 281 334 9.4 0.12 3.39 0.56 4.1 6.5 16.0 3.2<br />

Al2O3-10-24h 86 320 354 3.5 0.07 1.58 0.63 4.8 6.4 18.2 3.2<br />

Al2O3-10-200h 86 342 361 2.8 0.03 0.79 0.69 2.5 4.5 19.4 3.2<br />

Al2O3-15-RT 95 224 333 7.5 0.15 1.22 0.43 2.9 6.5 11.1 2.9<br />

Al2O3-15-8h 95 318 368 4.0 0.07 0.79 0.59 3.0 6.4 15.8 3.0<br />

Al2O3-15-24h 95 351 388 1.1 0.04 0.18 0.65 1.2 5.7 17.4 3.0<br />

Al2O3-15-200h 95 351 373 1.1 0.03 0.39 0.70 1.7 4.4 17.4 2.8<br />

Al2O3-20-RT 104 228 324 6.3 0.14 1.44 0.45 2.7 6.5 10.3 2.6<br />

Al2O3-20-8h 104 329 372 1.5 0.06 0.57 0.60 1.8 3.7 14.8 2.8<br />

Al2O3-20-24h 104 349 379 1.1 0.05 0.79 0.64 1.2 3.6 15.7 2.8<br />

Al2O3-20-200h 104 365 381 1.0 0.03 0.65 0.71 1.3 2.5 16.5 2.6<br />

Jic (P)<br />

[kN/m]


4.1.1.3. Fracture mechanics tests<br />

Section 4<br />

Compact tension (CT) specimens with a thickness <strong>of</strong> B = 12.5 mm, a width <strong>of</strong> W = 40 mm,<br />

<strong>and</strong> an initial crack length <strong>of</strong> a0 ≈ 20 mm are machined for the <strong>fracture</strong> mechanics tests <strong>of</strong> the<br />

cast MMCs (Fig. 4.4). The specimens have a longitudinal-transverse (LT) crack plane<br />

orientation.<br />

To insert the pre-crack in the CT specimens, they are subjected to cyclic compression loading<br />

with ∆K = 10 MPa m <strong>and</strong> further to cyclic tensile loading with ∆K = 5 MPa m . Fracture<br />

mechanics tests are conducted using the “ZWICK” testing machine which is equipped by a<br />

special computer program for <strong>fracture</strong> mechanics tests. The cross-head velocity is <strong>of</strong> 0.02<br />

mm/min for the reinforced materials, <strong>and</strong> 0.08 mm/min for unreinforced materials. A potential<br />

drop method is employed to determine the crack extension during the <strong>fracture</strong> mechanics<br />

tests. In this method, a constant electric current is sent through the specimen. An increase <strong>of</strong><br />

the crack length changes the electric resistance <strong>of</strong> the system <strong>and</strong> the measured potential.<br />

From the change <strong>of</strong> the potential, the crack extension can be calculated by the Johnson<br />

equation (4.1) [69],<br />

⎡<br />

⎤<br />

⎢<br />

⎥<br />

⎢<br />

⎥<br />

⎢<br />

π<br />

cosh<br />

⎥<br />

2W<br />

⎢<br />

⎥<br />

a = ⋅ arccos<br />

2W<br />

⎢<br />

⎥ , (4.1)<br />

π<br />

⎢<br />

⎡<br />

πy<br />

⎤<br />

⎥<br />

⎢<br />

⎢ cosh<br />

U<br />

⎥<br />

cosh<br />

⎥<br />

⎢<br />

⎢ ⋅ arccos<br />

2W<br />

⎥<br />

⎥<br />

⎢<br />

⎢U<br />

πa<br />

o<br />

o<br />

cos ⎥<br />

⎥<br />

⎣ ⎢⎣<br />

2W<br />

⎥⎦<br />

⎦<br />

where 2y is the distance between the points <strong>of</strong> the potential measurement, U is the current<br />

potential value, <strong>and</strong> Uo is the initial potential value.<br />

A schematic view <strong>of</strong> the potential drop method is presented in Figure 4.5. The electrical<br />

current intensity, I, is 10 A. The potential, U, is determined by a nanovoltmeter with an<br />

accuracy <strong>of</strong> ± 100 nV, which corresponds to an accuracy <strong>of</strong> the crack length <strong>of</strong> ± 0.01 mm. A<br />

clip-gage is used to determine the load line displacement <strong>of</strong> the specimen during the <strong>fracture</strong><br />

tests <strong>of</strong> the cast MMCs. The data on the load, load line displacement, <strong>and</strong> potential are saved<br />

in a file by the computer program every 10 s. The J-integral resistance curves were<br />

determined according to [73]. The J0.2/Bl -, Ji-values determined from these curves are given in<br />

Table 4.2, as well.<br />

32


True stress [MPa]<br />

a)<br />

True stress [MPa]<br />

b)<br />

400<br />

300<br />

200<br />

100<br />

0<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Section 4<br />

0 4 8 12 16 20<br />

Strain [%]<br />

0 2 4 6 8 10<br />

Strain [%]<br />

Fig. 4.3. a) Experimental <strong>and</strong> fitted stress-strain curve for the <strong>matrix</strong> material Al-0-RT; b)<br />

experimental, fitted, <strong>and</strong> theoretically calculated by the Mori-Tanaka approach stress-strain<br />

curves for the Al6061-10-RT.<br />

33<br />

experiment<br />

power-law<br />

Mori-Tanaka<br />

experiment<br />

power-law


Section 4<br />

Fig. 4.4. A view <strong>of</strong> a compact tension specimen <strong>of</strong> the cast Al6061 MMC.<br />

Fig. 4.5. A schematic illustration <strong>of</strong> the potential drop method.<br />

4.1.2. Powder <strong>metal</strong>lurgy MMC<br />

4.1.2.1. Materials characterization<br />

A powder <strong>metal</strong>lurgy MMC with the Al6061 <strong>matrix</strong> reinforced by 10% SiC particles is also<br />

used for this investigation. The mean particle size is about 100 µm. The <strong>matrix</strong> powder size is<br />

about 40 µm. The extrusion ratio was 10:1. The PM MMC was supplied in the shape <strong>of</strong> a rod<br />

with a diameter 25 mm by IFAM (Germany).<br />

The MMC was annealed at 530°C for 1h <strong>and</strong> quenched in water. Specimens were subjected to<br />

following heat-treatments:<br />

I=const<br />

34<br />

Clip-gage<br />

U


(1.) aging at 175°C for 15min;<br />

(2.) aging at 175°C for 8h;<br />

(3.) aging at 175°C for24h;<br />

(4.) aging at 175°C for 200h.<br />

Section 4<br />

Fig. 4.6. The microstructure <strong>of</strong> the Al6061-10%SiC.<br />

Analogously to the cast MMCs, the investigated specimens <strong>of</strong> the PM MMC are referred to<br />

by the volume percentage <strong>and</strong> the heat treatment, e.g., Specimen SiC-10-15min., Specimen<br />

SiC-10-8h, etc.<br />

4.1.2.2. Compression tests<br />

Because <strong>of</strong> the low ductility <strong>of</strong> these materials, the compression tests are used to determine<br />

the mechanical <strong>properties</strong>. The difficulty is the prevention <strong>of</strong> buckling, which is avoided by<br />

the use <strong>of</strong> small gauge length / gauge diameter ratios (2 - 3) [2]. The compression test<br />

specimens have a st<strong>and</strong>ard cylindrical shape with a diameter <strong>of</strong> 8 mm <strong>and</strong> a length <strong>of</strong> 20 mm.<br />

The compression tests are conducted on the mechanical testing machine “ZWICK” at a<br />

loading rate 5.6·10 -4 s -1 . No buckling <strong>of</strong> the tested specimens is observed. The strain<br />

hardening coefficient, N, <strong>and</strong> the coefficient, α, are determined as in the case <strong>of</strong> the cast<br />

MMCs. The results <strong>of</strong> the compression tests are collected in Table 4.3.<br />

35<br />

200 µm


Section 4<br />

Table 4.3. Mechanical <strong>properties</strong> <strong>of</strong> the PM-MMC Al6061-10%SiC determined by<br />

compression tests.<br />

It should be noted that tensile tests were performed, as well. In Figure 4.7, the true stress-<br />

strain curves for Specimen SiC-10-15min determined by compression <strong>and</strong> tensile tests are<br />

given. It is seen that these curves coincides. But since the tensile specimen fails at very low<br />

<strong>fracture</strong> strain, the N- <strong>and</strong> α-values cannot be determined from the tensile true stress-strain<br />

curve.<br />

Material<br />

Stress [MPa]<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

E<br />

[Gpa]<br />

σy<br />

[MPa]<br />

σUTS<br />

[MPa]<br />

0 5 10 15 20 25<br />

Strain [%]<br />

N α<br />

SiC-0-15min 71 197 380 0.27 5.22 7.1<br />

SiC-0-8h 71 340 474 0.09 0.99 2.9<br />

SiC-0-24h 71 315 434 0.08 0.95 1.2<br />

SiC-0-200h 71 232 333 0.09 1.07 1.9<br />

36<br />

compression<br />

tension<br />

power-law<br />

Fig. 4.7. The true stress-strain curves for Specimen SiC-10-15min.<br />

Ji<br />

[kN/m]<br />

Jic (HR)<br />

[kN/m]<br />

SiC-10-15min 88 210 420 0.21 2.94 5.5 11.9 3.4<br />

SiC-10-8h 88 360 470 0.09 1.43 3.6 20.4 4.0<br />

SiC-10-24h 88 348 440 0.08 2.04 2.9 19.8 3.8<br />

SiC-10-200h 88 295 375 0.09 2.02 3.9 16.8 3.4<br />

Jic (P)<br />

[kN/m]


4.1.2.3. Fracture mechanics tests<br />

Section 4<br />

Disk-shaped compact (DCT) specimens with a thickness B = 10 mm, a width <strong>of</strong> W = 17 mm,<br />

<strong>and</strong> an initial crack length <strong>of</strong> ao ≈ 8 mm were machined for the <strong>fracture</strong> mechanics tests (Fig.<br />

4.8).<br />

To insert the pre-crack in specimens, they are subjected to cyclic compression loading with<br />

∆K = 10 MPa m <strong>and</strong> further to cyclic tensile loading with ∆K = 5 MPa m . Fracture<br />

mechanics tests are conducted as for the cast MMCs. Due to the small size <strong>of</strong> the disk-shaped<br />

compact specimens, a video-extensometer is employed to measure the load line displacement.<br />

The J-integral resistance curves are evaluated according to ASTM st<strong>and</strong>ards E1320-1 [70]. It<br />

should be noted that due to the low ductility <strong>of</strong> the PM MMCs, the J0.2/Bl -values cannot be<br />

determined. Therefore, only the values <strong>of</strong> the <strong>fracture</strong> toughness, Ji, are given in Table 4.3.<br />

Fig. 4.8. A disk-shaped compact specimen made <strong>of</strong> the Al6061-10%SiC PM-MMC.<br />

4.1.3. The relations between the <strong>global</strong> mechanical <strong>properties</strong><br />

It is seen from Table 4.2 that for the cast MMCs, the yield strength, σy, <strong>and</strong> the ultimate<br />

tensile strength, σUTS, increase with increasing aging conditions <strong>and</strong> particle volume fraction,<br />

whereas the values <strong>of</strong> the <strong>fracture</strong> initiation toughness, the strain hardening coefficient, <strong>and</strong><br />

the <strong>fracture</strong> strain show an opposite behavior.<br />

A dependency between the mechanical <strong>properties</strong> similar to that observed in [16, 21] are<br />

found for investigated cast MMCs: εfr <strong>and</strong> N linearly decrease with increasing yield strength,<br />

37


εfr [%]<br />

εfr [%]<br />

N<br />

16<br />

12<br />

8<br />

4<br />

0<br />

150 250 350 450<br />

16<br />

12<br />

8<br />

4<br />

0<br />

0,25<br />

0,2<br />

0,15<br />

0,1<br />

0,05<br />

0<br />

σ y [MPa]<br />

Section 4<br />

10%<br />

15%<br />

20%<br />

a) b)<br />

0 0,1 0,2 0,3<br />

Ν<br />

c) d)<br />

150 250 350 450<br />

σy [MPa]<br />

10%<br />

15%<br />

20%<br />

10%<br />

e) f)<br />

Fig. 4.9. The relation between the mechanical <strong>properties</strong> <strong>of</strong>: a-b) the cast MMCs, c-d) the cast<br />

MMCs, e-f) the PM-MMC (to be continued).<br />

38<br />

N<br />

J 0.2 [kN/m]<br />

J i [kN/m]<br />

0,25<br />

0,2<br />

0,15<br />

0,1<br />

0,05<br />

0<br />

12<br />

8<br />

4<br />

150 250 350 450<br />

σy [MPa]<br />

0<br />

150 250 350 450<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

σ y [MPa]<br />

0<br />

150 200 250 300 350 400<br />

σ y [MPa]<br />

10%<br />

15%<br />

20%<br />

10%<br />

15%<br />

20%<br />

10%


(HR)<br />

Ji [kN/m]<br />

Section 4<br />

g) h)<br />

Fig. 4.9. The relation between the mechanical <strong>properties</strong>: g-h) the correlation between the<br />

Ji exp -, Ji (HR) -, <strong>and</strong> Ji (P) -values for the cast MMCs.<br />

σy (Figs 4.9a,b). The dependence <strong>of</strong> εfr on N has an inverse character (Fig. 4.9c). The <strong>fracture</strong><br />

toughness has an inverse dependence on the yield strength (Fig. 4.9d). As one can see from<br />

Table 4.3, in the PM-MMCs, σy increases with increasing aging time from 15min. to 8h <strong>and</strong><br />

slightly decreases with further increasing aging time. A <strong>fracture</strong> toughness shows the opposite<br />

behavior. The <strong>global</strong> mechanical <strong>properties</strong> demonstrate dependency to each other similar to<br />

the cast MMCs, excepting the values corresponding to the material aged for 8h (Figs 4.7e,f).<br />

The values <strong>of</strong> the <strong>fracture</strong> initiation toughness, Ji, from Tables (4.2, 4.3) can be compared to<br />

theoretical values estimated according to preliminary proposed models (see Section 2.2). For<br />

our materials, a significant difference between the Ji-values estimated according to Hahn-<br />

Rosenfield model, Jic (HR) , <strong>and</strong> the experimental values is found. It is seen that the Hahn-<br />

Rosenfield prediction tends to significantly overestimate the <strong>fracture</strong> toughness by a factor <strong>of</strong><br />

2…15 (Fig. 4.9g, Tables 4.2, 4.3) (as was reported in [17]). The values estimated by the<br />

P<strong>and</strong>ey et al. model, Jic (P) , show a better correlation with the experimentally determined<br />

<strong>fracture</strong> toughness, Ji exp (Fig. 4.9h, Tables 4.2, 4.3). They underestimate slightly the <strong>fracture</strong><br />

toughness for the cast MMCs with 10% <strong>of</strong> Al2O3 particles. For the aged cast MMCs with 20%<br />

<strong>of</strong> Al2O3 particles, the Jic (P) -values are higher <strong>of</strong> the Ji exp -values approximately by a factor <strong>of</strong><br />

1.5…2.<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

0 4 8 12 16 20<br />

exp<br />

J i [kN/m]<br />

39<br />

(P)<br />

Ji [kN/m]<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

0 4 8 12 16 20<br />

exp<br />

J i [kN/m]


Section 4<br />

4.2. The effect <strong>of</strong> the heat treatment on the mechanism <strong>of</strong> void initiation<br />

The <strong>fracture</strong> surfaces <strong>of</strong> the MMCs are studied in detail in the SEM LEO 440. The regions<br />

near the crack front in the mid-section <strong>of</strong> the specimens, where plane-strain condition<br />

prevails, are investigated. Ductile <strong>fracture</strong> by micro-void coalescence is observed in all<br />

materials.<br />

The images <strong>of</strong> the analysed regions for all investigated materials are given in Appendix 9<br />

(Figs. 9.1-9.25). For each investigated material, from 10 to 20 particles located closely in<br />

front <strong>of</strong> the crack tip are investigated. For each particle, the mechanism <strong>of</strong> void initiation<br />

either particle <strong>fracture</strong> or particle/<strong>matrix</strong> decohesion is determined from stereo image pairs.<br />

For instance, when the particle is observed only on a single half <strong>of</strong> the broken specimen, it is<br />

probable that the particle/<strong>matrix</strong> decohesion has occurred. On the contrary, if broken parts <strong>of</strong><br />

a particle are found on both specimen halves, particle <strong>fracture</strong> can be assumed.<br />

The study <strong>of</strong> the <strong>fracture</strong> surfaces <strong>of</strong> the casdt MMCs reveals a change <strong>of</strong> the mechanism <strong>of</strong><br />

void nucleation with increasing aging condition: a clear tendency to increasing probability <strong>of</strong><br />

particle/<strong>matrix</strong> decohesion mechanism can be noted (see Table 4.8 on p. 65). For instance, in<br />

Specimen Al2O3-10%-200h, 12 <strong>of</strong> 15 considered particles formed dimple by particle/<strong>matrix</strong><br />

decohesion, whereas in Specimen Al2O3-10%-RT, 13 <strong>of</strong> 16 considered particles are <strong>fracture</strong>d.<br />

Although only particles near the crack front are considered in this statistic, these data reflect<br />

the situation on whole <strong>fracture</strong> surface, as well.<br />

In Fig. 4.10a, a typical broken alumina particle is shown (Particle 4 <strong>of</strong> Specimen Al2O3-10-RT<br />

from Fig. 9.1 taken at higher magnification). It is seen that the particle <strong>fracture</strong> surface has a<br />

smooth view. Fig. 4.10b shows a debonded particle 13 in Specimen Al2O3-10-200h. The<br />

particle/<strong>matrix</strong> interface is rough <strong>and</strong> bumpy: very small dimples having a size less than 1 µm<br />

are observed at the debonded interface. These small dimples are initiated by the MgAl2O4<br />

spinel phases formed at the alumina particle/<strong>matrix</strong> interface [21, 71-72]. It has been<br />

suggested that the spinel phase can be formed at the interface both during the fabrication <strong>of</strong><br />

MMCs <strong>and</strong> during the following heat treatment. The following possible reactions with large<br />

enough thermodynamic driving forces to form the MgAl2O4 phase have been proposed in<br />

[71]:<br />

40


Section 4<br />

a)<br />

b)<br />

Figure 4.10. a) A <strong>fracture</strong>d particle on the <strong>fracture</strong> surface <strong>of</strong> the Specimen Al2O3-10-RT; b) a<br />

debonded particle/<strong>matrix</strong> interface on the <strong>fracture</strong> surface <strong>of</strong> the Specimen Al2O3-10-200h.<br />

Mg + 2Al + 2O2 = MgAl2O4<br />

MgO + Al2O3 = MgAl2O4<br />

Mg + 4/3Al2O3 = MgAl2O4 + 2/3Al<br />

2SiO2 + 2Al + Mg = MgAl2O4 + 2Si.<br />

In Figure 4.10b, some spinel phases are marked by dashed arrows.<br />

41<br />

3µm<br />

3µm


Section 4<br />

It should be noted that no dependence <strong>of</strong> the mechanism <strong>of</strong> the void initiation on the heat<br />

treatment condition is found for the PM MMC. In Fig. 4.11, it is seen that the broken SiC<br />

particles are observed on both halves <strong>of</strong> the broken Specimen SiC-10-200h (Fig. 4.11)<br />

(considered particles are marked <strong>and</strong> numbered). As well known, in MMCs reinforced by<br />

large ceramic particles, these particles are <strong>fracture</strong>d independently <strong>of</strong> the heat treatment<br />

conditions [2].<br />

Fig. 4.11. Fracture surface with marked particles <strong>of</strong> the Specimen SiC-10-200h.<br />

4.3. The effect <strong>of</strong> the particle location with respect to the crack tip on the <strong>local</strong> <strong>fracture</strong><br />

<strong>properties</strong>.<br />

To study the effect <strong>of</strong> the particle location with respect to the crack plane, the CODi- <strong>and</strong><br />

CODvi-values, the distance between crack tip <strong>and</strong> particle, r, <strong>and</strong> the angle <strong>of</strong> the particle<br />

42


Section 4<br />

location with respect to the crack tip, θ, as well as the particle size in two directions (parallel<br />

<strong>and</strong> perpendicular to the crack front) are measured for each investigated particle. The results<br />

<strong>of</strong> the measurements are collected in Tables 9.1-9.36.<br />

4.3.1. The relation between the particle location with respect to the crack tip <strong>and</strong> the CODi-<br />

values.<br />

The CODi-values show extremely high scatter in all investigated MMCs. In Figure 4.11, a<br />

distribution <strong>of</strong> the <strong>local</strong> CODi-values along the crack tip is presented for Specimen Al2O3-10-<br />

RT as an example. The top part <strong>of</strong> the top picture (corresponding to the first half <strong>of</strong> the broken<br />

CT specimen) <strong>and</strong> the bottom part <strong>of</strong> the bottom picture (corresponding to the second half <strong>of</strong><br />

the broken CT specimen) belong to ductile <strong>fracture</strong> region. To determine the CODi-<br />

distribution curve along the crack tip, the distribution <strong>of</strong> the stretch zone height along the<br />

crack tip was estimated for both halves <strong>of</strong> the broken specimen as a difference between the<br />

pr<strong>of</strong>ile along the stretched zone <strong>and</strong> micro-ductile <strong>fracture</strong> passed on the crack tip (marked by<br />

the red lines in Fig. 4.12) <strong>and</strong> the pr<strong>of</strong>ile at the end <strong>of</strong> the pre-fatigue region (blue lines in Fig.<br />

4.12). The obtained curves from both specimen halves are summarized. One can see that the<br />

<strong>local</strong> CODi-values at the regions between particles are <strong>of</strong>ten higher than the CODi estimated<br />

directly at investigated particles.<br />

It should be noted that the CODi-values determined by this procedure <strong>and</strong> the procedure<br />

described in Section 4.2 have the same values. In Figure 4.13, a crack pr<strong>of</strong>ile passed<br />

perpendicular to the crack front (marked by green lines in Figure 4.12) are presented. The<br />

CODi-value is equal to 14 µm. The same <strong>local</strong> CODi-value follows from Figure 4.12. Further,<br />

we will consider only the <strong>local</strong> CODi-values for individual particles determined by the<br />

procedure described in Section 3.2, since the polar coordinates <strong>of</strong> the particle distribution can<br />

be determined from these pr<strong>of</strong>iles (Section 4). In Figure 4.14, corresponding regions near the<br />

midsection on the two specimen halves <strong>of</strong> the Specimen Al2O3-10%-RT with drawn<br />

corresponding pr<strong>of</strong>iles are given as an example (the pre-fatigue region is at the midsection <strong>of</strong><br />

the page). The results <strong>of</strong> the analysis for each considered particle are collected in Table 4.4.<br />

From the data <strong>of</strong> the individual particles (Tables 9.1-9.17), it can be checked if the polar<br />

coordinates influence the CODi-values. The particle distance from the crack tip, r, <strong>and</strong> the<br />

particle location angle, θ, are considered for the investigated MMCs.<br />

43


CODi [µm]<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Section 4<br />

14µm<br />

-70 -50 -30 -10 10<br />

x [µm]<br />

30 50 70 90<br />

Fig. 4.12. The <strong>local</strong> CODi distribution along the crack front for Specimen Al2O3-10-RT<br />

44


height [µm]<br />

Section 4<br />

-80 -60 -40<br />

CODi =14µm<br />

-20<br />

0<br />

0<br />

-10<br />

20 40<br />

45<br />

30<br />

20<br />

10<br />

-20<br />

distance [µm]<br />

Particle<br />

Fig. 4.13. Crack pr<strong>of</strong>ile perpendicular to the crack front in Fig. 4.12.<br />

Fig. 4.14. Corresponding stereopaires with passes pr<strong>of</strong>iles for the Specimen Al2O3-10-RT.


