Download as pdf or txt
Download as pdf or txt
You are on page 1of 211

Introduction to Metal Matrix Composites

Yoshinori Nishida

Introduction to Metal Matrix


Composites
Fabrication and Recycling
Yoshinori Nishida
National Institute of Advanced Industrial Science
and Technology (AIST)
Nagoya, Japan (retired in 2002)

Original Japanese edition published by CORONA PUBLISHING CO., LTD.


KINZOKUKI FUKUGOZAIRYO NYUMON
Copyright # 2001 Yoshinori Nishida

ISBN 978-4-431-54236-0 ISBN 978-4-431-54237-7 (eBook)


DOI 10.1007/978-4-431-54237-7
Springer Tokyo Heidelberg New York Dordrecht London
Library of Congress Control Number: 2012954005

# Springer Japan 2013


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or
information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed. Exempted from this legal reservation are brief excerpts
in connection with reviews or scholarly analysis or material supplied specifically for the purpose of being
entered and executed on a computer system, for exclusive use by the purchaser of the work. Duplication
of this publication or parts thereof is permitted only under the provisions of the Copyright Law of the
Publisher’s location, in its current version, and permission for use must always be obtained from
Springer. Permissions for use may be obtained through RightsLink at the Copyright Clearance Center.
Violations are liable to prosecution under the respective Copyright Law.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
While the advice and information in this book are believed to be true and accurate at the date of
publication, neither the authors nor the editors nor the publisher can accept any legal responsibility for
any errors or omissions that may be made. The publisher makes no warranty, express or implied, with
respect to the material contained herein.

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)


Preface

All materials for aerospace, energy conversion systems, and preservation of the
environment are being assessed, and the demand for light materials is increasing.
Light metals have some drawbacks, however, and to improve these weak areas,
light and strong composites have been developed by strengthening light metals by
ceramic fibers or particles.
Most studies of metal matrix composites (MMCs) are related to the evaluation
of composite properties, and few are related to fabrication processes. Once a
new composite was developed, research on the evaluation of the composite became
very active. Some research on fabrication processes has been carried out, but it was
very difficult to discuss many fabrication processes fundamentally and systemati-
cally. At an early stage, MMCs such as sintered aluminum products (SAP) and
second-phase reinforced eutectic alloys obtained by unidirectional solidification
were developed without mixing reinforcements into matrix metals by mechanical
means. After the development of mixing techniques, by which ceramic fibers or
particles were dispersed in matrix metals, the research on evaluation of those
composites became very active and many papers about it have been published.
Many of those have been concerned with continuous fibers in particular.
The bonding strength at the fiber/matrix interface and properties of composites
are closely related to fabrication conditions. Therefore, the history of constituent
materials during fabrication should be taken into account to enhance the reliability
of property data of composites. Nevertheless, when the property of composites was
discussed, the fabrication process was made light of. In one case, when a composite
is commercialized, first a composite billet is fabricated and from the billet a
composite product is formed by a mechanical process such as machining, a plastic
deformation process, or casting. In another case, composite fabrication and shaping
of the product are performed at the same time by one operation. The latter will be
better than the former from the point of view of production efficiency and energy
efficiency. The fabrication process is very important for industrial application.
In addition, we can achieve high productivity and reliability of composite products
by determining the optimum production condition and by controlling that

v
vi Preface

condition. However, the fabrication process has been treated as only a technique
and has not been investigated scientifically.
To date, most monographs on MMCs have been edited for graduate students and
professional researchers. These monographs deal mainly with mechanics, which is
not easy for other students to understand. Now it is time to promote the commer-
cialization of MMCs. A book that will serve as a useful introduction to MMCs is
needed for researchers and industrial engineers who are engaged in developing new
materials.
Therefore, this book was written to systematically discuss in an easy-to-understand
manner the research work that has been carried out until now and to help
researchers in industry to develop composite products in expectation of further
advances in MMCs. The discussion is plainly written to engage the interest of many
people in MMCs and is not difficult even for undergraduate students to understand.
This book includes not only processing but also the properties of composites as an
introduction to MMCs for readers who want to explore the subject further.
The fundamental knowledge necessary to understand MMCs is explained in
Chap. 1. Major fabrication processes, except the pressure infiltration process by
squeeze casting and centrifugal casting, are described in Chap. 2. The fabrication
process by squeeze casting is explained in Chap. 3, because the process is very
important to understand the wetting phenomenon between reinforcement and
molten metal. The theory of pressure infiltration by squeeze casting is discussed
in Chap. 4. The centrifugal casting of metal matrix composites is discussed
in Chap. 5, and the properties of composites are discussed in Chap. 6.
The superplasticity of composites as a recent topic is introduced in Chap. 7. The
production processes of reinforcement materials for composites are briefly
explained in Chap. 8, and fundamental ideas of recycling of composites are
discussed in Chap. 9 to meet the demands of the times.
Thanks are due to Dr. Toshio Yamauchi, who was my co-worker on the pressure
infiltration process which occupies the most important part of Chap. 4. I am grateful
to Professors Shojiro Ochiai of Kyoto University (Fig. 6.15) and Sumio Nagata of
Fukuoka University (Figs. 3.3, 3.4 and 3.5) for permission to reproduce figures.
Finally, I would like to express my thanks to Professor Yutaka Kagawa of The
University of Tokyo, Professor Karl U. Kainer of the GKSS Research Centre
Geesthacht, and Professor Chitoshi Masuda and Professor Makoto Yoshida of
Waseda University for their very constructive discussions.
This book is based on the Japanese edition of my book Kinzokuki Fukugozairyo
Nyumon, published by Corona Publishing Co., Ltd., Tokyo. Most of the digital files
of the original figures were provided by the publisher.

Aichi, Japan Yoshinori Nishida


Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Classification and Characteristics of Composites . . . . . . . . . . . . . 2
1.1.1 Classification of Composites . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Characteristics of Metal Matrix Composites . . . . . . . . . . . 3
1.1.3 Examples of Metal Matrix Composite
Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Rule of Mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Surface Energy and Interface Energy . . . . . . . . . . . . . . . . . . . . . . 11
1.4 Thermodynamics on Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.5 Wettability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.6 Darcy’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2 Fabrication Processes for Composites . . . . . . . . . . . . . . . . . . . . . . . . 27
2.1 Aspects of Fabrication of Composites . . . . . . . . . . . . . . . . . . . . . 27
2.1.1 Energy of Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.1.2 Fabrication of Composites by Mechanical
Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.1.3 Mechanical Work and Wetting . . . . . . . . . . . . . . . . . . . . . 32
2.2 Classification of Fabrication Methods . . . . . . . . . . . . . . . . . . . . . 34
2.3 Comparison of Fabrication Techniques . . . . . . . . . . . . . . . . . . . . 34
2.4 Solid State Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.4.1 Powder Metallurgical Methods . . . . . . . . . . . . . . . . . . . . . 35
2.4.2 Mechanical Alloying . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.4.3 Diffusion Bonding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.4.4 Spark Plasma Sintering (SPS) . . . . . . . . . . . . . . . . . . . . . 40
2.5 Liquid State Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.5.1 Vortex Addition Technique . . . . . . . . . . . . . . . . . . . . . . . 41
2.5.2 Compo-Casting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.5.3 Pressureless Infiltration Process . . . . . . . . . . . . . . . . . . . . 44
2.5.4 Ultrasonic Infiltration . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

vii
viii Contents

2.6 Gaseous State Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46


2.6.1 CVD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.6.2 PVD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.7 In Situ Fabrication Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
2.7.1 Internal Oxidation Process . . . . . . . . . . . . . . . . . . . . . . . . 48
2.7.2 Unidirectional Solidification Process . . . . . . . . . . . . . . . . 49
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3 Fabrication by Squeeze Casting . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.1 Application of Pressure and Fabrication Energy . . . . . . . . . . . . . . 54
3.2 Threshold Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.2.1 Case of Random Reinforcement Distribution . . . . . . . . . . 56
3.2.2 Case of Unidirectional Fiber Distribution . . . . . . . . . . . . . 57
3.2.3 Measurement of Threshold Pressure . . . . . . . . . . . . . . . . . 58
3.3 Influence of Preheating of Preform . . . . . . . . . . . . . . . . . . . . . . . 58
3.4 Microscopic Analysis of Fluid Flow in Preforms . . . . . . . . . . . . . 61
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4 Theory of Pressure Infiltration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.1 Macroscopic Analysis of Fluid Flow in Preforms . . . . . . . . . . . . . 67
4.1.1 Analysis on the Assumption of Constant
Preform Surface Pressure . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.1.2 Infiltration Velocity Model . . . . . . . . . . . . . . . . . . . . . . . . 72
4.2 Infiltration Stop Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.3 Characteristics of Pressure Infiltration Method
and Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5 Centrifugal Casting of Metal Matrix Composites . . . . . . . . . . . . . . . 91
5.1 Infiltration of Molten Metal Using Centrifugal Force . . . . . . . . . . 91
5.1.1 Pressure Generated at the Surface of Preform . . . . . . . . . . 92
5.1.2 Infiltration Start Pressure . . . . . . . . . . . . . . . . . . . . . . . . . 95
5.1.3 Infiltration of Molten Metal into the Preform . . . . . . . . . . 95
5.1.4 Example Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.1.5 Examples of Composites Fabricated
Using Centrifugal Force . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.2 Centrifugal Casting of Particle Dispersed
Molten Metal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
6 Properties of Composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
6.1 Mechanical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
6.1.1 Elastic Modulus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
6.1.2 Strength of Composites . . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.1.3 Fracture of Composites . . . . . . . . . . . . . . . . . . . . . . . . . . 130
Contents ix

6.2 Physical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138


6.2.1 Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
6.2.2 Specific Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
6.2.3 Thermal Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
6.2.4 Thermal Expansion Coefficient . . . . . . . . . . . . . . . . . . . . 142
6.3 Interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
6.4 New Trends in MMCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
6.4.1 Carbon/Metal Composites with High
Thermal Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
6.4.2 Smart Composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
7 Superplasticity of Composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
7.1 Background for Superplasticity . . . . . . . . . . . . . . . . . . . . . . . . . . 155
7.2 Mechanism of Superplastic Forming . . . . . . . . . . . . . . . . . . . . . . 156
7.2.1 Superplastic Deformation Mechanism
of Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
7.2.2 Superplastic Deformation Mechanism
of Composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
7.3 Production Methods of Superplastic Materials . . . . . . . . . . . . . . . 159
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
8 Materials for the Fabrication of Composites . . . . . . . . . . . . . . . . . . . 165
8.1 Characteristics of Reinforcements
and Matrix Metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
8.2 Production Processes for Reinforcements . . . . . . . . . . . . . . . . . . . 168
8.2.1 Ceramic Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
8.2.2 Ceramic Fibers and Carbon Fibers . . . . . . . . . . . . . . . . . . 169
8.2.3 Ceramic Whiskers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
9 Recycling of Composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
9.1 Composite Ratio in Products and Re-melting . . . . . . . . . . . . . . . . 182
9.2 Separation of Fibers or Particles . . . . . . . . . . . . . . . . . . . . . . . . . 183
9.2.1 State of Composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
9.2.2 Mechanical Separation of Reinforcements . . . . . . . . . . . . 186
9.2.3 Chemical Separation of Reinforcements . . . . . . . . . . . . . . 187
9.3 Separation of Fiber and Metal from Composites . . . . . . . . . . . . . . 189
9.3.1 Chemical Method and Ratio of Separation . . . . . . . . . . . . 189
9.3.2 Phenomena Associated with Separation
Using Fluxes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
9.4 Entropy of Mixing by the Addition of Reinforcement
Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
x Contents

9.4.1 Entropy of Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195


9.4.2 Entropy Increase upon the Addition
of Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
9.5 Assessment of Metal Matrix Composites . . . . . . . . . . . . . . . . . . . 199
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
Chapter 1
Introduction

Abstract Metal matrix composites are discussed fundamentally from the micro- and
macroscopic viewpoints. A composite consists of matrix material and dispersoids such
as fibers or particles. These constituent materials have their own microstructure,
properties and shapes, even in the composite. These materials contact one another at
their interfaces. The microscopic structure of the composite is quite different from the
structures of alloys and intermetallic compounds, which are atomic-level mixtures.
The fundamentals needed to understand interfaces within composites, including
the formation energies of interfaces and surfaces, wettability, and contact angles
are discussed in this chapter. The classification of composites is described, and the
characteristic features of metal matrix composites are compared to those of other
composites. Images of typical commercial metal matrix composite products are also
shown in this chapter. Darcy’s law is introduced, because it is needed to understand the
phenomena which occur during the fabrication of metal matrix composites.

To understand composites, a wide range of knowledge is needed. The science of


composites is in fact a “composite” of many kinds of sciences, including funda-
mental sciences such as strength of materials, fluid dynamics and thermodynamics.
For example, to understand a ceramic fiber reinforced metal matrix composite, we
need to know the properties of both metals and ceramics, and also how they interact
with one another. In addition, because the motivation and objectives for fabricating
composites are many and varied, we need a wide range of knowledge to character-
ize the properties of the composites obtained. Therefore, composite science may be
rather difficult for a newcomer. However, as the science of composites is very much
an applied science, readers can enjoy the pleasure of fabricating new materials
applicable to industry by applying the science of composites.
Research into the development of composites intensified several decades ago.
After progress in the development of new metallic materials through alloying
techniques slowed down, research into intermetallic compounds and composites
became more active. It is now usually possible to estimate the properties of a
composite made from a combination of materials. Composites have the advantage

Y. Nishida, Introduction to Metal Matrix Composites: Fabrication and Recycling, 1


DOI 10.1007/978-4-431-54237-7_1, # Springer Japan 2013
2 1 Introduction

that their purpose and the direction required for their development are usually clear.
However, it is unlikely in composites research that we will come across a new
material with unexpected properties. In addition, production of composites to give
the expected properties is not simple. Although many kinds of fabrication processes
have been developed, whether a composite has the desired properties must still be
determined after fabrication, and the fabrication process modified, if necessary.
Therefore, fabrication methods are very important, and economical and easy
fabrication methods are needed.
This chapter describes the fundamental knowledge needed to understand the
science of composites. The more detailed discussion in the subsequent chapters
builds on this knowledge.

1.1 Classification and Characteristics of Composites

Composites in the literature are described as “materials fabricated by combining


constituent materials having different properties and shapes, to realize new
properties which each constituent does not have” [1]. The “combination” in this
definition is not an atomic or molecular level mixture. Composites are designed and
fabricated to realize characteristic or desired properties through mixing the constit-
uent materials; each constituent material retains its own microstructure and
properties at the microscopic level. Therefore, a large amount of interface between
the constituent materials is created in a composite after fabrication. If a composite is
held at a high temperature or under conditions where a chemical reaction between
the constituent materials occurs, the interfaces will become ambiguous and the
material tends towards the thermodynamically stable state at the high temperature.
Finally, the composite will become an atomic-level mixture, which cannot be called
a “composite”. Instead, this material is now an alloy or an intermetallic compound.
Therefore, generally, composites are thermodynamically unstable materials. Thus,
we must either choose a relatively stable combination of constituent materials or
use a coating technique to suppress the chemical reaction at the interface between
constituent materials.

1.1.1 Classification of Composites

As shown in Table 1.1, composites are classified into six categories based on the
type of matrix material: metal matrix composites [MMC or fiber reinforced metals
(FRM)], fiber reinforced plastics (FRP), fiber reinforced ceramics (FRC), fiber
reinforced glasses (FRG), intermetallic compound matrix composites (IMC), and
carbon fiber reinforced carbon (CC) composite.
Alternatively, as shown in Fig. 1.1, composites are also classified based on the
shape of the reinforcement used. These classes are continuous fiber reinforced
1.1 Classification and Characteristics of Composites 3

Table 1.1 Classification of composites by matrix


1. Metal matrix composites (MMC or FRM)
2. Plastic matrix composites (FRP)
3. Ceramic matrix composites (CMC or FRC) (continuous fiber reinforced ceramics, or fiber
reinforced ceramics)
4. Glass matrix composites (or FRG)
5. Intermetallic compound matrix composites (IMC)
6. Carbon fiber reinforced carbon (CC composites)

a b c

(a) Particulate reinforced composite


(b) Discontinuous fiber reinforced composite
(c) Continuous fiber reinforced composite

Fig. 1.1 Classification of composites by reinforcement shape

composites, discontinuous fiber reinforced composites and particulate reinforced


composites. A subset of discontinuous fiber reinforced composites is whisker
reinforced composites (i.e., composites strengthened by whiskers). A whisker is a
fine single crystal with a very small diameter which has the theoretical tensile
strength (for its crystal structure), because it has very few defects within its crystal
lattice. In addition, composites strengthened by very fine and hard particles are
referred to as dispersion strengthened metals.

1.1.2 Characteristics of Metal Matrix Composites

There are many characteristic features of metal matrix composites. The most
important characteristics are as follows:
1. Metal matrix composites have higher strength and elasticity than other metallic
materials.
4 1 Introduction

2. Metal matrix composites have high fracture toughness and are able to absorb
large impact fracture energy compared with other composites, because the
matrix metals have high ductility.
3. Generally, metal matrix composites have higher thermal conductivity than that
of other composites.
4. Soft matrix metal composites reinforced by hard ceramic particles have good
wear resistance.
5. Metal matrix composites have small thermal expansion coefficients.
6. Metal matrix composites have high electrical conductivity.
7. Plastic deformation occurs within metal matrix composites reinforced by
particles or short fibers, and these composites are further strengthened by plastic
deformation.

1.1.3 Examples of Metal Matrix Composite Products

1.1.3.1 Aluminum Alloy Matrix Composites

Aluminum alloys reinforced by continuous silicon carbide (SiC) fibers or carbon


fibers will be light composites having high strength (that is, they have high specific
strength—the value of tensile strength divided by specific gravity), because these
fibers are high strength, light materials. A vast amount of research has been
performed into these composites. However, very few composites have been
commercialized. An exception is continuous boron fiber reinforced aluminum
alloy composite, which was developed early on in research into continuous fiber
reinforced metal matrix composites and used for beams of the space shuttle orbiter.
The reason for this lack of commercialization may be that FRP already fills the need
for light composites with high strength at room temperature, and that continuous
fiber reinforced metal matrix composites are more difficult to produce than FRP,
and more expensive. In addition, composites reinforced one-dimensionally with
continuous fibers have a significant drawback in that strengthening is only effective
parallel to the fiber axis direction.
In the case of short fiber reinforced composites, pistons reinforced with alumina/
silica fibers have been developed and commercialized for diesel engines [2]. These
alumina/silica fibers were originally developed for thermal insulation of electric
furnaces. Only the part which surrounds the top piston ring was reinforced with
the short fibers. First, a preform (a porous material made of fibers used in the
pressure infiltration process) was made in the shape required for the final composite
in the piston. The preform was then set in a piston-shaped cavity in a mold and
infiltrated with molten metal by pressure casting. This process is therefore very
efficient, because the production of the piston and the fabrication of the composite
can be performed at the same time. The short fiber content in the composite part is
7–8 vol.%, which is sufficient for wear resistance against the piston ring.
1.1 Classification and Characteristics of Composites 5

Fig. 1.2 MMC cylinder block for an automobile (courtesy of Honda). The inside 3 mm of the
cylinder wall are MMC

A similar composite product reinforced by short fibers is shown in Fig. 1.2 [3].
This is an MMC cylinder block for an automobile. The inside surfaces of the
cylinders are reinforced by hybrid fibers (a mixture of alumina short fibers and
carbon fibers) to improve their wear properties, because the wear resistance of the
aluminum alloy was insufficient, and hence, previously, cast iron sleeves (pipes) had
been used for the inside surface parts. In the conventional production method for a
cylinder block with cast iron cylinders, the cast iron sleeves were placed in a mold
and molten aluminum alloy was poured and solidified to surround the outsides of the
sleeves. To produce an MMC cylinder block, a 3 mm thick fiber preform is used
instead of the cast iron sleeves. After molten die casting alloy (JIS-ADC12,
Al–12 wt.% Si–2.5 wt.% Cu) is poured, high pressure is applied by a die casting
machine to infiltrate the molten metal into the preform. By this process, fabrication
of the composite and production of the cylinder block are performed at the same
time. This composite part showed good wear properties (comparable to cast iron),
and allowed a significant reduction in the weight of the cylinder part. In this case, the
preform consists of a mixture of fibers (about 12 vol.% short alumina fibers, 3 mm
diameter, and about 9 vol.% chopped carbon fibers, 7 mm diameter). The role of
alumina fibers is to improve wear resistance, and carbon fibers are used to improve
lubrication and to prevent seizure between the piston and cylinder. The thickness of
the composite in the cylinder is only 3 mm from the inner surface, making the
composite much thinner and lighter than the cast iron sleeve.
6 1 Introduction

Fig. 1.3 MMC roller for belt


conveyer produced by die
casting (courtesy of Ryobi)

Another example of a short fiber reinforced composite product, a roller from a


belt conveyer, is shown in Fig. 1.3 [4]. The dark gray part of the cross-section of
the cylindrical product is the composite. The reinforcement is a hybrid preform
(a mixture of alumina short fibers and SiC particulates) with a total volume fraction
of 30 %. This composite roller is produced by infiltrating molten JIS ADC12 die
casting alloy (Al–12 wt.% Si–2.5 wt.% Cu) into the hybrid preform by a die casting
process, and resulted in both significant weight reduction and improved corrosion
resistance.
An example of a composite reinforced with ceramic whiskers is shown in
Fig. 1.4 [5]. This composite product is a piston for an outboard motor, and is
made of JIS-AC8A alloy (Al–12 wt.% Si–1 wt.% Cu–1 wt.% Mg–1 wt.% Ni)
reinforced by SiC whiskers; only the top part of piston, which faces into the
combustion chamber, is made of composite (to give high strength at high tempera-
ture). In this case, the fabrication of the composite part and formation of the piston
shape is performed at the same time by squeeze casting, which infiltrates molten
metal into the SiC whisker preform. Generally, composites reinforced by SiC
whiskers are so hard that diamond tools are needed to machine them. Thus, this
piston is produced at near-net shape, to reduce the amount of machining required as
much as possible. Recently, aluminum borate whiskers have been used instead of
SiC whiskers.
An example of a commercialized alumina short fiber reinforced aluminum
composite product, a large scale shower plate for liquid crystal production, is
shown in Fig. 1.5. The dimensions of the plate are 1,100 mm  1,000 mm. This
composite (Al/20 vol.% alumina fiber) was produced by squeeze casting. The
flexion of the product under its own weight at high temperature (400–500  C)
1.1 Classification and Characteristics of Composites 7

Fig. 1.4 MMC piston for outboard motor produced by pressure infiltration process (courtesy of
Suzuki Motor)

Fig. 1.5 Large (1,100 mm  1,000 mm) shower plate for liquid crystal production (Al/20 vol.%
alumina fiber composite) produced by squeeze casting (courtesy of Advanced Material
Technologies Co., Ltd., Japan)

was reduced compared with conventional materials, because the elastic modulus of
aluminum is improved by addition of alumina fibers.
Another example of a commercialized composite product is shown in Fig. 1.6,
which depicts a radiator for semiconductor production. This composite is Al/85
vol.% carbon produced by squeeze casting. The characteristic features of this
composite are light weight and high thermal conductivity.
A particulate reinforced composite product, a disk brake rotor for an automobile
driven by an electric motor, is shown in Fig. 1.7 [6]. Cast iron rotors are usually
used for automobiles. JIS-AC4C alloy (Al–7 wt.% Si) reinforced by about 20 vol.%
8 1 Introduction

Fig. 1.6 Radiator for the production of semiconductor (Al/85 vol.% carbon composite) produced
by squeeze casting (courtesy of Advanced Material Technologies Co., Ltd., Japan)

Fig. 1.7 (a) MMC disk brake rotor for electric vehicle, made from particulate reinforced
aluminum alloy; and (b) optical microstructure (courtesy of Aisin Takaoka Co.)

SiC particles with an average diameter of 13 mm (Duralcan, made in Canada) is


used to reduce the weight and the inertial moment of wheels. Aluminum composites
have higher thermal conductivity than cast iron, although their melting points are
lower than that of cast iron. Therefore, properties of aluminum composites are
sufficient for automobiles driven by electric motors, which usually do not run at
high speeds.

1.1.3.2 Copper Matrix Composites

The demand for high thermal conductivity materials for heat sinks for integrated
circuits (IC) is increasing, and active research is going on to develop such materials.
In addition, high thermal conductivity carbon fibers or particles have been devel-
oped, making Cu/C composites very attractive for this purpose. An example
1.2 Rule of Mixtures 9

Fig. 1.8 Heat sink (75 vol.%


C/25 vol.% Cu composite) for
base station of mobile phone
(courtesy of Advanced
Material Technologies Co.,
Ltd., Japan)

of a Cu/C composite heat sink (size: 24 mm  17 mm  1.4 mm) for a base station
of a mobile phone is shown in Fig. 1.8. This composite was produced by squeeze
casting, and contains 25 vol.% copper and 75 vol.% carbon. This product has been
commercialized as a heat sink material.
Several commercialized metal matrix composite products have been introduced
here. The advantage of these composite products is their light weight, and many of
them were developed for applications requiring wear resistance. High strength and
high elastic modulus are only expected for whisker reinforced aluminum alloys.
Interest in dispersion strengthened metals is based on the findings of the Swiss
researcher, Irmann [7], who developed sintered aluminum products (SAP). Since
then, the mechanism of strengthening via the dislocation pinning effect of fine and
hard particles has become well understood. Research into these kinds of metals
became very active, because this mechanism showed the potential for attaining very
high strength in metals. Then, DuPont Co. developed TD nickel, which is a very
fine ThO2 particle dispersed nickel. Afterwards, the mechanical alloying process
was developed by Benjamin [8] and dispersion strengthened metals became attrac-
tive materials. Few metals, however, have been commercialized. This may be
because dispersion strengthened metals are produced by powder metallurgical
methods, which are high cost and low productivity processes. Other automotive
and aerospace applications are introduced by Kainer [9].

1.2 Rule of Mixtures

Predicting the behavior of a composite from the known behavior of its component
parts is a crucial part of composite design. The simplest approach to doing this is the
“Rule of Mixtures”, so this important concept is presented below. Also briefly
discussed are situations where the rule is and is not applicable.
10 1 Introduction

Consider a mixture which consists of 50 vol.% metal, having density rm and


50 vol.% ceramic particles, having rf. It is apparent that the density rc of the
composite can be obtained by averaging densities rm and rf:

r f þ rm
rc ¼ : (1.1)
2

Then, if the volume fraction of ceramic particles is Vf, the volume fraction of
metal will be (1–Vf). The above relation is rewritten as:

rc ¼ Vf rf þ ð1  Vf Þrm : (1.2)

The more common form of this equation is given by:

Uc ¼ Vf Uf þ ð1  Vf ÞUm ; (1.3)

where Uc is the property of the composite, Uf is the property of the reinforcement, Vf


is the volume fraction of reinforcement and Um is the property of the matrix metal.
Equation (1.3) is called the “Rule of Mixtures”.
When there are two different reinforcements, Eq. (1.3) will be modified as
follows:

Uc ¼ V1 U1 þ V2 U2 þ ð1  V1  V2 ÞUm : (1.4)

Therefore, the general form of the rule of mixtures is given by:

X
n X
n
Uc ¼ Vn Un þ ð1  Vn ÞUm ; (1.5)
1 1

where U1, U2, U3, . . . Un are the properties of the reinforcements and V1, V2, V3, . . . Vn
are their volume fractions. Property U can be a variety of properties, such as thermal
expansion coefficient, elastic modulus, among others. If no chemical reaction and no
volume change occur at the interface between the matrix metal and reinforcements
during fabrication of the composite, the rule of mixtures applies well. However,
thermal expansion coefficients and thermal conductivities obtained by the rule of
mixtures are not so accurate, because these properties depend on the shape and
distribution of reinforcements in the matrix metal. Since tensile strength and elastic
modulus are also influenced by the distribution of reinforcements as well as the
adhesive strength at the interface between reinforcement and matrix metal, only
approximate values can be obtained from the rule of mixtures.
Aluminum is a high thermal conductivity material, which also has a high thermal
expansion coefficient. However, a high thermal conductivity material with a low
thermal expansion coefficient can be obtained by fabricating an aluminum com-
posite reinforced by SiC particles, which have a low thermal expansion coefficient.
1.3 Surface Energy and Interface Energy 11

In this case, the volume fraction of SiC to attain the required thermal expansion
coefficient is easily decided by the rule of mixtures. The rule of mixtures is a simple
equation, but is a very useful relationship when considering potential development
of composites.

1.3 Surface Energy and Interface Energy

Composites are materials containing at least two different constituent materials


which have quite different properties. The constituent materials are in contact at the
interfaces between them. Strong adhesion at the interfaces is required. In addition, it
is necessary to prevent changes in crystal structure and microstructure by chemical
reaction between the constituents. Therefore, in fabricating composites, it is impor-
tant to achieve ideal interfaces (without chemical reaction and with strong adhe-
sion). To understand how to do this, first, the surface energy of a metallic material
will be discussed, followed by discussion of the interface energy.
The surfaces of both solids and liquids have surface energy (or surface tension).
Why does this surface energy exist? As rigorous discussion is not the purpose of this
book, the essentials of the surface and interface energies will now be discussed
without rigor.
Figure 1.9 shows a schematic of atoms in a crystal, atoms at the surface and
atoms that exist solely in the vapor. They are the same type of atoms. This figure is
simply drawn; the actual crystal structures of metals are typically face-centered
cubic, body-centered cubic or hexagonal closed packed structures. When we
consider an individual atom (atom A) in the crystal, the number of nearest
neighbors around that atom (i.e., the coordination number) depends upon the crystal

A
Fig. 1.9 Schematic
description of atoms, (A) an
atom inside the crystal, (B) in
the surface, and (C) in the
vapor
12 1 Introduction

structure. For example, the coordination number is 12 for face-centered cubic and
8 for body-centered cubic structures. However, Fig. 1.9 has been simplified to make
it easier to understand.
First, consider atom A. This is surrounded by neighbors. Therefore, as atom A is
constrained strongly by bonds with its neighbors, the atom is unlikely to move to
another lattice point and is only vibrating at its specific lattice point. The energy that
atom A has is low. Next, consider atom B, which is in the surface of the crystal.
In this case, atom B does not have a top neighbor. As the vapor pressure of the same
atoms is very low, the attraction from those atoms in the vapor is negligible. The
constraint upon the atom in the surface resulting from the surrounding atoms (gas
atoms) is weaker than that upon atoms fully inside the crystal. Atom B is vibrating
more strongly than atom A, because atom B has a higher degree of motion than
atom A. Therefore, the energy level of the atoms in the surface is higher than that of
the atoms inside the crystal. The higher energy state causes the surface energy and
the surface tension of the crystal. In other words, a force towards the inside of the
crystal acts on the atoms in the surface. That force is the surface tension, which
tends to minimize the surface area of a material.
The gas atom C, in the vapor shown in Fig. 1.9, is barely constrained by its
surroundings (other gas atoms). All bonds of atom C are free. Thus atom C is at the
highest energy level. The energy difference between atom A and atom C
corresponds to the energy of all bonds of the atom.
We can calculate an approximate value of surface energy by the quasichemical
approach [10]. When the energy of a bond between two atoms is Hb and the
sublimation heat of 1 mole of the metal is DHv (J mol1) at room temperature,
the total energy to cut all bonds of 1 mole of the metal is given by the sublimation
heat, and so we obtain the following equation:

1
DHv ¼ ZN0 Hb ; (1.6)
2

where Z is the coordination number (the number of nearest neighbors) and N0 is the
Avogadro constant (6.02  1023 mol1). In the above equation, the number of
bonds is divided by 2, because one bond is made between two atoms.
One bond with no partner (as occurs for atoms in the surface of a crystal) has an
energy of half Hb. The number of atoms N per unit area in the surface of a crystal
depends on the crystal structure and crystal plane. For example, N of face-centered
cubic structures like aluminum or copper is given by:

4
f111gplane N ¼ pffiffiffiffiffiffiffi ; and (1.7)
3d2

2
f100gplane N ¼ ; (1.7a)
d2
1.3 Surface Energy and Interface Energy 13

where d is the lattice constant (for aluminum, d ¼ 4.0496  1010 m). In addition,
the number of bonds with no partner in the surface depends upon the crystal plane of
a face-centered cubic structure. For example, the number of bonds is 3 for a {1 1 1}
plane and 4 for a {1 0 0} plane. If the number of bonds with no partner is D, the
surface energy g of a crystal is given by:

Hb DHv DN
g¼ DN ¼ : (1.8)
2 ZN0

Next we will use the above equation to calculate the surface energy of aluminum.
As the experimental data for DHv were not found, DHv was obtained by the following
calculation. The sum of the enthalpy of vaporization of aluminum and the heat
required to heat the aluminum from room temperature to the boiling point (evaluated
from the specific heats of liquid and solid aluminum) is equal to the sum of the
sublimation heat at room temperature and the heat required to heat 1 mole gas from
room temperature to the boiling point (evaluated from the specific heat of the gas).
This relationship is given by:

DHv þ Cg ðTb  Ta Þ ¼ Cs ðTm  Ta Þ þ DHm þ Cl ðTb  Tm Þ þ DHb ; (1.9)

where Cg, Cs and Cl are the specific heats of gaseous, solid and liquid aluminum,
respectively, at atmospheric pressure. DHm, DHb are latent heats of melting and
boiling (i.e., enthalpies of melting and vaporization), respectively. Ta, Tm and Tb are
the room, melting and boiling temperatures, respectively. The calculation was
carried out assuming that the specific heats were independent of temperature. To
calculate DHv using Eq. (1.9), we need the value of Cg. We used the following
equation to estimate the specific heat of 1 mole of aluminum gas at atmospheric
pressure (assuming that aluminum gas is an ideal gas):

5R
Cg ¼ ; (1.10)
2

where R is the gas constant. The value of DHv obtained was 3.23  105 J mol1.
The surface energy values were: for {1 1 1} plane, gf111g ¼ 1:89 J m2 , and for
{1 0 0} plane, gf100g ¼ 2:18 J m2 .
The plane with the larger number of atoms per unit area, namely the {1 1 1} plane,
has lower surface energy. The experimentally obtained surface energy of aluminum
has been reported as 1.6 J m2 [11]. The surface energy calculated from Eq. (1.8) is
larger than the experimentally obtained value. This may be because the DHv obtained
from Eq. (1.9) is slightly larger than the real value. However, the agreement between
the experimental and theoretical values is still reasonable. The surface energies of
other solid materials are listed in Table 1.2 [12, 13].
To obtain rigorous values of surface energy and interaction force between atoms,
we need to discuss and calculate the atomic potential as expressed using quantum
14 1 Introduction

Table 1.2 Surface energy Material Surface energy (J m2) Temperature (K)
of solids [12, 13]
Ag 1.14  0.09 1,177
Al 1.6 298
Au 1.25  0.05 1,298
Cu 1.65 1,273
Fe–3% Si 1.36 –
Zn 0.11 –
Zn–0.1% Cd 5.77 298
Si 1.24 –
LiF 0.34 –
MgO 1.20 –
CaF2 0.45 –
BaF2 0.28 –
CaCO3 0.23 –
NaCl 0.31 –
Al2O3 1.00 –
Diamond 5.40 –

Fig. 1.10 Insufficient Grain boundary


bonding of atoms at a grain
boundary

mechanics. However, by the quasichemical approach discussed above, we can


understand the essential phenomenon of the formation of surface energy (or surface
tension). As in the above discussion, a bond with no partner, or insufficient bonding
at a grain boundary, as shown in Fig. 1.10, causes the formation of surface energy or
interface energy.
Two grains contacting at a grain boundary are shown in Fig. 1.10. As the crystal
direction of both grains is different, atoms from the grains on both sides of the
interface cannot bond sufficiently. The degree of freedom of the atoms at the interface
is high, compared with atoms inside the grains. Therefore, atoms at the interface are at
1.3 Surface Energy and Interface Energy 15

Fig. 1.11 As the curvature of


an extremely small crystal is
large, the insufficient bonding
of atoms in the curved surface
increases compared with the
bonding of atoms in the flat
surface of a large crystal

Fig. 1.12 Insufficient


bonding between materials A
and B, when A and B have Material A
different crystal structures Interface
and crystal parameters

Material B

higher energy levels than atoms inside the grains. The higher energy level causes
formation of the grain boundary energy.
In addition, a very interesting phenomenon is shown in Fig. 1.11. If a particle
consists of a small number of atoms, the curvature of the particle surface will be
very large. Atoms in the surface of the small particle have a greater number of
bonds with no partner than atoms in flat surfaces. Therefore, small particles have a
higher surface energy than large particles. We can imagine that if there is an
extreme decrease in particle diameter, the particle will behave like a liquid, and,
finally, behave like a gas.
Alternatively, when two different materials which have quite different crystal
structures (material A and material B), as shown in Fig. 1.12, come into contact, there
seems to be no theoretical way to estimate the interface energy quantitatively. It is
possible, however, to consider that the interface energy consists of two terms: the
geometrical energy term (DHg) and the chemical energy term (DHc) [10]. Then, the
interface energy gAB between material A and material B is given by:

gAB ¼ DHg þ DHc : (1.11)


16 1 Introduction

The geometrical energy term arises because of the difference in crystal structure,
lattice parameter and so on between the two materials. If the chemical affinity is
good, the chemical energy term will be negative and the interface energy gAB will
decrease. The geometrical energy term will be approximated by summing the
surface energies of both materials:

DHg ffi gA þ gB ; (1.12)

where gA is the surface energy of material A, and gB is the surface energy of material
B. Since the value of Eq. (1.12) will usually be larger than the real interface energy
between the two materials, the chemical energy term DHc should be negative, and
the interface energy gAB can be expressed by:

gAB  gA þ gB : (1.13)

If the affinity between materials A and B is very poor, in other words if the
contact angle between A and B approaches p rad, the energy of the interface formed
by the two materials will be close to the sum of the surface energies of both
materials, as given by Eq. (1.12).
When the affinity between the two materials is good, and the contact angle is
rather low, it is possible to estimate the interface energy using Girifalco and Good’s
equation [14, 15]:

gAB ¼ gA þ gB  2fðgA gB Þ1=2 ; (1.14)

where f is a constant, which is determined by the characteristics of the system. The


theoretical method to obtain f is discussed by Girifalco and Good [14]. However,
reliable values have been obtained experimentally. Therefore, to obtain the inter-
face energy, there is not much difference between the method using Eq. (1.14) and
the method using Young’s equation (which is described in Sect. 1.5). To use
Young’s equation to determine the interface energy, we require an experimentally
determined contact angle between the two materials. The real value of f is about
unity for a water/organic liquid system, about 0.7 for a mercury/nonmetallic liquid
system [14], and 0.4 for a molten salt/molten aluminum system [16]. When the
affinity of both materials is very poor, and the contact angle is very high, the value
of f should be close to zero. In this case, Eqs. (1.13) and (1.14) are almost equal.

1.4 Thermodynamics on Surfaces

When we break a crystal into two pieces, new surfaces will be formed and atoms
originally inside the crystal will appear on the surface. If we consider one particular
atom, some bonds from that atom to its nearest neighbors will be broken as the atom
1.4 Thermodynamics on Surfaces 17

becomes part of the surface. Assuming that the entire work dW given to the system
was used only for the formation of surface by the breaking operation, we get the
following equation:

dW ¼ gdA; (1.15)

where dA is the increased area of the crystal surface and g is the surface energy of
the crystal. The right-hand side of Eq. (1.15) has a negative symbol because the
work given to a system is customarily regarded as negative. By the first law of
thermodynamics, the energy change dE is given by:

dE ¼ dq  dW: (1.16)

In fact, breaking a crystal is an irreversible process. However, if we assume that


this process is reversible, we will be able to obtain the following equation:

dq ¼ TdS: (1.17)

This equation shows that even if the work given to the system is used only for the
formation of surface, some heat is absorbed by the system during new surface
formation, and the entropy of the surface will increase. Therefore, dE is given by:

dE ¼ TdS þ gdA: (1.18)

As enthalpy H is H ¼ E + PV,

dH ¼ dE þ PdV þ VdP: (1.19)

The above equation can be simplified because the system is a solid crystal of
constant volume at constant atmospheric pressure:

dH ¼ dE: (1.20)

The differential form of Gibb’s free energy (G ¼ H  TS) is given by:

dG ¼ dH  TdS  SdT (1.21)

By substituting Eqs. (1.18) and (1.20) into Eq. (1.21) and assuming constant
temperature, we obtain:

dG ¼ gdA: (1.22)

As discussed in the above section, the value of g depends upon the crystal plane.
However, if we take the average value as g, we can obtain the following relationship
from Eq. (1.22):

G ¼ gA: (1.23)
18 1 Introduction

This equation shows that the surface free energy is equal to the surface energy
per unit area.
Now we discuss the entropy change when a new surface of a solid is formed. The
relationship between entropy and free energy is known:
 
@G
S¼ : (1.24)
@T p

As the pressure range in this discussion is narrow and the surface energy change
of the solid with pressure is negligible, the relationship between the free energy
change DG per unit area and the entropy change DS becomes:
   
@DG dg
DS ¼  ¼ (1.25)
@T p dT

To obtain the increase in entropy upon the formation of a new surface, we need
to know Eq. (1.8) as a function of temperature. Once we determine how the
sublimation heat DHv and the lattice parameter d change with temperature, we
can express equation (1.8) as a function of temperature. Assuming that the subli-
mation heat at 25  C (298 K) is DHv0 and the specific heat of solid Cs is independent
of temperature, the sublimation heat as a function of temperature is given by

DHv ¼ DHv0  Cs ðT  Ta Þ: (1.26)

The lattice parameter as a function of temperature becomes:

d ¼ d0 f1 þ aðT  Ta Þg; (1.27)

where d0 is the lattice parameter at 298 K and a is the coefficient of linear thermal
expansion. By using Eqs. (1.26) and (1.27), the surface energy, Eq. (1.8), as a
function of temperature is given by:

DN 0 fDHv0  Cs ðT  Ta Þg
g¼ ; (1.28)
N0 Zf1 þ aðT  Ta Þg

where:

4
N 0 ¼ pffiffiffi 2 ðfor the f111g planeÞ; (1.29)
3d0

and

2
N0 ¼ ðfor the f100g planeÞ: (1.29a)
d02
1.5 Wettability 19

The derivative of Eq. (1.28) at room temperature is given by:


 
dg
¼ Cs  2aDHv0 : (1.30)
dT T¼Ta

DS ¼ 9.24  10-5 J m2 K1 at room temperature, as determined by Eq. (1.30),


which gives a value of TDS at room temperature of 0.0275 J m2. This value is very
small when compared with the free energy change of 1.6 J m2 per unit area of new
surface. This approximate calculation shows that the entropy term is much smaller
than the surface energy term.

1.5 Wettability

When two different materials come into contact, the term “wettability” is used to
express the degree of compatibility between the two materials. By this term, we
imagine that paper wets with water. This term is applicable to phenomena between
two kinds of liquids, as well as between a liquid and a solid. To describe wettability
quantitatively, the term “contact angle” is used.
If we put a small mass of liquid (a droplet) on a solid with a smooth surface, the
droplet becomes stable, with a particular shape. The shape depends on the degree of
compatibility between the solid and the liquid. An example of a droplet silhouette is
shown in Fig. 1.13. The angle y, made by the tangent to the liquid surface with the
solid surface at their point of intersection is the contact angle, as shown in Fig. 1.13.
The reason why this droplet keeps its shape is that the surface tensions (gLV, gSV)
(or surface energies) and interface tension (gSL) (or interface energy) are in equilib-
rium at the intersection. Their relationship is given by:

gSL ¼ gSV  gLV cos y; (1.31)

where gSL is the solid/liquid interface energy, gSV is the solid/vapor interface energy
and gLV is liquid/vapor interface energy.
This equation is called “Young’s equation”. The atmosphere should be saturated
with the vapor of the liquid of the droplet. If the atmosphere is not saturated, the

Fig. 1.13 Silhouette and


contact angle of a droplet
placed on a solid with γ
a smooth surface. The shape θ
of the droplet is controlled
by the balance of forces γ γ
arising from interface
energies
20 1 Introduction

Fig. 1.14 Shape of droplet a b


and contact angle

Fig. 1.15 Diagram showing a b


an advance of the droplet by
area dA at the liquid/solid
intersection
θ θ

liquid of the droplet evaporates continuously and the shape of the droplet changes
with time. If we take another combination of liquid and solid, the surface and
interface energies will be different and the contact angle in Eq. (1.31) will also be
different. As shown in Fig. 1.14, there are many different droplet shapes. In the case
of y ¼ 0, as we can imagine from Fig. 1.13, the shape of the droplet will be stable
when the droplet spreads infinitely. Then, in this case, the droplet spreads sponta-
neously. The range 0 < y < 90 means that the solid is easily wet by the liquid, as
shown in Fig. 1.14a. The range 90 < y <180 has the shape shown in Fig. 1.14b,
and the liquid does not wet the solid well.
The free energy has its minimum value at this equilibrium shape. Therefore, we
can set y0 to be the contact angle where the forces are in equilibrium. As in the
example shown in Fig. 1.15, if the liquid surface at the intersection point advances a
little, the contact angle will become y0–dy. As shown in (a) in the figure, an area on
the solid surface of dA will be covered with the liquid, and the increase of the liquid/
vapor surface of the droplet will be cos(y0–dy)dA. Then, the change of free energy
is given by:

dG ¼ ðgSL  gSV ÞdA þ gLV cosðy0  dyÞdA; and (1.32)

dG
¼ gSL  gSV þ gLV cosðy0  dyÞ: (1.33)
dA

Except when dy ¼ 0, the value of dG/dA is always positive in Eq. (1.33). When
dy ¼ 0, dG/dA ¼ 0 and the free energy has the minimum value. Young’s equation,
Eq. (1.31), is obtained by putting dG/dA ¼ 0 into Eq. (1.33). In contrast, if the
1.5 Wettability 21

Table 1.3 Contact angle between aluminum and various reinforcements


Matrix/reinforcement Contact angle, degree (K) References
Al/MgO 130 (1,173) [17]
120 (1,273)
110 (1,333)
Al–9.98 wt.% Mg/graphite 116 (923) [18]
Al–5 wt.% Cu/graphite 140 (923) [18]
Al–2.58 wt.% Si/graphite 155 (923) [18]
Al/a-SiC 155 (1,373) [19]
Al/Al2O3 155 (1,073) [20]
Al–1.1 wt.% Ca/Al2O3 110 (973) [21, 22]
Al–1.1 wt.% Ca/SiO2 110 (973) [21, 22]
Al/Si3N4 120 (1,173) [23]

liquid surface recedes in Fig. 1.15a and the decrease of the area of the liquid/solid
interface is dA, we can obtain an equation similar to Eq. (1.23). By setting dy ¼ 0,
Young’s equation is also obtained.
When the contact angle is higher than 90 , as shown in Fig. 1.15b, and the
increase in the liquid/sold interface is dA, the contact angle becomes y0–dy. The
solid surface becomes part of the solid/liquid interface, and the droplet surface
decreases as cos{p(y0–dy)}dA. Then, the free energy change of the system is
given by:

dG ¼ ðgSL  gSV ÞdA  gLV cosfp  ðy0  dyÞgdA: (1.34)

As cos{p(y0–dy)} ¼ cos(y0–dy), Eq. (1.34) coincides with Eq. (1.32).


The case shown in Fig. 1.15a is the one where the surface is easily wet. However,
the liquid does not advance any further, because the surface energies and interface
energy are in equilibrium, and thus the overall force is zero. The case shown in
Fig. 1.15b is that where the surface does not wet easily. However, the same
equilibrium means that the liquid does not recede any further.
For the fabrication of composites, we need experimentally obtained values of
contact angle. Some reported contact angle data are listed in Table 1.3 [17–23].
There are many papers and books relating to wettability [24–26].
We have just described how wettability can be characterized. We will now
describe the energy changes that occur when two different materials, having surface
energies of gA and gB come into contact and an interface having interface energy of
gAB is made. The work WAB done to the system is given by:

WAB ¼ gA þ gB  gAB : (1.35)

WAB is known as the “work of adhesion”. If materials A and B are the same liquid
material, after adhesion, the interface will disappear. In this case, the work of
adhesion WAA becomes:
22 1 Introduction

Fig. 1.16 Surface shapes and a b c


position inside a glass pipe
when the glass pipe is dipped
into three different liquids
with different contact angles

WAA ¼ 2gA : (1.36)

This WAA is called the “work of cohesion”.


Alternatively, if we dip one end of a glass pipe having a clean inside surface into
water, as shown in Fig. 1.16a, water will rise inside the pipe before stopping. The
water stops when the upward force exerted by water surface tension and the
downward force by gravity that acts on the water inside the pipe are in equilibrium.
The tangent at the intersection between the water surface and the glass inside
surface is the contact angle. The contact angle between clean glass and water is
close to zero. The top surface (meniscus) of water has a convex downward profile.
When liquid has a curved surface, the pressure inside the curved surface (in the air)
is higher than the pressure outside the curved surface (in the water). The pressure
difference is given by:

2gL cos y
DP ¼  ; (1.37)
r

where r is the inside radius of the glass pipe, gL is the surface tension of the liquid
and y is the contact angle. If we use a different liquid, having a contact angle of 90
with the glass pipe, the liquid surface inside the pipe will be at the same level as the
liquid outside the pipe as shown in Fig. 1.16b. This means that there is no pressure
difference across the liquid surface in the glass. If we use mercury as a liquid, the
contact angle between mercury and glass will be close to 180 and, as shown in
Fig. 1.16c, the mercury surface in the pipe will be pushed down and become curved.
In this case, the pressure inside the curved surface (in the mercury) is higher than
that outside the curved surface (in the air). The pressure difference is also given by
Eq. (1.37). This equation is useful to determine the minimum pressure needed to
infiltrate molten metal into channels between fibers of a fiber preform for a pressure
infiltration process.
1.6 Darcy’s Law 23

1.6 Darcy’s Law

Darcy’s law is introduced here, because the law is used to analyze the pressure
infiltration process used to fabricate composites. The flow of fluid through a porous
medium made of particles or fibers is complicated. However, if the flow rate is
sufficiently low, and the flow is laminar flow, Darcy’s law can be used to accurately
predict the fluid flow. As shown in Fig. 1.17, if we maintain a pressure difference
between both sides of a porous medium, fluid flows from the higher pressure side to
the lower pressure side, and stationary flow is attained. In this case, assuming that
the mass of fluid flow per unit time is proportional to the pressure difference, the
mass is given by:

kD ADP
Q¼ ; (1.38)
L

where Q is the mass of fluid flow per unit time (m3 s1), kD is the permeability
coefficient (m2 Pa1 s1), A is the cross-sectional area (m2), L is the length of the
porous medium, DP (¼ P1 P2) is the pressure difference between both sides of
the porous medium (Pa). This equation is known as Darcy’s law [27]. Because kD
changes with viscosity m, the specific permeability K is defined by K ¼ mkD.
Usually, specific permeability is called “permeability”. These equations were
introduced empirically.
Alternatively, it is possible to derive Darcy’s law theoretically if the fluid is
Newtonian (i.e., the shear stress is proportional to the strain rate of the fluid), the
flow rate is low enough, and the inertial force of the fluid is negligible [28].