Section 4<br />

Table 4.4. Results <strong>of</strong> the stereophotogrammetic analysis for the Specimen Al2O3-10-RT<br />

(Fig. 4.14).<br />

Figure 4.15 collects all the CODi-values <strong>of</strong> the cast MMCs specimens plotted against the<br />

particle location angle, θ. The data points scatter is seen between θ = 0° <strong>and</strong> 70°; no<br />

dependency <strong>of</strong> the CODi-values from the particle location angle can be recognized. CODi<br />

plotted against the particle distance, r, yields a similar picture, where r varies between 5.5 <strong>and</strong><br />

82 µm (Fig. 4.15).<br />

Particle Size<br />

[µm]<br />

r<br />

[µm]<br />

θ<br />

[°]<br />

The CODi-values in the PM MMCs demonstrate a high scatter, as well. In comparison with<br />

the CODi-values <strong>of</strong> the cast MMCs, they have considerably higher values (Tabs. 9.13-9.17).<br />

Again, no any dependency <strong>of</strong> the CODi-values on the particle location is found (Figs.<br />

4.17a,b). Near-zero CODi-values are found for particles located at high angles, θ ≥ 45°. The<br />

reasons for that are explained in next section. It should be noted that no any effect <strong>of</strong> the<br />

particle size in any direction is revealed for all cast <strong>and</strong> PM-MMCs.<br />

In Table 4.5, the average CODi- <strong>and</strong> CODvi-values for all cast <strong>and</strong> PM materials are given.<br />

The average values given on the first line are calculated without taking into account the near-<br />

zero values, <strong>and</strong> the values on the second line with it. One can see that the average values<br />

decrease with increasing aging conditions. For instance, for the PM MMC, an increase <strong>of</strong> the<br />

aging time from 15 minutes to 8 hours leads to decrease <strong>of</strong> both average CODi- <strong>and</strong> CODvi-<br />

values by a factor <strong>of</strong> 3. Only slight increase <strong>of</strong> these values after aging for 200h in comparison<br />

with the condition 24h can be noted for the cast MMCs. For the cast MMCs, the<br />

46<br />

CODvi<br />

[µm]<br />

CODi<br />

[µm]<br />

Mecha<br />

nicsm<br />

1 3x12 15.6 40 12 16.5 Dec.<br />

2 8x8 17.2 31 6 6.5 Fr.<br />

3 12x3 43.2 35 7 12 Fr.<br />

4 20x9 30.5 32 7 9 Fr.<br />

5 8x4 27.1 55 5 5 Fr.<br />

6 23x4 33.8 71 2 13 Fr.<br />

7 15x6 30.4 47 ≈0 ≈0 Fr.<br />

8 10x3 11.0 55 ≈0 ≈0 Fr.<br />

9 17x4 11.7 20 6 11.5 Fr.<br />

10 4x2 18 9 10 17 Dec.<br />

11 12x7 13.2 11 9.5 16 Fr.<br />

12 4x2 11.9 45 1.5 1.5 Fr.


COD i [µm]<br />

COD i [µm]<br />

CODi [µm]<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

Section 4<br />

0 20 40 60 80<br />

θ [°]<br />

a)<br />

0 20 40 60 80<br />

θ [µm]<br />

b)<br />

0 20 40 60 80<br />

θ [°]<br />

47<br />

10 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

15 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

20 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

c)<br />

Fig. 4.15. The CODi-values vs. the angle between crack tip <strong>and</strong> particle, θ, for different aging<br />

conditions <strong>of</strong> the cast MMCs: Specimens with (a) 10%, (b) 15%, (c) 20% Al2O3.


CODi [µm]<br />

COD i [µm]<br />

CODi [µm]<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

Section 4<br />

0 20 40 60 80 100<br />

r [µm]<br />

a)<br />

0 20 40 60 80 100<br />

r [µm]<br />

b)<br />

0 20 40 60 80 100<br />

r [µm]<br />

48<br />

10 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

15 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

20 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

c)<br />

Fig. 4.16. The CODi-values vs. the distance between crack tip <strong>and</strong> particle, r, for different<br />

aging conditions <strong>of</strong> the cast MMCs: Specimens with (a) 10%, (b) 15%, (c) 20% Al2O3.


COD i [µm]<br />

COD i [µm]<br />

80<br />

60<br />

40<br />

20<br />

0<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Section 4<br />

0 100 200 300 400 500<br />

r [µm]<br />

0 20 40 60 80<br />

θ [°]<br />

a)<br />

b)<br />

49<br />

15 min<br />

8h<br />

24h<br />

200h<br />

15 min<br />

8h<br />

24h<br />

200h<br />

Fig. 4.17. The CODi -values vs.: (a) the distance between the particle <strong>and</strong> the crack tip, r; (b)<br />

the angle between crack tip <strong>and</strong> particle, θ, for different aging conditions <strong>of</strong> the PM MMC.


Cast MMCs<br />

10%-RT<br />

10%-8h<br />

10%-24h<br />

10%-200h<br />

15%-RT<br />

15%-8h<br />

15%-24h<br />

15%-200h<br />

20%-RT<br />

20%-8h<br />

20%-24h<br />

20%-200h<br />

Section 4<br />

Table 4.5. Average CODi- <strong>and</strong> CODvi-values for investigated materials.<br />

CODi-values have a tendency to decrease with increasing particle volume fraction, as well.<br />

Fracture initiation occurs by void initiation at a particle close to the crack tip <strong>and</strong> void growth<br />

until the void <strong>and</strong> the blunted crack coalesce. It is reasonable, therefore, to investigate, as a<br />

next step, the relation between the <strong>local</strong> conditions for void initiation <strong>and</strong> the <strong>local</strong><br />

microstructure.<br />

average<br />

CODi-values<br />

[µm]<br />

12.2±5.1<br />

10.5±6.3<br />

9.1±2.9<br />

10.1±3.7<br />

9.5±4.3<br />

7.1±4.0<br />

6.4±4.3<br />

8.1±3.5<br />

6.4±4.6<br />

8.9±3.9<br />

4.0±2.3<br />

3.4±2.6<br />

5.3±2.6<br />

6.1±4.5<br />

7.3±2.8<br />

6.2±3.7<br />

3.5±1.6<br />

3.2±1.9<br />

4.2±3.7<br />

3.9±2.2<br />

average<br />

CODvi-values<br />

[µm]<br />

6.3±2.8<br />

5.5±3.4<br />

7.2±3.1<br />

3.9±2.0<br />

3.4±2.3<br />

7.1±3.7<br />

6.6±4.0<br />

5.6±1.8<br />

4.4±2.9<br />

4.0±1.6<br />

2.3±1.0<br />

1.1±1.4<br />

3.6±1.8<br />

2.4±1.5<br />

1.5±1.7<br />

2.8±1.4<br />

2.3±1.6<br />

1.8±1.2<br />

0.9±1.2<br />

3.3±1.8<br />

3.0±1.9<br />

50<br />

PM MMCs<br />

average<br />

CODi-values<br />

[µm]<br />

average<br />

CODvi-values<br />

[µm]<br />

10-15min 33.9±19.0 27.5±19.5<br />

10-8h 13.4±7.8 11.2±4.6<br />

10-24h 15.0±6.8 9.7±5.4<br />

10-200h 14.8±6.8 10.9±6.6<br />

St37 66.4±13.0 35.8±11.0


Section 4<br />

4.3.2. The relation between the particle location with respect to the crack tip <strong>and</strong> the CODvi-<br />

values<br />

As seen from Figures 4.18 - 4.20, the scatter <strong>of</strong> the CODvi-values is still high in all<br />

investigated specimens, but less than those <strong>of</strong> the CODi-values. The maximum CODvi-values<br />

seem to decrease with increasing aging conditions <strong>and</strong> with increasing particle volume<br />

fraction, leading to a narrowing <strong>of</strong> the scatter. Only for the cast MMCs, the CODi-values for<br />

the Specimens 160°C-200h are slightly higher in comparison with Specimens 160°C-24h.<br />

Contrary to the CODi-values, a slight tendency to decreasing CODvi-values with increasing<br />

particle location angle, θ, can be noted for the cast MMCs (Fig. 4.18). Whereas small CODvi-<br />

values appear at large angles between θ = 50 <strong>and</strong> 70° in the RT-condition, they are more<br />

equally distributed for the materials aged at 160°C.<br />

Very interesting are the crack pr<strong>of</strong>iles through alumina particles in cast MMCs with large<br />

particle location angle θ. For such particles, a step appears on both <strong>fracture</strong> surfaces, <strong>and</strong> the<br />

CODvi- <strong>and</strong> CODi-values are measured as the height differences <strong>of</strong> the two steps. Figure 4.21<br />

presents an example <strong>of</strong> the pr<strong>of</strong>iles through Particle 7 <strong>of</strong> the Specimen Al2O3-10-RT, located<br />

at θ = 47°. Here, void initiation <strong>and</strong> <strong>fracture</strong> initiation occur simultaneously at a very low<br />

value <strong>of</strong> CODvi ≈ CODi ≈ 0. One can conclude that Particle 7 <strong>fracture</strong>s at a low external<br />

loading, <strong>and</strong> the crack immediately extends.<br />

No dependency <strong>of</strong> the CODvi-values on the distance to the crack tip, r, is observed in both the<br />

cast <strong>and</strong> PM MMCs. The CODvi vs. r curves show a scatter similar to those <strong>of</strong> Figures 4.16,<br />

4.17a (Figs. 4.19, 4.20a). It should be noted that no near-zero CODvi-values are observed for<br />

the PM MMCs in comparison with the cast MMCs. What might be a reason for such different<br />

effect <strong>of</strong> the angle θ on the CODvi-values in different materials?<br />

It is plausible that the void initiation might be strongly influenced by plastic straining.<br />

Therefore, it can suggested that this finding has to do with the shape <strong>of</strong> the plastic zone which<br />

has its maximum extension at an angle <strong>of</strong> about θ = 70° with respect to the crack plane (Fig.<br />

4.22). For plane strain conditions, the plastic zone size in that direction, ω70, is by a factor 4.4<br />

larger than in the crack plane [73, 74]. As was shown in [74], ω70 can be determined directly<br />

from the COD as<br />

2<br />

K<br />

J E<br />

ω70 =<br />

0.<br />

157 ≈ 0.<br />

16 ≈<br />

σ<br />

σ<br />

2<br />

y<br />

51<br />

2<br />

y<br />

0.<br />

16<br />

d<br />

N<br />

E σ COD<br />

o<br />

2<br />

y<br />

σ<br />

(4.2)


COD vi [µm]<br />

COD vi [µm]<br />

CODvi [µm]<br />

15<br />

10<br />

5<br />

0<br />

15<br />

10<br />

5<br />

0<br />

15<br />

10<br />

5<br />

0<br />

Section 4<br />

0 20 40 60 80<br />

θ [ o ]<br />

a)<br />

0 20 40 60 80<br />

θ [°]<br />

b)<br />

0 20 40 60 80<br />

θ [ o ]<br />

52<br />

10 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

15 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

20 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

c)<br />

Fig. 4.18. The CODvi-values vs. the angle between crack tip <strong>and</strong> particle, θ, for different aging<br />

conditions <strong>of</strong> the cast MMCs: Specimens with (a) 10%, (b) 15%, (c) 20% Al2O3.


CODvi [µm]<br />

CODvi [µm]<br />

CODvi [µm]<br />

15<br />

10<br />

5<br />

0<br />

15<br />

10<br />

5<br />

0<br />

15<br />

10<br />

5<br />

0<br />

Section 4<br />

0 20 40 60 80 100<br />

r [µm]<br />

0 20 40 60 80 100<br />

a)<br />

r [µm]<br />

0 20 40 60 80 100<br />

b)<br />

r [µm]<br />

c)<br />

53<br />

10 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

15 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

20 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

Fig. 4.19. The CODvi-values vs. the distance between crack tip <strong>and</strong> particle, r, for different<br />

aging conditions <strong>of</strong> the cast MMCs: Specimens with (a) 10%, (b) 15%, (c) 20% Al2O3.


CODvi [µm]<br />

CODvi [µm]<br />

80<br />

60<br />

40<br />

20<br />

0<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Section 4<br />

0 100 200 300 400 500<br />

r [µm]<br />

a)<br />

0 20 40 60 80<br />

θ [°]<br />

b)<br />

54<br />

15 min<br />

8h<br />

24h<br />

200h<br />

15 min<br />

8h<br />

24h<br />

200h<br />

Fig. 4.20. The CODvi-values vs. (a) the distance between the particle <strong>and</strong> the crack tip, (b) the<br />

angle between crack tip <strong>and</strong> particle, θ, for different aging conditions <strong>of</strong> the PM-MMC.


z [µm]<br />

30<br />

20<br />

10<br />

0<br />

-10<br />

-20<br />

COD vi 0 µm<br />

Θ<br />

r<br />

particle<br />

Section 4<br />

-30<br />

x [µm]<br />

-80 -40 0 40 80<br />

Fig. 4.21. Crack pr<strong>of</strong>ile through the Particle 7 <strong>of</strong> Specimen Al2O3-10-RT in the moment <strong>of</strong><br />

void initiation: CODvi ≈ CODi ≈ 0.<br />

The values calculated by Eq. 4.2 for the specimens Al2O3-10-RT <strong>and</strong> SiC-10-24h in<br />

comparison with observed distances between particles <strong>and</strong> crack tip are given in Table 4.6.<br />

One can see that even for a very low COD <strong>of</strong> 0.2 µm, the maximum extension <strong>of</strong> the plastic<br />

zone for the cast MMC becomes comparable to the distances to nearest particles. As a result,<br />

the <strong>fracture</strong> <strong>of</strong> particles, that lie in the direction <strong>of</strong> the maximum extension <strong>of</strong> the plastic zone,<br />

can occur at very low COD. At the same time, for the PM MMC, the distance between the<br />

SiC particles <strong>and</strong> the crack tip, r, is significantly higher in comparison with ω70, i.e., the<br />

Fig. 4.22. The plastic zone shape for plane strain condition.<br />

55


Section 4<br />

Table 4.6. The values <strong>of</strong> the plastic zone size at angles θ =0°, 70° <strong>and</strong> the distance between<br />

particles <strong>and</strong> the crack tip at COD = 0.2 µm<br />

Material ω70 [µm] ω0 [µm] r[µm]<br />

cast Al2O3-10-RT 40 9.1 10…50<br />

PM SiC-10-24h 12 2.7 50…300<br />

maximum extension <strong>of</strong> the plastic zone at near-zero values is not comparable with the<br />

distance to the nearest SiC particle. So no near-zero CODvi-values can be observed for this<br />

material.<br />

4.4. The maximum particle stresses at the moment <strong>of</strong> void initiation<br />

The stress tensors for both the particles <strong>and</strong> the <strong>matrix</strong> were calculated for investigated MMCs<br />

according to the procedure described in Section 4. In Table 4.7, the results <strong>of</strong> calculation for<br />

Particle 4 <strong>of</strong> Specimen Al2O3-10-RT (Fig. 9.1, Tab. 9.1) are presented as example. Figures<br />

4.23-4.25, the maximum particle stresses at the moment <strong>of</strong> the void initiation are plotted<br />

against the angle to the crack plane, θ, <strong>and</strong> the distance to the crack tip, r, for the cast <strong>and</strong> PM<br />

MMCs. In comparison to the very large scatter <strong>of</strong> the CODi- <strong>and</strong> CODvi- values (Figs. 4.15-<br />

4.20), the scatter <strong>of</strong> the σ p max-values is much lower. The stresses vary between 800 <strong>and</strong> 1400<br />

MPa for the alumina particles in the cast MMCs <strong>and</strong> between 700 <strong>and</strong> 1100 MPa for SiC<br />

particles in the PM MMCs. The σ p max-values for alumina particles are nearly independent <strong>of</strong><br />

the angle to the crack plane, θ, (Fig. 4.23). A slight decrease may be found for particles which<br />

lie at angles larger than θ ≈ 55°. A slight decrease <strong>of</strong> the σ p max-values with increasing r is<br />

observed for the cast MMCs in Figure 4.24.<br />

For the PM MMCs, the stress tensors were calculated only for specimens SiC-10%-15 min.<br />

<strong>and</strong> SiC-10%-200h. In other specimens, the particle <strong>fracture</strong> occurs in the elastic region,<br />

where the mean-field approach fails <strong>and</strong> cannot be used to determine the stress states in<br />

constituent phases.<br />

56


Table 4.7. The detailed results <strong>of</strong> the calculation procedure for Particle 4 <strong>of</strong> Specimen Al2O3-<br />

10-RT . The particle coordinates are r = 30.5 µm, θ = 30°, <strong>and</strong> the COD at void initiation is<br />

CODvi = 7 µm. The particle <strong>and</strong> <strong>matrix</strong> stress <strong>and</strong> strain tensors are evaluated via an iterative<br />

Stress tensor [MPa]<br />

procedure. The final value <strong>of</strong> the secant modulus <strong>of</strong> the <strong>matrix</strong> was Es m = 14.46 GPa.<br />

⎛516<br />

⎜<br />

⎜ 37<br />

⎜<br />

⎝ 0<br />

Mesoscopic Particle Matrix<br />

37<br />

735<br />

0<br />

0 ⎞<br />

⎟<br />

0 ⎟<br />

623⎟<br />

⎠<br />

⎛385<br />

⎜<br />

⎜ 79<br />

⎜<br />

⎝ 0<br />

79<br />

845<br />

0<br />

0 ⎞<br />

⎟<br />

0 ⎟<br />

609⎟<br />

⎠<br />

⎛531<br />

33 0 ⎞<br />

⎜<br />

⎟<br />

⎜ 33 723 0 ⎟<br />

⎜<br />

⎟<br />

⎝ 0 0 624⎠<br />

⎛σ<br />

1 ⎞<br />

Principal stresses ⎜ ⎟<br />

⎜σ<br />

2 ⎟<br />

⎜ ⎟<br />

⎝σ<br />

3 ⎠<br />

[MPa]<br />

⎛ 741⎞<br />

⎜ ⎟<br />

⎜623⎟<br />

⎜ ⎟<br />

⎝510⎠<br />

⎛858⎞<br />

⎜ ⎟<br />

⎜609⎟<br />

⎜ ⎟<br />

⎝372⎠<br />

⎛728⎞<br />

⎜ ⎟<br />

⎜624⎟<br />

⎜ ⎟<br />

⎝525⎠<br />

Maximal Shear Stress<br />

[MPa]<br />

116 243 101<br />

Equivalent stress [MPa] 200 421 176<br />

Mean stress [MPa] 625 613 626<br />

Stress triaxiality 3.1 1.5 3.6<br />

Strain tensor<br />

⎛− 0 . 0088<br />

⎜<br />

⎜ 0.<br />

0031<br />

⎜<br />

⎝ 0<br />

0.<br />

0031<br />

0.<br />

0092<br />

0<br />

0 ⎞<br />

⎟<br />

0 ⎟<br />

-8<br />

- 7x10 ⎟<br />

⎠<br />

⎛ ×<br />

⎜<br />

⎜<br />

⎜<br />

⎝<br />

−5<br />

8 10<br />

0.<br />

00023<br />

0<br />

0.<br />

00023<br />

0.<br />

00141<br />

0<br />

0 ⎞<br />

⎟<br />

0 ⎟<br />

⎟<br />

0.<br />

00073⎠<br />

⎛−0.<br />

0098<br />

⎜<br />

⎜ 0.<br />

0034<br />

⎜<br />

⎝ 0<br />

0.<br />

0034<br />

0.<br />

0101<br />

Maximal shear strain 0.0095 0.0007 0.0105<br />

Equivalent strain 0.0106 0.0009 0.0117<br />

0<br />

0 ⎞<br />

⎟<br />

0 ⎟<br />

−5<br />

−8×<br />

10 ⎟<br />


σ p<br />

max [MPa]<br />

σ p max [MPa]<br />

σ p<br />

max [MPa]<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

Section 4<br />

0 20 40 60 80<br />

θ [ o ]<br />

a)<br />

0 20 40 60 80<br />

θ [ o ]<br />

b)<br />

0 20 40 60 80<br />

θ [ o ]<br />

58<br />

10 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

15 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

20 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

c)<br />

Fig. 4.23. The σ p max-values vs. the distance between crack tip <strong>and</strong> particle, r, for different<br />

aging conditions <strong>of</strong> the cast MMCs: Specimens with (a) 10%, (b) 15%, (c) 20% Al2O3.


p max [MPa]<br />

σ p<br />

max [MPa]<br />

σ p<br />

max [MPa]<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

Section 4<br />

0 20 40 60 80 100<br />

r [µm]<br />

a)<br />

0 20 40 60 80 100<br />

r [µm]<br />

b)<br />

0 20 40 60 80 100<br />

r [µm]<br />

59<br />

10 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

15 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

20 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

c)<br />

Fig. 4.24. The σ p max-values vs. the distance between crack tip <strong>and</strong> particle, r, for different<br />

aging conditions <strong>of</strong> the cast MMCs: Specimens with (a) 10%, (b) 15%, (c) 20% Al2O3.