Porous material

Fig. 1.17 Fluid flow through


a porous medium
24 1 Introduction

We consider an element of volume DADx in the x-direction in a field of fluid flow


through a porous medium. DA is the cross-sectional area perpendicular to the flow
direction and Dx is the length. The viscous drag Fm is proportional to the total inside
surface area of pores, fDADx, where f is the porosity. In addition, Fm is proportional
also to the viscosity m and to the volume of fluid ux that flows in the x-direction per
unit area and unit time. ux changes with time. Then, Fm is described by:

Fm ¼ Cfmux DADx: (1.39)

The right-hand side of Eq. (1.39) has a negative symbol, because the viscous
drag acts in a direction opposite to the flow direction. C in Eq. (1.39) is constant and
is equal to 1/K. There is a pressure gradient ∂p/∂x between both sides of element
DADx. The pressure gradient is the driving force for flow of fluid. The force Fp
arising from the pressure difference is given by:

@p
Fp ¼ fDADx : (1.40)
@x

The pressure difference acts only on the porous part within the area DA, so the
effective area is fDA. Gravity also acts on the fluid in the element DADx, with mass
rfDADx. If we take the x-axis as being upward as shown in Fig. 1.16, the force Fg
exerted by gravity is given by:

Fg ¼ rgx fDADx; (1.41)

where gx is gravity. Although gravity can only act in the x direction, the suffix x is
used to emphasize this point. r is the density of the fluid.
Generally, we should take into account the forces in the y-axis and z-axis
directions when determining Fm and Fp. However, as the flow in the x-direction is
driven by the sum of the above forces, Eqs. (1.39), (1.40) and (1.41), we obtain:
 
1 @p
 mux   rgx fDADx ¼ 0: (1.42)
K @x

The term inside the parentheses in Eq. (1.42) should be zero, so we obtain:
 
K @p
ux ¼  þ rgx : (1.43)
m @x

We can then extend equation (1.43) to a three-dimensional form, which includes


the y-axis and z-axis directions:

K
u¼ rp þ rg; (1.44)
m
References 25

where g denotes gravity in a vector form. If we neglect gravity, and take the one-
dimensional form, we can obtain the following simple equation:

K @p
u¼  : (1.45)
m @x

Alternatively, DeWiest showed that it is possible to derive equation (1.44) from


the Navier–Stokes equations [29]. His explanation is instructive, because he took
into account the fluid flow only in the pores, and used viscous drag and gravity as
the external force in the Navier–Stokes equations, while neglecting the inertial
force.
Darcy’s law is valid when the flow rate of fluid in a porous medium is low. This
flow rate is evaluated quantitatively by the Reynolds number. The Reynolds
number Re of a fluid in a fibrous medium is given by:

v 0 de
Re ¼ ; and (1.46)
n

2Rð1  Vf Þ
de ¼ ; (1.46a)
Vf

where de is the equivalent radius for a non-circular channel, n is the kinematic


viscosity, v0 is the real velocity of the fluid in the channel (not the velocity ux in
Darcy’s law), R is the radius of the fiber, and Vf is the volume fraction of fiber in the
composite [30, 31].
When the Reynolds number Re is smaller than unity, Darcy’s law is applicable.
This has been proven by many experiments [28]. However, if Re is larger than unity
and between about 1 and 10, turbulent flow will be involved. This zone is the
transition zone from laminar flow to turbulent flow and the accuracy of Darcy’s law
falls.

References

1. Hayashi, T. (ed.): Fukugozairyo Kogaku, p. 3. Nikkagiren Publisher, Tokyo (1971) (in Japanese)
2. Donomoto, T., Miura, N., Funatani, K., Miyake, N.: Ceramic fiber reinforced piston for high
performance diesel engines. SAE Paper No. 830252 (1983)
3. Hayashi, T., Ushio, H., Ebisawa, M.: The properties of hybrid fiber reinforced metal and its
application for engine block. SAE Paper No. 890557 (1989)
4. Ohmura, H., Murashima, I., Takahashi, Y.: Studies on composite processing using solidifica-
tion and casting. Report of Composite Processing Committee. Japan Foundry Engineering
Society, p. 42 (1997)
5. Yamauchi, T.: Development of SiC whiskers reinforced piston. SAE Paper No. 911284 (1991)
6. Hiraiwa, J.: Report of Nonferrous Metals Research Meeting. Tokai Branch of Japan Foundry
Engineering Society (September 1998)
26 1 Introduction

7. Irmann, R.: On a new sintered aluminum product with high strength at elevated temperatures.
Leichtmetall 3, 21–25 (1950)
8. Benjamin, J.S.: Dispersion strengthened superalloys by mechanical alloying. Metall. Trans. 1,
2943–2951 (1970)
9. Kainer, K.U. (ed.): Metal Matrix Composites: Custom-Made Materials for Automotive and
Aerospace Engineering. Wiley, New York (2006)
10. Swalin, R.A.: Thermodynamics of Solids. Wiley, New York (1976)
11. Mondolfo, L.F.: Aluminum Alloys: Structure and Properties, p. 61. Butterworths, London
(1976)
12. Gilman, J.J.: Direct measurements of the surface energies of crystals. J. Appl. Phys. 31,
2208–2218 (1960)
13. Kelly, A., Groves, G.W., Kidd, P.: Crystallography and Crystal Defects, revised ed. Wiley,
NewYork (2000)
14. Girifalco, L.A., Good, R.J.: A theory for the estimation of surface and interfacial energy.
1. Derivation and application to interfacial tension. J. Phys. Chem. 61, 904–909 (1957)
15. Adamson, A.W.: Physical Chemistry of Surfaces, 4th edn, p. 107. Wiley, New York (1982)
16. Utigard, T.A., Toguri, J.M.: Interfacial tension of aluminum in cryolite melts. Metall. Trans.
16B, 333–338 (1985)
17. Yoshimi, N., Nakae, H., Fujii, H.: Wettability of molten aluminum on MgO surface. J. Jpn.
Inst. Met. 52, 1179–1186 (1988)
18. Manning Jr., C.R., Gurganus, T.B.: Wetting of binary aluminum alloys in contact with Be, B4C
and graphite. J. Am. Ceram. Soc. 52, 115–118 (1969)
19. Koehler, W.: Wettability of SiC by pure aluminium. Aluminium 51, 443–447 (1975)
20. Brennan, J.J., Pask, J.A.: Effect of nature of surfaces on wetting of sapphire by liquid
aluminum. J. Am. Ceram. Soc. 51, 569–573 (1968)
21. Mori, N., Sorano, H., Kitahara, A., Ohgi, K., Matsuda, K.: Effect of Ca on the wettability of
liquid Al to graphite, Al2O3 and SiO2. J. Jpn. Inst. Met. 47, 1132–1139 (1983)
22. Mori, N., Miyahara, H., Koseki, M., Ohgi, K., Matsuda, K.: Wettability and interaction
between molten Al and Ni-coated Al2O3 or pure Ni substrate. J. Jpn. Inst. Met. 55, 444–451
(1991)
23. Naka, M., Kubo, M., Okamoto, I.: Wettability of silicon nitride by aluminium, copper and
silver. J. Mater. Sci. Lett. 6, 965–966 (1987)
24. Nakae, H.: Wettability of liquid aluminum to non-metallic materials. J. Jpn. Inst. Light Met.
39, 136–146 (1989)
25. Fujii, H., Nakae, H.: Precise measurement of the wettability of ceramics by molten aluminum.
Mater. Jpn. 34, 1269–1275 (1995)
26. Adamson, A.W.: Physical Chemistry of Surfaces, 4th edn, p. 433. Wiley, New York (1982)
27. Darcy, H.: Les fontaines publiques des la ville de Dijon. Dalmont, Paris (1856)
28. Collins, R.E.: Flow of Fluids Through Porous Materials, p. 47. Reinhold Publishing Corp.,
New York (1961)
29. DeWiest, R.J.M.: Flow Through Porous Media. Academic, New York (1969)
30. Ergun, S.: Fluid flow through packed columns. Chem. Eng. Prog. 48, 89–94 (1952)
31. Fukunaga, H., Goda, K.: Fabrication process of fiber reinforced metal by squeeze casting.
J. Jpn. Mech. Eng. (C) 49, 1808–1815 (1983)
Chapter 2
Fabrication Processes for Composites

Abstract Many processes have been developed for the fabrication of metal matrix
composites from constituent materials. These fabrication processes are classified
into four categories: solid state fabrication technique, liquid state fabrication
technique, gas state fabrication technique and in situ processing. Recent
developments in the major processes are introduced and their characteristic features
are described. Common phenomena of these processes are discussed in fundamen-
tal terms to obtain a systematic understanding of fabrication processes. Each
fabrication process is then discussed from the viewpoint of energy consumption.
The most important aspect of composite fabrication is making interfaces with
good bonding between the matrix metal and the reinforcements, without degrada-
tion by chemical reaction. Usually, the reinforcement/matrix interface is formed by
conversion from mechanical energy to interface energy. These important points are
discussed in more detail in this chapter.

2.1 Aspects of Fabrication of Composites

2.1.1 Energy of Fabrication

Metal matrix composites are materials in which fibers or particles as reinforcements


are dispersed in the matrix. Fabrication processes for composites are classified into
solid state, liquid state, and gas state fabrication techniques and in situ processing as
listed in Table 2.1. They are schematically shown in Fig. 2.1. Before fabrication, the
matrix and reinforcements are separate. After fabrication, the reinforcements are
dispersed in the matrix metal. The state difference, namely, the free energy differ-
ence, gives the minimum work required for the fabrication of the composite. In this
case, as the surface energy is the free energy per unit area as discussed in Eq. (1.23)
in Chap. 1, the free energy difference DG is given by:

Y. Nishida, Introduction to Metal Matrix Composites: Fabrication and Recycling, 27


DOI 10.1007/978-4-431-54237-7_2, # Springer Japan 2013
28 2 Fabrication Processes for Composites

Table 2.1 Classification of fabrication techniques for metal matrix composites


1. Solid state fabrication Powder metallurgy process (PM process), mechanical
technique alloying (MA), diffusion bonding, surface oxidation
process (e.g. SAP)
2. Liquid state fabrication Pressure infiltration process, vortex addition technique, pressureless
technique infiltration process (e.g. Lanxide), compo-casting, spray deposit
technique, centrifugal casting
3. Gas state fabrication CVD, PVD
technique
4. In situ process Unidirectional solidification, internal oxidation (nitridization)
technique

solid state
fabrication technique
fiber

liquid state
fabrication
technique metal
metal

fiber

gas state
fabrication
technique

in-situ process

Fig. 2.1 Schematic diagram showing the four categories of metal matrix composites processing
routes

DG ¼ g1  g0 ; (2.1)

where g1 is the reinforcement/matrix interface energy at room temperature and g0 is


the surface energy of the reinforcement surrounded by air at room temperature
before the composite is fabricated [1]. In addition, the fabrication energy W per unit
volume of composite becomes:

W ¼ ðg1  g0 ÞAf ; (2.2)

where Af is the total reinforcement/matrix interface area per unit volume of


composite.
There are two separate calculations for the value Af in Eq. (2.2). One is for the
case of fiber reinforcement and the other is for the case of particle reinforcement.
2.1 Aspects of Fabrication of Composites 29

If the fiber length is long enough that the cross-sectional area of both ends of the
fiber is negligible, Af is given by:

2Vf
Af ¼ ; (2.3)
rf

where Vf is the volume fraction of the fiber in the composite, and rf is the average
radius of the fiber.
When the reinforcement is particles, Af becomes:

3Vf
Af ¼ ; (2.3a)
rf

where Vf is the volume fraction of particles in the composite, and rf is the average
radius of the particle.
If we take a more rigorous approach to the process of fabrication of composites,
the total outer surface of the metal increases upon mixing fibers or particles into the
metal, because the volume of reinforcement is added into the metal and thus the total
volume increases. Supposing that no reinforcement is exposed on the surface of the
composite, only the surface area of the metal will increase. Then, a more rigorous
equation for the work for the fabrication of a composite is:

W ¼ ðg1  g0 ÞAf þ gam ðAc  Am Þ; (2.4)

where gam is the surface energy of the matrix metal surrounded by air. Ac and Am are
the total surface area of the matrix metal after and before the fabrication of
composite, respectively. The second term on the right-hand side of the equation
is very small and negligible compared with the first term. For example, when the
reinforcing fibers are long enough and the matrix metal is spherical before and after
the fabrication of the composite, the ratio of the total fiber/matrix interface area to
the increased total surface area of the matrix metal is given by:

Af rm
ffi (2.5)
Ac  Am r f

where rf is the radius of the fiber and rm is the radius of the spherical matrix metal.
When rf ¼ 1.5 mm and rm ¼ 0.05 m, the result is:

Af
ffi 33000: (2.5a)
Ac  Am

This result shows that the increased part of the total surface area of matrix metal
is very small and negligible compared with the total fiber/matrix interface area.
Therefore, we can use Eq. (2.2) as the fabrication energy of the composite.
If we consider solid-state fabrication, it is very difficult to introduce and disperse
reinforcements directly into a solid-state matrix metal. This has led to the develop-
ment of many fabrication processes. Usually, matrix metal powder or foil is mixed
30 2 Fabrication Processes for Composites

Fig. 2.2 Model of molten a Molten metal b


metal flow in a ceramic pipe,
(a) wetting by the advance of
molten metal, (b) no wetting
by the advance of molten θ
metal
Air
Pipe

Wetting No wetting

with the reinforcement. After mixing, the mixture is sintered under pressing.
Sometimes, the matrix metal is instead thickly coated on to the fibers and the
coated fibers are sintered under pressing. The work required to form the interface
energy is provided mainly by the sintering process as the composite is fabricated.
For these processes, additional energy is needed to produce the powder or foil and
then sinter the mixture.
The liquid-state fabrication technique is a route where reinforcements are mixed
with liquid matrix metal, and the composite is then solidified. Casting is an energy
saving and efficient process. However, usually wetting of ceramic fibers or particles
by molten matrix metals is poor. Many methods have been developed to wet the
reinforcements. The minimum work to wet reinforcements with molten matrix
metal and to disperse them in the metal is obtained from Eq. (2.2):

W ¼ ðgfl  gfa ÞAf ; (2.6)

where gfl is the reinforcement/molten matrix metal interface energy and gfa is the
surface energy of the reinforcement in air. Substituting Young’s equation into
Eq. (2.6), we obtain:

W ¼ Af gla cos y (2.7)

where gla is the surface energy of the molten metal. Equation (2.6) is illustrated
schematically in Fig. 2.2. A molten metal is in balance inside a ceramic pipe of
radius r. Both (a) and (b) in Fig. 2.2 shows poor wetting cases. If the molten metal
advances up to the indicated line and wets the pipe, as is the case in (a), a composite
will be formed. However, when wetting is extremely poor, and y  180 , as shown
in (b), the molten metal does not wet the ceramic pipe and an air film is formed
between the molten metal and the ceramic pipe.
When a molten metal wets the ceramic reinforcement easily (contact angle
y < p/2, which leads to the molten metal surface being concave downward in an
experiment similar to those shown in Fig. 2.2), the molten metal advances
2.1 Aspects of Fabrication of Composites 31

spontaneously. However, this is not usually the case, and there are two general ways
to improve the wetting of ceramic reinforcements so that composites can be
fabricated: (1) chemical method, and (2) mechanical method.
In the case of (1) chemical method, when a molten metal does not wet the ceramic
reinforcement (y > p/2), we need to add some elements into the molten metal or to
coat the surface of the reinforcement to improve wetting and to make the contact
angle lower (y < p/2). Then, the molten metal will advance spontaneously [2].
In the case of (2) mechanical method, we can effectively use mechanical work to
wet reinforcements, even if the contact angle is y > p/2 (Fig. 2.2a). Consider the
intersection of the three phases: molten metal, ceramic pipe and air. At this point,
the molten metal/ceramic pipe interface tension, the surface tension of the molten
metal and the surface tension of the ceramic pipe are well-balanced and the
relationship is given by Young’s equation as discussed before. If we break the
balance, the molten metal will move. If we hold the metal side at a pressure slightly
higher than the balanced pressure, the molten metal will advance in the direction
indicated by the arrow, and the ceramic surface will become molten metal/ceramic
interface. In contrast, if we hold the metal side at a pressure slightly lower than the
balanced pressure, the molten metal will recede and the molten metal/ceramic
interface will become ceramic surface.

2.1.2 Fabrication of Composites by Mechanical Methods

When a molten metal surface is well-balanced, as shown in Fig. 2.2a, the pressure
inside the curved molten metal is different from the air pressure outside the metal.
The pressure difference DP is given by Eq. (1.37) and becomes:

2gla cos y
DP ¼  ; (2.8)
r

where r is the inner radius of the ceramic pipe. Then, if we always hold the pressure
inside the curved surface at a value slightly higher than DP, the molten metal
surface will continuously advance in the direction indicated by the arrow and the
pipe surface will become molten metal/pipe interface as shown in Fig. 2.2a. This
means that the pipe surface has been wetted with molten metal as a result of
mechanically applied pressure. The work W of Eq. (2.6) has been done by the
applied pressure. This process is the mechanical method for fabrication of
composites.
If the distance of the advance of the molten metal is L, the newly formed molten
metal/pipe interface area will be 2prL and the work W of Eq. (2.6) will become:

W ¼ 2prLðgfl  gfa Þ: (2.9)


32 2 Fabrication Processes for Composites

Alternatively, as the force DP acts on the molten metal having cross-sectional


area pr2 and the metal advances for distance L, the work W done for the advance is
given by:

W ¼ pr 2 LDP: (2.10)

Substitution of Eq. (2.8) into Eq. (2.10) yields:

W ¼ 2prLgla cos y: (2.11)

Substituting Young’s equation (Eq. (1.31)), replacing gLV ! gla, gSV ! gfa and
gSL ! gfl into Eq. (2.11) yields Eq. (2.9). Equation (2.11) shows that the work W
depends upon contact angle y. If y < p/2, W will be negative and energy will be
released by the fabrication of the composite. For example, water wets paper easily.
In this example, the water is analogous to the molten metal and the paper consists of
fibers; wetting occurs spontaneously. Therefore, the paper/water interface energy is
much smaller than the sum of the water and paper surface energies before wetting.
Then, we can imagine that the contact angle y between water and paper is much
lower than p/2 and the work W is negative. Because W is negative, heat will be
released in the water and the temperature of the water will rise slightly. This
temperature rise of the water may be detected with a precise calorimeter.
When y > p/2, which corresponds to the combination of ceramic fiber and
molten metal, the work given by Eq. (2.11) is required for the molten metal to
advance. This work is positive and depends upon the contact angle between the
fiber and the metal. The minimum work to fabricate a unit volume of the composite
is given by Eq. (2.6).

2.1.3 Mechanical Work and Wetting

For high contact angle y, when we apply higher pressure than DP to the molten
metal in Fig. 2.2a, the metal advances. However, wetting of the ceramic pipe is
doubtful. If the pipe does not wet, a thin air film will be formed between the pipe
and the molten metal as shown in Fig. 2.2b. We will now discuss the free energy
change before and after the metal advances. Assuming that the fiber is a ceramic
having uniform surface energy gfa which does not depend on the location of the fiber
surface, the Gibbs free energy change DG per unit surface of fiber upon the
fabrication of a composite will be given by:

DG ¼ gfl  gfa : (2.12)

As this fabrication is carried out at atmospheric pressure, we can discuss it using


Gibbs free energy. There are some options for the standard point of the free energy
2.1 Aspects of Fabrication of Composites 33

to compare the free energy change of this phenomenon. However, we find that
Eq. (2.12) is reasonable for the free energy change for the fabrication of a compos-
ite, because only the material in contact with the surface of the fiber changes at
constant temperature and pressure. A comparison of Eqs. (2.6) and (2.12) shows
that the free energy change DG upon fabrication of a composite is the energy
needed for wetting the unit surface of the fiber. Then, we can calculate and compare
the work for the following two cases.
1. Molten metal wets ceramic pipe.
The work W(A) of this case is given by Eq. (2.11):

WðAÞ ¼ 2prLgla cos y: (2.13)

Therefore, the free energy change DG(A) of the whole system concerned is given by:

DGðAÞ ¼ 2prLgla cos y: (2.13a)

2. Molten metal does not wet ceramic pipe.


As shown in Fig. 2.2b, an air film will be formed between the pipe and the
molten metal. Then, the difference between before and after advancing of the
metal is only the increase of the molten metal surface area which has formed
along the air film. The increase in the metal surface area is 2prL and the work of
this case W(B) is given by:

WðBÞ ¼ 2prLgla (2.14)

In this case, the free energy change DG(B) of the whole system concerned is
given by:

DGðBÞ ¼ 2prLgla (2.14a)

When we compare Eq. (2.13a) for the wetting case and Eq. (2.14a) for the
non-wetting one, we find that when y < p, DG(A) is always smaller than DG(B):

DGðAÞ <DGðBÞ ðy<pÞ; (2.15)

DGðAÞ ¼ DGðBÞ ðy ¼ pÞ (2.15a)

Generally, the lower free energy case occurs. In the range of y < p, when
slightly higher pressure than DP is applied to the molten metal in the pipe and the
work does not exceed W(B), the metal advances slowly and wetting will definitely
occur between the metal and pipe. If the molten metal advances rapidly, turbulent
flow may appear and, locally, work larger than W(B) will be given and non-wetted
regions will be formed. In the case of y ¼ p, the free energy of the wetting case is
34 2 Fabrication Processes for Composites

equal to that of the non-wetting case. Usually, in these circumstances, non-wetting


occurs. In addition, when y is close to p, we can expect that a significant part of the
fiber will not wet.
The most important aspect of the discussion in this section is that we can convert
mechanical energy into interface energy. The energy conversion is realized by
pressure infiltration methods.

2.2 Classification of Fabrication Methods

The methods to disperse fibers or particles in metals (i.e., fabrication methods for
composites) are classified into:
1. Solid state fabrication:
Mixing is performed with the matrix metal in the solid state (this includes both
powder metallurgical methods and methods using metal foils);
2. Liquid state fabrication:
Mixing is performed with liquid matrix metal;
3. Gas state fabrication:
Fabrication by chemical vapor deposition (CVD), physical vapor deposition
(PVD) or similar processes;
4. In situ fabrication:
Methods which use external treatments such as heating or cooling to make fibers
or particles from within a matrix.
These techniques are listed in Table 2.1. Some processes are associated with
more than one of the categories in the list. For example, in the diffusion bonding
method (see Sect. 2.4.3), to make continuous fiber reinforced composites, fibers
coated by thermal spraying are bundled and sintered under pressing at high temper-
ature. This method uses techniques from categories (1) and (2).

2.3 Comparison of Fabrication Techniques

Ideally, a composite would be fabricated using only the energy given by Eq. (2.5).
If we can do work only to the molten metal in the ceramic pipe (reinforcement) as
shown in Fig. 2.2a, we can wet the ceramic and produce a composite effectively.
This is the pressure infiltration process. However, molten metal agitation tech-
niques such as the vortex process or the compo-casting process are processes which
mix reinforcements (particles) by adding them to agitating molten metal. In these
processes, most of the energy is consumed in making and maintaining the vortex or
metal flow. Only a small fraction of the total energy is used for wetting. In this
respect, these processes are not efficient. In the molten metal agitation technique,
2.4 Solid State Fabrication 35

sometimes shear stress acts on the reinforcement/molten metal interface as shown


in Fig. 2.2a and wetting progresses intermittently, although the reinforcements and
the metal move together.
Alternatively, for the processes associated with metal powder or foil, a mixture
of metal powder or foil and reinforcements must be sintered and deformed under
high pressure at high temperature to make good reinforcement/matrix interfaces.
However, these processes are suitable for producing a uniform mixture.
The effect of ultrasonic waves on the infiltration of molten metal has been
reported by Nakanishi et al. [3]. The ultrasonic wave improves wetting between
reinforcements and molten metal. When an ultrasonic wave propagates through a
material, the energy travels through the atomic structure by a series of compression
(dense zone) and expansion movements (rough zone). As the dense zone is under
high pressure, when the dense zone passes through the reinforcement/molten metal
interface, the equilibrium of forces may be broken and the molten metal may
advance as shown in Fig. 2.2a.
Andrews et al. reported that the Lorentz force induced by an intense high
frequency magnetic pulse is also effective for infiltration of molten aluminum
into a preform [4]. The equilibrium of forces, as shown in Fig. 2.2a, may be broken
and the molten metal may advance.

2.4 Solid State Fabrication

It is extremely difficult to introduce fibers or particles as reinforcements directly


into a solid matrix metal and disperse them uniformly. Instead, metal powder or foil
is used as the matrix metal. Particles or fibers of reinforcements are mixed with
metal powder or foil and the mixture is sintered to fabricate a composite. Many
processes have been proposed to carry out these steps. Some important processes
are introduced here.