σ p<br />

max [MPa]<br />

σ p<br />

max [MPa]<br />

1200<br />

800<br />

400<br />

1200<br />

800<br />

400<br />

0<br />

0<br />

Section 4<br />

0 100 200 300 400 500<br />

r [µm]<br />

a)<br />

0 20 40 60 80<br />

θ [°]<br />

60<br />

15 min<br />

200h<br />

15 min<br />

200h<br />

b)<br />

Fig. 4.25. The σ p max-values vs. (1) the distance between crack tip <strong>and</strong> particle, r, (2) the angle<br />

between the crack tip <strong>and</strong> the particle for different aging conditions <strong>of</strong> the PM-MMC.<br />

For the present model, i.e., spherical particles <strong>and</strong> isotropic material behavior, the maximum<br />

normal interface tractions are identical to the maximum principal stresses in the particle [88].<br />

However, the interfacial stresses for the Al2O3 <strong>and</strong> SiC particles in the MMCs can be<br />

estimated according to Argon et al (Eq. 2.3), as well. These values are also listed in Tables<br />

9.1-9.17. It is seen that they are slightly less in comparison with the σ p max-values. The<br />

dependence <strong>of</strong> these interfacial stress values on the r <strong>and</strong> θ is similar to dependence <strong>of</strong> the<br />

σ<br />

p max-values (Figs. 4.25 - 4.27).


σ interf<br />

max [MPa]<br />

σ interf max [MPa]<br />

σ interf<br />

max [MPa]<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

Section 4<br />

0 20 40 60 80<br />

θ [ o ]<br />

a)<br />

0 20 40 60 80<br />

θ [ o ]<br />

b)<br />

0 20 40 60 80<br />

θ [ o ]<br />

61<br />

10 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

15 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

20 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

c)<br />

Fig. 4.26. The σ interf max-values vs. the angle between crack tip <strong>and</strong> particle, θ, for different<br />

aging conditions <strong>of</strong> the cast MMCs: Specimens with (a) 10%, (b) 15%, (c) 20% Al2O3.


σ interf<br />

max [MPa]<br />

σ interf max [MPa]<br />

σ interf<br />

max [MPa]<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

1600<br />

1200<br />

800<br />

400<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

0<br />

Section 4<br />

0 20 40 60 80 100<br />

r [µm]<br />

a)<br />

0 20 40 60 80 100<br />

r [µm]<br />

b)<br />

0 20 40 60 80 100<br />

r [µm]<br />

62<br />

10 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

15 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

20 %<br />

RT<br />

8h<br />

24h<br />

200h<br />

c)<br />

Fig. 4.27. The σ interf max-values vs. the distance between crack tip <strong>and</strong> particle, r, for different<br />

aging conditions <strong>of</strong> the cast MMCs: Specimens with (a) 10%, (b) 15%, (c) 20% Al2O3.


σ interf<br />

max [MPa]<br />

σ interf<br />

max [MPa]<br />

1200<br />

800<br />

400<br />

0<br />

1200<br />

800<br />

400<br />

0<br />

Section 4<br />

0 100 200 300 400 500<br />

r [µm]<br />

a)<br />

0 20 40 60 80<br />

θ [°]<br />

b)<br />

63<br />

15 min<br />

200h<br />

15 min<br />

200h<br />

Fig. 4.28. The σ interf max-values vs. (a) the distance between crack tip <strong>and</strong> particle, r, (b) the<br />

angle between the crack tip <strong>and</strong> the particle for different aging conditions <strong>of</strong> the PM-MMC.


Section 4<br />

4.5. The relation between the <strong>local</strong> conditions for void initiation <strong>and</strong> <strong>global</strong> mechanical<br />

<strong>properties</strong><br />

The average values <strong>of</strong> the maximum principal stresses in the particles are calculated for all<br />

specimens <strong>and</strong> inserted in Tables 4.8 <strong>and</strong> 4.9. For calculating the mean values, the near-zero<br />

CODvi-particles have been excluded. In Figure 4.29, the average σ max p fr - <strong>and</strong> σ max interf dec-<br />

values are plotted against the yield stress <strong>of</strong> the composites. One can see that these values are<br />

far from being equal in all specimens. The average <strong>of</strong> values <strong>of</strong> the maximum particle stresses<br />

<strong>and</strong> maximum stresses at the interface increase with increasing yield stress <strong>of</strong> the composites.<br />

The large differences in the average σ p max-values <strong>of</strong> about 400 MPa cannot be explained by<br />

methodogical uncertainties <strong>and</strong> computational inaccuracies <strong>of</strong> the analyzing procedure. This<br />

leads to the question: “What might be a reason for such a behavior?”<br />

A possible explanation could be that the onset <strong>of</strong> plastic deformation has a decisive influence<br />

on the particle <strong>fracture</strong>. Assume that void initiation would occur at the onset <strong>of</strong> gross plastic<br />

yielding in the <strong>matrix</strong> (not at the onset <strong>of</strong> <strong>local</strong> plastic yielding near some corners or edges <strong>of</strong><br />

the particles). This would require that, considered at the meso-scale, the equivalent stress <strong>of</strong><br />

the composite around the considered particle, σ c eq , reaches the composite yield stress, σy. For<br />

σ max [MPa]<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

0 100 200 300 400<br />

σ y [MPa]<br />

64<br />

Al2O3 particles<br />

Al2O3 interface<br />

SiC particles<br />

Fig. 4.29. The dependence <strong>of</strong> the average σ max interf dec- <strong>and</strong> σ max p fr-values on the composite<br />

yield strength, σy, for all investigated MMCs.


Section 4<br />

plane strain conditions, a similar stress triaxiality, σ c m/σ c eq, can be assumed for all specimens<br />

in front <strong>of</strong> the crack tip. Therefore, the composite mean stress, σm c , <strong>and</strong>, consequently, the<br />

mesoscopic maximum principal stress at void initiation, σ HRR max, will be proportional to σy.<br />

As the microscopic maximum principal particle stresses lie between 150 <strong>and</strong> 200 MPa higher<br />

than the meso-scale values (Tables 9.1-9.16), a linear relation between the σ p max-values <strong>and</strong><br />

σy can be expected as is observed in Figure 4.29.<br />

What could be the physical reason for an increase <strong>of</strong> the maximum interfacial stresses? It is<br />

plausible that this finding has to do with the presence <strong>of</strong> the spinel phase MgAl2O4 at the<br />

particle/<strong>matrix</strong> interface. A spinel phase potentially can form strong chemical bonds with both<br />

<strong>metal</strong>s <strong>and</strong> ceramics [71]. It has been suggested that this spinel phase may enhance the<br />

interfacial bonding. Thus, the interface bonding may be enhanced by a longer aging time, as<br />

the percentage <strong>of</strong> the interface covered by the spinel phase increases with increasing aging<br />

time [21].<br />

Table 4.8. Conditions for void initiation in the cast MMCs (average values).<br />

Fractured particles Debonded particles<br />

Specimen Number average<br />

<strong>of</strong><br />

particles<br />

σ max p average<br />

fr<br />

[MPa]<br />

σ max interf Number average<br />

fr<br />

[MPa]<br />

<strong>of</strong><br />

particles<br />

σ max p dec<br />

[MPa]<br />

Al2O3-10-RT 13 867±84 823±89 3 984±88 943±66<br />

65<br />

average<br />

σ max interf dec<br />

[MPa]<br />

Al2O3-10-8h 10 993±53 947±55 4 977±37 935±25<br />

Al2O3-10-24h 10 1090±78 1036±86 7 1173±41 1120±32<br />

Al2O3-10-200h 3 1198±89 1115±97 12 1275±47 1195±47<br />

Al2O3-15-RT 14 974±58 936±60 0 - -<br />

Al2O3-15-8h 11 1157±71 1107±73 5 1203±21 1151±21<br />

Al2O3-15-24h 16 1331±31 1267±29 4 1367±36 1302±32<br />

Al2O3-15-200h 3 1264±46 1209±14 13 1245±28 1214±48<br />

Al2O3-20-RT 9 818±94 808±93 1 834 828<br />

Al2O3-20-8h 12 1180±46 1147±47 0 - -<br />

Al2O3-20-24h 5 1175±125 1142±132 5 1265±25 1231±28<br />

Al2O3-20-200h 8 1217±36 1185±41 7 1263±24 1245±25


Section 4<br />

Table 4.9. Conditions for void initiation in the PM-MMCs.<br />

Specimen Number <strong>of</strong><br />

particles<br />

σ max p fr<br />

[MPa]<br />

SiC-10-15min 15 803±86 730±80<br />

66<br />

σ max interf fr<br />

[MPa]<br />

SiC-10-8h 15 - -<br />

SiC-10-24h 15 - -<br />

SiC-10-200h 19 961±80 927±54<br />

4.6. Application <strong>of</strong> Weibull distribution for <strong>fracture</strong>d particles<br />

As the MMCs are reinforced by ceramic particles, a Weibull distribution can be assumed for<br />

the distributions <strong>of</strong> particle strengths. Such a distribution is <strong>of</strong>ten used for brittle materials<br />

[75-77]:<br />

F(σ) = 1-exp[-(σ /σ0) m ], (4.3)<br />

where F(σ) is the failure probability at the stress σ, σ0 is the characteristic stress, <strong>and</strong> m is the<br />

Weibull modulus which describes the distribution <strong>of</strong> the measured strengths. The higher the<br />

Weibull modulus, the smaller is the scatter <strong>of</strong> the strengths. The maximum principal stresses<br />

<strong>of</strong> the particles at the moment <strong>of</strong> void initiation are taken as σ. In Figures 4.30a <strong>and</strong> 4.31a,<br />

Weibull diagrams for alumina particles in the cast MMC with 10% <strong>of</strong> the Al2O3 particles <strong>and</strong><br />

for SiC particles in some specimens <strong>of</strong> the PM-MMC are given. As seen, the Weibull<br />

modulus lies in the range 16..21 for Al2O3 <strong>and</strong> 9…10 for SiC particles. The high scatter <strong>of</strong> the<br />

Weibull modulus for the alumina particles can be explained by the small number <strong>of</strong> particles<br />

which has been available. As was shown in [75], the uncertainties <strong>of</strong> the calculated parameter<br />

strongly depend on the sample size, N. The uncertainties can be very high in the case <strong>of</strong> small<br />

sample size. For instance, the 95% confidence interval for the determined Weibull modulus m<br />

can lie for N = 10 between 70% <strong>and</strong> 230% <strong>of</strong> the real modulus. The accuracy <strong>of</strong> determination<br />

<strong>of</strong> m can be increased through increase <strong>of</strong> number <strong>of</strong> specimens (in our case, particles).<br />

Indeed, if all analysed Al2O3 particles are taken from all cast specimens (Fig. 4.30b), one gets<br />

the Weibull modulus m = 9, which is in a good accordance with experimentally determined<br />

values for Al2O3 [77].


Ln(Ln(1/(1-F σ)))<br />

Ln(Ln(1/(1-F σ)))<br />

2<br />

1<br />

2<br />

1<br />

0<br />

-2<br />

-3<br />

-4<br />

-5<br />

a)<br />

Section 4<br />

0<br />

6,4 6,6 6,8 7 7,2<br />

-1<br />

-2<br />

-3<br />

m = 16<br />

σ 0 = 926 MPa<br />

Ln(σ)<br />

m = 9<br />

σ 0 = 1105 MPa<br />

6,4 6,6 6,8 7 7,2<br />

-1<br />

Ln (σ)<br />

b)<br />

Fig. 4.30. The Weibull diagram for: a) the alumina particles in investigated cast MMCs<br />

specimens; b) all considered alumina particles in all cast MMCs.<br />

67<br />

m = 18<br />

σ 0 = 1019 MPa<br />

m = 21<br />

σ 0 = 1140<br />

MPa<br />

10%-RT<br />

10%-160°C-8h<br />

10%-160°C-24h<br />

all particles


Ln(Ln(1/(1-F σ)))<br />

Ln(Ln(1/(1-F σ)))<br />

2<br />

1<br />

0<br />

-2<br />

-3<br />

-4<br />

Section 4<br />

6,4 6,5 6,6 6,7 6,8 6,9 7<br />

-1<br />

2<br />

1<br />

-2<br />

-3<br />

-4<br />

m = 10<br />

σ 0 = 821 MPa<br />

Ln(σ)<br />

m = 10<br />

σ 0 = 1032 MPa<br />

m = 9<br />

σ 0 = 930 MPa<br />

0<br />

6,4 6,6 6,8 7<br />

-1<br />

Ln(σ)<br />

a)<br />

b)<br />

Fig. 4.31. The Weibull diagram for the SiC particles: a) for particles in the Specimens SiC-<br />

10%-15min <strong>and</strong> SiC-10%-200h; b) for all particles.<br />

68<br />

10%-175°C-15min<br />

10%-175°C-200h<br />

all particles<br />

Linear (all<br />

particles)


Section 5<br />

5. The relation between the inclusion size <strong>and</strong> the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong><br />

in a mild steel<br />

The knowledge <strong>of</strong> the maximum particle stresses or the maximum decohesion stresses at the<br />

moment <strong>of</strong> void initiation would be interesting for other materials, too. To check whether our<br />

procedure can be used to more generally used engineering materials, it shall be applied in this<br />

section for MnS-inclusions in a mild steel St37.<br />

5.1. Material <strong>and</strong> mechanical <strong>properties</strong><br />

5.1.1 Materials characterization<br />

The chemical composition <strong>of</strong> the St37 is given in Table 5.1. Test pieces were cut transverse to<br />

the rolling direction. They were austenized at 870°C for 45 minutes <strong>and</strong> quenched in salt<br />

water to achieve a uniform distribution <strong>of</strong> the carbon content. Then they were tempered at<br />

700°C for 4 hours <strong>and</strong> furnace cooled down to 500°C at a rate <strong>of</strong> 2° per min. to spheroidize<br />

the iron carbides.<br />

The resulting microstructure after heat treatment consist <strong>of</strong> large MnS-inclusions with a<br />

Table 5.1. Chemical composition <strong>of</strong> the mild steel St37.<br />

C Mn Si Ni Mo P S Cr Cu As<br />

0.17 0.54 0.01 0.04 0.01 0.019 0.018 0.01 0.01 0.002<br />

00 µm<br />

Fig. 5.1. The microstructure <strong>of</strong> the mild steel St37.<br />

69<br />

50 µm


Section 5<br />

diameter <strong>of</strong> 2…8 µm <strong>and</strong> a length <strong>of</strong> 4…100 µm (Fig. 5.1). The MnS-inclusions are elongated<br />

in the rolling direction. Very small Fe3C particles embedded in the ferrite <strong>matrix</strong> are observed,<br />

as well.<br />

5.1.2. Tensile tests<br />

A st<strong>and</strong>ard tensile specimen with a gage length <strong>of</strong> 30 mm <strong>and</strong> a diameter <strong>of</strong> 6 mm is used.<br />

The mechanical test is conducted using a “ZWICK” machine at a the loading rate <strong>of</strong> 6.9×10 -4<br />

s -1 . The experimental true stress - strain <strong>and</strong> theoretical power-law curves are given in Figure<br />

5.2. The results <strong>of</strong> mechanical tensile tests are listed in Table 5.2.<br />

True stress [MPa]<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

5.1.3. Fracture mechanics tests<br />

0 4 8 12 16<br />

Strain [%]<br />

Fig. 5.2. The true stress-strain curve for the mild steel St37.<br />

70<br />

experiment<br />

power-law<br />

St<strong>and</strong>ard CT specimens were machined, with a thickness <strong>of</strong> 25 mm <strong>and</strong> a short-transverse<br />

crack plane orientation. Elongated MnS inclusions lie parallel to the crack front. The<br />

specimens are pre-fatigued <strong>and</strong> <strong>fracture</strong> mechanics tests are conducted using a “ZWICK” test<br />

machine [78]. The value <strong>of</strong> the <strong>fracture</strong> toughness is given.<br />

Large elongated dimples formed by debonded MnS inclusions are observed on the <strong>fracture</strong><br />

surface (Fig. 5.3). These dimples are surrounded by very small equiaxed cavities generated by<br />

iron carbide inclusions Fe3C. The width <strong>of</strong> the primary dimples is larger than the diameter <strong>of</strong>


Section 5<br />

Table 5.2. The mechanical <strong>properties</strong> <strong>of</strong> the mild steel St37.<br />

the MnS-inclusions, on average, by about a factor <strong>of</strong> (5÷7). The dimple length is comparable<br />

to the length <strong>of</strong> the MnS inclusions.<br />

It should be noted that we consider in our analysis only particles forming primary voids near<br />

the crack tip, therefore, Fe3C particles, forming very small equiaxed cavities during the void<br />

coalescence will not be taken into account.<br />

5.2. The effects <strong>of</strong> the particle location <strong>and</strong> the inclusion size on the CODi- <strong>and</strong> CODvi-<br />

values<br />

Material<br />

E<br />

[GPa]<br />

σy<br />

[MPa]<br />

Nine MnS-inclusions located in front <strong>of</strong> the crack tip are investigated. The procedure<br />

described in Section 3 is used to estimate the CODi- <strong>and</strong> CODvi-values <strong>and</strong> the polar<br />

coordinates <strong>of</strong> the particle location with respect to the crack plane. Pr<strong>of</strong>iles perpendicular to<br />

the crack front are analyzed. In Figure 5.4, the pr<strong>of</strong>iles for Inclusion 4 in Table 9.17 are<br />

Fig. 5.3. Corresponding stereopaires from the <strong>fracture</strong> surface <strong>of</strong> the mild steel St37.<br />

71<br />

σUTS<br />

[MPa]<br />

N α<br />

Steel 37 200 299 426 0.2 7.95 51<br />

JIC<br />

[kN/m]


MnS-inclusion<br />

r<br />

θ<br />

MnS-inclusion<br />

Section 5<br />

a)<br />

b)<br />

72<br />

42 µm<br />

Fig. 5.4. Pr<strong>of</strong>iles for MnS-inclusion from Fig. 6.3: a) at the moment <strong>of</strong> the <strong>fracture</strong> initiation,<br />

presented as an example.<br />

CODi = 70 µm, b) at the moment <strong>of</strong> void initiation, CODvi = 42 µm.<br />

There is no clear dependence <strong>of</strong> the CODi- <strong>and</strong> CODvi-values on the angle θ <strong>and</strong> the distance<br />

r (Fig. 5.5a,b). No influence <strong>of</strong> the MnS-inclusion diameter on the CODi-values is observed,<br />

as well (Fig. 5.5c). But it is seen that the MnS-inclusions with larger diameters are more<br />

favorable for void initiation: the CODvi-values clearly decrease with increasing inclusion<br />

diameter, d, in the range <strong>of</strong> 2…8 µm.<br />

The maximum interfacial stresses, σ interf max, for MnS-inclusions in the steel St37 are estimated<br />

by Eq. 2.3. They vary between 964 <strong>and</strong> 1478 MPa. A weak influence <strong>of</strong> the distance, r, <strong>and</strong><br />

the angle, θ, on the σ interf max -values is found (Fig. 5.6a,b), as well as a strong dependence <strong>of</strong><br />

the σ interf max-values on the inclusion size. As one can see from Fig. 5.6c, they decrease with<br />

increasing diameter <strong>of</strong> the MnS-inclusion, d.