2.4.1 Powder Metallurgical Methods

Processes to produce metal products from metal powders are known as “powder
metallurgy” and so the processes which use metal powder to fabricate composites
are “powder metallurgical” methods. Some portion of the as-received metal powder
will always be agglomerates. It is important to loosen these agglomerates so that the
metal powder and reinforcements can be mixed uniformly. There are two methods
to mix metal powder and reinforcements: the wet method, which uses a solvent for
mixing and the dry method which does not use any solvent.
36 2 Fabrication Processes for Composites

Ceramic particle

Metal powder

Solvent

Mixing Drying by water Sintering Hot extrusion


bath

Fig. 2.3 Fabrication of particulate dispersed composites by a powder metallurgical method

1. Wet method:
Suitable solvents are liquid materials which have low surface energy (surface
tension) and low latent heat for evaporation (so that the mixture can be dried
easily). Industrial ethanol is an easily obtainable solvent which satisfies the
above conditions. The matrix metal powder and reinforcement are mixed and
agitated well in the solvent, the mixed powders are then allowed to settle and the
solvent is decanted. The mixture is then dried by either heating or by vacuum
evaporation. A water bath held at the boiling point of water is very effective for
drying an ethanol-based mixture by heating, because the boiling point of ethanol
is around 80  C. This process is shown schematically in Fig. 2.3.
One might expect that water would also be a desirable solvent. However,
water is not easy to use as a solvent for this purpose, because it has high surface
tension and a large latent heat of evaporation. In addition, the surfaces of some
reinforcements are strongly acidic or alkaline, so surface treatment of these
reinforcements is needed before water can be used as a solvent. Other
reinforcements do not disperse well in water, so additives may be required to
decrease the surface tension of water and improve dispersion.
2. Dry method:
This is a process to mix and agitate reinforcements and matrix metal without
solvent. It is not easy to disperse reinforcements uniformly, but it does have the
advantage that drying is not required. Typically, a ball mill (without solvent) is
used to attain uniform dispersion of reinforcements. The mixture obtained is
degassed, canned and then sintered using a hot press, hot isostatic pressing or a
hot press under vacuum to produce a composite.
The advantage of powder metallurgical methods is that as the matrix metal
and reinforcements are mixed in the solid state, the effect of non-wetting
between the reinforcements and liquid metal is not an issue, and relatively
good dispersion of reinforcements is easily obtained. However, it is difficult to
force a ceramic particle into a metal particle. Therefore, the uniformity of the
dispersion depends upon the particle size of the matrix metal. If we want to
disperse very fine ceramic particles (less than 1 mm in diameter) uniformly in a
matrix metal, we need to use a matrix metal powder of comparable particle size.
2.4 Solid State Fabrication 37

Fig. 2.4 Comparison of the


strengths of SAP and
aluminum alloys

At this point, metal powder production methods should be mentioned. There


are many processes, including decomposition or deoxidization of inorganic
peroxides. Of these processes, gas atomization is one of the most efficient
processes. In gas atomization, molten metal is subjected to an inert gas stream.
The gas is forced through high energy jets, which cause the molten metal to
disintegrate and rapidly solidify into the desired spherical shape. Mechanical
milling of matrix metal is also useful, and can be applied to the products of
gas atomization to produce finer metal powder. These processes produce
powders with a relatively broad particle size distribution. The percentage of
very fine particles is low in these products, and therefore the productivity of very
fine powder is limited.
SAPs, introduced in Sect. 1.1.3, are produced by a powder metallurgical
method. These composites are aluminum alloys containing dispersed fine
alumina particles. They are produced by sintering aluminum powder followed
by hot extrusion to break up the aluminum particles, which have become
covered with aluminum oxide, and then disperse this oxide (i.e. the alumina)
in aluminum. The tensile strength of the SAP developed by Irmann is shown in
Fig. 2.4 [5]. The characteristic feature of this alloy is that the tensile strength
decreases gradually with increasing temperature, compared with the relatively
rapid decrease of other aluminum alloys. For example, the high strength of
duralumin is due to fine precipitates which were formed in the matrix metal by
heat treatments. Unfortunately, this means that the tensile strength drops
quickly with increasing temperature as the precipitates grow large or redis-
solve in the matrix above the aging temperature. In contrast, the SAP is
aluminum strengthened by alumina particles, which are stable at high temper-
ature. Therefore, the strength of the SAP decreases much more gradually with
increasing temperature.
38 2 Fabrication Processes for Composites

Fig. 2.5 Relationship


between tensile strength
and volume fraction of oxides
dispersed in aluminum,
compared with SAP
(data from Benjamin [7])

Fig. 2.6 Example of high


energy ball mill used for
mechanical alloying
Inert gas
Outlet of water

Hard ball

Inlet of water

Blade

2.4.2 Mechanical Alloying

Mechanical alloying (MA) was developed by Benjamin in 1970 to strengthen heat


resisting alloys such as superalloys for jet engines [6, 7]. By this process, new
dispersion strengthened superalloys have been developed. In addition, he applied
this process to aluminum alloys and developed oxide dispersion strengthened
aluminum alloys. An example is shown in Fig. 2.5. The oxide particles produced
by an MA process are very fine and have a dramatic strengthening effect. The MA
process resembles the “dry method” of dispersion discussed in the powder metal-
lurgical method above, but uses a higher energy ball mill (known as an attritor, and
shown in Fig. 2.6) to mill for a long time (20–30 h).
2.4 Solid State Fabrication 39

Fig. 2.7 Schematic Metal powders Welding by ball


description of repeated
pressure welding and
Metal A
crushing by MA process.
Different metals (materials)
become an atomic-level
mixture as a result of this Metal B
mechanical process

Welding again Crush by ball

The sequence for MA processing is shown in Fig. 2.7, where two kinds of metals
(metal A and metal B) are milled. First, particles of metals A and B are compressed,
breaking their oxide surface layers and causing fresh metal surfaces to appear.
Then, the metal A and metal B particles are pressure welded. Once the particles
become larger after pressure welding, the particles are crushed and broken again.
This sequence of pressure welding and crushing is repeated many times over a long
period. Although this process is carried out in the solid state, we can obtain alloys
which are atomic-level mixtures. If we use oxide particles as reinforcements instead
of metal B, and mill the oxide particles with metal A powder and consolidate the
mixture, a very fine particle dispersion strengthened alloy will be obtained. Because
some alloys such as aluminum are very soft, these alloys can stick to the milling
balls and container wall during milling. To prevent this problem, methanol or
stearic acid is added into the powder mixture. If too much additive is used, pressure
welding becomes difficult [8].
Some elements have no solubility in a metal, even when the metal is liquid. In
such cases, we cannot produce an alloy by casting of liquid metal. However, it is
possible to mix powders of these elements and the matrix metal powder in the solid
state by the MA process to produce an alloy. Amorphous metals can also be
produced for some combinations of elements using the MA process. This process
is widely used for fundamental research.

2.4.3 Diffusion Bonding

This process was developed for the fabrication of continuous fiber reinforced
composites. An example of a diffusion bonding process is shown schematically in
Fig. 2.8. First, fibers are placed side by side on a metal foil and covered with another
metal foil as shown in Fig. 2.8a. This operation is repeated until a fiber/foil stack is
constructed. The stack is inserted into a container and degassed. Then, the stack is
40 2 Fabrication Processes for Composites

Fig. 2.8 Diffusion bonding a b


process by using metal foil
Degassing
(a) units of fibers and foils, Aluminum foil
(b) fiber stack in a capsule for
degassing, (c) sintering under
high pressure, (d) continuous
fiber reinforced composite

Fiber
Capsule

d c

hot-pressed and sintered under high pressure to produce a composite. In another


example, fibers are arranged and then molten metal is sprayed onto them to bind the
fibers together and make monolayer fiber sheets. The sheets are then piled up. The
stack is inserted into a container, degassed and hot-pressed to sinter it and produce a
composite.
In this process, sintering is vital to thoroughly remove voids from the stack. This
process is applicable to fabrication of unidirectionally reinforced composites, but is
difficult to use for fabrication of multidirectionally reinforced composites.

2.4.4 Spark Plasma Sintering (SPS)

Recently, spark plasma sintering (SPS) has attracted attention as an efficient solid state
fabrication technique for highly thermally conductive materials [9, 10]. SPS is a
pressure assisted sintering technique which is based on electrical spark discharge,
where a pulsed DC current passes through a conductive powder compact (the applica-
tion to non-conductive materials is also possible, but limited) [11]. In the case of
conventional hot pressing, the heat is provided by external heating. In contrast, in SPS,
the heat is generated internally between the powder particles. The sintering times for
SPS are very short, approximately 5–20 min. When carbon fibers are included in the
metal powder, heat will be generated between the fibers and metal particles by both
spark discharge and Joule heating, because carbon fiber is a conductive material. This
means that only the surfaces of the metal particles will be melted, resulting in good
2.5 Liquid State Fabrication 41

bonding with the carbon fibers, and a decrease in thermal resistance between the
carbon fibers and metal particles.
Recently, scale-like graphite particles which have high thermal conductivity in
two dimensions have been developed. Scale-like graphite particle/aluminum
composites were fabricated by SPS and showed very high thermal conductivity
[12]. A key aspect of fabricating these composites is how to distribute the scale-like
graphite particles horizontally on a plane. This scale-like graphite particle/alumi-
num composite is likely to perform better than the carbon fiber/aluminum
composites, because the carbon fiber/aluminum composites only have high thermal
conductivity in one dimension.

2.5 Liquid State Fabrication

This category contains processes where reinforcements are mixed into liquid matrix
metal. Ceramic reinforcements are usually used. These reinforcements are not wet
well by molten metal. When the contact angle between molten metal and the
reinforcement is higher than p/2, it is difficult to mix and disperse the
reinforcements in the molten metal. There are two general types of processes to
achieve this: chemical and mechanical. Chemical processes include application of
coatings to the surface of the reinforcements to improve wetting between
reinforcements and molten metal. Addition of certain elements such as Ca, Mg or
Li to aluminum alloys has also been found to improve wetting. In the mechanical
processes, the energy necessary for wetting between reinforcements and molten
metal is supplied by mechanical means, as discussed in Sect. 2.1.2.

2.5.1 Vortex Addition Technique

As shown in Fig. 2.9, a vortex forms when molten metal is strongly agitated.
Reinforcements are then added into the vortex, which is maintained until the
reinforcements wet well. The mixture is then poured into a mold. This process is
called the “vortex addition technique”, because the reinforcements are incorporated
using a vortex. In conjunction with this technique, chemical techniques such as the
addition of some elements into the matrix or coating the reinforcement surface are
used to improve the wetting of ceramic particles by molten aluminum. The effective
additive elements are Ca, Mg or Li at levels of about 1 wt.% [13].
As an example of chemical modification, ceramic particles chemically plated
with nickel or copper are wetted easily and incorporated into molten aluminum
alloys by agitating. In particular, graphite particles coated with nickel can be easily
dispersed in aluminum alloys or copper alloys. The key to the effectiveness of the
nickel chemical plating is a reaction product, nickel tetracarbonyl (Ni(CO)4),
between the carbon and the nickel coating. Ni(CO)4 mediates wetting between
42 2 Fabrication Processes for Composites

Fig. 2.9 An example of


equipment used for composite motor
production by the vortex ceramic
addition technique particles

molten metal
crucible

furnace

blades

carbon particles and aluminum. However, if long agitation times are used with
aluminum alloys, aluminum carbide (Al4C3) will be formed on the surface of
graphite particles, which means that care must be taken when these particles are
mixed into molten metal. These graphite particle dispersed alloys show good oil-
less wear resistance [14, 15]. A crucial aspect of the vortex addition technique is the
atmosphere. If the ceramic particles or fibers are added into molten aluminum in air,
the surface oxide film of the molten aluminum will be mixed into the composite
along with the fibers and the properties of the composite will be degraded. Nitrogen
or argon atmospheres are preferred.
Ceramic particles are wet easily by magnesium alloys, so it is possible to
disperse these particles into magnesium alloys to produce composites without
needing a vortex [16–18]. However, an inert gas atmosphere is still required.
By this process, when molten metal is agitated, the metal and ceramic particles
move together. It is difficult to fix only the particles and allow only the metal to
move. Therefore, most of the mechanical energy added by agitation is exhausted
in making the metal form a vortex, and a small fraction of the energy is used for
wetting of the particle/metal interface. This process is not energy efficient. If
we make a narrow space between the crucible wall and the agitator to allow shear
stress to do work in the molten metal, wetting will be improved. The addition
of some elements to the metal is also effective at increasing viscosity and
shear stress.
Duralcan, which was introduced in Sect. 1.1.3, is an example of a composite
fabricated by the vortex addition technique. This composite has been
commercialized, and a commercial product was shown in Fig. 1.7a.
2.5 Liquid State Fabrication 43

2.5.2 Compo-Casting

When an alloy is held between its liquidus and solidus temperatures, liquid coexists
with solid (i.e. the alloy is a semi-solid state). At this temperature, if the alloy is
strongly agitated, dendrites of the solids will be broken and become round particles.
By this change of solid shape, the fluidity of the alloy is improved and a good slurry
will be obtained. This slurry can be die cast, and some products have been
commercialized. This process is called “Rheocasting” [19, 20], and has the follow-
ing characteristics:
1. Homogeneous grain structure is obtained and macro-segregation decreases
compared with other die cast products.
2. Before casting, some percentage of the alloy is solid. This means that the
shrinking upon solidification is less than that for a completely liquid alloy.
Therefore the dimensional accuracy of the products increases.
3. As some percentage of the latent heat has already been released, the heat shock
to the mold is lower than that for casting a completely liquid alloy. This
increases mold life.
4. As flow stress is low compared with an entirely solid alloy, deformation is
easier; and
5. As the semi-solid alloy has high viscosity, mixing of ceramic particles is easy
and homogeneous dispersion is obtainable.
In the case of pure metal, melting occurs over a very narrow temperature range;
in other words there is no temperature span for solidification. However, we can
produce a similar semi-solid state by controlling the heat extraction rate. This heat
controlling is not straightforward.
Two different types of metal agitation techniques have been used in compo-
casting. One is direct mechanical agitation using a blade such as a screw, while the
other is indirect agitation by magnetic force [21–24]. However, when magnetic
force is used, there is a limit of 30–40 % solid fraction for effective agitation.
The apparent viscosity of the slurry depends on the fraction and shape of the solid
[25–28]. When the solid is spherical, flow is possible up to a solid fraction of 60 %.
The apparent viscosity of this semi-solid slurry is higher than that of entirely
liquid metal. By agitating the slurry, it is possible to easily mix and disperse
ceramic particles into the slurry as shown in Fig. 2.10 [29]. By adding the desired
amount of ceramic particles to the molten metal and solidifying it in a mold,
composites can be produced. This process is called the “compo-casting process”
[30]. The effect of gases (mainly H2O) adsorbed onto the surface of ceramic
particles (especially fine particles, around 1 mm in diameter) is very large when
those particles are added into the slurry. To remove the gases, heating at 873 K in
vacuum for 1 h is effective [31]. When ceramic particles are added into the slurry,
the volume of solid will increase by the volume of the particles. This increase is
offset by heating the slurry and thus decreasing the solid fraction of the matrix
metal. This allows the total solid fraction in the slurry to be held constant.
44 2 Fabrication Processes for Composites

Fig. 2.10 An example of DC motor


equipment used for the
Joint
compo-casting process
Torque meter

Thermo-couple
SiC whisker
Furnace
Insulator
Crucible

Rotor

Slurry
Blade

A narrow space between the wall of the crucible and the agitator works best for
effective crushing of the metal dendrites in the semi-solid state into spherical
particles, because shear stress works well in the narrow space. In addition, after
adding ceramic particles, the high shear stress resulting from the mechanical energy
supplied by agitation is very effective to wet the ceramic particles with the matrix
metal. The addition of magnesium into the slurry of aluminum alloys is also used in
this process to improve wetting between the ceramic particles and matrix metal.

2.5.3 Pressureless Infiltration Process

If molten metal infiltrates spontaneously into a preform, a porous material made of


ceramic fibers or particles, a composite will be easily produced. Such a pressureless
infiltration process has been developed in the USA and is named the “Lanxide
process” [32]. In this process, a preform made of SiC particles which do not wet
well with molten aluminum is treated with magnesium. The preform is set on top of
molten aluminum under a nitrogen atmosphere. Then, magnesium nitride is formed
on the surface of the SiC particles and molten aluminum spontaneously infiltrates
into the preform without pressurization. The infiltration continues until the preform
fills with molten aluminum. This process is schematically shown in Fig. 2.11. The
preform is made in the same shape as the desired product, and its outer surface
(except for the base which will be in contact with the molten aluminum) is coated so
that infiltration stops once the molten aluminum reaches the outer surface. There-
fore, after infiltration, the composite has the near-net shape of the product. The
application of this process is not wide, because appropriate combinations of
reinforcements and molten matrix metal are limited. In addition, it is not easy to
achieve the same extent of filling with molten metal in this process as can be
obtained with pressurized infiltration processes.
2.5 Liquid State Fabrication 45

700 900˚C
N2 atmosphere
Reinforcement Coating to stop infiltration

Composite part

Molten metal

Product

Fig. 2.11 Schematic description of pressureless infiltration by the Lanxide process

2.5.4 Ultrasonic Infiltration

The term “ultrasonic wave” brings to mind the ultrasonic cleaner which is very
useful to remove stains. This cleaning mechanism is an interesting phenomenon,
because ultrasonic vibration may be related to the wetting between water and stains.
Nakanishi et al. investigated the effect of ultrasonic vibration on the contact angle
between liquid and solid [33]. They showed that the apparent contact angle of a
water droplet on a paraffin coated substrate is decreased greatly by ultrasonic
vibration and reaches a stable value. Further, they applied the ultrasonic vibration
to the infiltration of molten aluminum into an alumina fiber preform and showed the
feasibility of ultrasonic infiltration for the fabrication of composites [34]. The
ultrasonic vibration reduced the threshold pressure for the infiltration of molten
aluminum into the alumina fiber preform. The reduction of the threshold pressure
was about 140 kPa for an alumina fiber preform with Vf ¼ 0.17 at a resonant
frequency of 20.5 kHz [34]. When an ultrasonic wave propagates through a
material, the energy travels through the atomic structure by a series of compression
(dense zone) and expansion movements (rough zone). As the dense zone is under
high pressure, when the dense zone passes through the reinforcement/molten metal
interface, the equilibrium of forces may be broken and the molten metal may
advance as shown in Fig. 2.2a.
In another application of ultrasonic waves, Deming et al. [35] developed
“preform wires” (which are aluminum composite wires reinforced with continuous
SiC fibers), by ultrasonic infiltration. They showed that it is possible to infiltrate
molten aluminum into the fiber bundle using ultrasonic vibration without fiber
coating or fiber pretreatment. Preform wires of carbon fiber/aluminum composites
46 2 Fabrication Processes for Composites

by ultrasonic infiltration without fiber surface treatments were developed by Cheng


et al., although the wettability of molten aluminum with carbon fiber is very
poor [36]. Matsunaga et al. determined the optimum conditions (such as preheating
temperature of fibers, fabrication speed, ultrasonic power, among others) for the
fabrication of carbon fiber/Al–Mg alloy preform wires by ultrasonic infiltration
[37–40]. The effect of high temperature holding on the carbon fiber/Al–Mg
alloy preform wire composites fabricated by supersonic infiltration was reported
by Mizoguchi et al. [41] and Yamaguchi et al. [42]. Research on the interface
reaction of carbon fiber/Al alloy preform wire composites was performed by
Mikuni et al. [43]. The effect of Mg content on the tensile strength of the carbon
fiber/Al–Mg alloy preform wire composites was clarified by Matsunaga et al. [44].

2.6 Gaseous State Fabrication

A thin film can be produced on the surface of a material (substrate) by depositing a


substance from a gaseous state. The material coated with another material is a kind
of composite. It takes a very long time to make a thick layer of coating material
using these processes. Therefore, the application of gaseous state fabrication
techniques is limited to thin film coating. There are many processes which have
been developed for the production of thin films. Those processes are roughly
classified into CVD and PVD processes.

2.6.1 CVD

During CVD (chemical vapor deposition), flowing gas phase materials react on the
surface of a hot solid material (substrate) and solid reaction products deposit on the
surface of the substrate. Three major processes have been used: (1) reduction of
metal halides by hydrogen with a catalyst, (2) thermal decomposition of gas phases
and (3) the reaction of the substrate with gas phases.
The deposition rate or crystallization rate depends mainly on the temperature of
the substrate. Therefore, the microstructure of the deposited layer also depends on
the substrate temperature. Generally, at high temperature, several kinds of single
crystals grow and at lower temperature, polycrystalline or amorphous phases grow.
Fine polycrystalline deposits having dense and homogeneous mechanical
properties are desirable for coating purposes.
An example of the apparatus for this process, which has a heating system around
the furnace inner wall, is shown in Fig. 2.12 This equipment is suitable for coating
of many small substrates at the same time and is called a “hot wall type” apparatus.
Alternatively, when high frequency induction heating is used, the furnace wall is
not hot; this is the “cold wall type” of apparatus. When the difference in thermal
expansion coefficient between the coating material and substrate is large, cracks
2.6 Gaseous State Fabrication 47

Fig. 2.12 An example of a


hot wall CVD system
Substrate

Furnace

Thermocouple
Raw material gases Gas outlet

may form at the substrate/coating layer interface after CVD. To prevent such
cracking, the following methods are employed:
1. Using a thinner coating and sacrificing some of the properties of the coating
material.
2. Using an intermediate phase or diffusion layer containing both materials
between the coating material and the substrate.

2.6.2 PVD

The simple image of PVD (physical vapor deposition) is that a material is


evaporated by electrical resistive heating or electron beam heating in a vacuum
and is deposited onto the surface of a substrate. Many different PVD processes
exist; sputtering is a typical example. In the sputtering process, atoms or clusters are
ejected from a material (target) by ion bombardment and deposited onto the
surfaces of substrates. Another variation on this process, which includes additional
materials (such as C2H2 or NH3) in the atmosphere, has been developed to produce
carbides or nitrides [45]. Here we introduce two types of sputtering.
48 2 Fabrication Processes for Composites

2.6.2.1 Conventional Sputtering

In an argon, oxygen or nitrogen plasma at a pressure of 1–10 Pa, atoms are ejected
from the negatively charged target (the material to be used for the coating) by
bombardment of positive ions from the plasma and then deposited onto the surface
of a substrate. The potential difference used in this technique is several thousand
volts. During this process, the substrate temperature rises to several hundred  C,
because the substrate is exposed to high speed electrons. The deposition rate is
about 0.1 mm/min.

2.6.2.2 Ion Beam Sputtering

An ion beam is injected from a separately equipped ion source chamber at a


pressure of 0.1–1 Pa and irradiates the target. The ejected atoms from the target
are deposited onto a substrate. With this equipment, because the substrate is not
exposed to an electron beam, the substrate temperature can be lower, allowing
better control of the deposition conditions.

2.7 In Situ Fabrication Process

In some cases, another material can be generated and grown within an alloy by
controlling the atmosphere and temperature of the alloy. If the conditions are
suitable, the generated material can reinforce the alloy. Such processes are “in
situ fabrication”. Particles generated during these processes by chemical reactions
are relatively stable at high temperature compared with reinforcements mixed into
the matrix in other composite fabrication processes. Nearly uniform, fine particles
can be produced by in situ fabrication processes.

2.7.1 Internal Oxidation Process

When a copper alloy including a small amount of aluminum is held at a high


temperature in an oxidizing atmosphere, the oxygen diffusing into the alloy from
the surface reacts preferentially with aluminum atoms and thus alumina particles
are formed in the alloy. The final product is a fine alumina particle dispersed copper
alloy. Thus, if an alloy includes a small amount of an element that is more reactive
towards oxygen than the matrix metal, a similar oxide particle dispersed metal will
be obtained when the alloy is held at high temperature in an oxidizing atmosphere.
However, as oxygen diffuses from the surface of the alloy, oxidization proceeds
near the surface of the alloy and decreases with increasing distance from the
2.7 In Situ Fabrication Process 49

surface. The oxidization usually occurs only in a thin area near the surface, making
this process a type of surface treatment.
As an example, the fabrication process for a SiO2 (or TiO2) dispersed Ni
composite is introduced here [46, 47].
Starting material: Ni–0.5 wt.% Si alloy (or Ni–0.5 wt.% Ti alloy).
#
Insert the alloy into a mixture of powders of Ni and NiO (mole ratio of 1:1) and
degas.
#
Hold at about 1,400 K for 48 h (will be under oxygen pressure as a result of
decomposition of NiO).
#
Ni–SiO2 (or Ni–TiO2) dispersion strengthened alloy.
By this process, internal oxidization occurs for a depth of about 1 mm from the
surface. Generally, near the surface, particles are finer and the particle size becomes
larger deeper into the alloy. The average particle size is 1–5 mm. A similar process
to produce nitride particles in an alloy using a nitrogen atmosphere is called the
“internal nitridation process”.

2.7.2 Unidirectional Solidification Process

When a eutectic alloy (a and b phases) is solidified at a unidirectional solidification


rate R under conditions of constant temperature gradient, lamellar type crystals of
two phases grow at the same time. This produces a kind of composite of lamellar
crystals of the two phases stacked alternately. The thickness of the lamellae and the
solidification rate R have the following relationship:

l2 R ¼ constant; (2.16)

where l is the lamellar spacing (from the center of a lamellar to the center of b
lamellar) [48, 49]. It is clear that if the solidification rate increases, the thickness of
the lamellae decreases and a composite containing thin a phase and b phase is
obtained.
In addition, when the volume fraction of one phase in the eutectic alloy is
relatively small, the phase grows as rod-like crystals under certain conditions. It
is known that in the case of Al–Al3Ni alloy, the rod-like crystals are the same as
whiskers [50].
50 2 Fabrication Processes for Composites

References

1. Nishida, Y.: Development of pressure infiltration method for fabrication of metal matrix
composites. Materia Jpn. 36, 40–46 (1997)
2. Rocher, J.P., Quenisset, J.M., Naslain, R.: Wetting improvement of carbon or silicon carbide
by aluminium alloys based on a K2ZrF6 surface treatment: application to composite material
casting. J. Mater. Sci. 24, 2697–2703 (1989)
3. Nakanishi, H., Tsunekawa, Y., Okumiya, M., Mori, N., Niimi, I., Sato, M.: Ultrasonic
infiltration in alumina particle/molten aluminum system assisted by exothermic reaction of
titanium aluminide formation. J. Jpn. Inst. Met. 57, 81–87 (1993)
4. Andrews, R.M., Mortensen, A.: Lorentz-force-driven infiltration by aluminum. Mater. Sci.
Eng. A 144, 165–168 (1991)
5. Irmann, R.: On a new sintered aluminum product with high strength at elevated temperatures.
Leichtmetall 3, 21–25 (1950)
6. Benjamin, J.S.: Dispersion strengthened superalloys by mechanical alloying. Metall. Trans. 1,
2943–2951 (1970)
7. Benjamin, J.S., Bomford, M.J.: Dispersion strengthened aluminum made by mechanical
alloying. Metall. Trans. 8A, 1301–1305 (1977)
8. Horiuch, R., Kohara, Y.: Aluminum alloy made by mechanical alloying and fiber reinforced
aluminum. J. Jpn. Inst. Light Met. 32, 688–695 (1982)
9. Imanishi, T., Sasaki, K., Katagiri, K., Kakitsuji, A.: Thermal and mechanical properties of
VGCF-containing aluminum. Trans. Jpn. Soc. Mech. Eng. A 74, 655–661 (2008)
10. Imanishi, T., Sasaki, K., Katagiri, K., Kakitsuji, A.: Effect of CNT addition on thermal
properties of VGCF/aluminum composites. Trans. Jpn. Soc. Mech. Eng. A 75, 27–33 (2009)
11. Tokita, M.: Trend in advanced SPS spark plasma sintering systems and technology. J. Soc.
Powder Technol. Jpn. 30, 790–804 (1993)
12. Ueno T, Yoshioka H.: Japanese Patent JP 4441768
13. Hikosaka, T., Miki, K., Nishida, Y.: Mechanical properties of aluminum-alumina particle
composites fabricated by vortex method. Imono (J Jpn. Foundry Eng. Soc.) 61, 780–786
(1989)
14. Badia, F.A., Rohatgi, P.K.: Dispersion of graphite particles in aluminium castings through
injection of melt. Trans. AFS 77, 402–406 (1969)
15. Suwa, M., Komuro, K., Soeno, K.: Mechanical properties and wear resistance of graphite-
dispersed Al–Si alloys. J. Jpn. Inst. Met. 40, 1074–1081 (1976)
16. Lim, S.-w., Cho, T.: Effect of alloying elements on SiC particulate dispersion behavior in
molten magnesium. J. Jpn. Inst. Met. 56, 210–217 (1992)
17. Lim, S.-w., Cho, T.: Mechanical properties of SiC particulate reinforced magnesium matrix
composites fabricated by melt stirring method. J. Jpn. Inst. Met. 56, 1101–1107 (1992)
18. Lim, S.-w., Cho, T.: Effect of alloying elements on particulate dispersion behavior and
mechanical properties in TiC particulate reinforced magnesium matrix composites. J. Jpn.
Inst. Light Met. 42, 772–778 (1992)
19. Spencer, D.B., Mehrabian, R., Flemings, M.C.: Rheological behavior of Sn-15 pct Pb in the
crystallization range. Metall. Trans. 3, 1925–1932 (1972)
20. Flemings, M.C., Mehrabian, R.: Casting in the liquid–solid region. Trans. AFS 81, 81–88
(1973)
21. Flemings, M.C.: Behavior of metal alloys in the semisolid state. Metall. Mater. Trans. 22A,
957–981 (1991)
22. Nannba, A.: Semi-solid metal processing. J. Jpn. Inst. Light Met. 45, 346–354 (1995)
23. Vives, C.: Elaboration of semisolid alloys by means of new electromagnetic rheocasting
processes. Metall. Mater. Trans. 23B, 189–206 (1992)
24. Ichikawa, R.: Present status of rheocast process. Tetsu-to-Hagane 74, 51–60 (1988)
25. Ichikawa, R., Miwa, K.: Apparent viscosity and structure in partially solidified Al–Cu alloys. J.
Jpn. Inst. Met. 42, 1023–1028 (1978)
References 51

26. Mori, N., Ohgi, K., Matsuda, K.: On the apparent viscosity and structure of partially solidified
Al–Cu alloys under stirring. J. Jpn. Inst. Met. 48, 936–944 (1984)
27. Shibuya, A., Arihara, K., Nakamura, Y.: Measurement of apparent viscosity of ferrous and
non-ferrous alloys in liquid/solid coexisting state-Fe–C, Sn–Pb, Al–Cu and Fe–Cr–Ni–C
alloys. Tetsu-to-Hagane 66, 1550–1556 (1980)
28. Hirai, M., Takebayashi, K., Yoshikawa, Y., Yamaguchi, R.: Apparent viscosity of semi-solid
metals. Tetsu-to-Hagane 78, 902–909 (1992)
29. Nishio, T., Kobayashi, K., Miwa, K., Ozaki, K., Asano, S.: Effect of rotor shape on flow slurry
in compocasting process. Rep. Natl. Ind. Res. Inst. Nagoya Jpn. 44, 75–81 (1995)
30. Sato, A., Mehrabian, R.: Aluminum matrix composites: fabrication and properties. Metall.
Trans. 7B, 443–451 (1976)
31. Miwa, K.: Fabrication of SiCp reinforced aluminum matrix composites by compocasting
process. Imono (J. Jpn. Foundry Eng. Soc.) 62, 423–428 (1990)
32. Nagelberg, A.S., Antolin, S., Urquhart, A.W.: Formation of Al2O3/metal composites by the
directed oxidation of molten aluminum–magnesium–silicon alloys: part II, growth kinetics. J.
Am. Ceram. Soc. 75, 455–462 (1992)
33. Nakanishi, H., Tsunekawa, Y., Mohri, N., Okumiya, M., Niimi, I.: Ultrasonic infiltration in
alumina particle/molten aluminum system. J Jpn. Inst. Light Met. 43, 14–19 (1993)
34. Nakanishi, H., Tsunekawa, Y., Okumiya, M., Mohri, N.: Ultrasonic infiltration in alumina
fiber/molten aluminum system. Mater. Trans. JIM 34, 62–68 (1993)
35. Deming, Y., Xinfang, Y., Jin, P.: Continuous yarn fibre-reinforced aluminium composites
prepared by the ultrasonic liquid infiltration method. J. Mater. Sci. Lett. 12, 252–253 (1993)
36. Cheng, H.M., Lin, Z.H., Zhou, B.L., Zhen, Z.G., Kobayashi, K., Uchiyama, Y.: Preparation of
carbon fibre reinforced aluminum via ultrasonic liquid infiltration technique. Mater. Sci.
Technol. 9, 609–614 (1993)
37. Matsunaga, T., Matsuda, K., Hatayama, T., Shinozaki, K., Amanuma, S., Jin, P., Yoshida, M.:
Development in manufacturing of carbon fiber reinforced aluminum preform wires using
ultrasonic infiltration method. J. Jpn. Inst. Light Met. 56, 28–33 (2006)
38. Matsunaga, T., Ogata, K., Hatayama, T., Shinozaki, K., Yoshida, M.: Infiltration mechanism
of molten aluminum alloys into bundle of carbon fibers using ultrasonic infiltration method. J.
Jpn. Inst. Light Met. 56, 226–232 (2006)
39. Matsunaga, T., Ogata, K., Hatayama, T., Shinozaki, K., Yoshida, M.: Effect of acoustic
cavitation on ease of infiltration of molten aluminum alloys into carbon fiber bundles using
ultrasonic infiltration method. Composites Part A 38, 771–778 (2007)
40. Matsunaga, T., Matsuda, K., Hatayama, T., Shinozaki, K., Yoshida, M.: Fabrication of
continuous carbon fiber-reinforced aluminum–magnesium alloy composite wires using ultra-
sonic infiltration. Composites Part A 38, 1902–1911 (2007)
41. Mizoguchi, I., Yamaguchi, S., Yachi, S., Yoshida, M.: Influence of high temperature holding
on tensile strength of pitch-based carbon fiber reinforced Al–Mg alloy composites fabricated
by ultrasonic infiltration method. J. Jpn. Inst. Light Met. 60, 396–402 (2010)
42. Yamaguchi, S., Mikuni, J., Mizoguchi, I., Matsunaga, T., Shinozaki, K., Yoshida, M.: Influ-
ence of high temperature holding on tensile strength of PAN-based carbon fiber reinforced
aluminum–magnesium alloy composites fabricated by ultrasonic infiltration method. J. Jpn.
Inst. Light Met. 59, 241–247 (2009)
43. Mikuni, J., Nonokawa, K., Matsunaga, T., Shinozaki, K., Yoshida, M.: Influence of interfacial
chemical reaction for tensile strength of carbon fiber reinforced aluminum–magnesium alloy
composites. J. Jpn. Inst. Light Met. 58, 27–32 (2008)
44. Matsunaga, T., Matsuda, K., Hatayama, T., Shinozaki, K., Amanuma, S., Yoshida, M.: Effect
of magnesium content on tensile strength of carbon-fiber-reinforced aluminum–magnesium
alloy composite wires fabricated by ultrasonic infiltration method. J. Jpn. Inst. Light Met. 56,
105–111 (2006)
45. Solzbacher, F.: Physical vapor deposition. In: Semiconductor Manufacturing Handbook.
McGraw-Hill, New York (2005) (Chapter 13)
52 2 Fabrication Processes for Composites

46. Goto, S., Mori, K., Yoshinaga, H.: High-temperature hardness of dispersion-hardened Ni–SiO2
alloys made by internal oxidation method. J. Jpn. Inst. Met. 46, 764–772 (1982)
47. Matsuda, N., Matsuura, K.: Work hardening of a dispersion hardened Ni–TiO2 alloy. J. Jpn.
Inst. Met. 48, 362–370 (1984)
48. Chalmers, B.: Principles of Solidification, p. 204. Wiley, New York (1964)
49. Flemings, M.C.: Solidification Processing, p. 94. McGraw-Hill Book Co., New York (1974)
50. Lemkey, F.D., Hertzberg, R.W., Ford, J.A.: The microstructure, crystallography and mechani-
cal behavior of unidirectionally solidified Al–Al3Ni eutectic. Trans. AIME 233, 334–341
(1965)
Chapter 3
Fabrication by Squeeze Casting

Abstract The history of squeeze casting as a fabrication process for MMCs is


briefly described in this chapter. Squeeze casting is one of the liquid state fabrica-
tion techniques. We can learn essentials of the fabrication of composites from
studying squeeze casting, because, during squeeze casting, mechanical energy is
converted into interface energy at the reinforcement/matrix interface. This energy
conversion is economical and efficient and means that composites can be fabricated
with minimum energy using squeeze casting. In this chapter, the threshold pressure
equation for infiltration into preforms is introduced, and the infiltration energy,
effect of preform preheat temperature, and microscopic phenomena occurring
during infiltration of molten metal are discussed theoretically.

Research on a high pressure casting began around 1960 at the Government Indus-
trial Research Institute, Nagoya, Japan [present name: National Institute of
Advanced Industrial Science and Technology (AIST)] [1–3]. In this high pressure
casting, molten metal was solidified under high hydrostatic pressure, supplied
mechanically, to improve the properties of the metal and to form a near-net shape
product at the same time. This process is now called “squeeze casting”. In around
1965, after the technique for application of high hydrostatic pressure was
established, attempts were made to use a sand core in a metallic mold under high
hydrostatic pressure. However, molten metal infiltrated into the sand core, so the
sand cores did not work and those particular experiments were failures. However,
this was in fact the first time we noticed what would become the pressure infiltration
process for the fabrication of metal matrix composites.
After that, we studied the segregation which often occurred during the high
pressure casting. In these experiments, heat insulating cups made of asbestos were
used to hold the molten aluminum at uniform temperature in the mold before
pressure application. After applying high pressure, the infiltration of the molten
metal into asbestos, to produce what appeared to be an aluminum/asbestos
composite, was observed [4, 5]. At AIST in Kyushu, Japan, aluminum alloy/
Shirasu-balloon composites were produced by a pressure infiltration process [6].

Y. Nishida, Introduction to Metal Matrix Composites: Fabrication and Recycling, 53


DOI 10.1007/978-4-431-54237-7_3, # Springer Japan 2013
54 3 Fabrication by Squeeze Casting

The Shirasu-balloons were made by heating volcanic ash at 1,000  C. Because the
diameter of the Shirasu-balloons was 200–500 mm, molten aluminum alloy
infiltrated easily into the packed Shirasu-balloon layer. Subsequently, finer fibers
such as glass fibers, carbon fibers and SiC whiskers were actively used in attempts
to fabricate composites reinforced with those fibers [7]. This showed the feasibility
of producing high strength composites, and pressure infiltration has since become
an important process for the fabrication of metal matrix composites.
For example, the fabrication of continuous fiber reinforced metals by the pres-
sure infiltration process has been extensively studied by Kagawa et al. and the
excellent properties of those composites have been demonstrated [8–12]. Towata
et al. showed that the strength of continuous fiber reinforced composites can be
drastically improved by using preforms made of fibers covered with fine ceramic
powder or whiskers to prevent the fibers from contacting each other [13, 14]. After
new whiskers such as potassium titanate whisker (K2O6TiO2) and silicon nitride
whisker (Si3N4) were developed, fabrication of whisker reinforced aluminum alloys
was studied [15–18]. SiC whisker reinforced magnesium alloy composites were
also developed by Kim et al. using the pressure infiltration process [19, 20].
Research into fabrication processes for MMC products has actively continued,
and some MMC products are now commercialized [21–23]. The fabrication process
used for most commercial composites is “squeeze casting”, a type of high pressure
casting. However, there are many possible means by which high pressure can be
generated for infiltration. If the generated pressure is high enough to infiltrate
molten metal into preforms made of fibers or particles, composites can be
fabricated.
After about a decade of active research into MMC fabrication, the theory of the
process became established around 1995. Before then, many researchers worked to
understand the phenomena that governed the fabrication of MMCs [24, 25]. Com-
puter simulations were used by Lacoste et al. to understand infiltration (and the
associated heat transfer) in fibrous preforms [26]. The simulation technique is a
useful one to understand infiltration into preforms with complicated shapes. Alter-
natively, if we can obtain a rigorous analytical description of infiltration into a
preform, we can determine the general relationships between the infiltration
parameters.

3.1 Application of Pressure and Fabrication Energy

A schematic drawing of the pressure infiltration process for the fabrication of


MMCs is shown in Fig. 3.1. A preform made of fibers or particles is set into a
metallic mold, and molten metal is poured onto the preform. High pressure is then
applied using a mechanical punch to make the molten metal infiltrate into the
preform and solidify under the hydrostatic pressure to give an MMC product. In
this case, one can produce composites by just using compacted fibers or particles
in the mold instead of a preform. However, if the fibers are fixed to each other with
3.1 Application of Pressure and Fabrication Energy 55

Punch

Metal mold

Heater

Molten metal

Preform

Bottom plate

Fig. 3.1 Schematic of apparatus for pressure infiltration process

Set up pressure
Pressure [MPa]

End of infiltration

Threshold pressure Work used for infiltration

Punch displacement [m]

Fig. 3.2 Applied pressure versus punch displacement during pressure infiltration

binder to make a preform, as discussed in Fig. 2.2, the fibers have a rigid structure
and high pressure can be effectively applied to the molten metal. In contrast, simply
compacted fibers are easily compressed and deformed by pressure. This deforma-
tion of the compacted fibers causes the channels between fibers to narrow, and
sometimes the infiltration then stops because the channels are too narrow to
infiltrate.
Figure 3.2 schematically shows the relationship between the displacement of the
punch and the applied pressure during the pressure infiltration process. The pressure
starts to rise at the moment when the punch contacts the molten metal. When
the pressure reaches the threshold pressure (or the infiltration start pressure),
56 3 Fabrication by Squeeze Casting

infiltration begins. As shown in the figure, the punch displacement continues up to


the end point of infiltration. The total volume occupied by the punch displacement
roughly agrees with the volume of the molten metal that has been infiltrated into the
preform. After the infiltration finishes, pressure rises quickly up to the set up
pressure and the punch displacement stops. Therefore, the integration of this
pressure curve along the punch displacement axis (i.e., the hatched area below
the pressure curve) corresponds to the work needed for the infiltration (i.e., the
fabrication energy of the composite).

3.2 Threshold Pressure

3.2.1 Case of Random Reinforcement Distribution

A higher pressure than the threshold pressure is necessary to start infiltration into a
preform made of ceramic fibers or particles that are barely wet by the molten metal.
We will now introduce the equation for the threshold pressure. Assume that the
distribution of discontinuous fibers is random and the threshold pressure at any place
in the preform is a constant Pc. If the pressure of the molten metal is held slightly
higher than the threshold pressure, Pc + dP using the apparatus shown in Fig. 3.1,
infiltration will start and advance slowly. When the area of the top surface of the
preform is A, and the infiltrated distance is L, the work Wc done for the infiltration is:

Wc ¼ ðPc þ dPÞAð1  Vf ÞL: (3.1)

A is multiplied by (1–Vf), because the work is done only to the molten metal
portion of the preform surface area A and therefore the volume fraction of the fibers
should be excluded. Setting A ¼ 1 and L ¼ 1 makes Wc the work necessary for the
fabrication of a unit volume of composite. Wc should therefore coincide with the
work of Eq. (2.6), so that:

ðPc þ dPÞð1  Vf Þ ¼ ðgfl  gfa ÞAf : (3.2)

Since dP has an infinitely small value and can be neglected, Pc becomes:

ðgfl  gfa ÞAf


Pc ¼ : (3.3)
1  Vf

The dimensions in the transformation of Eqs. (3.2) and (3.3) are important. To
give Eq. (3.3), Eq. (3.2) must be divided by the dimension of volume as well as
(1–Vf), because the left-hand side of Eq. (3.2) is multiplied by the dimension of unit
volume. Substitution of Young’s equation into Eq. (3.3) gives:
3.2 Threshold Pressure 57

gla Af cos y
Pc ¼  : (3.4)
1  Vf

(1) Af for a preform made from long fibers


When the fiber diameter is df, the average fiber length is l and the number of
fibers is n, the volume fraction of fibers in unit volume of the composite is:
 2
df
p nl ¼ Vf ; (3.5)
2

where

4Vf
n¼ : (3.5a)
plðdf Þ2

The areas of both ends of the fibers can be neglected. Af is given by:

4Vf
Af ¼ pldf n ¼ : (3.6)
df

(2) Af for a preform made from particles


When the particle diameter is df and the number of particles is n, the volume
fraction of particles in unit volume of the composite is:
 
4 df 3 6Vf
p n ¼ Vf ; where n¼ (3.7)
3 2 pðdf Þ3

Af in this case is:


 2
df 6Vf
Af ¼ 4p n¼ : (3.8)
2 df

The threshold pressures obtained by using Eqs. (3.6) and (3.8) give the values for
a random distribution of fibers or particles. However, real local threshold pressures
in the preform are a little higher or a little lower than the value given by Eq. (3.4).
The threshold pressure has been discussed by Carman [27], White [28] and
Mortensen et al. [29].

3.2.2 Case of Unidirectional Fiber Distribution

When the fiber distribution is unidirectional, the threshold pressure depends on the
flow direction of the molten metal against the fiber axis. If the flow direction is
58 3 Fabrication by Squeeze Casting

perpendicular to the fiber axis direction, the molten metal should flow through the
narrowest space between the fibers. In that situation, the threshold pressure is
determined by the narrowest space. This model was discussed by Nakanishi et al.
[30–33], and the threshold pressure obtained is:

2g cos y
Pc ¼   la : (3.9)
1=2
p
df 4Vf 1

The threshold pressure calculated by Eq. (3.9) is about double that calculated
using Eq. (3.4). Apart from the narrowest space between the fibers, the actual
threshold pressure will be lower than the value calculated from Eq. (3.9). However,
the pressure given by Eq. (3.9) is that needed to start infiltration.

3.2.3 Measurement of Threshold Pressure

Threshold pressures have been obtained experimentally by Oh et al. [34, 35]. They
measured threshold pressures when molten aluminum alloys infiltrated into
compacted particle layers of SiC and B4C having 52 vol.% of theoretical density.
The diameters of the SiC and B4C particles were 9.6 and 8.9 mm, respectively. For
the SiC particles, the threshold pressure near the melting temperature of the alloys
was about 900 kPa and decreased with increasing temperature for all alloys except
for Al–2 wt.% Mg. The threshold pressure for the Al–2 wt.% Mg alloy was lower
than those of the other alloys. For the B4C particles, the threshold pressure near the
melting temperature was about 800 kPa and also decreased with increasing temper-
ature. The decrease in threshold pressure for B4C particles with increasing temper-
ature was noticeably steeper for the Al–2 wt.% Mg alloy compared to the other
alloys studied.

3.3 Influence of Preheating of Preform

As shown in Fig. 3.3, molten metal is poured onto a compacted particle layer and
pressure is applied to force the molten metal to infiltrate into the particle layer. The
influence of the preheat temperature of the particle layer on the infiltration of
molten metal was investigated by Nagata and Matsuda [36–40]. When a molten
pure metal at its melting point contacts the particle layer, some part of the metal
solidifies, covering the particles, as shown in Fig. 3.4. At the same time, the
particles are heated up to the melting point of the metal. When the initial
temperatures of the molten metal and particles are Tm and Tf, respectively, and
3.3 Influence of Preheating of Preform 59

Fig. 3.3 Production method


of particulate reinforced
metal matrix composites by
pressure infiltration process
[36] (courtesy of Prof. S. Punch
Nagata)

Molten metal

Metal mold
Particles
Heater

Air outlet

Fig. 3.4 Covering particles


with solidified metal during
infiltration process [36]
(courtesy of Prof. S. Nagata)

the solid fraction of the metal solidified as a result of supplying heat to the particles
is F, we can obtain the following relationship:

Vf rf Cf ðTm  Tf Þ ¼ Fð1  Vf Þrm H; (3.10)

where rf, rm, Cf and H are the density of particles, the density of solid metal, the
heat capacity of particles and the latent heat of the metal, respectively. This
phenomenon is schematically shown in Fig. 3.4.
60 3 Fabrication by Squeeze Casting

Fig. 3.5 Dependence of infiltration length on the preheat temperature of particles; a critical
preheat temperature is clearly observed for the molten metal/particles system [36] (courtesy of
Prof. S. Nagata)

We assume that the particles and the molten metal exchange heat and that the
particles are covered uniformly with the solid metal. Then, the apparent diameter of
particles increases. From this assumption and Eq. (3.10), the solid fraction F and the
liquid fraction Lm are given by:

Vf rf Cf ðTm  Tf Þ
F¼ (3.11)
ð1  Vf Þrm H

Lm ¼ 1  F (3.12)

Nagata et al. carried out many experiments on several kinds of particles. An


example of their results is shown in Fig. 3.5 which shows the relationship between
the preheat temperature and the infiltration distance. In the figure, the notation
“SnM–PbP system” means that molten Sn was infiltrated into Pb particles. When the
preheat temperature of the Pb particles is lower than 403 K, the infiltration distance
of molten tin is shorter than 40 mm. The infiltration distance suddenly increases
above 403 K. In systems containing other particles the curves are very similar and
all feature a particular temperature above which infiltration advances suddenly.
Nagata et al. called this temperature the “critical preheat temperature”. In addition,
to continue infiltration of molten metal, Lm must be higher than a critical value. The
critical value of Lm depends on the material used for the particles and ranges from
0.3 to 0.49.
3.4 Microscopic Analysis of Fluid Flow in Preforms 61

Solidified metal

Original fiber

Fig. 3.6 Schematic description of fibers covered with solidified metal and their increased apparent
diameter

This idea can be applied to fiber reinforcements. In a similar way, fibers and
molten metal exchange heat and the fibers become covered with solid metal of
uniform thickness. The apparent diameter of the fibers increases as shown in
Fig. 3.6, where the white part inside the fiber is the original fiber and the outside
part (hatched part) is the solidified metal. In this model, all the solidified metal
covers the fibers and no other solid is present within the molten metal. When the
fiber length is l and the fiber diameter increases from df to d0 f, the following
relationships are obtained:
 2
df
p l ¼ Vf (3.13)
2

 0 2
df
p l ¼ Vf0 (3.13a)
2

Therefore, the apparent fiber diameter d0 f and the apparent volume fraction of
fibers V0 f are:
   1=2
1
df0 ¼ df 1þ 1 F (3.14)
Vf

Vf0 ¼ Vf þ ð1  Vf ÞF: (3.15)

3.4 Microscopic Analysis of Fluid Flow in Preforms

There are two methods to mathematically analyze the infiltration of molten metal
into a preform under pressure. One method takes the microscopic viewpoint, while
the other is from the macroscopic viewpoint. The microscopic method is helpful to
62 3 Fabrication by Squeeze Casting

Fig. 3.7 Molten metal flow


assuming that pores are
Molten metal
channels with elliptical cross-
section having a major radius
of R1 and a minor radius of R2
Preform
Channel in
the preform

visualize the infiltration. However, this method is not useful for industrial
applications, where the shapes of the pores in the preform are very complicated,
making it difficult to use a microscopic model of the pore to analyze infiltration. If
we assume some specific shape of pores in the preform and calculate the infiltration,
the agreement with experimental results will usually be poor. The alternative is the
macroscopic method, where the permeability of the preform to fluid flow is used to
analyze infiltration without taking the microscopic pore shape into account. The
permeability is given by an empirical equation which is applicable for most pore
shapes, and the calculated results agree well with experimental ones. In addition,
we can obtain a lot of information which is needed for industrial applications.
First, we will obtain the threshold pressure from the microscopic viewpoint.
Assume that the fibers or particles are not wet well with molten metal, and that the
pores in the preform are channels having an elliptical cross-section, where molten
metal flows through the channels under the applied pressure, as shown in Fig. 3.7.
The front of fluid flow (infiltration front) has a curved surface, as shown in Fig. 3.7,
and the pressure inside the curved surface (in this case, in the fluid) is always higher
than the pressure outside the curved surface. The pressure difference DP is
given by:
 
1 1
DP ¼  þ g cos y; (3.16)
R1 R2 la

where R1 and R2 are the radii of the major and minor axes of the ellipse, respec-
tively. This equation gives the threshold pressure from the microscopic viewpoint.
When R1 and R2 are equal (and therefore, the cross-section of the channel is a circle
with radius R), Eq. (3.16) agrees with Eq. (2.8) and the pressure difference is:

2gla cos y
DP ¼  : (3.17)
R

The pressure differences between the inside and outside of the curved fluid
calculated using Eq. (2.8) are shown in Fig. 3.8. Curves are given for different
values of contact angle y. It is apparent that the threshold pressure increases steeply
as the radius of the channel becomes small. As y increases, the threshold pressure
3.4 Microscopic Analysis of Fluid Flow in Preforms 63

Fig. 3.8 Curves of


infiltration starting pressure
(threshold pressure) versus
radius of a channel for
different values of contact
angle

increases greatly. Therefore, this result clearly shows that very high pressure is
needed to push molten metal into a very narrow space between fibers in a preform.
When higher pressure than the threshold pressure is applied, infiltration starts.
The shear stress applied to ordinary fluids such as molten metal is proportional to
the strain rate of the fluid. Such viscous fluids are called “Newtonian fluids”. When
the viscous fluid is flowing through a long cylindrical pipe and the flow is laminar in
steady state, there is a pressure difference DP between both ends of the pipe arising
from the resistance against the flow. Incompressible viscous fluid flow through a
pipe is expressed by the Poiseuille equation:

dV pDPr 4
¼ ; (3.18)
dt 8Lm

where V is the volumetric flow, L is the length of the pipe, r is the radius of the pipe,
m is the viscosity and t is time. In this equation, gravity is neglected. Equation (3.18)
shows that the flow rate through the pipe is proportional to the pressure difference
and the fourth power of r, and is inversely proportional to the length of the pipe and
viscosity. As the flow is viscous, the velocity of the fluid is zero at the wall of the
pipe and highest in the center of the pipe (i.e., on the axis of the pipe). Then, by
using the average velocity V, the Hagen–Poiseuille equation is given by:

8mLV
DP ¼ : (3.19)
r2
64 3 Fabrication by Squeeze Casting

Fig. 3.9 Curves of pressure


difference between both ends
of a channel (length: 5 cm)
versus radius of the channel.
The viscous fluid is molten
aluminum, and curves are
shown for two different flow
velocities, V

By using this equation, the pressure difference between both ends of the pipe is
calculated and plotted against the radius of the pipe in Fig. 3.9, for different values
of flow velocity V . The results show that to attain a high flow velocity, a high
pressure difference must be applied between the ends of the pipe.
As discussed above, it is possible to calculate the threshold pressure and infiltra-
tion of molten metal by assuming that the pores in a preform are channels with
circular cross-section. However, the agreement between the calculated results and
experimental results is usually very poor. Therefore, when considering threshold
pressure, Eq. (3.4), introduced in the discussion on interface energy, is more useful.
In addition, as we will discuss in the next chapter, the analysis of the infiltration of
molten metal using Darcy’s law provides more important information than is
obtained using the Hagen–Poiseuille equation.