COD vi, COD i [µm]<br />

COD vi, COD i [µm]<br />

COD vi, COD i [µm]<br />

100<br />

75<br />

50<br />

25<br />

0<br />

100<br />

75<br />

50<br />

25<br />

0<br />

100<br />

75<br />

50<br />

25<br />

0<br />

Section 5<br />

0 20 40 60<br />

θ [°]<br />

a)<br />

0 50 100 150<br />

r [µm]<br />

b)<br />

0 2 4 6 8<br />

d [µm]<br />

73<br />

CODvi<br />

CODi<br />

CODvi<br />

CODi<br />

CODvi<br />

CODi<br />

c)<br />

Fig. 5.5. The dependence <strong>of</strong> the CODi- <strong>and</strong> CODvi-values on: a) the angle <strong>of</strong> the inclusion<br />

location with respect to the crack plane, θ; b) the distance between the inclusion <strong>and</strong> the crack<br />

tip, r; c) the diameter <strong>of</strong> the MnS-inclusion, d.


σ interf [MPa]<br />

σ interf [MPa]<br />

σ interf [MPa]<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

Section 5<br />

0 10 20 30 40 50<br />

θ [°]<br />

a)<br />

0 50 100 150<br />

b)<br />

74<br />

r [µm]<br />

0 2 4 6 8<br />

d [µm]<br />

c)<br />

Fig. 5.6. The dependence <strong>of</strong> the σ interf max-values on: a) the angle <strong>of</strong> the inclusion location with<br />

respect to the crack plane, θ; b) the distance between the inclusion the crack tip, r; c) the<br />

diameter <strong>of</strong> the MnS-inclusion, d.


Section 6<br />

6. Comparison <strong>of</strong> the <strong>local</strong> <strong>and</strong> <strong>global</strong> <strong>fracture</strong> <strong>properties</strong><br />

In theoretical models to predict the <strong>fracture</strong> toughness in MMCs, the <strong>fracture</strong> toughness was<br />

estimated as a function <strong>of</strong> <strong>global</strong> material <strong>properties</strong>, such as average particle size, particle<br />

volume fraction, etc. (Eqs 2.1, 2.2) [2, 17, 54]. However, the results <strong>of</strong> our investigation<br />

clearly show that the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> significantly vary along the crack front. It can<br />

be assumed that the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> determine in some unknown way the <strong>global</strong><br />

<strong>fracture</strong> <strong>properties</strong>. Therefore, it is interesting to compare the individual <strong>local</strong> values <strong>of</strong> the<br />

<strong>fracture</strong> initiation toughness along the crack front to the <strong>global</strong> <strong>fracture</strong> initiation toughness<br />

measured in the <strong>fracture</strong> mechanics experiments. In Figure 6.1, experimentally determined Ji-<br />

values (taken from Tables 4.2, 4.3) are compared to the Ji-values calculated from the average<br />

value <strong>of</strong> the CODi (including near-zero values) (Table 4.5) through the Eq. 3.2. In spite <strong>of</strong> the<br />

high scatter <strong>of</strong> the values <strong>of</strong> the <strong>local</strong> <strong>fracture</strong> toughness observed in our investigated<br />

materials, their average values correlate well with experimentally determined values <strong>of</strong> the<br />

<strong>global</strong> <strong>fracture</strong> toughness. A very good correlation is found for the cast MMCs with 10% <strong>and</strong><br />

20% Al2O3 particles in all aging conditions (Fig. 6.1c) <strong>and</strong> for the mild steel St37 (Fig. 6.2b).<br />

For the cast MMCs with 15%, a good correlation is observed for the specimen aged at room<br />

temperature; at other aging conditions the correlation becomes worse but still remains in the<br />

range <strong>of</strong> the st<strong>and</strong>ard deviation (Fig. 6.1a,b).<br />

For the PM MMC, Ji-values which are determined by the <strong>fracture</strong> mechanics experiment are<br />

close to the bottom boundary <strong>of</strong> the st<strong>and</strong>ard deviation <strong>of</strong> the Ji-values calculated from the<br />

average value <strong>of</strong> the CODi (Fig. 6.2a). This can be explained when it is assumed that the<br />

<strong>fracture</strong> <strong>of</strong> one particle might trigger the <strong>fracture</strong> <strong>of</strong> the neighboring particles within a certain<br />

region so that the area <strong>of</strong> <strong>local</strong> crack extension is larger than the resolution <strong>of</strong> the potential<br />

drop technique. The larger particle size (100 µm, compared to 10 µm for the cast MMC)<br />

facilitates this effect.<br />

So on, for the cast MMCs <strong>and</strong> the steel St37, it can be proposed that the <strong>global</strong> <strong>fracture</strong><br />

toughness <strong>of</strong> any material can be estimated as<br />

J<br />

n<br />

∑<br />

<strong>global</strong><br />

i = 1<br />

75<br />

J<br />

n<br />

<strong>local</strong><br />

i<br />

, (6.1)


Section 6<br />

where n is the number <strong>of</strong> considered <strong>local</strong> Ji-values. The Eq. 6.1. is not valid for the PM<br />

MMCs due to their brittleness.<br />

From these results it can be concluded, that the process <strong>of</strong> <strong>fracture</strong> in materials should be<br />

considered not only at the <strong>global</strong> level (as has been done so far), but at the <strong>local</strong> level, as well.<br />

Of course, in the first approach, the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> seem to be not so important for<br />

the <strong>fracture</strong> mechanics operating at the <strong>global</strong> level. But results obtained in current research<br />

can be very useful not only for improvement <strong>of</strong> application <strong>of</strong> damage mechanics at the <strong>local</strong><br />

level, but they might have even industrial application for the cases where a high level <strong>of</strong><br />

safety <strong>of</strong> structures is required. For instance, from results <strong>of</strong> this research, it can be advised to<br />

estimate the critical conditions for void initiation for such structures not in terms <strong>of</strong> the<br />

average particle size (as has been done so far), but in terms <strong>of</strong> the maximum particle size,<br />

since the largest particles are critical.<br />

76


Ji [kN/m]<br />

J i [kN/m]<br />

Ji [kN/m]<br />

10<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

8<br />

6<br />

4<br />

2<br />

0<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Section 6<br />

RT 8h 24h 200h<br />

Heat treatment<br />

a)<br />

RT 8h 24h 200h<br />

Heat treatment<br />

b)<br />

RT 8h 24h 200h<br />

Heat treatment<br />

c)<br />

77<br />

<strong>local</strong> from CODi<br />

<strong>global</strong><br />

<strong>local</strong> from CODil<br />

<strong>global</strong>l<br />

<strong>local</strong> from CODi<br />

<strong>global</strong><br />

Fig. 6.1. A comparison <strong>of</strong> the theoretical <strong>and</strong> experimental Ji-values for the cast MMCs with:<br />

10%<br />

15%<br />

20%<br />

a) 10% <strong>of</strong> Al2O3, b) 15% <strong>of</strong> Al2O3, c) 20% <strong>of</strong> Al2O3.


Ji [kN/m]<br />

J i [kN/m]<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

60<br />

40<br />

20<br />

0<br />

Section 6<br />

15min 8h 24h 200h<br />

Heat treatment<br />

St37<br />

a)<br />

b)<br />

78<br />

<strong>local</strong> from CODi<br />

<strong>global</strong><br />

<strong>local</strong> from CODi<br />

<strong>global</strong><br />

Fig. 6.2. A comparison <strong>of</strong> the theoretical <strong>and</strong> experimental Ji-values for: a) the PM-MMC;<br />

b) the mild steel St37<br />

10%<br />

St37


Section 7<br />

7. Application <strong>of</strong> ECAP to improve the homogeneity <strong>of</strong> the particle<br />

distribution in MMCs<br />

This work is a part <strong>of</strong> a big project on the deformation <strong>and</strong> <strong>fracture</strong> behavior <strong>of</strong> MMCs. It was<br />

originally planned to investigate (in addition to the cast MMCs) powder <strong>metal</strong>lurgy MMCs<br />

with particle sizes between 100 µm <strong>and</strong> 1 µm. We found, however, that the as-received<br />

material with small particle sizes could not be used for our investigations since the material<br />

<strong>properties</strong> were extremely bad due to the appearance <strong>of</strong> large particle clusters.<br />

This part <strong>of</strong> the work deals with the problem <strong>of</strong> the homogeneity <strong>of</strong> the particle distribution in<br />

powder <strong>metal</strong>lurgy MMCs: a method <strong>of</strong> equal channel angular pressing is proposed to solve<br />

this problem, <strong>and</strong> the effect <strong>of</strong> this method on the <strong>global</strong> <strong>and</strong> <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> is<br />

studied.<br />

7.1. Material<br />

An Al6061 based powder <strong>metal</strong>lurgy MMC reinforced by 20% <strong>of</strong> Al2O3 particles is chosen as<br />

a material for this investigation. The alumina particle size scatters in the range between 1 <strong>and</strong><br />

5 µm. A few, single particles having larger sizes are observed. The size <strong>of</strong> the <strong>matrix</strong> powder<br />

is 40 µm. The composite was supplied in a form <strong>of</strong> a rod with a diameter <strong>of</strong> 25 mm: the<br />

extrusion ratio was 10:1. A <strong>metal</strong>lographic section <strong>of</strong> the rod shows that the as-fabricated<br />

material has a strongly clustered particle distribution (Fig. 7.1).<br />

Fig. 7.1. Microstructure <strong>of</strong> the as-fabricated PM MMC.<br />

79<br />

30µm


Section 7<br />

7.2. A model by Tan <strong>and</strong> Zhang for the case <strong>of</strong> combination <strong>of</strong> extrusion <strong>and</strong> equal<br />

channel angular pressing<br />

Tan <strong>and</strong> Zhang [23] proposed a model to predict the homogeneity <strong>of</strong> the particle distribution<br />

in PM-MMCs. They suggest that a homogeneous particle distribution can be expected only<br />

when the reinforcement size is not less than a critical value, dc, which is determined as a<br />

function <strong>of</strong> the <strong>matrix</strong> powder size, dm, the particle volume fraction, f, <strong>and</strong> the reduction ratio<br />

<strong>of</strong> extrusion, R, namely<br />

d<br />

c<br />

d<br />

=<br />

1 ⎡⎛<br />

π ⎞<br />

⎢⎜<br />

⎟<br />

⎢ f<br />

⎣⎝<br />

6 ⎠<br />

In the case <strong>of</strong> rolling, the criterion becomes [23]<br />

d<br />

c<br />

m<br />

/ 3<br />

80<br />

⎤<br />

−1⎥<br />

⎥⎦<br />

R<br />

. (7.1)<br />

d m<br />

=<br />

, (7.2)<br />

1/<br />

3<br />

⎡⎛<br />

π ⎞ ⎤ 1<br />

⎢⎜<br />

⎟ −1⎥<br />

⎢ 6 f ⎥ 1−<br />

R'<br />

⎣⎝<br />

⎠ ⎦<br />

where R’ is the percentage <strong>of</strong> reduction in thickness. In the case <strong>of</strong> a combination <strong>of</strong> extrusion<br />

with rolling, the critical value <strong>of</strong> reinforcement size can be derived as [23]<br />

d<br />

c<br />

d m<br />

=<br />

. (7.3)<br />

1/<br />

3<br />

⎡⎛<br />

π ⎞ ⎤ R<br />

⎢⎜<br />

⎟ −1⎥<br />

⎢ 6 f ⎥ 1−<br />

R'<br />

⎣⎝<br />

⎠ ⎦<br />

One can see from all three criteria that the extrusion or (<strong>and</strong>) reduction ratio for a given<br />

material is (are) the only parameter(s) responsible for the homogeneity <strong>of</strong> the particle<br />

distribution in the PM-MMCs during its processing: its (their) increase decreases the dc-value.<br />

Theoretically, a homogeneous distribution <strong>of</strong> reinforcements could be achieved in any<br />

material, independently <strong>of</strong> the particle <strong>and</strong> <strong>matrix</strong> powder size or the particle volume fraction,<br />

if enough deformation is induced into the sample. However, a great shortcoming <strong>of</strong> extrusion<br />

<strong>and</strong> rolling is a significant decrease <strong>of</strong> the sample section with increasing extrusion or (<strong>and</strong>)<br />

reduction ratio. Thus, the extrusion or (<strong>and</strong>) reduction ratio is (are) <strong>of</strong>ten limited by a fixed<br />

final dimension <strong>of</strong> the sample required for industrial application.


Section 7<br />

Fig. 7.2. The principle <strong>of</strong> the equal channel angular pressing.<br />

For this reason, a deformation method would be advantageous which induces intense strains<br />

into the sample without any change <strong>of</strong> the sample size.<br />

The equal channel angular pressing (ECAP) might be an effective tool to solve this problem.<br />

In this method, intense plastic strains are induced into massive billets without changing their<br />

cross section [79]. The ECAP facility consists <strong>of</strong> two channels which are equal in cross<br />

section (Fig. 7.2). The channels are intersected at a certain angle, ϕ. Another important<br />

parameter describing the ECAP facility is the angle subtended by the arc curvature, ψ. A<br />

sample is usually subjected to a multiple pressing through a die.<br />

An analysis performed in [80] has shown that the strains introduced into a sample during<br />

ECAP are a function <strong>of</strong> the angles ϕ <strong>and</strong> ψ, <strong>and</strong> the number <strong>of</strong> ECAP passes, N,<br />

ε<br />

Die<br />

ECAP<br />

=<br />

Plunger<br />

N ⎡ ⎛ ϕ ψ ⎞ ⎛ ϕ ψ ⎞⎤<br />

⎢2cot⎜<br />

+ ⎟ + ψ ⋅ csc⎜<br />

+ ⎟⎥<br />

. (7.4)<br />

3 ⎣ ⎝ 2 2 ⎠ ⎝ 2 2 ⎠⎦<br />

The extrusion ratio, R, in Eq (7.1) is considered as a measure <strong>of</strong> strain induced into the<br />

sample. Actually, it is equal to the ratio <strong>of</strong> the final length <strong>of</strong> the sample after extrusion, lf, to<br />

81<br />

Sample


Section 7<br />

its initial length, li. On the other h<strong>and</strong>, this ratio can be also presented as a function <strong>of</strong> the true<br />

strain induced into the sample by extrusion (Eq. 7.5) which follows from the general<br />

determination <strong>of</strong> the true strain, ε = ln (lf / li),<br />

R = exp (ε ). (7.5)<br />

In the case <strong>of</strong> two step extrusion with extrusion ratios R1 <strong>and</strong> R2, a total extrusion ratio can be<br />

determined as R = R1⋅ R2. So, if we combine extrusion <strong>and</strong> ECAP, a total extrusion ratio can<br />

be presented as<br />

R = Rext ⋅ RECAP . (7.6)<br />

According to Eq. 7.5, RECAP can be determined as RECAP = exp (εECAP) or, taking into account<br />

Eq. 7.4, as<br />

R ECAP<br />

⎡ N ⎡ ⎛ ϕ ψ ⎞ ⎛ ϕ ψ ⎞⎤⎤<br />

= exp⎢<br />

⎢2cot⎜<br />

+ ⎟ + ψ ⋅ csc⎜<br />

+ ⎟⎥⎥<br />

. (7.7)<br />

⎣ 3 ⎣ ⎝ 2 2 ⎠ ⎝ 2 2 ⎠⎦⎦<br />

Inserting Eq. 7.7 into Eq. 7.6, <strong>and</strong> obtained result into Eq. 7.1, we get the model for the case<br />

<strong>of</strong> combination <strong>of</strong> extrusion with ECAP<br />

d<br />

c<br />

d m<br />

=<br />

. (7.8)<br />

1/<br />

3 ⎡⎛<br />

π ⎞ ⎤ ⎡ N ⎡ ⎛ ϕ ψ ⎞ ⎛ ϕ ψ ⎞⎤⎤<br />

⎢⎜<br />

⎟ −1⎥<br />

Rext<br />

exp⎢<br />

⋅ ⎢2cot⎜<br />

+ ⎟ + ψ ⋅ csc⎜<br />

+ ⎟⎥⎥<br />

⎢⎣<br />

⎝ 6 f ⎠ ⎥⎦<br />

⎣ 3 ⎣ ⎝ 2 2 ⎠ ⎝ 2 2 ⎠⎦⎦<br />

Here, both the strains introduced during extrusion <strong>and</strong> ECAP are taken into account.<br />

It should be noted that ECAP has been used so far just for processing <strong>of</strong> the bulk<br />

nanostructured materials [79]. There have been already attempts to subject MMCs to ECAP in<br />

order to refine grains. In [81], a cast Al6061 MMC reinforced by 10% Al2O3 particles having<br />

a size in the range <strong>of</strong> 6…13 µm was subjected to ECAP. The samples were pressed for 8<br />

passes at 400°C plus additional 2 passes at 200°C in order to give a total strain <strong>of</strong> ∼10.<br />

Detailed measurements indicated that there was little or no breaking <strong>of</strong> the Al2O3 particles<br />

during the ECAP. Significant grain refinement <strong>and</strong> increase <strong>of</strong> the composite strength by<br />

almost a factor <strong>of</strong> 2 was observed. But the material did not have any particle clusters initially.<br />

Therefore, the evolution <strong>of</strong> uniformity <strong>of</strong> particle distribution was not investigated.<br />

82


Section 7<br />

Table 7.1. Data on the critical particle size for the PM-MMC Al6061-20%Al2O3 before <strong>and</strong><br />

after different number <strong>of</strong> ECAP passes.<br />

Number <strong>of</strong> ECAP passes 0 4 7<br />

dc [µm] 33.4 4.1 0.9<br />

The rod <strong>of</strong> investigated material is cut into samples with a length <strong>of</strong> 100 mm. These samples<br />

are heated to 370°C <strong>and</strong> pressed repeatedly through the ECAP die. The parameters <strong>of</strong> the<br />

ECAP die are ϕ = 90° <strong>and</strong> ψ = 20°. The material is subjected to 4 <strong>and</strong> 7 ECAP passes. The<br />

choice <strong>of</strong> these numbers <strong>of</strong> the ECAP passes can by explained by results <strong>of</strong> calculations given<br />

in Table 8.1. Here, the critical values <strong>of</strong> the particle size determined by Eq. 7.8 are presented<br />

for the as-fabricated MMC, <strong>and</strong> the MMC after 4 <strong>and</strong> 7 ECAP passes. It is seen that the<br />

increase <strong>of</strong> number <strong>of</strong> ECAP passes significantly decreases the dc-value. dc = 33.4 µm for the<br />

as-fabricated MMCs, thus, the strongly clustered particle distribution is expected for this<br />

material. After 4 ECAP passes, the dc-value decreases up to 4.1 µm. Taking into account that<br />

the particle size in investigated material is in the range <strong>of</strong> 1…5 µm, particle clusters should<br />

start to scatter. After 7 ECAP passes, the dc = 0.9 µm. Since this value is less than the lowest<br />

bound <strong>of</strong> the particle size, a homogeneous particle distribution after 7 ECAP passes can be<br />

expected.<br />

7.3. A quadrat method to estimate the homogeneity <strong>of</strong> the particle distribution in the<br />

<strong>metal</strong> <strong>matrix</strong> composites<br />

Different methods to estimate the homogeneity <strong>of</strong> the particle distribution in the MMCs have<br />

been proposed: the mean free path, the nearest neighbour distance, the radial distribution<br />

function, <strong>and</strong> the quadrat method [82]. As was found in [82], for the MMCs with strongly<br />

clustered particle distribution, the quadrat method is more effective to detect pronounced<br />

changes in the MMC microstructure in comparison to other methods. With the quadrat<br />

method, the image to be studied is divided into a grid <strong>of</strong> square cells <strong>and</strong> the number <strong>of</strong><br />

particles in each cell is counted [83]. In general, an ordered particle distribution would be<br />

expected to generate a large number <strong>of</strong> quadrats containing approximately the same number<br />