References

1. Suzuki, S., Ochiai, M., Itoh, M., Shirayanagi, I., Awano, T.: Effect of high pressure applied
during the solidification of alloys (I), Al-Si binary system. Rep. Natl. Ind. Res. Inst. Nagoya 10,
299–307 (1961)
2. Suzuki, S., Ochiai, M., Shirayanagi, I., Itoh, M., Kurahashi, S., Awano, T.: Effect of high
pressure applied during the solidification of alloys (II), Al-Si binary system. Rep. Natl. Ind.
Res. Inst. Nagoya 11, 10–614 (1962)
References 65

3. Nishida, Y., Matsubara, H., Shirayanagi, I., Suzuki, S.: Fundamental study on the squeeze
casting. Bull. Jpn. Inst. Met. 19, 895–902 (1980)
4. Suzuki, S., Shirayanagi, I., Izawa, N., Nishida, Y., Ochiai, M.: 31st Fall Meeting of Japan
Institute of Light Metals, p. 23 (1966). Preprint
5. Suzuki, S., Nishida, Y., Shirayanagi, I., Izawa, N., Matsubara, H.: Segregation in aluminium
alloys solidified under high pressure. Aluminium 59, 544–546 (1983)
6. Imagawa, K., Nagata, S., Kitahara, A., Akiyama, S., Ueno, H.: 41st Fall Meeting of Japan
Institute of Light Metals, p. 1 (1971). Preprint. How to process shirasu balloon-metal compos-
ite. Jpn. Inst. Light Met. 23, 282–284 (1973)
7. Suzuki, S., Shirayanagi, I., Matsubara, H., Izawa, N., Kobayashi, N.: Aluminum/glass-fiber
composites by squeeze casting (i) (Fabrication condition). In: 52nd Spring Meeting of Japan
Institute of Light Metals, p. 3 (1977). Preprint
8. Nakata, E., Kagawa, Y.: Evaluation of the toughness of high volume fraction W/Al
composites. J. Mater. Sci. Lett. 3, 968–970 (1984)
9. Kagawa, Y., Oishi, Y., Yoshida, S., Nakata, E.: Workability of helical fiber reinforced
composite metal. J. Jpn. Soc. Compos. Mater. 7, 140–146 (1981)
10. Nakata, E., Kagawa, Y., Terao, H.: Fabrication and properties of tubular type composite by
squeeze casting method. J. Jpn. Soc. Compos. Mater. 9, 115–117 (1983)
11. Kagawa, Y.: Application of casting technology to fiber reinforced metals. Imono (J. Jpn.
Foundry Eng. Soc.) 58, 614–621 (1986)
12. Kagawa, Y.: Fiber Reinforced Metals, p. 147. CMC, Tokyo (1985) (in Japanese)
13. Towata, S., Yamada, S.: Interaction between SiC fibers and aluminum alloys. J. Jpn. Inst. Met.
47, 159–165 (1983)
14. Towata, S., Ikuno, H., Yamada, S.: Mechanical properties of carbon fiber-reinforced aluminum
alloys with whiskers and particulates of silicon-carbide. Trans. JIM 29, 314–321 (1988)
15. Nishida, Y., Imai, T., Yamada, M., Matsubara, H., Shirayanagi, I.: Fabrication of potassium
titanate whisker/aluminum composites and some their properties. J. Jpn. Inst. Light Met. 38,
515–521 (1988)
16. Matsubara, H., Nishida, Y., Shirayanagi, I., Yamada, M.: Fabrication of silicon nitride
whisker/aluminum alloy composites and some their properties. J. Jpn. Inst. Light Met. 39,
338–343 (1989)
17. Suganuma, K., Sasaki, G., Fujita, T., Suzuki, N.: Interfacial reaction between aluminum borate
whisker and AC8A and 6061 aluminum alloys. J. Jpn. Inst. Light Met. 41, 297–303 (1991)
18. Saito, N., Nakanishi, M., Nishida, Y.: Effect of heat treatment on the mechanical properties of
aluminum-borate whisker reinforced 6061 aluminum alloy. J. Jpn. Inst. Light Met. 44, 86–90
(1994)
19. Kim, J.-s., Sugamata, M., Kaneko, J.: Effect of hot extrusion on the mechanical properties of
SiC whisker/AZ91 magnesium alloy composites. J. Jpn. Inst. Met. 55, 521–528 (1991)
20. Kim, J.-s., Kaneko, J., Sugamata, M.: High temperature deformation of SiC whisker/AZ91
magnesium alloy and SiC whisker/2324 aluminum alloy composites. J. Jpn. Inst. Met. 56,
819–827 (1992)
21. Donomoto, T., Miura, N., Funatani, K., Miyake, N.: Ceramic fiber reinforced piston for high
performance diesel engines. SAE Paper No. 830252 (1983)
22. Hayashi, T., Ushio, H., Ebisawa, M.: The properties of hybrid fiber reinforced metal and its
application for engine block. SAE Paper No. 890557 (1989). Wear properties of hybrid fiber
reinforced aluminum matrix composites and application to an automotive engine block. J. Jpn.
Inst. Light Met. 40, 787–792 (1990)
23. Komatsubara, T., Okajima, M., Koyasukata, Y., Hoshino, H.: Sanyo Tech. Rev. 20, 107 (1988)
24. Clyne, T.W., Withers, P.J.: An Introduction to Metal Matrix Composites. Cambridge Univer-
sity Press, Cambridge (1993)
25. Clyne, T.W., Bader, M.G., Cappleman, G.R., Hubert, P.A.: The use of a d-alumina fibre for
metal-matrix composites. J. Mater. Sci. 20, 85–96 (1985)
66 3 Fabrication by Squeeze Casting

26. Lacoste, E., Aboulfatah, M., Danis, M., Girot, F.: Numerical simulation of the infiltration of
fibrous preforms by a pure metal. Metall. Trans. 24A, 2667–2678 (1993)
27. Carman, P.C.: Capillary rise and capillary movement of moisture in fine sands. Soil Sci. 52,
1–14 (1941)
28. White, L.R.: Capillary rise in powders. J. Colloid Interface Sci. 90, 536–538 (1982)
29. Mortensen, A., Cornie, J.A.: On the infiltration of metal matrix composites. Metall. Trans.
18A, 1160–1163 (1987)
30. Nakanishi, H., Tsunekawa, Y., Okumiya, M., Higashi, M., Niimi, I.: Influence of fiber array on
the threshold pressure of infiltration in alumina fiber/aluminum composite system. J. Jpn. Inst.
Light Met. 41, 325–330 (1991)
31. Nakanishi, H., Tsunekawa, Y., Okumiya, M., Niimi, I.: Influence of processing parameters on
the threshold pressure of infiltration in alumina fiber/aluminum composite system. J. Jpn. Inst.
Light Met. 41, 576–581 (1991)
32. Nakanishi, H., Tsunekawa, Y., Okumiya, M., Higashi, M., Niimi, I.: Threshold pressure for
infiltration in mica-ceramic particle/aluminum composite. J. Jpn. Inst. Light Met. 42, 92–97
(1992)
33. Nakanishi, H., Tsunekawa, Y., Okumiya, M., Mohri, N., Niimi, I., Satoh, M.: Ultrasonic
infiltration in alumina particle/molten aluminum system assisted by exothermic reaction of
titanium aluminide formation. J. Jpn. Inst. Met. 57, 81–87 (1993)
34. Oh, S.-Y., Cornie, J.A., Russell, K.C.: Wetting of ceramic particulates with liquid aluminum
alloys: Part I. Experimental techniques. Metall. Trans. 20A, 527–532 (1989)
35. Oh, S.-Y., Cornie, J.A., Russell, K.C.: Wetting of ceramic particulates with liquid aluminum
alloys: Part II. Study of wetting. Metall. Trans. 20A, 533–541 (1989)
36. Nagata, S., Matsuda, K.: Effects of particle preheating temperature on the length of metal-
particle composite in pressure casting. Imono (J. Jpn. Foundry Eng. Soc.) 53, 300–304 (1981)
37. Nagata, S., Matsuda, K.: Pressure casting conditions of metal-hybrid particle composites and
their applications. Imono 54, 657–663 (1982)
38. Nagata, S., Matsuda, K.: Effects of some factors on the critical preheating temperature of
particles in producing metal-particle composites by pressure casting. Imono 53, 686–691
(1981)
39. Nagata, S., Matsuda, K.: On the condition of the pressure infiltration method for metal
composites. Bull. Jpn. Inst. Met. 25, 1026–1033 (1986)
40. Nagata, S., Kitahara, A., Akiyama, S., Ueno, H.: Making metal composite by pressure casting.
AFS Trans. 85–08, 49–54 (1985)
Chapter 4
Theory of Pressure Infiltration

Abstract In this chapter, the theory of pressure infiltration into preforms, which
has been developed for squeeze casting, is discussed and compared with experi-
mental results. The infiltration theory is based on Darcy’s law, which was
introduced in Chap. 1. It allows us to predict most aspects of pressure infiltration:
the start point of preform deformation, the distribution of preform compression,
preform breakage, infiltration stop mechanism, infiltration stop point and the
influence of preheat temperature of the preform. In addition, the theory enables
us to easily determine suitable pressure infiltration conditions for the development
of new MMCs. The characteristic features of the pressure infiltration process are
also summarized in this chapter.

4.1 Macroscopic Analysis of Fluid Flow in Preforms

This analytical method for the infiltration of molten metal into a preform is based on
Darcy’s law, which we discussed in Chap. 1. The infiltration is analyzed by consid-
ering the permeability of the molten metal in a porous preform made of fibers or
particles, without modeling the shape of pores in the preform. This analysis is simple
and the results usually agree well with experimental data. Darcy’s law is originally
applicable to laminar flow at a steady state. However, experimental results show that
even at a non-steady state, it is applicable to the case of a flow with low velocity.
The application limit is that the Reynolds number is less than unity [1].
The one-dimensional differential equation of Darcy’s law at a non-steady state is
given by:

@p mu
¼ ; (4.1)
@x K

where u is the volume of fluid flowing per unit time and unit cross-sectional area of
porous media (i.e., it has the dimensions of velocity), p is the pressure of fluid in the

Y. Nishida, Introduction to Metal Matrix Composites: Fabrication and Recycling, 67


DOI 10.1007/978-4-431-54237-7_4, # Springer Japan 2013
68 4 Theory of Pressure Infiltration

infiltrated area, m is the viscosity, K is the permeability and x is the distance from the
top surface of the preform. p decreases as x increases, which is why the right-hand
side of Eq. (4.1) has a negative sign. In addition, as the flow of molten metal is
laminar flow, the continuity equation is applicable and given by:

@u
¼ 0: (4.2)
@x

Equation (4.2) shows that u differentiated with respect to x is always zero. This
equation means that u is independent of x and has the same value at any point of x;
in other words, u is only time-dependent.

4.1.1 Analysis on the Assumption of Constant Preform


Surface Pressure

The analysis of infiltration will be simplified if we use the assumption that a


constant set-up pressure is applied to the preform top surface. That is, the molten
metal pressure is constant throughout infiltration. As molten metal is considered to
be an incompressible fluid and the continuity equation, Eq. (4.2), is applicable, we
can regard Eq. (4.1) as an ordinary differential equation with respect to x at a fixed
time. In this case, the solution is:

mm u
p¼ x þ P0 ; (4.3)
Km

where P0 is the set-up pressure which is applied to the preform surface.


Alternatively, the velocity of the infiltration front is:

dxf u
¼ ; (4.4)
dt f

where xf is the location of the infiltration front at time t, and f (¼ 1–Vf) is the
porosity of the preform. The pressure p of equation (4.3) when x ¼ xf is the
threshold pressure Pc. Therefore, u is given by:

Km ðP0  Pc Þ
u¼ : (4.5)
mm xf

By substituting the above equation into Eq. (4.4), Eq. (4.4) becomes an ordinary
differential equation:
4.1 Macroscopic Analysis of Fluid Flow in Preforms 69

dxf Km ðP0  Pc Þ
¼ : (4.6)
dt fmm xf

The solution of Eq. (4.6) is:


sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2Km ðP0  Pc Þ
xf ¼ t: (4.7)
fmm

This equation gives the change of infiltration front with time.


By substituting Eq. (4.5) into Eq. (4.3), the pressure distribution curve of molten
metal in the infiltrated region at a fixed infiltration front is given by:

P0  Pc
p ¼ P0  x: (4.8)
xf

The pressure distribution is a straight line; the pressure decreases linearly from
the maximum pressure P0 at the top surface of the preform to the minimum pressure
Pc at the infiltration front.
In a slightly different approach, Fukunaga and Goda have analyzed the infiltra-
tion of molten metal into a preform for the condition that infiltration advances by
pushing out the air that is initially filling the preform [2–5]. Darcy’s law, Eq. (4.1),
was applied to both molten metal and air to give:

@pm m u
¼ m ð0  x  xf Þ (4.9)
@x Km

and

@pa mu
¼ a ðxf  x  LÞ; (4.9a)
@x Ka

where L is the thickness of the preform, and subscripts m and a indicate metal and
air, respectively. The analysis is based on the assumption that air is an incompress-
ible fluid and threshold pressure is negligible. The boundary conditions used were:

p m ¼ P0 ðx ¼ 0Þ (4.10)

pm ¼ pa ðx ¼ xf ðtÞÞ; (4.10a)

Km @pm Ka @pa
¼ ðx ¼ xf ðtÞÞ; (4.10b)
mm @x ma @x

and

pa ¼ P1 ðx ¼ LÞ; (4.10c)
70 4 Theory of Pressure Infiltration

where P0 and P1 are the pressures at the preform surface and the atmosphere,
respectively, and are constants. At a fixed time, we can regard Eqs. (4.9) and (4.9a)
as ordinary differential equations with respect to x, and solve them to give:

mm u
pm ¼ P 0  x; (4.11)
Km

and

ma u
pa ¼ P1  xðx  LÞ: (4.12)
Ka

Using these solutions and the above boundary conditions, the molten metal
pressure pm at location x in the preform is:

ðP0  P1 Þ
p m ¼ P0  x (4.13)
ð1  mÞxf þ mL

Km ma
m¼ : (4.13a)
Ka mm

As pm is a linear function of x, it is apparent that the pressure of molten metal in


the preform decreases linearly from the preform surface to the infiltration front.
Eq. (4.4) gives the relationship between the velocity of the infiltration front and u.
Combining Eqs. (4.11), (4.12) and (4.4) at x ¼ xf (t) gives:

dxf Km ðP0  P1 Þ
¼ ; (4.14)
dt fmm ð1  mÞxf þ mL

which is a differential equation relating xf and t. The solution to Eq. (4.14) is:

x2f Km ðP0  P1 Þ
ð1  mÞ þ mLxf ¼ t (4.15)
2 fmm

This equation gives the relationship between the location of the infiltration front
and time. If we regard this equation as a quadratic equation with respect to xf, the
solution in the case where (1–m) > 0 becomes:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 
mL mL 2 2Km ðP0  P1 Þ
xf ¼  þ þ t: (4.16)
1m 1m fmm ð1  mÞ

In addition, at any specific time, the pressure distribution from the preform
surface to the infiltration front x ¼ xf is given by Eq. (4.13).
4.1 Macroscopic Analysis of Fluid Flow in Preforms 71

Fig. 4.1 Pressure distributions and variations in Reynolds number (Re) with advancing infiltration
front (on the assumption that a constant preform surface pressure of 100 kPa is given initially) for
two different volume fractions of reinforcement

We discussed the infiltration for the case where a constant preform surface
pressure is supplied initially and does not change with time. When the volume
fraction of the preform Vf is large and the infiltration velocity is low, the results
calculated using the above model agree roughly with experimental results. How-
ever, for the initial stages of infiltration, when xf is very small and near zero,
Eqs. (4.6) and (4.14) give extremely high infiltration velocities.
As shown by Eq. (1.46), the Reynolds number is proportional to the fluid flow
velocity. Thus, the Reynolds number of the above model, assuming constant
preform surface pressure, is very large at the beginning of infiltration (at small
xf), although Darcy’s law only applies when the Reynolds number is less than unity.
Therefore, the above model does not describe real phenomena at the beginning of
infiltration, but, as will now be shown, is approximately applicable to the cases of
high Vf.
The pressure distribution in a preform as calculated by Eqs. (4.7) and (4.8) is
shown in Fig. 4.1. The constant preform surface pressure is 100 kPa (about 1
atmosphere) and the volume fractions of fibers are 6% and 25%. In this case, the
threshold pressure Pc is ignored. The Reynolds numbers Re for Vf ¼ 6% are 260
when the infiltration front, xf, is at 10 mm, and 27 when it is at 100 mm. These
Reynolds numbers are far beyond the Re < 1 application limit of Darcy’s law.
However, the broken lines in the same figure correspond to Vf ¼ 6%. In this case,
the Reynolds numbers are 0.25 for t ¼ 3.1 (xf ¼ 30 mm) and 0.15 for t ¼ 8.5
(xf ¼ 50 mm). Here, the infiltration velocity is not high, and these Reynolds
numbers satisfy the application limit of Re < 1. Therefore, Darcy’s law is almost
applicable to preforms with this volume fraction, except for at the beginning of
infiltration.
72 4 Theory of Pressure Infiltration

4.1.2 Infiltration Velocity Model

The analysis based on the assumption of constant preform surface pressure does not
satisfy the application limit of Darcy’s law in many cases, especially when Vf is
low. If we assume constant preform surface pressure, the compressive deformation
of the preform should occur instantaneously and homogeneously before infiltration
starts. However, experimentally obtained samples show quite different results. An
alternative analysis using the infiltration velocity model solves these problems and
satisfies the application limit of Darcy’s law.
For production of composites in industry by the pressure infiltration method, a
mechanical punch or compressed gas is usually used to apply high pressure to the
molten metal. The pressure of the molten metal begins to rise from zero when
pressure is transferred to molten metal by one of these means. In the case of low Vf,
infiltration starts just after the pressure overcomes the threshold pressure of the
preform. For a thin preform (a couple of centimeters thick), infiltration finishes
before the pressure of the molten metal reaches the set-up pressure. The pressure
change of molten metal with time for a thin preform is shown schematically in
Fig. 4.2a. The pressure distribution in the infiltrated preform is also shown in
Fig. 4.2b.
The real pressure change was measured by the equipment shown schematically
in Fig. 4.3. The pressure change of the molten metal is shown in Fig. 4.4. These
curves were measured for the infiltration of molten pure aluminum into a preform
made of 15 vol.% SiC whiskers. The punch speed was constant at 10 mm/s and the
force (pressure) of the molten metal in equilibrium with the punch was measured.
As predicted in Fig. 4.2, the pressure of molten metal began to rise from zero and a
small curvature change is apparent near the start point of the pressure curve. The
curvature change point corresponds to the threshold pressure. The infiltration end
point is the point where the pressure starts to increase linearly (at around t ¼ 6 s on
the figure).
In addition, pressure curves for the infiltration of molten pure aluminum without
a preform and with preforms containing different volume fractions of SiC whiskers
are shown in Fig. 4.5. As predicted by Eq. (4.4), these experimental curves show
that the threshold pressure increases as Vf increases. When the pressure curves
become parallel to the curve without the preform, infiltration has finished. There-
fore, it has been experimentally shown that the infiltration end pressure increases as
Vf increases.

4.1.2.1 Location of Infiltration Front and Preform Surface Pressure

We need to obtain an analytical solution which can explain all the experimental
results mentioned above while satisfying the Reynolds number of less than unity
throughout infiltration condition needed for Darcy’s law to be applicable.
Yamauchi and Nishida introduced an analytical solution for infiltration by making
4.1 Macroscopic Analysis of Fluid Flow in Preforms 73

Pressure

Time

b
Deformation start point
Pressure

Infiltration distance, xf

Fig. 4.2 (a) Change of pressure applied to preform surface with time, and (b) pressure distribution
in the infiltrated region. ts: start point of infiltration; te: end point of infiltration; tmax: arrival point
of set-up pressure; Pmax: set-up pressure; Pc: threshold pressure

Oil pump

Pressure controller
Speed controller
Cylinder Pressure senser

Printer
Data recorder
Punch Displacement
senser
Mould
Molten metal

Preform
Heater

Fig. 4.3 Schematic of equipment used for pressure infiltration process


74 4 Theory of Pressure Infiltration

Fig. 4.4 Experimentally


obtained punch displacement,
punch speed and pressure Pressure

Displacement, L [mm]
applied to a preform made

Velocity, V [mm·s-1]

Pressure, P [MPa]
of 15 vol.% SiC whiskers
during the infiltration of
pure aluminum (set-up
punch speed: 10 mm/s) Punch displacement

Punch velocity

start

Time, t [s]

Fig. 4.5 Experimentally


No preform Vf:30% Vf:20% Vf:15%
obtained pressure curves
for the infiltration of molten
pure aluminum without
Pressure, P [MPa]

preform and with preforms


made of different volume
fractions of SiC whiskers

Pure Al

Time, t [s]

u in Eqs. (4.1), (4.2) and (4.4) a known function of time. They demonstrated that
almost all experimental results can be explained by this infiltration velocity model
[6–9]. In their discussion, the following assumptions were made:
1. The pressure which will be applied on the top surface of the preform begins to
rise from zero.
2. The compressive stress which acts on the uninfiltrated area is equal to the
pressure applied to the top surface of the preform.
3. Until the applied pressure reaches the compressive strength of the preform,
compressive deformation does not occur. When the pressure exceeds the com-
pressive strength, deformation begins.
4. Further deformation of the preform does not occur in the region which has
already been infiltrated.
5. The back pressure of air is ignored to simplify the calculation; and
6. The thickness of the preform is infinite.
4.1 Macroscopic Analysis of Fluid Flow in Preforms 75

We express Ps as the pressure applied to the top surface of the preform. Then, we
can regard Eq. (4.1) as an ordinary differential equation with respect to x at a fixed
time. The solution is given by:

mm u
p¼ x þ Ps ; (4.17)
K0

where K0 is the initial permeability of the preform. In this discussion, we consider


the compressive deformation of the preform. If compressive deformation occurs,
the permeability of the preform decreases. As the pressure at the infiltration front
x ¼ xf is the threshold pressure Pc, Eq. (4.17) becomes:

mm u
Ps ¼ x f þ Pc : (4.18)
K0

Alternatively, by integrating Eq. (4.4), the infiltration front xf is given by:


ðt
1
xf ¼ udt; (4.19)
1  Vf 0 0

where Vf 0 is the volume fraction of the preform before compressive deformation occurs.
By substituting Eq. (4.19) into Eq. (4.18), the preform surface pressure Ps is given by:
ðt
mm u
Ps ¼ udt þ Pc : (4.20)
K0 ð1  Vf 0 Þ 0

Thus, if u is given as a function of time, the preform surface pressure and the
location of the infiltration front can be simply expressed using u. Pc is given by
Eq. (3.4). Af in Eq. (3.4) takes different values depending on the shape of reinforce-
ment (fibers or particles) as discussed before.
This may be illustrated using a worked example: a constant punch velocity is
usually employed for pressure infiltration by mechanical methods. If we assume u
to be a constant u0, Eqs. (4.19) and (4.20) become:

u0 t
xf ¼ ; (4.21)
1  Vf 0

and

mm u20 t
Ps ¼ þ Pc : (4.22)
K0 ð1  Vf 0 Þ

Apparently, xf and Ps are simply proportional to time t. This corresponds to the


assumption where the pressure curve Ps between times ts and te in Fig. 4.2a is linear.
76 4 Theory of Pressure Infiltration

Fig. 4.6 Molten metal Liquid metal Fibrous preform


pressure Pm in the preform
during infiltration and stress
Pf applied to the fibers in the
preform (Pm + Pf is a
constant)

front

Pressure in preform
before deformation

after deformation

start point of deformation

Next, we discuss the compressive deformation of the preform. A schematic drawing


of force balance in this case is shown in Fig. 4.6. Pressure Ps acts on the top surface of
the preform, and the same pressure Ps acts on the bottom of the preform in the opposite
direction, as shown in the same figure. Before deformation of the preform begins, the
pressure Pm of molten metal in the preform at a fixed time decreases linearly from the
top surface of the preform to the infiltration front. As the threshold pressure is ignored
in this figure, the pressure at infiltration front xf is zero. However, the distribution of the
stress Pf applied to the fibers is the opposite to that of Pm, as shown in the figure. The
relationship between the pressure Pm and the stress Pf is:

Ps ¼ Pm þ Pf : (4.23)

Therefore, depending on the preform surface pressure, compressive deformation


of the preform occurs homogeneously in the uninfiltrated region, (i.e., between the
infiltration front and the bottom of the preform). If the deformation begins when
the infiltration front xf arrives at x1 at time t1, then the location x1 can be calculated
from Eq. (4.19):
ð t1
1
x1 ¼ udt: (4.24)
1  Vf 0 0

The preform surface pressure at time t1, Ps0, is given from Eq. (4.20):
ð t1
mm u
Ps0 ¼ udt þ Pc : (4.25)
K0 ð1  Vf 0 Þ 0
4.1 Macroscopic Analysis of Fluid Flow in Preforms 77

When the preform surface pressure exceeds Ps0, deformation begins and Vf
increases with increasing preform surface pressure Ps. Because the permeability
K is a function of Vf, K is also a function of pressure. If we choose x1 to be the fixed
location where deformation initiates, and make P1 the time-dependent pressure at
this location, then, from Eq. (4.17) we obtain:

mm u
P s ¼ P1 þ x1 : (4.26)
K0

After deformation starts, by integrating Eq. (4.1) from x ¼ x1 to x ¼ xf at time t,


we obtain:
ð xf
dx
Pc  P1 ¼ mm u : (4.27)
x1 K

The pressure at x ¼ xf is the threshold pressure Pc. In addition, by integrating


Eq. (4.4) from x1 to xf and dividing by K and using Vf (a function of time) instead of
Vf 0, Eq. (4.4) becomes:
ð xf ðt
dx u
¼ dt: (4.28)
x1 K t1 Kð1  Vf Þ

Therefore, Eq. (4.27) is rewritten as:


ðt
u
P1 ¼ mm u dt þ Pc : (4.29)
t1 Kð1  Vf Þ

By substituting Eqs. (4.24) and (4.29) into Eq. (4.26), we obtain the preform
surface pressure:
ð t1 ðt
mm u u
Ps ¼ udtþmm u dt þ Pc : (4.30)
K0 ð1  Vf 0 Þ 0 t1 Kð1  Vf Þ

It is important to note that when the deformation of the preform starts, Vf in the
equation for the threshold pressure [Eq. (3.4)] begins to increase, and Pc also begins
to increase. Further, xf is given by:
ð t1 ðt
1 u
xf ¼ udtþ dt: (4.31)
1  Vf 0 0 t1 1  Vf

To calculate real values of Ps and xf, we need to obtain the relationship between
the compressive stress applied to a preform and Vf, and also the relationship
between Vf and the permeability of the preform. The experimentally obtained
compressive stress versus Vf curve for SiC whiskers is shown in Fig. 4.7 [6]. The
78 4 Theory of Pressure Infiltration

Fig. 4.7 Relationship between applied stress and volume fraction Vf of preforms when compres-
sive stress is applied to SiC whisker preforms

Fig. 4.8 Relationship between applied stress and volume fraction Vf of preforms when compres-
sive stress is applied to preforms made of aluminum borate whiskers or alumina short fibers

data were obtained by simple compression tests of SiC whisker preforms.


The curves for aluminum borate whiskers and alumina short fibers are shown in
Fig. 4.8 [8]. For example, the curve for SiC whiskers shows that a compressive
stress of 20 MPa is needed to produce a dense SiC whisker preform with Vf ¼ 30%.
Therefore, for the case of a Vf ¼ 30% SiC whisker preform, compressive deforma-
tion does not occur when the molten metal pressure Ps is lower than 20 MPa. When
Ps exceeds 20 MPa, compressive deformation begins, and Vf increases as described
by the stress versus Vf curve. The stress versus Vf curves for aluminum borate
whiskers and alumina short fibers in Fig. 4.8 can be understood in the same way.
4.1 Macroscopic Analysis of Fluid Flow in Preforms 79

Fig. 4.9 Experimentally


obtained variation in hardness
of composites measured
along infiltration distance

Fig. 4.10 Hardness change


with Vf for SiC whisker
reinforced aluminum
composites

To test the theory described above, pure aluminum was poured and infiltrated
into SiC whisker preforms with initial Vf ¼ 15%, 20% and 30%, at a punch speed of
10 mm/s, and composites were produced. The preheat temperature of preform,
pouring temperature and mold temperature were 1,023 K, 1,073 K and 523 K,
respectively. Figure 4.9 gives the distributions of Vickers hardness which were
measured along the infiltration direction from the center of the top surface to the
bottom of the preforms [6]. In the Vf ¼ 30% case, the hardness does not change up
to a depth of 23 mm, and then begins to rise from this point. This result shows that
the compressive deformation began at a depth of 23 mm. For the Vf ¼ 20% and
15% preforms, deformation began at depth of 15 and 0 mm, respectively, from the
top surface. Figure 4.10 shows the relationship between Vickers hardness and Vf,
measured for SiC whiskers reinforced pure aluminum composite. The hardness
increased linearly with increasing Vf of SiC whiskers within this range measured.
These hardness values were then used to determine Vf across a range of samples
with different starting Vf values, and the relationship between Vf and infiltration
distance is shown in Fig. 4.11. Solid lines correspond to experimentally obtained
values and dotted lines to results obtained theoretically. This figure shows that the
80 4 Theory of Pressure Infiltration

Fig. 4.11 Change in volume fraction Vf because of compressive deformation of SiC whisker
preforms. Initial Vf values of: 15 %, 20 % and 30 %

Fig. 4.12 Volume fraction Vf change versus infiltration distance for preforms made of aluminum
borate whiskers or alumina short fibers

theoretical results agree quite well with experimental results. In particular, the
agreement of the depths at which deformation begins is very good.
In a similar way, pure aluminum was infiltrated into aluminum borate whisker
and alumina short fiber preforms with a punch speed of 10 mm/s, and the Vickers
hardness values of the resulting composites were measured from the top surface of
the preforms to the bottom. The hardness values were again used to determine Vf,
and the relationships between Vf and infiltration distance are plotted in Fig. 4.12 [8].
Figure 4.13 gives the relationships between hardness and Vf for composites
reinforced with aluminum borate whiskers and alumina short fibers. Figure 4.12
shows that the deformation begins about 20 mm from the top surface of the
4.1 Macroscopic Analysis of Fluid Flow in Preforms 81

Fig. 4.13 Linear relationship between hardness and Vf for composites made using preforms
containing aluminum borate whiskers or alumina short fibers

Fig. 4.14 Volume fraction Vf versus infiltration distance for different punch speeds. Initial
Vf ¼ 15% SiC whiskers

aluminum borate whisker preforms for Vf ¼ 30%, 14 mm for Vf ¼ 25% and 12 mm


for Vf ¼ 21%. However, because the hardness for the composite reinforced with
alumina short fibers of Vf ¼ 22% is almost constant, the deformation start point
does not appear within this thickness of preform. The theoretical curves (dotted
lines) are also shown in Fig. 4.12. The theoretical curves agree reasonably well with
experimental ones for all cases, demonstrating that the theory discussed above can
usefully predict real phenomena.
The infiltration theory also shows that if the infiltration velocity is high, the
deformation of the preform will be large. Then, the influence was examined experi-
mentally, and the results are shown in Fig. 4.14. An SiC whisker preform with
Vf ¼ 15% was infiltrated with pure Al at punch speeds of 10, 20 and 40 mm/s. For
punch speeds of 10 and 20 mm/s, the theoretically calculated lines agree very well
with experimental plots. However, the agreement between theoretical line and
experimental result for punch speed of 40 mm/s is poor. This is because although
the set-up punch speed was 40 mm/s, the actual punch speed was slower and changed
with time. This speed is too high for industrial application and was beyond the
capability of the equipment.
82 4 Theory of Pressure Infiltration

We need to obtain the relationship between punch speed and infiltration velocity
u. Generally, the cross-section of the preform is made to be slightly smaller than
that of the punch, as shown in Fig. 4.3. The inflow velocity of molten metal u is
usually not constant. Thus, a real curve of u is obtained as follows:

up Sr
u¼  D; (4.32)
Sp

where up ¼ punch speed obtained by experiments, Sr ¼ cross-sectional area of the


punch, Sp ¼ preform top surface area and D ¼ velocity of the decrease in preform
thickness arising from compressive deformation. We assume that all molten metal
enters the top surface of the preform and that there is no inflow from the side surface
of the preform.
To calculate infiltration phenomena using Eqs. (4.30) and (4.31), we need to
know the relationship between permeability K and Vf. There are many theoretical
equations for the permeability of porous media [10]. Among them, useful equations
are Langmuir’s Equation (4.33) for flow parallel to fiber alignment and Happel’s
equation (4.34) for flow normal to the fibers. Supposing that the fibers are
distributed randomly in the preform, we can estimate the permeability of the
preform and obtain good agreement with experimental results by averaging the
values from both equations:
 
K 1 3 1 2
¼  ln Vf  þ 2Vf  Vf ; (4.33)
R2 4Vf 2 2
!
K 1 Vf2  1
¼  ln Vf þ ; (4.34)
R 2 8Vf Vf2 þ 1

where R is the radius of the fiber.

4.1.2.2 Infiltration Involving Solidification

When the preheat temperature of a preform is lower than the freezing temperature
of a metal (pure metal in this case) and the molten metal at its freezing tempe-
rature contacts fibers, some portion of the metal solidifies and deposits on the
surface of the fibers. If the molten metal is poured at its melting point, as discussed
in Sect. 3.3 in Chap. 3, heat exchange between the molten metal and the fibers will
be given by Eq. (3.10). In this case, the fraction of solid, apparent diameter of fibers
and apparent volume fraction of fibers are given by Eqs. (3.11), (3.14) and (3.15),
respectively. Based on this model, we will now obtain the infiltration theory which
gives the distribution of molten metal pressure in the preform, the top surface
pressure of the preform and the location of the infiltration front.
4.1 Macroscopic Analysis of Fluid Flow in Preforms 83

First, by regarding Eq. (4.1) as an ordinary differential equation at a fixed time,


integration of Eq. (4.1) with respect to x gives:
mm u
p¼ x þ Ps ; (4.35)
K00

where K00 is the apparent permeability of the preform after some portion of molten
metal has solidified and deposited on the fiber surface, resulting in an increase in the
effective fiber diameter. Ps is the preform top surface pressure. This equation gives
the pressure distribution of the molten metal in the infiltrated region in a preform at
a fixed time. If uF out of the molten metal inflow u solidifies at the infiltration front,
the velocity of the infiltration front xf will be given by:

dxf uð1  FÞ
¼ : (4.36)
dt 1  Vf0 0

Therefore, xf as a function of time is given by:


ðt
1F
xf ¼ udt; (4.37)
1  Vf 0 0 0

where Vf 00 is the apparent Vf after the molten metal has contacted the fibers and the
fibers have become covered with some solidified metal.
Alternatively, as the molten metal pressure at the infiltration front x ¼ xf is Pc,
from Eq. (4.35), the preform surface pressure Ps is given by:
ðt
m uð1  FÞ 4Vf 0 g cos y
Ps ¼ 0m udt : (4.38)
K0 ð1  Vf 0 0 Þ 0 df ð1  Vf 0 Þ

The second term of the right-hand side of this equation is threshold pressure Pc.
Vf 0 is the initial Vf before compressive deformation occurs. Parameters related to
infiltration before compressive deformation occurs: the location of the infiltration
front xf, the preform surface pressure Ps and the pressure distribution of molten
metal in infiltrated region are given by Eqs. (4.37), (4.38) and (4.35), respectively.
The depth at which the compressive deformation starts at the time t1 is denoted
by x1, and is:
ð t1
1F
x1 ¼ udt: (4.39)
1  Vf 0 0 0

If Ps1 is the preform surface pressure at time t1, Eq. (4.38) becomes:
ð t1
m uð1  FÞ 4Vf 0 g cos y
Ps1 ¼ 0m udt : (4.40)
K0 ð1  Vf 0 0 Þ 0 df ð1  Vf 0 Þ
84 4 Theory of Pressure Infiltration

When the preform surface pressure exceeds Ps1, compressive deformation


occurs. For a preform made from SiC whiskers, Vf increases along the curve of
pressure (stress) shown in Fig. 4.7. The permeability K of a preform is also a
function of Vf, and therefore a function of Ps. If P1 is the time-dependent pressure at
x1, where deformation began, then the preform surface pressure Ps after deforma-
tion begins is:

mm ux1
P s ¼ P1 þ : (4.41)
K00

After the start of deformation, integration of Eq. (4.1) between x1 and xf gives:
ð xf
1
Pc  P1 ¼ mm u dx: (4.42)
x1 K

The pressure at xf is always Pc. By dividing Eq. (4.36) by K and integrating from
x1 to xf, Eq. (4.36) becomes:
ð xf ðt
1 uð1  FÞ
dx ¼ dt: (4.43)
x1 K t1 Kð1  Vf 0 Þ

In this equation, Vf0 is used instead of Vf 00 , because deformation has already


started. Therefore, Eq. (4.42) can be rewritten as follows:
ðt
uð1  FÞ
P1 ¼ mm u 0 dt þ Pc : (4.44)
t1 Kð1  Vf Þ

Substitution of Eqs. (4.39) and (4.44) into Eq. (4.41) yields:


ð t1 ðt
m uð1  FÞ uð1  FÞ 4Vf g cos y
Ps ¼ 0m udtþmm u 0 dt  : (4.45)
K0 ð1  Vf 0 0 Þ 0 t1 Kð1  Vf Þ df ð1  Vf Þ

However, Pc begins to increase when compressive deformation starts, because Vf


also begins to increase. Integrating Eq. (4.36) from t1 to t gives:
ð t1 ðt
1F uð1  FÞ
xf ¼ udt þ 0 dt: (4.46)
1  Vf 0 0 0 t1 1  Vf

In Eqs. (4.45) and (4.46), K and Vf are functions of time after deformation starts
(after t1) and are very complicated. Therefore numerical integration must be used.
To verify the theory discussed above, further experiments were carried out.
Figure 4.15 shows the appearance of 20 vol.% SiC whisker reinforced aluminum
composites which were infiltrated at preform preheat temperatures of 933, 673, 473
and 303 K (room temperature) at a pouring temperature of 933 K (the melting point
4.1 Macroscopic Analysis of Fluid Flow in Preforms 85

Fig. 4.15 Appearance of SiC whisker reinforced composites showing the relationship between
preheat temperature and infiltration distance

of aluminum) with a punch speed of 10 mm/s. The darker parts in the images are the
composite. In the case of the 933 K preheat temperature, the infiltrated molten
aluminum reached the bottom of the 40 mm thick preform. However, when the
preheat temperature was below the melting point of aluminum, infiltration stopped
before the aluminum arrived at the bottom of the preform. Infiltration barely
occurred in the case of the room temperature preform. In the photos of the 473
and 303 K samples in Fig. 4.15, the uninfiltrated parts of the preforms were broken
and lost during polishing of the samples.
The Vickers hardness of these composites was measured along the infiltration
direction from the center of the top surface to the bottom of preform. The hardness
values were used to determine Vf, and the resulting relationship between Vf and
infiltration distance is shown in Fig. 4.16. The Vf near the top surface of the preforms
is almost the same value for all preheat temperatures. However, Vf begins to rise at a
certain depth, which increases with increasing preheat temperature. For the sample
where the preform is at room temperature, Vf rises abruptly from the top surface of
the preform. This result shows that the preform was compressed from the top
surface. The rise of Vf becomes smaller with increasing preheat temperature. How-
ever, even if the preheat temperature is the melting point of aluminum, Vf begins to
rise from a certain depth during infiltration. This means that compressive deforma-
tion occurs even when the preform is heated to the melting point of aluminum.
86 4 Theory of Pressure Infiltration

Fig. 4.16 Vf curves for different preheat temperatures. Initial Vf ¼ 20 % SiC whiskers

Fig. 4.17 Vf curves for different preheat temperatures. Initial Vf ¼ 25 % aluminum borate
whiskers

The values of Vf as a function of distance calculated using Eqs. (4.45) and (4.46)
are also plotted on the same figure (the dotted lines). The calculation used the
relationship of Vf versus pressure shown in Fig. 4.7. Assuming that the whiskers are
distributed randomly in the preforms, the average value of permeability was
obtained from Eqs. (4.33) and (4.34). The theoretically obtained deformation start
points and the gradients of Vf curves agree well with the experimental values for
each preheat temperature.
Another example is shown in Fig. 4.17. These composite samples were produced
by infiltrating molten aluminum at its melting point into preforms made of alumi-
num borate whiskers at preheat temperatures of 933, 673, 473 and 303 K.
4.1 Macroscopic Analysis of Fluid Flow in Preforms 87

The hardness was measured and used to determine Vf. These results are very similar
to those obtained for the SiC whisker reinforced composites. For a preheat temper-
ature of 303 K (room temperature), infiltration barely advanced, and Vf increased
steeply from the top surface of the preform. At higher preheat temperatures,
infiltration distance increases and the gradient of the Vf curves becomes less
steep. The values calculated using Eqs. (4.45) and (4.46) are also plotted as dotted
lines in Fig. 4.17. In this case, the theoretical values also agree with the experimen-
tal results. Especially, at low preheat temperatures, the start points of compression
and the gradients of the increase in Vf coincide well with the experimental curves.
These results demonstrate that the infiltration theory discussed in this section can
predict essential phenomena of infiltration in compressible fibrous preforms.
Equations (4.45) and (4.46) were introduced for the case of pure metals. How-
ever, these equations are applicable to alloys when the solidification range is not
particularly wide. Molten Al—12 wt.% Si alloy was infiltrated into SiC whisker
preforms of initial Vf ¼ 15%, 20% and 30% by the pressure infiltration process [7].
The curves for the preform surface pressure Ps obtained experimentally were very
similar as those of pure aluminum. The hardness of the composites was measured
along the infiltration direction in the same manner. However, the measured hard-
ness values were influenced by Si crystals that had precipitated in the preform, and
therefore the hardness values were slightly more scattered.
As discussed here, if we regard infiltration velocity u as a known function, we
can explain theoretically the infiltration phenomena. In addition, the Reynolds
number obtained using Eq. (1.46) is always less than unity at the punch speeds
used in industry.
Alternatively, the permeability of preforms is needed for the theoretical
calculations. As discussed earlier, Eqs. (4.33) and (4.34) are available. To confirm
the permeability values obtained using these equations, experiments to directly
measure the permeability of preforms were carried out using air instead of molten
metal as the fluid [11]. The experimentally obtained permeabilities agreed well with
the values from Eqs. (4.33) and (4.34). These results show that measurement using
air is a simple and useful method to obtain permeability. For industrial applications,
whisker granules with a diameter of about 50 mm have been developed by
Yamauchi et al. to improve the permeability of preforms [12]. If preforms are
made using these granules, large channels are formed between the granules, and
permeability increases. By this technique, infiltration becomes very easy and
deformation or breakage of the preform during infiltration is effectively prevented,
although the tensile strength of the composites decreases slightly.
In contrast, if we use very fine fibers such as carbon nanotubes or carbon
nanofibers, the threshold pressure Pc of the preform will be very high. For example,
when the matrix metal, fiber, gla, contact angle and Vf are molten aluminum, carbon
fiber (diameter 10 mm), 0.914 J/m2, 127 and 0.25, respectively, the threshold
pressure Pc obtained is 0.072 MPa as calculated using Eq. (3.4). However, the
calculated threshold pressure when carbon nanotubes (diameter of 50 nm) are used
is 14.4 MPa, which is 200 times higher than for the carbon fiber. This result was
confirmed experimentally by Uozumi et al. [13].
88 4 Theory of Pressure Infiltration

The thermal conductivity of a carbon nanotube along its axis is very high; a value
of about 3,500 W/(m K) was obtained by Pop et al. [14]. However, there is thermal
resistance at the interface between the nanotube and the matrix metal. Usually, the
total area of the interface between the fibers and the matrix increases with decreasing
radius of the fiber if Vf is constant. Therefore, the total interface area for nanotubes in
a unit volume of matrix metal will be very large compared with that of carbon fibers
at constant Vf. This makes it extremely difficult to fabricate a composite with high
thermal conductivity using carbon nanotubes. One way of dealing with this problem
may be to use very long nanotubes aligned in one direction.

4.2 Infiltration Stop Mechanism

We have often observed that the infiltration of the molten metal stopped before
reaching the bottom of the preform, even when using squeeze casting. Fukunaga
et al. tried to explain the stop mechanism. They insisted that infiltration stopped
because no heat was supplied to the infiltration front [4]. Mortensen et al.
investigated theoretically the infiltration phenomena, including the stop mecha-
nism, using long cylindrical model preforms containing 25 vol.% alumina fibers,
and taking into account the radial heat flow to the mold. However, their theory
ignored the compressive deformation of preforms during infiltration. They also
performed experiments to verify their theory [15–17]. Their theory may well be
applicable to their situation of long, thin cylindrical preforms containing 25 vol.%
alumina fiber at a maximum infiltration pressure of 3.5 MPa supplied by gas
pressure. However, because wide and thin preforms are most often used in industry,
heat transfer in the radial direction or toward the side wall is not important.
In industry, a pressure of 50–100 MPa is usually applied to the molten metal to
produce composites using the pressure infiltration process. At these pressures,
deformation of the preform occurs and Vf increases. This is the main infiltration
stop mechanism, and is described below. When the preheat temperature of a
preform is lower than the freezing temperature of the metal, solid metal deposits
on the surface of fibers and the fibers grow. This increases both the apparent
diameter of fibers and the apparent Vf, and decreases the permeability of the
preform. Therefore, to continue infiltration, we need to apply higher pressure to
the molten metal. However, as the surface pressure of the preform is increased,
compressive deformation of the preform continues, and the permeability continues
to decrease. Finally, the applied pressure reaches the set-up pressure and infiltration
stops. This is the infiltration stop mechanism when the preheat temperature is lower
than the freezing temperature of a metal. Even if the preheat temperature is higher
than the freezing temperature, as can be seen in Fig. 4.16, as infiltration proceeds,
compressive deformation continues and Vf increases. This increase in Vf causes an
increase in Pc. When Pc becomes higher than the set-up pressure, infiltration stops.
These phenomena can be explained by the infiltration velocity model [Eqs. (4.45)
and (4.46)].
References 89

4.3 Characteristics of Pressure Infiltration Method


and Products

Fabrication techniques for metal matrix composites have been extensively devel-
oped, and it is well known that MMCs have excellent properties which can meet
many application requirements. The number of commercialized examples, however,
is small, because of the high cost of the products. The high cost of these products is
caused by the high price of reinforcements and high production costs. Casting is an
excellent way to reduce the fabrication costs. In addition, the use of smaller amounts
of cheaper reinforcements will also effectively reduce the cost of products. Overall,
if the composite is only placed where required in the metal product, the product costs
will be lower. We call such reinforcing “partial reinforcement”. Partial reinforce-
ment works very well, as shown in Figs. 1.2, 1.3 and 1.4, and most commercialized
products are composites with partial reinforcement [18–20]. Casting processes
which allow lower cost production of partially reinforced composites are squeeze
casting and die casting [21]. Any other casting process which can generate some
hydrostatic pressure may be able to be used for the pressure infiltration process. The
benefits of using pressure within the pressure infiltration process are:
1. It is easy to produce fiber or particulate reinforced composites even if the
reinforcement is barely wet by the molten metal.
2. Chemical reaction between reinforcements and molten metal can be suppressed,
because the freezing time of the molten metal is short under high pressure.
3. Because the molten metal solidifies rapidly under high pressure, the microstruc-
ture of the matrix metal is very fine.
4. It is possible to produce composite products which contain reinforcement only in
parts of the products.
5. As the production of the composite and forming of the article are performed at
the same time, the production process is more economical; and
6. Composite products can be produced at near-net shape.