<strong>of</strong> particles, whereas a clustered distribution would be expected to produce a combination <strong>of</strong><br />

empty quadrats, quadrats with a small number <strong>of</strong> particles, <strong>and</strong> quadrats with many particles.<br />

83


1 µm<br />

Section 7<br />

5 µm<br />

Fig. 7.3. A schematic view <strong>of</strong> the quadrat method for the PM-MMC Al6061-20%Al2O3.<br />

The major problem <strong>of</strong> the quadrat method is to determine the optimum quadrat size. Non-<br />

r<strong>and</strong>omness is highly dependent on the size <strong>of</strong> the sample quadrat. In this study, the problem<br />

becomes more complicated due to the difference in the size <strong>of</strong> the reinforcements. According<br />

to [82], the optimum quadrat size is twice the size <strong>of</strong> the mean area per particle. If for our<br />

material the particle size is set to 3 µm (which lies in the middle <strong>of</strong> the range <strong>of</strong> observed<br />

particle sizes), the quadrat size becomes a = 8.5 µm. Qualitative analysis <strong>of</strong> the<br />

microstructures shows that this quadrat size seems to be the optimum value for this<br />

investigation. In Figure 7.3, a schematic view <strong>of</strong> the quadrat method for our case is presented.<br />

The particles that belong to different quadrats are painted in different colours. One can see<br />

that a lower a-value could artificially decrease the number <strong>of</strong> particles in the quadrats<br />

corresponding to clusters containing large particles (5 µm). At the same time, higher a-values<br />

could significantly decrease the number <strong>of</strong> empty quadrats <strong>and</strong> tend to give the same number<br />

<strong>of</strong> particles in each quadrat, irrespective <strong>of</strong> the presence <strong>of</strong> clusters.<br />

For the microstructural investigation, the rods are sectioned in the longitudinal direction,<br />

grinded, polished, <strong>and</strong> investigated in the SEM. 9 micrographs from the center region <strong>of</strong> the<br />

specimens are taken in the SEM, each covering an area <strong>of</strong> 130×100 µm. Each image is<br />

divided into 140 quadrats. The histograms <strong>of</strong> the particle per quadrat distribution for each<br />

material condition are plotted. These histograms are compared with theoretical distributions:<br />

84<br />

a =8.5 µm


Section 7<br />

1) The Poisson distribution which corresponds to the homogeneous particle distribution<br />

[83]<br />

r<br />

µ<br />

P(<br />

r)<br />

= exp( −µ<br />

) , (7.9)<br />

r!<br />

where P(r) is the probability, µ the mean value <strong>of</strong> the number <strong>of</strong> particles per quadrat, <strong>and</strong> r a<br />

variable (the number <strong>of</strong> particles per quadrat);<br />

2) The negative binomial distribution which corresponds to the clustered particle<br />

distribution [83]<br />

85<br />

r<br />

⎛ ( k + r −1)!<br />

⎞⎛<br />

p ⎞ ⎛ 1 ⎞<br />

P( r)<br />

= ⎜ ⎟⎜<br />

⎟ ⎜ ⎟ , (7.10)<br />

⎝ ( k −1)!<br />

r!<br />

⎠⎝1<br />

+ p ⎠ ⎝1<br />

+ p ⎠<br />

where k <strong>and</strong> p are parameters determined according to [83].<br />

7.4. The effect <strong>of</strong> ECAP on the particle distribution, the <strong>global</strong> <strong>and</strong> <strong>local</strong> <strong>fracture</strong><br />

<strong>properties</strong> in MMCs with severe clustered particle distribution<br />

7.4.1. The effect <strong>of</strong> ECAP on the microstructure<br />

The uniformity <strong>of</strong> the particle distribution increases with the number <strong>of</strong> ECAP passes (Fig.<br />

7.4). For the as-fabricated MMC, clusters elongated in the extrusion direction are observed.<br />

The clusters have a size <strong>of</strong> up to 40 µm in the direction perpendicular to the extrusion<br />

direction. Large particle free zones, having a size <strong>of</strong> up to 200 µm in the extrusion direction<br />

<strong>and</strong> up to 40 µm in the perpendicular direction, are located between particle clusters (Fig.<br />

7.4a). From Fig. 7.5a, it is seen that the histogram corresponding to the particle distribution<br />

for the as-fabricated material seems to follow the negative binomial distribution indicating a<br />

clustering. After 4 ECAP passes, particle free zones become smaller <strong>and</strong> the particle clusters<br />

start to scatter (Fig. 7.4b). For this material condition, a deviation from the negative binomial<br />

distribution to the Poisson distribution can be noted in the histogram (Fig. 7.5b). Almost<br />

complete declustering <strong>and</strong> an absence <strong>of</strong> particle free zones are observed for the MMC after 7<br />

ECAP passes (Fig. 7.4c). The histogram <strong>of</strong> the particle distribution follows the Poisson<br />

distribution confirming the particle declustering, as well (Fig. 7.5c). Thus, these observations<br />

prove the validity <strong>of</strong> proposed Eq. 7.8.<br />

k


Section 7<br />

a)<br />

b)<br />

c)<br />

Fig. 7.4. Microstructure <strong>of</strong> the Al6061-20%Al2O3 PM-MMC: a) in as-fabricated condition, b)<br />

after 4 ECAP passes, c) after 7 ECAP passes.<br />

86<br />

30µm<br />

30µm<br />

30µm


a)<br />

b)<br />

Frequency [%]<br />

Frequency [%]<br />

Frequency [%]<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Section 7<br />

0 5 10 15 20<br />

N q<br />

0 5 10 15 20<br />

N q<br />

0 5 10 15 20<br />

N q<br />

87<br />

Poisson<br />

distrubution<br />

experimental data<br />

Negative binomial<br />

distribution<br />

Poisson<br />

distribution<br />

experimental data<br />

Negative binomial<br />

distribution<br />

Poisson<br />

distribution<br />

experimental data<br />

Negative binomial<br />

distribution<br />

c)<br />

Fig. 7.5. Theoretical distribution curves <strong>and</strong> experimental results from the quadrat analysis <strong>of</strong><br />

the PM MMC Al6061-20%Al2O3: a) in as-fabricated condition, b) after 4 ECAP passes, c)<br />

after 7 ECAP passes.


Section 7<br />

It is important that no particle breaking during the ECAP is observed: the number <strong>of</strong> particles<br />

per quadrat, µ, are 6.2, 5.9, <strong>and</strong> 6.7 for the initial state, after 4, <strong>and</strong> 7 ECAP passes,<br />

respectively. It should be further noted that no voids are observed in the investigated material<br />

after 7 ECAP passes.<br />

7.4.2. The effect <strong>of</strong> ECAP on the <strong>global</strong> <strong>fracture</strong> <strong>properties</strong><br />

Grinded <strong>and</strong> polished tensile specimens with a gage length <strong>of</strong> 10 mm <strong>and</strong> a quadratic section<br />

<strong>of</strong> 2x2 mm are used for tensile tests. Disk-shaped compact specimens similar to that presented<br />

in Fig. 4.8 is used to perform <strong>fracture</strong> tests. Specimens for mechanical tests are annealed at<br />

530°C for 1h, quenched in water, <strong>and</strong> aged at 175°C for 15min. which corresponds to an<br />

under-aged condition. This heat treatment is chosen because <strong>of</strong> the low <strong>fracture</strong> toughness <strong>of</strong><br />

the MMC. As was already mentioned in Section 2, MMCs in the under-aged condition have<br />

the highest <strong>fracture</strong> toughness. In our case, a high <strong>fracture</strong> toughness would be advantageous<br />

to compare the <strong>fracture</strong> <strong>properties</strong> <strong>of</strong> the material.<br />

In Table 7.2, the results <strong>of</strong> mechanical tensile tests are listed. No significant difference<br />

between the yield strength, σy, ultimate tensile strength, σu, the strain hardening coefficient,<br />

N, <strong>and</strong> the <strong>fracture</strong> strain, εfr, is found for considered materials.<br />

In Figure 7.6, the J-∆a curves determined according to [70] for all tested disk compact<br />

specimens are given. The values <strong>of</strong> the <strong>fracture</strong> toughness, J0.2, the maximum extension <strong>of</strong> the<br />

stable crack propagation, ∆astab, the slope <strong>of</strong> the J-∆a-curve, dJ/d(∆a), in the range between<br />

the ∆a = 0.2 mm <strong>and</strong> ∆astab are listed in Table 7.2, as well. A tendency to increase <strong>of</strong> the J0.2-,<br />

the dJ/d(∆a)-, <strong>and</strong> the ∆astab-values with increasing homogeneity <strong>of</strong> the particle distribution<br />

can be noted.<br />

Evolution <strong>of</strong> the slope <strong>of</strong> the J-∆a curve, dJ/d(∆a), is especially interesting, as this parameter<br />

is a measure <strong>of</strong> the total crack growth resistance in material, Rtot, [50, 84]. In [50], the relation<br />

Number <strong>of</strong><br />

ECAP passes<br />

Table 7.2. Data on mechanical <strong>properties</strong> <strong>of</strong> the PM-MMC Al6061-20%Al2O3.<br />

σy<br />

[MPa]<br />

σu<br />

[MPa]<br />

N εfr<br />

[%]<br />

88<br />

J0.2/Bl<br />

[kN/m]<br />

dJ/d(∆a) ∆astab<br />

[mm]<br />

0 225 287 0.11 3.0 1.5 1.93 0.42 7.7<br />

4 230 315 0.12 4.2 1.7 2.50 0.72 10.3<br />

Rtot<br />

[kJ/m 2 ]<br />

7 230 306 0.12 3.5 2.7 3.88 1.05 15.4


J [kN/m]<br />

J [kN/m]<br />

J [kN/m]<br />

8<br />

6<br />

4<br />

2<br />

0<br />

8<br />

6<br />

4<br />

2<br />

0<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Section 7<br />

0 0,4 0,8 1,2 1,6<br />

∆a [mm]<br />

a)<br />

0 0,4 0,8 1,2 1,6<br />

∆a [mm]<br />

b)<br />

0 0,4 0,8 1,2 1,6<br />

∆a [mm]<br />

c)<br />

Fig. 7.6. The J-integral resistance curves for the MMC: a) in the as-fabricated condition; b)<br />

after 4 ECAP passes; c) after 7 ECAP passes<br />

89


Section 7<br />

between the Rtot <strong>and</strong> the slope <strong>of</strong> the J-∆a curve was derived<br />

dJ η<br />

≈ R<br />

d(<br />

∆ a)<br />

b<br />

90<br />

tot<br />

, (7.11)<br />

where η is the pre-factor in the J-evaluation formula <strong>and</strong> b the ligament length b = W-a [70].<br />

The Rtot data, calculated by Eq. 7.11 in the region from ∆a = 0.2 mm up to ∆astab, are listed in<br />

Table 7.2, as well. A significant increase <strong>of</strong> the Rtot-values with increasing homogeneity <strong>of</strong> the<br />

particle distribution is observed: the Rtot-value <strong>of</strong> the MMC after 7 ECAP passes is higher<br />

than the Rtot-value <strong>of</strong> the as-fabricated material by a factor <strong>of</strong> 2. As was shown in [50], for the<br />

flat <strong>fracture</strong> region, Rtot is determined by the plastic strain energy to form the micro-ductile<br />

<strong>fracture</strong> surface <strong>and</strong> the energy spent below the <strong>fracture</strong> surface.<br />

7.4.3. The effect <strong>of</strong> ECAP on the <strong>fracture</strong> surface morphology<br />

A different morphology <strong>of</strong> the <strong>fracture</strong> surfaces <strong>of</strong> tested disk compact specimens is revealed.<br />

In Figure 7.7a, the pre-fatigued region <strong>and</strong> the <strong>fracture</strong> surface in front <strong>of</strong> the crack tip for the<br />

specimen from the as-fabricated MMC is shown. A vast amount <strong>of</strong> particle clusters on the<br />

<strong>fracture</strong> surfaces as well as on the pre-fatigued region is seen. Regions with relative<br />

homogeneous particle distribution on the <strong>fracture</strong> surface can be also observed. In Figure<br />

7.7b, an image <strong>of</strong> the broken particle cluster is given at higher magnification. One can see that<br />

there is no <strong>matrix</strong> between the alumina particles in clusters. On the contrary, the <strong>fracture</strong><br />

surface <strong>of</strong> the specimen after 7 ECAP passes looks more homogeneous (Fig. 7.7c). Typical<br />

micro-ductile mechanism <strong>of</strong> <strong>fracture</strong> prevails, a few particle clusters on the <strong>fracture</strong> surface<br />

are observed. Almost no particle clusters can be found on the pre-fatigued region. Voids are<br />

initiated by a particle/<strong>matrix</strong> decohesion mechanism. In Figure 7.8, corresponding pictures<br />

from both halves <strong>of</strong> the broken specimen from the MMC after 7 ECAP passes are given. It is<br />

clearly seen, that most <strong>of</strong> particles are observed only on a single half <strong>of</strong> the broken specimen.<br />

Some <strong>of</strong> them are marked by arrows: the position <strong>of</strong> the particle on the half <strong>of</strong> the broken<br />

specimen, where the particle is observed, is marked by solid arrow; whereas the other dimple<br />

initiated by this particle on the other half <strong>of</strong> the broken specimen by a dashed arrow.


Section 7<br />

a) b)<br />

c)<br />

Fig. 7.7. (a, b) the <strong>fracture</strong> surface <strong>of</strong> the PM-MMC Al6061-20%Al2O3 in as-fabricated<br />

condition; (c) the <strong>fracture</strong> surface <strong>of</strong> the PM-MMC after 7 ECAP passes.<br />

91


Section 7<br />

Fig. 7.8. The <strong>fracture</strong> surface <strong>of</strong> the PM-MMC Al6061-20%Al2O3 after 7 ECAP passes.<br />

7.4.4. The effect <strong>of</strong> ECAP on the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong><br />

As was mentioned in previous section, the <strong>fracture</strong> surface <strong>of</strong> the specimen with strongly<br />

clustered particle distribution is non-homogeneous. In Figure 7.9, a schematic view <strong>of</strong> this<br />

<strong>fracture</strong> surface is presented. This scheme is correlated with the stereophotogrammetic pairs<br />

subjected to analysis by the automatic <strong>fracture</strong> surface analysis system (Fig. 7.7a). Table 7.3<br />

provides the results <strong>of</strong> measurements <strong>of</strong> the <strong>local</strong> CODi-values. The <strong>local</strong> CODi-values were<br />

measured at four different points at the particle cluster <strong>and</strong> at four different points at the area<br />

with homogeneous particle distribution on the <strong>fracture</strong> surface. A significant scatter <strong>of</strong> the<br />

92


Section 7<br />

particle cluster area area without particle clusters<br />

CODi = 0.2...0.5µm<br />

Fig. 7.9. Schematic view <strong>of</strong> the <strong>fracture</strong> surface <strong>of</strong> the the PM-MMC Al6061-20%Al2O3 in as-<br />

fabricated condition.<br />

<strong>local</strong> CODi-values is found. In front <strong>of</strong> the particle cluster, they have near-zero values<br />

0.2…0.5 µm. In Figure 7.10, such a pr<strong>of</strong>ile is presented. In front <strong>of</strong> the region with<br />

homogeneous particle distribution on the <strong>fracture</strong> surface, the CODi-values lie in the range<br />

2…3 µm (Fig. 8.11). On the contrary, in the specimen from the MMC after 7 ECAP passes,<br />

no significant scatter <strong>of</strong> the CODi-values is observed: the <strong>local</strong> CODi-values measured at six<br />

different positions vary in the range <strong>of</strong> 2…3 µm (Table 7.3).<br />

An improvement <strong>of</strong> the <strong>global</strong> <strong>fracture</strong> toughness <strong>and</strong> the total crack growth resistance <strong>of</strong> the<br />

MMC with increasing homogeneity <strong>of</strong> the particle distribution might be explained by results<br />

<strong>of</strong> estimation <strong>of</strong> the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong>. Such a difference in <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> at<br />

regions with different <strong>local</strong> particle distribution can be caused by a different amount <strong>of</strong> energy<br />

required for crack propagation in this region. It is clear that at particle clusters, where no<br />

<strong>matrix</strong> between the particles is observed (Fig. 7.7b), no energy is needed for the void<br />

initiation <strong>and</strong> void growth: the <strong>local</strong> values <strong>of</strong> the crack growth resistance are extremely low,<br />

since the <strong>local</strong> values <strong>of</strong> the void initiation <strong>fracture</strong> toughness is almost equal zero. At the<br />

same time, in regions with homogeneous particle distribution, some amount <strong>of</strong> energy is<br />

required for the void initiation <strong>and</strong> following void growth that results in relative high values<br />

<strong>of</strong> the <strong>local</strong> <strong>fracture</strong> toughness.<br />

CODi = 2...3 µm<br />

93<br />

<strong>fracture</strong> surface<br />

crack tip<br />

pre-fatigued<br />

region


10<br />

z [mm]<br />

5<br />

-10<br />

-15<br />

Section 7<br />

0<br />

-20 -10 0 10 20 30 40<br />

x [mm]<br />

-5<br />

COD i = 0.5 µm<br />

crack tip<br />

Fig. 7.10. Pr<strong>of</strong>iles through a particle cluster at specimen from the as-fabricated material:<br />

CODi = 0.5 µm (Pr<strong>of</strong>ile 1 in Table 7.3).<br />

94


Section 7<br />

10<br />

5<br />

0<br />

-40 -30 -20 -10 0 10 20 30 40<br />

COD i = 3 µm<br />

z [µm]<br />

-5<br />

-10<br />

crack<br />

tip<br />

-15<br />

95<br />

x [µm]<br />

Fig. 7.11. Pr<strong>of</strong>iles through a region with homogeneous particle distribution at Specimen from<br />

the as-fabricated material: CODi = 3 µm (Pr<strong>of</strong>ile 1 in Table 7.3).


Section 7<br />

Table 7.3. Data on measured CODi-values for the PM-MMC Al6061-20%Al2O3 in as-<br />

fabricated condition <strong>and</strong> after 7 ECAP.<br />

Data for the as-fabricated MMC<br />

A particle cluster area<br />

Number<br />

<strong>of</strong> pr<strong>of</strong>ile<br />

CODi-values<br />

[µm]<br />

An area with <strong>local</strong><br />

homogeneous particle<br />

Number<br />

<strong>of</strong> pr<strong>of</strong>ile<br />

distribution<br />

96<br />

CODi-values<br />

[µm]<br />

Data for the MMC after 7<br />

Number <strong>of</strong><br />

pr<strong>of</strong>ile<br />

ECAP passes<br />

1 0.5 1 3.0 1 2.5<br />

2 0.2 2 2.5 2 2.0<br />

3 0.5 3 2.0 3 3.0<br />

4 0.2 4 2.0 4 2.0<br />

CODi-values<br />

[µm]<br />

5 2.5<br />

6 2.0<br />

7.5. About possible application <strong>of</strong> ECAP to improve the particle distribution in different<br />

MMCs<br />

As one can see from the results <strong>of</strong> our investigation, the ECAP can be useful as an additional<br />

operation during the secondary processing <strong>of</strong> MMCs produced by powder <strong>metal</strong>lurgy<br />

technology to increase the uniformity <strong>of</strong> particle distribution. The required number <strong>of</strong> ECAP<br />

passes can be easily calculated by Eq. 7.8. The ECAP can be employed as a last procedure <strong>of</strong><br />

fabrication <strong>of</strong> MMCs as well as the ECAP <strong>and</strong> extrusion can be combined during the<br />

fabrication <strong>of</strong> MMCs. It should be noted that the ECAP does not degrade the microstructure<br />

<strong>of</strong> the investigated material: no void formation <strong>and</strong> no particle breaking in the case <strong>of</strong> small<br />

particles are observed in the sample interior, but due to the occurrence <strong>of</strong> surface defects, the<br />

number <strong>of</strong> ECAP passes is limited. The application <strong>of</strong> heat treatment between the consecutive<br />

ECAP passes to relax the stresses in the sample could solve this problem.<br />

Although no investigations <strong>of</strong> the influence <strong>of</strong> ECAP on the uniformity <strong>of</strong> particle distribution<br />

in cast MMCs have been undertaken, our results suggest that the ECAP might be useful in this<br />

case too. Indeed, the difficulty <strong>of</strong> achieving a homogeneous distribution <strong>of</strong> reinforcement in


Section 7<br />

the <strong>matrix</strong> is one <strong>of</strong> the problems associated with the production <strong>of</strong> cast MMCs, as well [85,<br />

86]. In cast MMCs, the distribution <strong>of</strong> the reinforcement particles in the <strong>matrix</strong> alloy is<br />

influenced by several factors, such as the rheological behavior <strong>of</strong> the <strong>matrix</strong> melt, the particle<br />

incorporation method, interactions <strong>of</strong> the particles <strong>and</strong> the <strong>matrix</strong> before, during <strong>and</strong> after<br />

mixing, the changing particle distribution <strong>of</strong> during solidification [85, 86]. As was shown in<br />

[87], the post solidification processing <strong>of</strong> the cast MMCs extrusion or rolling can modify the<br />

particle distribution, but complete declustering can not be achieved even at high extrusion<br />

ratios. Therefore, an application <strong>of</strong> ECAP could be useful in this case, as well. But additional<br />

investigations are necessary to be performed in order to create a model analogously to the<br />

case <strong>of</strong> the PM-MMCs.<br />

97


8. Summary<br />

Section 8<br />

This work deals with the <strong>fracture</strong> <strong>properties</strong> <strong>of</strong> cast <strong>and</strong> powder <strong>metal</strong>lurgy (PM) <strong>metal</strong> <strong>matrix</strong><br />

composites (MMCs) with various particle volume fractions <strong>of</strong> Al2O3 <strong>and</strong> SiC - particles. The<br />

age hardening conditions <strong>of</strong> the <strong>matrix</strong> are also varied. The average particle size is 10 µm for<br />

the Al2O3 <strong>and</strong> 100 µm for the SiC particles. The <strong>global</strong> tensile <strong>and</strong> <strong>fracture</strong> <strong>properties</strong> are<br />

measured by st<strong>and</strong>ard tests, but the work is focused on the determination <strong>of</strong> <strong>local</strong> <strong>fracture</strong><br />

<strong>properties</strong>, i.e. the <strong>fracture</strong> <strong>properties</strong> at different positions along the crack front.<br />

A new procedure was developed to investigate the <strong>local</strong> conditions for void initiation near the<br />

crack tip. It consists <strong>of</strong>:<br />

1) the determination <strong>of</strong> the crack tip opening displacement at the moment <strong>of</strong> void<br />

initiation, CODvi, for considered particles near the crack tip;<br />

2) the estimation <strong>of</strong> the stress tensor at the position <strong>of</strong> a particle by the HRR-theory<br />

through the determined CODvi-values;<br />

3) the estimation <strong>of</strong> the stress tensors in both the particle <strong>and</strong> in the <strong>matrix</strong> by the Mori-<br />

Tanaka approach.<br />

An extremely high scatter <strong>of</strong> the critical crack tip opening displacement, CODi, which is a<br />

measure <strong>of</strong> the <strong>local</strong> <strong>fracture</strong> initiation toughness, is found for all MMCs. Neither the particle<br />

coordinates, r <strong>and</strong> θ, nor the particle dimensions have a decisive influence on the value <strong>of</strong><br />

CODi. The crack tip opening displacement at the moment <strong>of</strong> void initiation, CODvi, which is<br />

considered to be the measure <strong>of</strong> the <strong>local</strong> void initiation toughness, shows also a very high<br />

scatter in all investigated specimens. The CODvi-values slightly decrease with increasing<br />

angle to the crack plane, θ. Near-zero CODvi-values are found for alumina particles in cast<br />

MMCs. It is shown that it might be related with the shape <strong>of</strong> the crack tip plastic zone which<br />

has its maximum extension at θ ≈ 70° comparable with the distance between the particle <strong>and</strong><br />

the crack tip. No marked influence <strong>of</strong> the distance, r, or the particle size on the values <strong>of</strong><br />

CODi <strong>and</strong> CODvi is revealed for the MMCs.<br />

In the cast MMCs, a shifting <strong>of</strong> the main mechanism <strong>of</strong> the void initiation from particle<br />

<strong>fracture</strong> to particle/<strong>matrix</strong> decohesion with increasing aging time is revealed. In the PM<br />

MMC, SiC particles are only <strong>fracture</strong>d, independently on the aging condition.<br />

The analysis shows that the maximum principal stresses in particles at the moment <strong>of</strong> void<br />

initiation, σ p max, have a rather small scatter. The σ p max-values, e.g., 974±58 MPa for a Al2O3-<br />

98


Section 8<br />

15-RT Specimen decrease slightly with increasing distance, r. A slight decrease is also<br />

observed for particles lying at a large angles, θ > 50 °. The particle volume fraction has no<br />

simple effect on the σ p max-values.<br />

For the different aging conditions <strong>and</strong> particle volume fractions, the computed average values<br />

<strong>of</strong> the maximum principal stresses in the particles for <strong>fracture</strong>d particles <strong>and</strong> at the interface<br />

for debonded particles show a roughly linear dependency on the composite yield strength.<br />

This can be explained, if it is assumed that the void initiation occurs at the onset <strong>of</strong> gross<br />

plastic yielding in the <strong>matrix</strong> at the location <strong>of</strong> the first particles in front <strong>of</strong> the crack tip. A<br />

rise <strong>of</strong> the σ max interf dec-values might be also related with increasing volume fraction <strong>of</strong> the<br />

spinel phase with increasing aging time at the particle/<strong>matrix</strong> interface, leading to enhanced<br />

interfacial bonding.<br />

It is shown that the distributions <strong>of</strong> strengths <strong>of</strong> ceramic reinforcements in the cast <strong>and</strong> PM<br />

MMCs is not in contradiction to the Weibull distribution. The values <strong>of</strong> the Weibull modulus,<br />

m, determined from the values <strong>of</strong> the particles strength, are in good accordance with values<br />

given in literature.<br />

A good correlation is found between the <strong>local</strong> Ji-values evaluated from the average <strong>of</strong> the<br />

<strong>local</strong> CODi-values <strong>and</strong> the <strong>global</strong> Ji-values determined from the <strong>fracture</strong> mechanics tests for<br />

all investigated materials, excepting the PM MMCs. This is explained when it is assumed that<br />

the <strong>fracture</strong> <strong>of</strong> one particle might trigger the <strong>fracture</strong> <strong>of</strong> the neighboring particles within a<br />

certain region so that the area <strong>of</strong> <strong>local</strong> crack extension is larger than the resolution <strong>of</strong> the<br />

potential drop technique. The larger particle size (100 µm, compared to 10 µm for the cast<br />

MMC) facilitates this effect.<br />

The proposed procedure is applied to the MnS-inclusions in the mild steel St37 to study the<br />

effect <strong>of</strong> the inclusion size on the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong> <strong>and</strong> <strong>local</strong> conditions for void<br />

initiation, as well. It is shown that the CODvi-values strongly decrease with increasing<br />

diameter <strong>of</strong> MnS-inclusions. A clear reduction <strong>of</strong> the σ interf max-values with increasing diameter<br />

<strong>of</strong> the MnS-inclusions is found, as well.<br />

It is shown that equal channel angular pressing (ECAP) leads to an increase <strong>of</strong> the uniformity<br />

<strong>of</strong> the particle distribution in PM MMCs with small particle sizes <strong>and</strong> clustered particle<br />

distribution. The model proposed by Tan <strong>and</strong> Zhang [23] about the estimation <strong>of</strong> the critical<br />

value <strong>of</strong> the particle size which is re quired for a homogeneous particle distribution in the<br />

<strong>matrix</strong> is extended to the case <strong>of</strong> a combination <strong>of</strong> extrusion <strong>and</strong> ECAP. A good accordance<br />

between the modified Tan <strong>and</strong> Zhang model <strong>and</strong> the quantitative analysis <strong>of</strong> the particle<br />

distribution by the quadrat method is found.<br />

99


Section 8<br />

The effect <strong>of</strong> the homogeneity <strong>of</strong> the particle distribution in MMCs on the <strong>global</strong> <strong>and</strong> <strong>local</strong><br />

<strong>fracture</strong> <strong>properties</strong> is investigated. An increase <strong>of</strong> the <strong>global</strong> <strong>fracture</strong> <strong>properties</strong> both the<br />

<strong>fracture</strong> initiation toughness <strong>and</strong> the slope <strong>of</strong> the J-∆a-curve, dJ/d(∆a) with increasing<br />

homogeneity <strong>of</strong> the particle distribution is revealed. The scatter <strong>of</strong> the <strong>local</strong> <strong>fracture</strong> <strong>properties</strong><br />

depends on the <strong>local</strong> homogeneity <strong>of</strong> the particle distribution, A high scatter is found in the<br />

material with a clustered particle distribution. At the same time, in the material subjected to 7<br />

ECAP passes (with homogeneous particle distribution), the scatter <strong>of</strong> the CODi-values is<br />

significantly less. It is shown that the values <strong>of</strong> the <strong>local</strong> <strong>fracture</strong> toughness at the particle<br />

clusters are extremely low, whereas, in the regions with homogeneous particle distribution,<br />

higher values <strong>of</strong> the <strong>local</strong> <strong>fracture</strong> toughness appear.<br />

100


Section 9<br />

9. Appendix<br />

101


Table 9.1. Data on stress for Specimen Al2O3-10-RT.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 3x12 15.6 40 12 16.5 901 279 739 1084 600 736 880 266 739 1005 Dec.<br />

2 8x8 17.2 31 6 6.5 795 212 672 926 451 664 780 205 672 877 Fr.<br />

3 12x3 43.2 35 7 12 700 200 584 822 424 245 686 192 585 776 Fr.<br />

4 20x9 30.5 32 7 9 741 200 625 863 424 616 727 193 626 818 Fr.<br />

5 8x4 27.1 55 5 5 678 260 528 848 558 525 659 246 528 774 Fr.<br />

6 23x4 33.8 71 2 13 516 239 378 671 512 375 499 225 378 603 Fr.<br />

7 15x6 30.4 47 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

8 10x3 11.0 55 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

9 17x4 11.7 20 6 11.5 829 188 720 940 397 709 817 184 721 904 Fr.<br />

10 4x2 18 9 10 17 822 173 720 919 364 706 811 171 722 893 Dec.<br />

11 12x7 13.2 11 9.5 16 860 183 754 966 386 741 848 180 755 934 Fr.<br />

12 4x2 11.9 45 1.5 1.5 660 221 532 799 470 526 644 210 532 742 Fr.<br />

13 10x6 17.2 22 6 12 782 181 676 887 382 665 770 177 678 853 Fr.<br />

14 6x6 9.3 16 5.5 17 841 184 734 948 388 721 829 181 735 915 Dec.<br />

15 18x6 18.4 45 6 17 772 258 623 940 554 619 754 246 623 869 Fr.<br />

16 13x8 21.6 56 5 17 701 272 544 880 585 541 682 257 544 801 Fr.


Table 9.2. Data on stress for Specimen Al2O3-10-8h.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 12x5 36.9 42 3.5 6 795 254 647 952 539 639 777 243 648 890 Fr.<br />

2 6x2 42.3 28 8 8 862 235 725 1000 495 712 846 227 726 952 Dec.<br />

3 10x6 70.5 49 6.5 9 774 270 617 945 575 611 755 256 618 873 Fr.<br />

4 3x15 53.8 42 8 8 834 267 679 1002 568 672 815 255 680 934 Fr.<br />

5 5x27 64.5 4 13 13 844 214 719 959 444 700 831 207 721 926 Fr.<br />

6 8x7 35 30 6 10 855 237 717 996 500 705 839 229 718 946 Fr.<br />

7 4x13 50.4 40 3.5 5 772 241 632 917 508 622 756 230 633 862 Fr.<br />

8 17x6 33.5 8 14 16 915 232 779 1049 488 765 900 225 781 1004 Fr.<br />

9 7x6 42.3 29 8 8 863 237 725 1004 500 712 847 229 726 954 Fr.<br />

10 6x6 49 33 6.5 10 833 239 694 976 504 683 818 229 696 923 Dec.<br />

11 4x6 15.2 10 9 10 951 242 810 1095 511 798 935 234 812 1044 Fr.<br />

12 7x5 34.3 44 5.5 11 836 274 677 1011 585 671 817 261 678 938 Fr.<br />

13 10x8 38.9 46 6 8 827 278 666 1005 594 661 808 265 667 931 Dec.<br />

14 3.5x2.5 35 18 4.5 5.5 813 210 691 925 435 671 801 204 693 895 Dec.


Table 9.3. Data on stress for Specimen Al2O3-10-24h.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 3x2 74.7 64 ≈0 10 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Dec.<br />

2 3x10 82 60 3 10 825 318 640 1030 681 636 802 641 160 941 Fr.<br />

3 6x16 19.2 51 3 3…5 950 336 755 1174 725 754 925 755 169 1074 Dec.<br />

4 4x14 28.5 10 4 8…9 971 281 807 1107 570 774 956 811 143 1077 Dec.<br />

5 1.5x1.5 21.1 2 9 17 1039 300 864 1218 632 851 1019 866 152 1152 Dec.<br />

6 5x11 16.8 17 5.5 10 1027 298 854 1204 628 839 1007 855 151 1140 Fr.<br />

7 5x12 21.1 5 4 13 988 285 822 1135 584 794 971 825 145 1096 Fr.<br />

8 6x13 32 29 6 13 1002 298 829 1178 626 814 983 831 151 1114 Dec<br />

9 10x8 50.4 65 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Dec.<br />

10 6x6 47.1 58 3 9...10 864 328 675 1080 704 672 840 675 165 985 Fr.<br />

11 15x6 50.2 54 3 11 881 321 695 1090 688 691 858 695 162 999 Fr.<br />

12 4x14 42 69 0..1 3...13 724 306 547 913 646 539 703 548 155 835 Fr.<br />

13 6x12 42.7 50 2 8 886 310 706 1082 661 699 864 707 157 1001 Fr.<br />

14 11x16 36.8 9 5 13 969 280 805 1102 566 771 954 809 143 1074 Fr.<br />

15 3x5 25.6 30 5 16 1006 300 831 1186 633 818 986 833 152 1118 Dec.<br />

16 4x15 28.3 41 3 5…7 964 309 784 1159 659 776 942 785 156 1079 Fr.<br />

17 8x19 30.3 31 2.5 4.5 954 286 788 1103 586 761 937 791 146 1062 Fr.


Table 9.4. Data on stress for Specimen Al2O3-10-200h.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 5x11 31,1 68 1,5 2 868 356 662 1107 769 662 841 336 662 998 Fr.<br />

2 4x4 40,4 60 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Dec.<br />

3 24x5 35,0 57 4 4 963 358 756 1204 775 756 936 339 756 1095 Fr.<br />

4 5x4 36,9 56 4 4 968 356 761 1208 771 761 941 338 761 1099 Dec.<br />

5 11x4 28,6 32 5 6 1058 333 865 1264 707 854 1035 318 866 1183 Dec.<br />

6 11x4 19,0 46 3,5 3,5 1039 352 836 1275 760 835 1013 334 836 1170 Dec.<br />

7 6x6 27,5 4 5 10 1050 325 861 1230 675 837 1030 311 863 1172 Dec.<br />

8 6x4 32,3 46 6 6 1040 352 836 1276 760 835 1013 334 836 1170 Dec.<br />

9 6x4 50,0 52 10 10 1011 360 803 1255 779 803 984 341 803 1144 Dec<br />

10 2x5 15 30 13 13 1071 336 876 1284 717 868 1047 321 877 1197 Fr.<br />

11 8x15 12 25 12 12 1115 348 914 1348 752 912 1090 332 914 1246 Dec.<br />

12 4x7 27,4 26 7,5 7,5 1070 334 876 1279 711 867 1047 319 877 1195 Dec.<br />

13 7x17 12,5 4 10 10 1101 341 903 1324 733 899 1077 326 904 1229 Dec.<br />

14 12x4 48,7 54 6 6 984 356 778 1224 771 778 957 338 778 1116 Dec.<br />

15 2x2 15,1 32 12 12 1111 350 908 1345 756 907 1085 334 908 1242 Dec.


Table 9.5. Data on stress for Specimen Al2O3-15-RT.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 6x13 13.7 -24 4 4 899 232 764 1029 466 758 876 218 765 982 Fr.<br />

2 5x13 50.5 +58 6 6 753 293 583 924 592 581 722 269 583 852 Fr.<br />

3 7x8 31.2 +68 7 7 770 335 576 969 680 575 735 306 576 882 Fr.<br />

4 4x16 37.6 +50 5 13 792 280 630 955 566 627 763 258 630 888 Fr.<br />

5 7x6 23.8 +68 5 5…6 764 333 571 961 674 570 729 304 571 875 Fr.<br />

6 2x8 20.5 +13 6.5 8 892 219 764 1012 439 757 871 207 765 971 Fr.<br />

7 13x5 13 +22 6 14 949 242 808 1086 486 803 925 227 809 1035 Fr.<br />

8 7x4 26.2 +61 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

9 21x8 30.8 +39 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

10 27x5 50.8 +51 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

11 16x8 23.3 +2 9 12 908 220 780 1029 441 772 886 208 781 988 Fr.<br />

12 10x3 33.1 -27 4 4 808 213 683 924 427 676 787 200 685 883 Fr.<br />

13 18x6 49.8 -24 7 7 819 211 696 934 422 688 799 198 697 894 Fr.<br />

14 14x6 30.6 -32 2.5 9 773 216 647 892 433 640 752 202 649 849 Fr.


Table 9.6. Data on stress for Specimen Al2O3-15-8h.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 4x10 31.7 16 4 8 1006 292 837 1173 588 831 977 272 838 1109 Dec.<br />

2 1x1 25.2 36 3.5 4 1027 316 843 1210 639 840 995 293 844 1136 Dec.<br />

3 16x15 33.4 14 3 5 984 285 818 1145 573 812 956 265 819 1083 Fr.<br />

4 4x8 23.7 17 4.5 10.5 1033 301 858 1205 606 853 1002 280 859 1138 Fr<br />

5 6x19 21 16 5 5…6 1047 304 871 1222 613 866 1016 283 872 1154 Fr<br />

6 3x18 35.5 14 5 12 1012 293 842 1179 590 836 983 273 843 1115 Fr<br />

7 2x3 20.6 29 4 8 1045 310 865 1224 626 861 1013 288 866 1153 Dec.<br />

8 4x16 34.7 42 7 15 1041 337 846 1239 683 843 1006 312 846 1158 Fr<br />

9 6x13 33.9 41 5 12 1024 329 833 1216 664 830 990 304 833 1137 Dec.<br />

10 13x5 39.3 53 4 10 952 344 753 1155 696 751 917 316 754 1069 Fr<br />

11 3x16 40.6 37 5 18 1018 316 835 1202 638 831 986 293 836 1128 Fr<br />

12 12x5 34.7 58 1.5 6 877 332 685 1072 672 682 843 305 685 990 Fr<br />

13 4x2 15.2 53 2.5 7 981 354 776 1190 717 775 945 326 776 1102 Dec.<br />

14 10x5 25.4 6 6 9 1042 301 868 1215 606 863 1012 280 868 1148 Fr<br />

15 5x10 44.5 56 1 8 852 317 669 1036 639 665 820 291 669 960 Fr<br />

16 5x23 51.5 62 2.5 4 862 340 665 1062 687 663 827 311 666 976 Fr


Table 9.7. Data on stress for Specimen Al2O3-15-24h.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 6x6 18.9 9 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

2 1x1 12.8 39 2 3 1145 369 931 1367 751 932 1106 341 931 1272 Dec.<br />

3 2x1 8.8 26 4 4 1186 370 971 1409 754 972 1146 343 971 1314 Dec.<br />

4 10x8 36.2 38 1 4 1079 346 878 1284 702 877 1043 320 879 1198 Fr.<br />

5 15x9 31.1 32 3 9 1127 355 921 1339 721 921 1089 329 921 1250 Fr.<br />

6 11x11 25.3 38 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

7 10x5 10 25 2 3 1151 359 943 1367 730 943 1113 333 943 1276 Fr.<br />

8 13x10 21.4 28 1.5 7…9 1112 348 910 1319 706 909 1076 323 910 1233 Fr.<br />

9 5x8 17.6 43 ≈0 2…4 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

10 6x24 34 12 ≈0 8 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

11 4x2 16.6 16 1.5 5 1117 347 916 1322 704 914 1080 321 916 1237 Dec.<br />

12 12x6 14.7 38 ≈0 0…1 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

13 14x7 17.9 42 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

14 4x7 19.1 45 2.5 3 1119 376 901 1346 765 902 1079 347 901 1248 Fr.<br />

15 7x8 14.2 10 3.5 4.5 1155 358 947 1369 728 947 1117 332 947 1279 Dec.<br />

16 6x20 64.6 67 ≈0 0…5 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

17 8x9 66 72 ≈0 0…5 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

18 6x27 79.7 68 ≈0 0…5 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

19 5x15 37.9 74 ≈0 0…6 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

20 3x7 34.2 74 ≈0 0…5 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.


Table 9.8. Data on stress for Specimen Al2O3-15-200h.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 6x12 24 2 3 5 1072 332 878 1225 643 849 1045 309 883 1187 Fr.<br />

2 15x5 10 16 4.5 5 1113 346 912 1307 694 904 1079 321 914 1233 Dec.<br />

3 4x5 10.5 12 4 9 1107 344 907 1296 686 897 1074 318 909 1225 Dec.<br />

4 1x1 18.0 20 6.5 6.5 1108 345 908 1300 690 898 1074 320 909 1228 Dec.<br />

5 6x6 21.0 51 1 2 1007 355 801 1215 718 798 971 327 802 1128 Dec.<br />

6 13x9 23.0 56 1 4 976 360 768 1188 728 766 939 330 768 1098 Dec.<br />

7 6x17 32.0 52 3.5 3.5 1025 365 814 1243 741 814 987 336 814 1150 Dec.<br />

8 8x13 38.2 54 2 2 993 360 785 1205 728 783 956 331 785 1116 Dec.<br />

9 5x10 17.2 12 4 7 1092 339 895 1269 670 878 1061 314 898 1209 Dec.<br />

10 6x10 28.4 52 5 5 1039 370 825 1261 752 825 1000 340 825 1165 Dec.<br />

11 7x14 34.6 40 3 3 1069 347 868 1264 695 860 1035 321 869 1189 Fr.<br />

12 6x22 32.2 44 3 4 1061 353 855 1266 714 852 1024 326 856 1181 Dec.<br />

13 3x7 19.3 1 3 4.5 1079 334 884 1240 653 859 1050 311 888 1195 Dec.<br />

14 11x10 8 32 4 5 1124 354 919 1331 716 915 1087 328 919 1247 Dec.<br />

15 12x14 13 33 8 12 1130 357 923 1340 722 921 1093 330 924 1253 Dec.


Table 9.9. Data on stress for Specimen Al2O3-20-RT.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 22x8 16.9 19 5 12 871 218 744 976 411 736 845 200 745 944 Fr.<br />

2 11x10 52 65 2 3 620 261 468 753 496 466 587 232 469 700 Fr.<br />

3 23x8 52.3 45 ≈0 4.5 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

4 8x7 29 13 1.5 3 699 172 598 772 319 584 681 158 601 756 Fr.<br />

5 11x9 30.2 24 ≈0 12.5 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

6 4x13 30.3 8 3 13 753 183 645 834 343 633 732 169 648 814 Dec.<br />

7 7x21 21.1 21 ≈0 3 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

8 12x18 26.8 10 2.5 5 748 183 641 830 341 629 728 168 644 809 Fr.<br />

9 9x2.5 14.5 11 ≈0 2 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

10 11x11 11.6 15 0.5 3 685 169 586 757 314 572 668 156 589 742 Fr.


Table 9.10. Data on stress for Specimen Al2O3-20-8h.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 14x25 54.2 57 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

2 4x21 46.2 68 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

3 18x13 44 33 4 6 1046 319 861 1210 606 857 1006 287 862 1148 Fr.<br />

4 6x22 52.1 36 3 10 1025 318 840 1188 606 836 984 287 841 1127 Fr.<br />

5 22x9 53 50 2 9 961 335 766 1134 639 764 918 300 767 1066 Fr.<br />

6 6x8 36.6 58 1 5 902 341 704 1079 651 701 857 305 704 1009 Fr.<br />

7 5x17 35.5 49 3 5 1009 350 806 1191 669 803 963 314 806 1120 Fr.<br />

8 6x20 25.8 14 5 7 1080 316 896 1242 602 892 1040 286 897 1182 Fr.<br />

9 6x16 23.2 50 1.5 7 990 345 790 1170 658 787 945 309 790 1099 Fr.<br />

10 7x23 47.1 48 4 12 1015 350 812 1198 669 809 969 314 812 1126 Fr.<br />

11 11x18 41 43 4 9 1041 340 843 1217 649 841 997 306 844 1149 Fr.<br />

12 5x14 30.3 50 2 8 991 345 791 1171 659 788 946 309 791 1100 Fr.


Table 9.11. Data on stress for Specimen Al2O3-20-24h.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 3x6 13.5 -3 ≈0 5 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Dec.<br />

2 7x8 20 -6 ≈0 2.5 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

3 3x6 7.6 -18 0..1 1 1089 327 898 1251 620 890 1048 296 900 1194 Dec.<br />

4 2x2 17.3 +56 3.5 3.5 1043 386 820 1249 740 820 992 344 819 1164 Dec.<br />

5 6x10 37.3 +64 1 2…5 898 358 691 1086 684 690 851 319 691 1010 Fr.<br />

6 25x7 15.7 +11 ≈0 3 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

7 1.5x1.5 7.4 +28 1.5 2 1119 340 921 1294 649 918 1075 307 922 1228 Dec.<br />

8 4x5 31.9 -18 ≈0 5 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Dec.<br />

9 7x12 33.2 -45 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Fr.<br />

10 6x10 16.5 -20 2.5 6 1098 330 905 1263 627 899 1056 298 907 1203 Fr.


Table 9.12. Data on stress for Specimen Al2O3-20-200h.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 6x9 45 48 4 4 1067 367 854 1258 700 851 1019 329 854 1183 Fr.<br />

2 12x25 25,0 9 8 8 1127 349 923 1290 655 908 1086 315 927 1238 Dec.<br />

3 9x12 24,8 50 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 ≈0 Dec.<br />

4 8x19 47,0 52 3 4 1035 368 821 1228 703 819 987 329 822 1150 Fr.<br />

5 14x25 75,0 58 4 5 993 372 777 1189 712 776 944 332 777 1109 Fr.<br />

6 20x9 15,9 12 2 4 1100 341 899 1239 626 871 1065 309 906 1208 Fr.<br />

7 4x8 32,0 50 1 1 1027 359 818 1208 681 812 981 321 819 1139 Fr.<br />

8 12x18 26,0 53 3 3.5 1046 376 828 1245 719 828 996 336 828 1164 Dec.<br />

9 10x26 39,0 52 1 2 1010 359 801 1191 681 795 965 322 803 1123 Fr.<br />

10 10x12 17,0 28 3,5 5 1121 351 916 1288 660 903 1079 316 919 1232 Dec.<br />

11 19x24 22,4 26 2 2 1095 342 894 1237 629 867 1059 309 901 1203 Dec.<br />

12 11x23 35,0 34 2,5 4 1092 346 890 1245 642 869 1054 312 895 1202 Dec.<br />

13 12x9 17,8 54 4 5 1058 383 836 1263 735 837 1007 342 836 1178 Fr.<br />

14 19x8 40,6 64 3 3 961 380 741 1163 728 741 911 338 741 1079 Fr.<br />

15 19x14 36,5 34 4,5 8 1108 351 904 1275 659 891 1067 316 907 1220 Dec.


Table 9.13. Data on stress for Specimen SiC-10-15min.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 135x85 174 47 10 12 610 210 488 736 443 479 596 199 490 687 Fr.<br />

2 100x60 114 54 15 15 689 261 538 856 558 532 670 247 538 785 Fr.<br />

3 85x35 280 58 42 42 693 276 533 871 591 528 673 261 533 794 Fr.<br />

4 80x65 215 53 21 21 657 245 514 812 524 508 640 232 515 746 Fr.<br />

5 160x80 461 74 60 60 617 296 446 811 635 443 596 278 446 724 Fr.<br />

6 85x130 308 61 64 64 722 298 550 917 640 546 701 281 550 831 Fr.<br />

7 160x100 204 55 28 30 692 266 537 862 570 532 673 252 538 789 Fr.<br />

8 150x90 458 69 59 59 636 288 469 824 619 465 615 271 469 740 Fr.<br />

9 170x55 288 70 20 20 568 260 417 735 557 412 549 245 417 662 Fr.<br />

10 100x45 180 71 10 12 543 251 398 704 537 393 525 237 398 635 Fr.<br />

11 140x105 291 65 26 26 610 264 457 780 565 453 591 249 458 706 Fr.<br />

12 170x70 266 71 25 28 595 275 435 773 590 431 575 259 436 694 Fr.<br />

13 60x60 235 56 8 10 540 210 418 667 443 410 526 198 419 616 Fr.<br />

14 125x75 169 72 10 10 545 256 397 709 546 393 527 240 398 637 Fr.<br />

15 105x120 51 48 14 16 800 279 638 981 599 633 780 265 639 903 Fr.


Table 9.14. Data on stress for Specimen SiC-10-8h.<br />

Mechanism<br />

HRR-theory<br />

Composite<br />

σ eq vi σ<br />

[MPa]<br />

mean vi<br />

[MPa]<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ max vi<br />

[MPa]<br />

1 80x105 85 29 11 11 1114 325 919 Fr.<br />

2 50x95 85 43 8 8 1077 351 868 Fr.<br />

3 70x110 158 24 12 12 1062 304 877 Fr.<br />

4 130x85 116 26 10 10 1076 310 888 Fr.<br />

5 70x100 165 31 10 10 1052 311 863 Fr.<br />

6 105x80 168 33 10 10 1050 311 861 Fr.<br />

7 120x150 163 44 6 6 995 328 799 Fr.<br />

8 110x125 91 4 8 8 1061 296 883 Fr.<br />

9 80x90 101 7 12 12 1087 304 905 Fr.<br />

10 120x85 155 62 23 23 1018 404 783 Fr.<br />

11 150x75 115 48 8 8 1032 355 823 Fr.<br />

12 80x80 134 6 13 15 1070 299 890 Fr.<br />

13 75x65 172 10 20 22 1087 304 905 Fr.<br />

14 175x70 93 54 7 7 1009 369 793 Fr.<br />

15 150x105 130 10 10 14 1051 294 874 Fr.


Table 9.15. Data on stress for Specimen SiC-10-24h.<br />

Mechanism<br />

HRR-theory<br />

Composite<br />

σ eq vi σ<br />

[MPa]<br />

mean vi<br />

[MPa]<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ max vi<br />

[MPa]<br />

1 130x105 163 0 18 20 1033 290 855 Fr.<br />

2 120x140 157 31 10 10 1013 300 830 Fr.<br />

3 85x50 78 21 7 10 1029 295 849 Fr.<br />

4 210x120 100 18 7 12 1006 286 831 Fr.<br />

5 145x115 120 10 13 16 1034 291 855 Fr.<br />

6 160x120 137 60 15 15 962 374 742 Fr.<br />

7 130x90 101 24 19 22 1089 312 900 Fr.<br />

8 125x50 82 28 12 15 1074 311 889 Fr.<br />

9 150x75 60 3 7 7 1037 289 862 Fr.<br />

10 125x120 73 37 6 8 1034 320 844 Fr.<br />

11 130x95 239 57 2 5 809 306 627 Fr.<br />

12 145x140 181 55 3 3 859 317 671 Fr.<br />

13 170x50 113 26 15 18 1064 306 881 Fr.<br />

14 95x95 120 5 6 9 972 271 808 Fr.<br />

15 90x90 82 5 5 8 986 275 820 Fr.


Table 9.16. Data on stress for Specimen SiC-10-200h.<br />

Mechanism<br />

σ interf max<br />

[MPa]<br />

HRR-theory Mean-Field theory<br />

Composite Particle Matrix<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Particle Size<br />

[µm]<br />

σ mean vi<br />

[MPa]<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

σ max vi<br />

σ mean vi<br />

σ eq vi<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

[MPa]<br />

σ max vi<br />

[MPa]<br />

1 150x80 121 3 7 10 820 228 685 845 421 595 818 226 695 911 Fr.<br />

2 130x130 158 14 4 4 776 218 647 763 381 536 778 219 659 866 Fr.<br />

3 130x90 127 41 3 6 782 251 634 883 500 590 770 241 639 875 Fr.<br />

4 230x150 125 37 8 10 845 261 693 971 534 659 831 250 696 943 Fr.<br />

5 150x145 109 38 8 10 853 264 699 986 544 669 839 253 702 952 Fr.<br />

6 175x150 61 7 18 20 925 257 774 1039 523 733 911 248 779 1022 Fr.<br />

7 160x105 50 8 13 16 917 256 768 1028 518 725 905 247 772 1015 Fr.<br />

8 80x45 134 72 13 13 732 322 546 944 692 543 709 303 546 1015 Fr.<br />

9 170x105 139 43 8 10 828 268 671 968 554 645 812 256 674 849 Fr.<br />

10 95x140 143 39 10 14 849 267 693 988 552 666 834 255 696 927 Fr.<br />

11 190x100 215 37 3 5 757 234 619 802 437 544 752 228 627 948 Fr.<br />

12 220x110 123 49 17 17 863 300 689 1052 640 681 842 285 690 847 Fr.<br />

13 160x140 137 56 18 20 831 310 651 1031 665 645 808 294 652 974 Fr.<br />

14 125x125 126 55 10 10 801 299 627 989 638 619 780 283 628 945 Fr.<br />

15 150x70 108 54 8 8 776 218 647 763 381 536 778 219 659 866 Fr.<br />

16 145x60 165 54 30 30 869 311 688 1070 668 683 846 295 689 983 Fr.<br />

17 115x95 310 42 5 5 757 244 614 838 475 559 749 235 620 849 Fr.<br />

18 145x110 122 1 12 14 853 237 713 908 454 640 847 232 721 945 Fr.<br />

19 115x75 188 51 13 13 813 288 646 987 609 633 794 273 647 919 Fr.


Table 9.17. Results <strong>of</strong> the stereophotogrammetic analysis<br />

<strong>and</strong> estimation <strong>of</strong> the interfacial stresses for individual MnS-inclusion.<br />

σ interf max<br />

[MPa]<br />

CODi<br />

[µm]<br />

CODvi<br />

[µm]<br />

θ<br />

[°]<br />

r<br />

[µm]<br />

Diameter<br />

<strong>of</strong><br />

inclusion<br />

[µm]<br />

Inclusion Length<br />

<strong>of</strong><br />

inclusion<br />

[µm]<br />

1 72 5 60 1 29 38 1170<br />

2 160 7 119 7 18 58 964<br />

3 130 3 67 1 26 68 1129<br />

4 8 4 98 24 42 70 1202<br />

5 28 6 100 16 27 73 1081<br />

6 80 3,5 96 42 45 70 1260<br />

7 60 2,5 41 36 51 71 1478<br />

8 80 4 77 47 42 86 1283<br />

9 80 4 59 12 42 64 1261


1 2<br />

1 2<br />

3<br />

3 4 5 6<br />

5<br />

9<br />

119<br />

7<br />

7<br />

8<br />

9<br />

6<br />

10 11<br />

12<br />

50 µm<br />

10<br />

11<br />

Fig. 9.1. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-10-RT.<br />

12


1 2<br />

1 2<br />

3<br />

3 4 5 6<br />

5<br />

9<br />

119<br />

7<br />

7<br />

8<br />

9<br />

6<br />

10 11<br />

12<br />

50 µm<br />

10<br />

11<br />

Fig. 9.1. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-10-RT.<br />

12


1<br />

1<br />

2<br />

2<br />

3<br />

3<br />

121<br />

4<br />

4<br />

3<br />

5<br />

6<br />

50 µm<br />

Fig. 9.3. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-10-8h<br />

5<br />

6<br />

7


8 9 10<br />

6<br />

10<br />

122<br />

11 12 13 14<br />

11<br />

12<br />

13 14<br />

50 µm<br />

Fig. 9.4. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-10-8h.


1 2<br />

2<br />

3<br />

4 5<br />

5 7<br />

123<br />

6 7 8<br />

50 µm<br />

Fig. 9.5. Fracture surfaces with marked particles <strong>of</strong> the specimen Al2O3-10-24h.<br />

8


9<br />

10 11 12 13<br />

124<br />

50 µm<br />

Fig. 9.6. The <strong>fracture</strong> surface with marked particles <strong>of</strong> the Specimen Al2O3-10-24h.


14 15<br />

14 15<br />

16<br />

16<br />

17<br />

17<br />

125<br />

50 µm<br />

Fig. 9.7. The <strong>fracture</strong> surface with marked particles <strong>of</strong> the Specimen Al2O3-10-24h.


1 2<br />

1<br />

2<br />

3<br />

3<br />

4<br />

4<br />

126<br />

5<br />

5<br />

6<br />

9<br />

7<br />

50 µm<br />

Fig. 9.8. The <strong>fracture</strong> surface with marked particles <strong>of</strong> the Specimen Al2O3-10-200h.<br />

7


8<br />

9 10 11<br />

8 9 10 11<br />

12<br />

12<br />

127<br />

13<br />

50 µm<br />

13<br />

14<br />

14<br />

Fig. 9.9. The <strong>fracture</strong> surface with marked particles <strong>of</strong> the Specimen Al2O3-10-200h.<br />

15<br />

15


1<br />

1<br />

2<br />

2<br />

3 4<br />

3 4<br />

5<br />

5<br />

128<br />

6<br />

7<br />

6 7<br />

50 µm<br />

Fig. 9.10. The <strong>fracture</strong> surface with marked particles <strong>of</strong> the Specimen Al2O3-15-RT.


8 9 10<br />

8 9 10<br />

11<br />

11<br />

129<br />

12 13 14<br />

12 13 14<br />

50 µm<br />

Fig. 9.11. The <strong>fracture</strong> surface with marked particles <strong>of</strong> the Specimen Al2O3-15-RT.


Fig. 9.12. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-15-8h.<br />

130


Fig. 9.13. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-15-8h.<br />

131


1 2 3<br />

1<br />

2 3<br />

4 5 6 7 8 9 10<br />

4 5<br />

132<br />

6 7<br />

50 µm<br />

8 9 10<br />

Fig. 9.14. The <strong>fracture</strong> surface with marked particles <strong>of</strong> the Specimen Al2O3-15-24h.


Fig. 9.15. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-15-24h.<br />

133


Fig. 9.16. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-15-200h.<br />

134


Fig. 9.17. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-15-200h.<br />

135


Fig. 9.18. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-20-RT.<br />

136


Fig. 9.19. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-20-RT.<br />

137


Fig. 9.20. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-20-8h.<br />

138


Fig. 9.21. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-20-8h.<br />

139


Fig. 9.22. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-20-24h.<br />

140


Fig. 9.23. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-20-24h.<br />

141


Fig. 9.24. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-20-200h.<br />

142


Fig. 9.25. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-20-200h.<br />

143


Fig. 9.26. Fracture surface with marked particles <strong>of</strong> the Specimen Al2O3-20-200h.<br />

144


Fig. 9.27. Fracture surface with marked particles <strong>of</strong> the Specimen SiC-10-15min.<br />

145


Fig. 9.28. Fracture surface with marked particles <strong>of</strong> the Specimen SiC-10-15min.<br />

146


Fig. 9.29. Fracture surface with marked particles <strong>of</strong> the Specimen SiC-10-8h.<br />

147


Fig. 9.30. Fracture surface with marked particles <strong>of</strong> the Specimen SiC-10-8h.<br />

148


Fig. 9.31. Fracture surface with marked particles <strong>of</strong> the Specimen SiC-10-24h.<br />

149


Fig. 9.32. Fracture surface with marked particles <strong>of</strong> the Specimen SiC-10-24h.<br />

150


Fig. 9.33. Fracture surface with marked particles <strong>of</strong> the Specimen SiC-10-200h.<br />

151


Fig. 9.34. Fracture surface with marked particles <strong>of</strong> the Specimen SiC-10-200h.<br />

152


Fig. 9.35. Fracture surface with marked particles <strong>of</strong> the Specimen SiC-10-200h.<br />

153


Fig. 12.36. Fracture surface <strong>of</strong> the mild steel St37.<br />

154<br />

50 µm


References<br />

[1] Suresh S, Mortensen A, Needleman A. Fundamentals <strong>of</strong> Metal Matrix Composites.<br />

Butterworth – Heinemann, Boston, USA, 1993.<br />

[2] Clyne TV, Withers PJ. An Introduction to Metal Matrix Composites. Cambridge<br />

University Press, UK, 1993.<br />

[3] Shyong JH, Derby B. The deformation characteristics <strong>of</strong> SiC particulate-reinforced<br />

aluminium alloy 6061. Materials Science <strong>and</strong> Engineering. 1995, A197, pp. 11-18.<br />

[4] Chang-You Chen, Chuen-Guang Chao. Effect <strong>of</strong> particle-size distribution on the<br />

<strong>properties</strong> <strong>of</strong> high-volume-fraction SiC-Al-based composites. Metallurgical <strong>and</strong><br />

Materials Transactions. 2000, 31A, pp. 2351-2359.<br />

[5] Kapoor R, Vecchio KS. Deformation behavior <strong>and</strong> failure mechanisms in particulate<br />

reinforced 6061 Al <strong>metal</strong> <strong>matrix</strong> composites. Materials Science <strong>and</strong> Engineering. 1995,<br />

A202, pp. 63-75.<br />

[6] Kaczmar JW, Pietrzak K, Wlosinski W. The production <strong>and</strong> application <strong>of</strong> <strong>metal</strong> <strong>matrix</strong><br />

composite materials. Journal <strong>of</strong> Materials Processing Technology. 2000, 206, pp. 58-67.<br />

[7] Doel TJA, Bowen P. Effect <strong>of</strong> particle size <strong>and</strong> <strong>matrix</strong> aging condition on toughness <strong>of</strong><br />

particle reinforced aluminium based <strong>metal</strong> <strong>matrix</strong> composites. Materials Science <strong>and</strong><br />

Technology. 1996, 12, pp. 586-594.<br />

[8] McDaniels DL. Analysis <strong>of</strong> stress-strain, <strong>fracture</strong>, <strong>and</strong> ductility behavior <strong>of</strong> aluminum<br />

<strong>matrix</strong> composites containing discontinuous silicon carbide reinforcements. Metallurgical<br />

Transactions. 1985, 16A, pp. 1105-1115.<br />

[9] Stephens JJ, Lucas JP, Hosking FM. Cast Al-7Si composites: effect <strong>of</strong> particle type <strong>and</strong><br />

size on mechanical <strong>properties</strong>. Scripta Metallurgica. 1988, 22, pp.1307-1312.<br />

[10] Arsenault RJ, Wang L, Feng CR. Strengthening <strong>of</strong> composites due to microstructural<br />

changes in the <strong>matrix</strong>. Acta Metallurgica et Materialia. 1991, 39, pp. 47-57.<br />

[11] Flom Y, Arsenault RJ. Effect <strong>of</strong> particle size on <strong>fracture</strong> toughness <strong>of</strong> SiC/Al composite<br />

material. Acta Metallurgica. 1989, 37, pp. 2413-2423.<br />

[12] Lew<strong>and</strong>owski JJ, Singh PM. In: Fatigue <strong>and</strong> Fracture, ASM H<strong>and</strong>book, vol. 19. ASM<br />

International, 1997, p. 895.<br />

[13] Kamat SV, Hirth JP, Mehrabian R. Mechanical <strong>properties</strong> <strong>of</strong> particulate-reinforced<br />

aluminum-<strong>matrix</strong> composites. Acta Metallurgica et Materialia. 1989, 37, 2395-2402.<br />

155


[14] Lew<strong>and</strong>owski JJ, Liu C, Hunt WH. Effects <strong>of</strong> <strong>matrix</strong> microstructure <strong>and</strong> particle<br />

distribution on <strong>fracture</strong> <strong>of</strong> an aluminum <strong>metal</strong> <strong>matrix</strong> composite. Materials Science <strong>and</strong><br />

Engineering. 1989, A107, pp. 241-255.<br />

[15] Manoharan M, Lew<strong>and</strong>owski JJ. Crack initiation <strong>and</strong> growth toughness <strong>of</strong> an aluminum<br />

<strong>metal</strong>-<strong>matrix</strong> composite. Acta Metallurgica et Materialia. 1990, 38, pp. 489-496.<br />

[16] P<strong>and</strong>ey AB, Majumdar BS, Miracle D.B. Deformation <strong>and</strong> <strong>fracture</strong> <strong>of</strong> a particle-<br />

reinforced aluminum alloy composites: Part I. Experiments. Metallurgical <strong>and</strong> Materials<br />

Transactions. 2000, 31, pp.921-936.<br />

[17] Majumdar BS, P<strong>and</strong>ey AB. Deformation <strong>and</strong> <strong>fracture</strong> <strong>of</strong> a particle-reinforced aluminum<br />

alloy composites: Part II. Modelling. Metallurgical <strong>and</strong> Materials Transactions. 2000, 31,<br />

pp.937-950.<br />

[18] Hahn GT, Rosenfield AR. Metallurgical factors affecting <strong>fracture</strong> toughness <strong>of</strong> aluminum<br />

alloys. Metallurgical <strong>and</strong> Materials Transactions. 1975, 6A, pp.653-670.<br />

[19] Rice JR, Johnson MA. In: Inelastic behavior <strong>of</strong> solids, Kanninen MF, Adler WF,<br />

Rosenfield AR, Jaffe R.I., editors. McGraw-Hill, New York, Battelle Memorial Institute,<br />

Columbus, USA, 1999, pp. 641-672.<br />

[20] Shih CF. Relationships between the J-integral <strong>and</strong> the crack opening displacement for<br />

stationary <strong>and</strong> extending cracks. Journal <strong>of</strong> the Mechanics <strong>and</strong> Physics <strong>of</strong> Solids. 1981,<br />

29, pp. 305-326.<br />

[21] Shakeri HR., Wang Z. Effect <strong>of</strong> alternative aging process on the <strong>fracture</strong> <strong>and</strong> interfacial<br />

<strong>properties</strong> <strong>of</strong> particulate Al2O3 reinforced Al(6061) <strong>metal</strong> <strong>matrix</strong> composites.<br />

Metallurgical <strong>and</strong> Materials Transactions. 2002, 33, pp. 1699-1713.<br />

[22] Lloyd DJ. Aspects <strong>of</strong> <strong>fracture</strong> in particulate reinforced <strong>metal</strong> <strong>matrix</strong> composites. Acta<br />

Metallurgica et Materialia. 1991, 39, pp. 59-71.<br />

[23] Tan MJ, Zhang X. Powder <strong>metal</strong> <strong>matrix</strong> composites: selection <strong>and</strong> processing. Materials<br />

Science <strong>and</strong> Engineering. 1998, A244, pp. 80-85.<br />

[24] Prangnell PB, Barnes SJ, Roberts SM, Withers PJ. The effect <strong>of</strong> particle distribution on<br />

damage formation in particulate reinforced <strong>metal</strong> <strong>matrix</strong> composites deformed in<br />

compression. Materials Science <strong>and</strong> Engineering. 1996, A220, pp. 41-56.<br />

[25] Lloyd DJ. In: Lew<strong>and</strong>owski JJ, Hunt W, editors. Intrinsic Fracture Mechanisms in<br />

Inorganic Composite Systems. TMS Annual Meeting, 1994, Las Vegas, USA, pp. 39-47.<br />

[26] Corbin SF, Wilkinson DS. Influence <strong>of</strong> <strong>matrix</strong> strength <strong>and</strong> damage accumulation on the<br />

mechanical response <strong>of</strong> a particulate <strong>metal</strong> <strong>matrix</strong> composite. Acta Metallurgica et<br />

Materialia. 1994, 42, pp. 1329-1335<br />

156


[27] Corbin SF, Wilkinson DS. The influence <strong>of</strong> particle distribution on the mechanical<br />

response <strong>of</strong> a particulate <strong>metal</strong> <strong>matrix</strong> composite. Acta Metallurgica et Materialia. 1994,<br />

42, pp. 1311-1318<br />

[28] Christman T, Needleman A, Suresh S. An experimental <strong>and</strong> numerical study <strong>of</strong><br />

deformation in <strong>metal</strong>-ceramic composites. Acta Metallurgica. 1989, 37, pp. 3029-3050.<br />

[29] Wang Z, Chen TK, Lloyd DJ. A stress distribution in particulate reinforced <strong>metal</strong> <strong>matrix</strong><br />

composites subjected to external load. Metallurgical Transactions. 1993, 24A, pp. 197-<br />

207.<br />

[30] Tao S, Townley N, Boyd JD. Effects <strong>of</strong> particulate spacing distribution on internal stress<br />

<strong>and</strong> damage in MMCs. Microstructural Science. 1994, 22, pp 249-259.<br />

[31] Goods SH, Brown LM. The nucleation <strong>of</strong> cavities by plastic deformation. Acta<br />

Metallurgica. 1979, 27, pp. 1-15.<br />

[32] Tanaka B, Mori T, Nakamura T. Cavity formation at the interface <strong>of</strong> a spherical inclusion<br />

in a plastically deformed <strong>matrix</strong>. Philosophical Magazine. 1970, 21, pp. 267-279.<br />

[33] Ashby MF. Work hardening <strong>of</strong> dispersion-hardened crystals. Philosophical Magazine.<br />

1966, 14, pp. 1157-1178.<br />

[34] Argon AS, Im J, Safoglu R. Cavity formation from inclusions in ductile <strong>fracture</strong>.<br />

Metallurgical Transactions. 1975, 6, pp. 825-838.<br />

[35] Argon AS, Im J. Separation <strong>of</strong> second phase particles in spheroidized 1045 steel, Cu-<br />

0.6Pct Cr alloy, <strong>and</strong> maraging steel in plastic straining. Metallurgical Transactions. 1975,<br />

6, pp. 839-851.<br />

[36] Argon AS, Im J, Needleman A. Distribution <strong>of</strong> plastic strain <strong>and</strong> negative pressure in<br />

necked steel <strong>and</strong> copper bars. Metallurgical Transactions. 1975, 6, pp. 815-824.<br />

[37] Flom Y, Arsenault RJ. Interfacial bond strength in an aluminum alloy 6061–SiC<br />

composite. Materials Science <strong>and</strong> Engineering. 1986, A77, pp. 191-197.<br />

[38] Beremin FM. Cavity formation from inclusions in ductile <strong>fracture</strong> <strong>of</strong> A508 steel.<br />

Metallurgical Transactions. 1981, 12, pp. 723-729.<br />

[39] Berveiller M, Zaoui A. An extension <strong>of</strong> the self-consistent scheme to plastically-flowing<br />

polycrystals. Journal <strong>of</strong> the Mechanics <strong>and</strong> Physics <strong>of</strong> Solids. 1978, 26, pp. 325-344.<br />

[40] Toda H, Kobayashi T, Takahashi A. Micromechanism <strong>of</strong> <strong>fracture</strong> in wrought aluminum<br />

alloy containing coarse inclusion particles. Aluminum Transactions. 1999, 1, pp. 109-116.<br />

[41] Hutchinson JW. Singular behavior at the end <strong>of</strong> a tensile crack tip in a hardening material.<br />

Journal <strong>of</strong> the Mechanics <strong>and</strong> Physics <strong>of</strong> Solids. 1968, 16, pp. 13-31.<br />

157


[42] Rice JR, Rosengren GF. Plane strain deformation near a crack tip in a power-law<br />

hardening material. Journal <strong>of</strong> the Mechanics <strong>and</strong> Physics <strong>of</strong> Solids. 1968, 16, pp. 1-12.<br />

[43] Chan KS. A new criterion for brittle-to-ductile <strong>fracture</strong> transition. Acta Metallurgica.<br />

1989, 37, pp. 1217-1226.<br />

[44] Hutchinson J, Paris P. Stability analysis <strong>of</strong> J-controlled crack growth. American Society<br />

for Testing <strong>and</strong> Materials. 1979, 668, pp. 37-64.<br />

[45] Scherer S, Werth P, Pinz A, Tatschl A, <strong>and</strong> Kolednik O. Automatic surface reconstruction<br />

using SEM images based on a new computer vision approach. In: Kiely CJ, editor.<br />

Electron Microscopy <strong>and</strong> Analysis. Institute <strong>of</strong> Physics Publishing. Bristol, U.K. 1999,<br />

pp. 107-110.<br />

[46] Scherer S, Kolednik O. A New System for Automatic Surface Analysis in SEM.<br />

Microscopy <strong>and</strong> Analysis. 2001, 70, pp. 15-17.<br />

[47] Kolednik O, Schwarzböck P. Quantitative fractography by means <strong>of</strong> a new digital image<br />

analysis system. In: Fuentes M, Elices M, editors. Fracture Mechanics: Applications <strong>and</strong><br />

Challanges, Proceedings <strong>of</strong> ECF13, 2000, Elsevier, Amsterdam, Paper 1U.60.<br />

[48] Stampfl J, Scherer S, Berchthaler M, Gruber M, Kolednik O. Determination <strong>of</strong> the<br />

<strong>fracture</strong> toughness by automatic image processing. International Journal <strong>of</strong> Fracture.<br />

1996, 78, pp. 35-44.<br />

[49] Tatschl A, Kolednik O. A new tool for the experimental characterization <strong>of</strong> micro-<br />

plasticity. Materials Science <strong>and</strong> Engineering. 2003, 339, pp. 265-280.<br />

[50] Stampfl J, Kolednik O. The separation <strong>of</strong> the <strong>fracture</strong> energy in <strong>metal</strong>lic materials.<br />

International Journal <strong>of</strong> Fracture. 2000, 101, pp. 321-345.<br />

[51] Semprimoschnig COA, Stampfl J, Pippan R, Kolednik O. A new powerful tool for<br />

surveying cleavage <strong>fracture</strong> surfaces. Fatigue <strong>and</strong> Fracture <strong>of</strong> Engineering Materials <strong>and</strong><br />

Structures. 1997, 20, pp.1541-1550.<br />

[52] Tatschl A, Gilbert CJ, Schroeder V, Pippan R, Ritchie O. Stereophotogrammetric<br />

investigation <strong>of</strong> overload <strong>and</strong> cyclic fatigue <strong>fracture</strong> surface morphologies in a Zr-Ti-Ni-<br />

Cu-Be bulk <strong>metal</strong>lic glass. Journal <strong>of</strong> Materials Resolution. 2000, 15, pp. 898-903.<br />

[53] Hebesberger T, Semprimoschnig COA, Pippan R, Kolednik O. In: Kim YM, Dimidik<br />

DM, Loretto MH. editors. 2 nd International Symposium on Gamma Titanium Aluminides.<br />

The Minerals, Metals <strong>and</strong> Materials Society. 1999, Warrendale, PA, USA, pp. 573.<br />

[54] Qiu H, Enoki M, Kawaguchi Y, Kishi T. A model for the static <strong>fracture</strong> toughness <strong>of</strong><br />

ductile structural steel. Engineering Fracture Mechanics. 2003, 70, pp. 599-609.<br />

158


[55] McMeeking RM. Finite deformation analysis <strong>of</strong> crack tip opening in elastic-plastic<br />

materials <strong>and</strong> implication for <strong>fracture</strong>. Journal <strong>of</strong> the Mechanics <strong>and</strong> Physics <strong>of</strong> Solids.<br />

1977, 25, pp. 357-381.<br />

[56] Jun S, Zeng-Jie D, Ming-Jing TU. Effect <strong>of</strong> stress triaxiality levels in crack tip regions on<br />

the characteristics <strong>of</strong> void growth <strong>and</strong> <strong>fracture</strong> criteria. Engineering Fracture Mechanics.<br />

1991, 39, pp. 1051-1060.<br />

[57] Baaser H, Gross D. Analysis <strong>of</strong> void growth in a ductile material in front <strong>of</strong> a crack tip.<br />

Computational Materials Science. 2003, 26, pp. 26-35.<br />

[58] Shih CF. Relationships between the J-integral <strong>and</strong> the crack opening displacement for<br />

stationary <strong>and</strong> extending cracks. Journal <strong>of</strong> the Mechanics <strong>and</strong> Physics <strong>of</strong> Solids. 1981,<br />

29, pp. 305-326.<br />

[59] Shih CF. Tables <strong>of</strong> HRR Singular Field Quantities, MRL E-147. Materials Research<br />

Laboratory, Brown University. Providence, RI, USA, 1983.<br />

[60] Eshelby JD. The Determination <strong>of</strong> the Elastic Field <strong>of</strong> an Ellipsoidal Inclusion <strong>and</strong><br />

Related Problems. Proceedings <strong>of</strong> Royal Society. London. 1957, A241, pp. 376-396.<br />

[61] T<strong>and</strong>on GP, Weng GJ. The effect <strong>of</strong> the aspect ratio <strong>of</strong> inclusions on the elastic <strong>properties</strong><br />

<strong>of</strong> unidirectionally aligned composites. Polymer Composites. 1984, 5(4), pp. 327-333.<br />

[62] Hill R. Continuum micro-mechanics <strong>of</strong> elastoplastic polycrystals. Journal <strong>of</strong> the<br />

Mechanics <strong>and</strong> Physics <strong>of</strong> Solids. 1965, 13, pp. 89-101.<br />

[63] Benveniste Y. A new approach to the application <strong>of</strong> Mori-Tanaka's theory in composite<br />

materials. Mechanics <strong>of</strong> Materials. 1987, 6, pp. 147-157.<br />

[64] Mori T, Tanaka K. Average stress in a <strong>matrix</strong> <strong>and</strong> average elastic energy <strong>of</strong> materials with<br />

misfitting inclusions. Acta Metallurgica. 1973, 23, pp. 571-574.<br />

[65] Pettermann HE, Böhm HJ, Rammerstorfer FG. Some direction-dependent <strong>properties</strong> <strong>of</strong><br />

<strong>matrix</strong>-inclusion type composites with given reinforcement orientation distributions.<br />

Composites Part B: Engineering. 1997, 28, pp. 253-265.<br />

[66] Ponte Castaneda P, Suquet P. In: E.van der Giessen, Wu Ty, editors. Advances in Applied<br />

Mechanics. 34, Academic Press, New York, NY, 1998. p. 171.<br />

[67] Hu G. A method <strong>of</strong> plasticity for general aligned spheroidal void <strong>of</strong> fiber-reinforced<br />

composites. International Journal <strong>of</strong> Plasticity. 1996, 12(4), pp. 439-449.<br />

[68] Qiu YP, Weng GJ. A theory <strong>of</strong> plasticity for porous materials <strong>and</strong> particle reinforced<br />

composites. Journal <strong>of</strong> Applied Mechanics. 1992, 59, pp. 261-268.<br />

[69] Brantner HP. Bruchzähigkeit teilchenverstärkter Aluminiumlegirungen. Diploma Thesis.<br />

University <strong>of</strong> Leoben, 2004.<br />

159


[70] ASTM E1820-01, “St<strong>and</strong>ard Test Method for Measurement <strong>of</strong> Fracture Toughness”,<br />

Annual Book <strong>of</strong> St<strong>and</strong>ards, American Society <strong>of</strong> Testing <strong>and</strong> Materials, Philadelphia,<br />

USA, 2001.<br />

[71] Levi CG, Abbaschian GJ, Mehrabian R. Interface interactions during fabrication <strong>of</strong><br />

aluminum alloy-alumina fiber composites. Metallurgical Transactions. 1978, 9, pp. 697-<br />

711.<br />

[72] Munitz A, Metzger M, Mehrabian R. The interface phase in Al-Mg / Al2O3 composites.<br />

Metallurgical Transactions. 1979, 10, pp. 1491-1497.<br />

[73] Schwalbe KH. Bruchmechanik <strong>metal</strong>lischer Werkst<strong>of</strong>fe. München, Wien: Hanser, 1980.<br />

[74] Kolednik O, Sabirov I. Local <strong>fracture</strong> <strong>properties</strong> in MMCs. In: Ravi-Ch<strong>and</strong>ar K,<br />

Karihaloo BL, Kishi T, Ritchie RO, Yokobori AT Jr, Yokobori T, editors. Advances in<br />

Fracture Research, Proceedings <strong>of</strong> ICF10, 2001, Elsevier, Oxford, UK, paper 410.<br />

[75] Danzer R, Lube T, Supancic P. Monte Carlo simulations <strong>of</strong> strength distributions <strong>of</strong> brittle<br />

materials – type <strong>of</strong> distribution, specimen, <strong>and</strong> sample size. Zeitschrift für Metallkunde.<br />

2001, 92, pp. 773-783.<br />

[76] Danzer R. A general strength distribution function for brittle materials. Journal <strong>of</strong> the<br />

European Ceramic Society. 1992, 10, pp. 461-472.<br />

[77] Lawn B. Fracture <strong>of</strong> brittle solids. Cambridge University Press, UK, 1993.<br />

[78] Kolednik O. Stereophotogrammetrische Untersuchungen zur elastisch-plastischen<br />

Bruchmechanik. PhD thesis. University <strong>of</strong> Leoben, 1984.<br />

[79] Valiev RZ, Islamgaliev RK, Alex<strong>and</strong>rov IV. Bulk nanostructured materials from severe<br />

plastic deformation. Progress in Materials Science. 2000, 45, pp. 103-189.<br />

[80] Iwahashi Y, Wang J, Horita Z, Nemoto M, Langdon TG. Principle <strong>of</strong> equal-channel<br />

angular pressing for the processing <strong>of</strong> ultra-fine grained materials. Scripta Materialia.<br />

1996, 35, pp. 143-146.<br />

[81] Valiev RZ, Islamgaliev RK, Kuzmina NF, Yong L, Langdon T. Strengthening <strong>and</strong> grain<br />

refinement in an Al-6061 <strong>metal</strong> <strong>matrix</strong> composite through intense plastic straining.<br />

Scripta Materialia. 1999, 40, pp. 117-122.<br />

[82] Karnezis PA, Durrant G, Cantor B. Characterization <strong>of</strong> reinforcement distribution in cast<br />

Al-Alloy/SiCp composites. Materials Characterization. 1998, 40, pp. 97.<br />

[83] Rogers A. Statistical analysis <strong>of</strong> spatial dispersions: The quadrat method. Pion, London,<br />

1974.<br />

[84] Kolednik O. On the physical meaning <strong>of</strong> the J-∆a-curves. Engineering Fracture<br />

Mechanics. 1991, 38, pp. 403-412.<br />

160


[85] Hashim J, Looney L, Hashmi MSJ. Particle distribution in cast <strong>metal</strong> <strong>matrix</strong> composites-<br />

Part I. Journal <strong>of</strong> Materials Processing Technology. 2002, 123, pp. 251-257.<br />

[86] Hashim J, Looney L, Hashmi MSJ. Particle distribution in cast <strong>metal</strong> <strong>matrix</strong> composites-<br />

Part II. Journal <strong>of</strong> Materials Processing Technology. 2002, 123, pp. 258-263.<br />

[87] Lloyd DJ, Lagace H, McLeod A, Morris PL. Microstructural aspects <strong>of</strong> aluminum-silicon<br />

carbide particulate composites produced by a casting method. Materials Science <strong>and</strong><br />

Engineering. 1989, A107, pp. 73-80.<br />

[88] Pettermann HE. Numerical simulations <strong>of</strong> a compositionally graded structure using a<br />

hierarchical approach. Materials Science Engineering. 2000, A276, pp. 277-282.<br />

161

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!