References

1. Collins, R.E.: Flow of Fluids Through Porous Materials, p. 47. Reinhold Publishing Corp.,
New York (1961)
2. Fukunaga, H., Goda, K.: The study on fabrication process of fiber reinforced metal by squeeze
casting. J. Jpn. Soc. Mech. Eng. (C) 49(1808–1815) (1983)
3. Fukunaga, H., Goda, K.: Experimental study on fabrication of SiC fiber reinforced aluminum
by squeeze casting. J. Jpn. Soc. Mech. Eng. (C) 47(1207–1215) (1981)
4. Fukunaga, H., Goda, K.: Formation and role of the solidified layer on a fiber during the
fabrication of fiber reinforced metal by the liquid process. J. Jpn. Inst. Met. 49, 78–83 (1985)
5. Fukunaga, H., Goda, K.: Squeeze casting process for fiber reinforced metal and its mechanical
properties. J. Jpn. Soc. Compos. Mater. 12, 198–205 (1986)
90 4 Theory of Pressure Infiltration

6. Yamauchi, T., Nishida, Y.: Infiltration kinetics of aluminum into fibrous preform. J. Jpn. Inst.
Met. 58, 552–558 (1994)
7. Yamauchi, T., Nishida, Y.: Infiltration kinetics of Al-12%Si alloy into SiC whisker preform.
J. Jpn. Inst. Met. 58, 1436–1443 (1994)
8. Yamauchi, T., Nishida, Y.: Pressurized infiltration of aluminum melt into aluminum borate
whisker preform. J. Jpn. Inst. Met. 59, 564–570 (1995)
9. Yamauchi, T., Nishida, Y.: Infiltration kinetics of fibrous preforms by aluminum with solidifi-
cation. Acta Metall. Mater. 43, 1313–1321 (1995)
10. Jackson, G.W., James, D.F.: The permeability of fibrous porous media. Can J. Chem. Eng. 64,
364–374 (1986)
11. Yamauchi, T., Nishida, Y.: Permeability measurement of fibrous preform for MMC by
pressurized air. J. Jpn. Inst. Met. 60, 84–91 (1996)
12. Yamauchi, T., Nishida, Y., Nakae, H.: Pressure infiltration of molten aluminum into preform
of granulated whiskers. J. Jpn. Inst. Met. 61, 158–165 (1997)
13. Uozumi, H., Kobayashi, K., Nakanishi, K., Matsunaga, T., Shinozaki, K., Sakamoto, H.,
Tsukada, T., Masuda, C., Yoshida, M.: Fabrication process of carbon nanotube/light metal
matrix composites by squeeze casting. Mater. Sci. Eng. A 495, 282–287 (2008)
14. Pop, E., Mann, D., Wang, Q., Goodson, K., Dai, H.: Thermal conductance of an individual
single-wall carbon nanotube above room temperature. Nano Lett. 6, 96–100 (2006)
15. Mortensen, A., Masur, L.J., Cornie, J.A., Flemings, M.C.: Infiltration of fibrous preforms by a
pure metal: part I. Theory Metall. Trans. 20A, 2535–2547 (1989)
16. Masur, L.J., Mortensen, A., Cornie, J.A., Flemings, M.C.: Infiltration of fibrous preforms by a
pure metal: part II Experiment. Metall. Trans. 20A, 2549–2557 (1989)
17. Mortensen, A., Wong, T.: Infiltration of fibrous preforms by a pure metal: part III Capillary
phenomena. Metall. Trans. 21A, 2257–2263 (1990)
18. Hayashi, T., Ushio, H., Ebisawa, M.: The properties of hybrid fiber reinforced metal and its
application for engine block. SAE Paper. No. 890557 (1989); Wear properties of hybrid fiber
reinforced aluminum matrix composites and application to an automotive engine block. J. Jpn.
Inst. Light Met. 40, 787–792 (1990)
19. Komatsubara, T., Okajima, M., Koyasukata, Y., Hoshino, H.: Sanyo Tech. Rev. 20, 107 (1988)
20. Yamauchi, T.: Development of SiC whiskers reinforced piston. SAE Paper No. 911284 (1991)
21. Jarry, P., Michaud, V.J., Mortensen, A., Dubus, A., Tirard-Collet, R.: Infiltration of fiber
preform by an alloy: part III Die casting experiments. Metall. Trans. 23A, 2281–2289 (1992)
Chapter 5
Centrifugal Casting of Metal Matrix Composites

Abstract There are three major types of centrifugal casting used for the infiltration
of molten metal into fibrous preforms. In this chapter, the infiltration of molten
metal into fibrous preforms using centrifugal force is discussed theoretically and the
predictions of the theory are compared with experimental results. We discuss
the rotational speed necessary for infiltration to start, the pressure distribution in
the preforms, the velocity of the infiltration front and other important parameters.
When the volume fraction of fibers is not high, the pressure necessary for the
infiltration of molten metal is low. This process is suitable for fabricating products
which are symmetrical around a rotational axis, and uses simple and economical
casting equipment. In addition to discussion of fibrous preforms, centrifugal casting
of molten metal including ceramic particles is also discussed, focusing on the
theory of the behavior of a ceramic particle in molten metal in the centrifugal
force field.

5.1 Infiltration of Molten Metal Using Centrifugal Force

If molten metal is rotated, centrifugal force will act on the molten metal. The force
increases radially from zero at the rotation center. This force generates pressure in
the molten metal, and the pressure increases with distance along the radius. This
pressure can be used to infiltrate molten metal into a porous preform made of
ceramic fibers or particles and thus produce composites [1–4].
In this process, when the volume fraction Vf of the preform is relatively low, the
pressure needed for infiltration is not high. The apparatus is simple and economical,
although the process is only suitable for the symmetrical products. However, when
Vf of the preform is high, a high rotation number is needed, and the apparatus
required to produce the rate will be large.

Y. Nishida, Introduction to Metal Matrix Composites: Fabrication and Recycling, 91


DOI 10.1007/978-4-431-54237-7_5, # Springer Japan 2013
92 5 Centrifugal Casting of Metal Matrix Composites

Fig. 5.1 Three major cases a


for the infiltration of molten container
metal by centrifugal casting
molten metal

preform

rotation center

preform

container

rotation molten metal


center

preform

c A

container

rotation
center molten metal

5.1.1 Pressure Generated at the Surface of Preform

There are three cases for the production of composites using centrifugal force as
shown in Fig. 5.1. In Case 1, the preform is shaped like a pipe. Infiltration begins
from the inner surface of the pipe. The infiltration front spreads to the outer surface
of the pipe along the rotation radius. In Case 2, the volume of the preform is very
small compared with the total volume of molten metal, and therefore the displace-
ment of the inner surface of molten metal is negligible during infiltration. This
means that the pressure generated on the inner surface of the preform is almost
constant throughout infiltration. In addition, as the cross-section of the preform in
Case 2 is constant, it is possible to apply a one-dimensional infiltration model to this
case. In Case 3, the hatched area outside the fan-shaped molten metal also presses
the preform, which means that the pressure generated in Case 3 is higher than in the
other two cases. Therefore, the pressure which acts on the preform surface depends
on the type of infiltration.
A good starting point to develop the theory of centrifugal casting is to consider
the infiltration of molten metal in the small angle a at the rotation center during
5.1 Infiltration of Molten Metal Using Centrifugal Force 93

Fig. 5.2 Locations of the preform


infiltration front and inner
front
surface of molten metal
molten metal
during infiltration

rotation
container
center

Case 1, as shown in Fig. 5.1a. The mass of this part is rmAdr and its acceleration is
o2r. The centrifugal force, dFc, which acts on the region r to r + dr shown in
Fig. 5.1a, is given by:

dFc ¼ rm Ao2 rdr; (5.1)

where rm is the density of the molten metal, A is the cross-sectional area of the fan-
shaped molten metal with small angle a, and o is the angular velocity (¼ 2pN,
where N is the number of revolutions per second or revolution number).

5.1.1.1 Case 1

Case 1 corresponds to Fig. 5.1a. By integration of Eq. (5.1) between r0 and r1, the
force, Fc, is obtained as:

1
Fc ¼ rm ao2 ðr13  r03 Þ; (5.2)
3

where r0 and r1 are the locations of the inner surfaces of the molten metal and the
preform during rotation. The pressure P1, which acts on the inner surface of the
preform, is obtained by dividing Fc by the preform inner surface area ar1 at r ¼ r1.
When the infiltration of molten metal into the preform starts, the location of the
inner surface of the molten metal, rs(t), moves as shown in Fig. 5.2 and time-
dependent P1(t) during infiltration is given by:

1
P1 ðtÞ ¼ r o2 ðr13  ðrs ðtÞÞ3 Þ: (5.3)
3r1 m

We suppose that solidification of the molten metal does not occur during
infiltration and that the molten metal is an incompressible liquid. When the location
of the infiltration front in the preform once infiltration starts is rf (t), rs(t) is given by:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n offi
rs ðtÞ ¼ r02 þ f ðrf ðtÞÞ2  r12 ; (5.4)
94 5 Centrifugal Casting of Metal Matrix Composites

where f is the porosity of the preform (¼ 1  Vf). Substitution of Eq. (5.4) into
Eq. (5.3) yields:

rm o2 h 3 3=2
i
P1 ðtÞ ¼ r1  fr02 þ fððrf ðtÞÞ2  r12 Þg : (5.5)
3r1

This is the pressure which is applied to the inner surface of the preform during
infiltration.

5.1.1.2 Case 2

Case 2 is shown in Fig. 5.1b. The volume of the preform is so small compared with
the whole volume of the molten metal that the volume change of molten metal is
negligible during infiltration. In this case, the location of the inner surface of the
molten metal does not move during infiltration. Therefore, the pressure P1(t) in
Eq. (5.3) is constant throughout infiltration, and by putting rs(t) ¼ r0 in Eq. (5.3),
P1(t) becomes:

1
P1 ðtÞ ¼ r o2 ðr13  r03 Þ: (5.6)
3r1 m

5.1.1.3 Case 3

This case is shown in Fig. 5.1c. The cross-section of the molten metal column is
constant and does not change with increasing r. In other words, as well as the fan-
shaped metal with angle a, the hatched part of the molten metal also presses against
the inner surface of the preform. By putting A ¼ constant in Eq. (5.1), the force
which acts on the inner surface of the preform is given by:

1
Fc ¼ rm Ao2 ðr12  r02 Þ: (5.7)
2

In this case, as infiltration proceeds, the location of the inner surface of the liquid
column moves. When the location is rs(t), the time-dependent P1(t) is given by:

1
P1 ðtÞ ¼ rm o2 fr12  ðrs ðtÞÞ2 g; (5.8)
2

and

rs ðtÞ ¼ r0 þ fðrf ðtÞ  r1 Þ: (5.9)


5.1 Infiltration of Molten Metal Using Centrifugal Force 95

Substituting Eq. (5.9) into Eq. (5.8), we obtain:

1
P1 ðtÞ ¼ rm o2 ½r12  fr0 þ fðrf ðtÞ  r1 Þg2 : (5.10)
2

5.1.2 Infiltration Start Pressure

The contact angle of a ceramic fiber with molten aluminum is usually higher than
90 . A certain pressure must be applied to the molten metal to induce infiltration. The
pressure is called the “threshold pressure”, and was discussed in detail in Chap. 3.
When the distribution of fibers is random, the threshold pressure Pc is given by:

4Vf g cos y
Pc ¼  ; (5.11)
df ð1  Vf Þ

where g is the surface energy of the molten metal, y is the contact angle and df is the
diameter of fibers. When P1(t) is smaller than Pc, infiltration cannot begin. The
minimum revolution number necessary for infiltration is given by Eqs. (5.12) and
(5.13):
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 3r1 Pc
N ðfor Case 1 and Case 2Þ (5.12)
2p rm ðr13  r03 Þ

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 2Pc
N ðfor Case 3Þ (5.13)
2p rm ðr12  r02 Þ

5.1.3 Infiltration of Molten Metal into the Preform

Ignoring fluid inertial contributions and preform compression, the flow of the fluid
(molten metal) in the preform in the radial direction is governed by Darcy’s law in a
centrifugal force field [5, 6]:

@pðr; tÞ m uðr; tÞ
¼ m þ rm o2 r; (5.14)
@r K
96 5 Centrifugal Casting of Metal Matrix Composites

where p(r, t) is the pressure in the preform, mm is the viscosity of the molten metal, K
is the permeability, and u(r, t) is the radial superficial velocity which is the mass of
molten metal which flows into the preform per unit time and unit area.

5.1.3.1 Case 1

The continuity equation for the radial direction in cylindrical coordinates is given by:

rm @fruðr; tÞg
¼ 0: (5.15)
r @r

From Eq. (5.15), we obtain the following relationship:

ruðr; tÞ ¼ CðtÞ ði:e:; it is a function of time onlyÞ: (5.16)

Using Eq. (5.16), Eq. (5.14) is rewritten as:

@pðr; tÞ m CðtÞ
¼ m þ rm o2 r: (5.17)
@r K r

At a fixed time, it is possible to regard Eq. (5.17) as an ordinary differential


equation with respect to r. Then, we can obtain the following solution:

fmm CðtÞ r 1
pðr; tÞ  P1 ðtÞ ¼  ln þ rm o2 ðr 2  r12 Þ; (5.18)
K r1 2

where P1(t) is the pressure at r ¼ r1. As the pressure at r ¼ rf (t) (infiltration front)
is Pc (the threshold pressure):

fmm CðtÞ rf ðtÞ 1


Pc  P1 ðtÞ ¼  ln þ rm o2 fðrf ðtÞÞ2  r12 g: (5.19)
K r1 2

Alternatively, we have the following relationship between infiltration front rf (t)


and the superficial velocity uf (t) at the front:

drf ðtÞ uf ðtÞ


¼ : (5.20)
dt f

If mm, K, rm and f are independent of t, substitution of Eqs. (5.16) and (5.19) into
Eq. (5.20) yields the following ordinary differential equation:
ð rf ðtÞ
mm f2 rf ðtÞ lnðrf ðtÞ=r1 Þ
t¼ drf ðtÞ: (5.21)
r1 K½P1 ðtÞ  Pc þ ð1=2Þrm o2 fðrf ðtÞÞ2  r12 g
5.1 Infiltration of Molten Metal Using Centrifugal Force 97

It is difficult to obtain an analytical solution of this differential equation, but


numerical integration can be used.
The pressure distribution in the infiltrated region is obtained by substitution of C
(t) from Eq. (5.18) into Eq. (5.19):
 
1 2 lnðr=r1 Þ
pðr; tÞ ¼ P1 ðtÞ  P1 ðtÞ  Pc þ rm o fðrf ðtÞÞ  r1 g
2 2
2 lnðrf ðtÞ=r1 Þ
1
þ rm o2 ðr 2  r12 Þ (5.22)
2
Equation (5.22) gives the pressure distribution in the infiltrated region when the
infiltration front is located at position rf(t). This relationship shows that the pressure
is independent of f and K except for the threshold pressure Pc which depends on f.
When this process is applied industrially, the thickness of the preform is finite.
When the infiltration front reaches the outer surface of the preform (r ¼ rt), the
infiltration stops and the pressure distribution in the infiltrated region changes
instantaneously. Supposing that the framework of the fibrous preform is sufficiently
rigid and does not deform in the infiltrated region, the centrifugal force, dFc, which
acts on the metal of the region r to r + dr in the preform, is given by:

dFc ¼ frm Ao2 rdr: (5.23)

When the preform thickness is L (¼ rt  r1), the pressure in the preform p(r) at r
(r1  r  rt) is given by:

1 1
pðrÞ ¼ r o2 ½r13  fr02 þ fðrt2  r12 Þg3=2  þ frm o2 ðr 3  r13 Þ: (5.24)
3r1 m 3r

When the amount of molten metal is small and all molten metal enters the
preform before the infiltration front arrives at the outer surface of the preform, the
pressure curve is given only by the second term of Eq. (5.24).

5.1.3.2 Case 2

This is the case where P1(t) is constant (¼ P0). The cross section of the preform is
independent of r and constant. Therefore, the superficial velocity u(r, t) in
Eq. (5.14) is also independent of r and expressed by u(t). The one-dimensional
form of the continuity equation, which applies in this situation, is:

@uðtÞ
¼ 0: (5.25)
@r

At a fixed time, it is also possible to regard Eq. (5.14) as an ordinary differential


equation with respect to r. Then, we can obtain the following solution:
98 5 Centrifugal Casting of Metal Matrix Composites

mm uðtÞ 1
pðr; tÞ ¼ P0  ðr  r1 Þ þ rm o2 ðr 2  r12 Þ: (5.26)
K 2
As the pressure at the infiltration front r ¼ rf (t) is Pc, we obtain:

mm uðtÞ 1
P0 ¼ ðrf ðtÞ  r1 Þ  rm o2 fðrf ðtÞÞ2  r12 g þ Pc : (5.27)
K 2

Alternatively, the velocity of the infiltration front is given by:

drf ðtÞ uðtÞ


¼ : (5.28)
dt f

Substitution of u(t) from Eq. (5.28) into Eq. (5.27) yields an ordinary differential
equation:

m3 þ m4 rf ðtÞ
drf ðtÞ ¼ dt; (5.29)
m1 þ m2 frf ðtÞg2

where
 
r o2 r12 Krm o2
m 1 ¼ K P 0  Pc  m ; m2 ¼ ; m3 ¼ mm fr1 ; m4 ¼ mm f:
2 2

The solution of Eq. (5.29) depends on the sign (negative or positive) of m1.
However, since we know P0 from Eq. (5.6), m1 is always negative. Then, the
following solution is obtained:

rf ðtÞ  n3
t ¼ n1 lnfðrf ðtÞÞ2  n23 g þ n2 ln þ n4 ; (5.30)
rf ðtÞ þ n3

where
rffiffiffiffiffiffiffiffiffiffi
m4 m3 m1
n1 ¼ ; n2 ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; n3 ¼  ; and
2m2 2 m2 m1 m2
 
r 1  n 3
n4 ¼ n1 lnðr12  n23 Þ  n2 ln :
r 1 þ n3

By substituting uðtÞ from Eq. (5.26) into Eq. (5.27), we obtain:


   
rf ðtÞ  r 1 r  r1
pðr; tÞ ¼ P0 þ rm o ðr  r1 Þðr  rf ðtÞÞ þ Pc
2
: (5.31)
rf ðtÞ  r1 2 rf ðtÞ  r1

This equation gives the pressure distribution in the infiltrated region when the
infiltration front is specified. The characteristic features of this equation are that the
equation does not include f, and that the pressure distribution pattern is
5.1 Infiltration of Molten Metal Using Centrifugal Force 99

independent of the volume fraction of fibers in the preform except for the depen-
dence of the threshold pressure, Pc, on f.
When the preform thickness is L, the pressure p(r) in the infiltrated region after
the infiltration front reaches the outer surface of the preform is given by:

1
pðrÞ ¼ P0 þ frm o2 ðr 2  r12 Þ; ðr1  r  rt Þ: (5.32)
2

5.1.3.3 Case 3

This is the case shown in Fig. 5.1c. Since cross sectional variations can be neglected
because of the cylindrical configuration, the superficial velocity u(r, t) in Eq. (5.14)
is actually u(t) and the continuity equation in the one-dimensional form, Eq. (5.25),
should also be valid in this case. At a fixed time, it is possible to regard Eq. (5.14) as
an ordinary differential equation with respect to r, if mm, K and rm are independent
of r. Then, we can get the following solution:

mm uðtÞ 1
pðr; tÞ ¼ P1 ðtÞ  ðr  r1 Þ þ rm o2 ðr 2  r12 Þ; (5.33)
K 2

where P1(t) is the pressure at r ¼ r1. As the pressure p(r, t) at r ¼ rf (t) in Eq. (5.33)
is the threshold pressure Pc, we obtain:

mm uðtÞ 1
P1 ðtÞ ¼ ðrf ðtÞ  r1 Þ  rm o2 ððrf ðtÞÞ2  r12 Þ þ Pc : (5.34)
K 2

Substitution of Eqs. (5.10) and (5.28) into Eq. (5.34) provides an ordinary
differential equation, if mm, K, rm and f are independent of t. The solution is:
   
1 rf ðtÞ þ m6 m4
t¼ ln fðrf ðtÞÞ2 þ m1 rf ðtÞ þ m2 g þ m7 ; (5.35)
2m3 rf ðtÞ þ m5

where

2fðfr1  r0 Þ r02  f2 r12 þ 2fr0 r1  2Pc =ðrm o2 Þ


m1 ¼ ; m 2 ¼ ;
1  f2 1  f2
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rm o2 Kð1  f2 Þ m1 þ 2r1 1 m21
m3 ¼ ; m4 ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi ; m5 ¼ m1   m2 ;
2mm f 2 m21 =4  m2 2 4
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi    
1 m21 1 r1 þ m6 m4
m6 ¼ m1 þ  m2 ; and m7 ¼  ln ðr12 þ m1 r1 þ m2 Þ
2 4 2m3 r1 þ m5
100 5 Centrifugal Casting of Metal Matrix Composites

Equation (5.35) gives the relationship between the location of the infiltration
front rf (t) and time t.
Alternatively, by substituting Eq. (5.34) into Eq. (5.33), we obtain the
pressure distribution in the infiltrated region when the infiltration front is specified
at rf (t):

ðrf ðtÞ  rÞP1 ðtÞ þ ðr  r1 ÞPc 1


pðr; tÞ ¼ þ rm o2 ðr  r1 Þðr  rf ðtÞÞ: (5.36)
rf ðtÞ  r1 2

When the preform thickness is L (¼ rt  r1), the pressure in the preform after
infiltration reaches the outer surface of the preform is:

1 h i 1
pðrÞ ¼ rm o2 r12  fr0 þ fðrt  r1 Þg2 þ frm o2 ðr 2  r1 2 Þ: (5.37)
2 2

5.1.4 Example Calculations

The relationship between Vf and the permeability of the preform is needed for the
calculation of infiltration. There are many theoretical equations for the permeability
of porous media [7]. Among them, Langmuir’s equation [Eq. (5.38)] which applies
to flow parallel to the fiber alignment and Happel’s equation [Eq. (5.39)] which
applies to flow normal to fibers can be used to obtain the permeability of the
preform. Generally the permeability calculated using these equations agrees well
with experimental results. The equations are:
 
K 1 3 1
¼  ln Vf  þ 2Vf  Vf 2 ; (5.38)
R2 4Vf 2 2

and
!
K 1 Vf2  1
¼  ln Vf þ ; (5.39)
R2 8Vf Vf2 þ 1

where R is the radius of the fiber. As the fiber distribution we used was random, the
permeability of the preform was obtained by averaging the values of permeability
calculated for flows parallel and normal to the fiber alignment.

5.1.4.1 Case 1

The advance of the infiltration front of molten aluminum (without solidification)


calculated using Eq. (5.21) is shown in Fig. 5.3. The calculation was performed for
5.1 Infiltration of Molten Metal Using Centrifugal Force 101

Fig. 5.3 Relationship between infiltrated distance and time for a range of rotational speeds for
Case 1

Fig. 5.4 Change with time of pressure distribution in the infiltrated region for Case 1 when
N ¼ 50 rev/s

a preform made from alumina short fibers. The following data were also used in this
calculation: r0 ¼ 0.07 m, r1 ¼ 0.12 m, f (¼ 1  Vf) ¼ 0.87, rm ¼ 2.38  103 kg/
m3, R (¼ df/2) ¼ 1.85  106 m, g ¼ 0.893 Pa m, y ¼ 160 and mm ¼ 0.984  10
3
Pa s. The minimum revolution number N for the infiltration is 19.74 rev/s, as
calculated from Eq. (5.12). Figure 5.3 shows that if the revolution rate is slightly
higher than the minimum value, infiltration will be continuous. The pressure
distribution in the infiltrated region obtained from Eq. (5.22) is shown in Fig. 5.4
for N ¼ 50 rev/s. The pressure distribution curves are convex towards lower
pressures, because the second derivative of Eq. (5.22) is always positive and
102 5 Centrifugal Casting of Metal Matrix Composites

Fig. 5.5 Pressure distribution in the infiltrated region for Case 1 after the infiltration front reaches
the outer surface of the preform when N ¼ 50 rev/s

proportional to o2. The preform surface pressure decreases quickly as the infiltra-
tion advances, and the intermediate part of the pressure distribution curve falls
below the threshold pressure. Nevertheless, infiltration continues, because the
threshold pressure is needed only to wet fibers with molten metal, and the pressure
at the infiltration front always remains constant at the threshold pressure. After
wetting, the molten metal does not separate from the preform, even if the pressure
becomes lower than the threshold pressure, unless enough work for the separation is
supplied to the molten metal/fiber interface.
When the preform thickness is L and the infiltration front reaches the location r
(t) (the outer surface of the preform), the pressure distribution curve shown in
Fig. 5.4 for N ¼ 50 rev/s is changed instantaneously to that of Fig. 5.5. Only the
preform surface pressure remains constant, and the pressure in the infiltrated region
rises quickly and becomes stable. In this calculation, in the cases of L ¼ 0.04 and
0.05 m, all liquid metal enters into the preform and a cavity is formed near the inner
surface of the preform, because of the lack of liquid metal.

5.1.4.2 Case 2

Figure 5.6 shows the relationship between time and the advance of the infiltration front
obtained from Eq. (5.30). The front advances faster than in Case 1, because the
preform surface pressure remains constant throughout the infiltration. The pressure
distribution curves are shown in Fig. 5.7 for N ¼ 50 rev/s. The pressure distribution
curves in this case are also convex toward lower pressure, because the second
derivative of Eq. (5.31) is also positive and proportional to o2. The pressure distribu-
tion curves for finite preform thickness cases are shown in Fig. 5.8 for N ¼ 50 rev/s. In
Case 2, even if the preform thickness is different, the pressure distribution curve is the
same, and the curve depends only on the revolution number.
5.1 Infiltration of Molten Metal Using Centrifugal Force 103

Fig. 5.6 Relationship between infiltrated distance and time for a range of rotational speeds for
Case 2

Fig. 5.7 Change with time of pressure distribution in the infiltrated region for Case 2 when
N ¼ 50 rev/s

5.1.4.3 Case 3

Figure 5.9 shows the relationship between time and the advance of the infiltration
front obtained from Eq. (5.35) for Case 3. The front advances faster than in Case 1
and slower than in Case 2 for a particular revolution number. The pressure distri-
bution curves obtained from Eq. (5.36) are shown in Fig. 5.10 for N ¼ 50 rev/s.
These curves are also convex toward lower pressure. Regions having lower pressure
than the threshold pressure appeared at t ¼ 0.163 s in Fig. 5.10. The pressure
104 5 Centrifugal Casting of Metal Matrix Composites

Fig. 5.8 Pressure distribution in the infiltrated region for Case 2 after the infiltration front reaches
the outer surface of the preform when N ¼ 50 rev/s

Fig. 5.9 Relationship between infiltrated distance and time for a range of rotational speeds for
Case 3

distribution curves obtained using Eq. (5.37) are shown in Fig. 5.11 for various
finite preform thickness at N ¼ 50 rev/s. Until the infiltration front arrives at the
outer surface of the preform, the preform surface pressure changes as given in
Fig. 5.10. When the front arrives at the preform outer surface, the change in the
preform surface pressure stops and the pressure distribution in the infiltrated region
becomes that shown in Fig. 5.11.
5.1 Infiltration of Molten Metal Using Centrifugal Force 105

Fig. 5.10 Change with time of pressure distribution in the infiltrated region for Case 3 when
N ¼ 50 rev/s

Fig. 5.11 Pressure distribution in the infiltrated region for Case 3 after the infiltration front
reaches the outer surface of the preform when N ¼ 50 rev/s

5.1.5 Examples of Composites Fabricated Using


Centrifugal Force

In this section we summarize a previous study where the feasibility of centrifugal


force for the infiltration of molten metal into preforms was examined and the
composites obtained were characterized [8]. A schematic of the apparatus used is
106 5 Centrifugal Casting of Metal Matrix Composites

pouring device

graphite pipe
graphite container

preform

Fig. 5.12 Schematic of the apparatus used to generate centrifugal force for the infiltration of
aluminum

Fig. 5.13 Relationship between the pressure generated and infiltrated distance. The initial
thicknesses of the preforms are 20 mm for Vf ¼ 6 %, 19 mm for Vf ¼ 9 %, and 18 mm for
Vf ¼ 13 %. The compressive deformation of preforms is neglected

shown in Fig. 5.12. A preheated graphite container with a uniform cross section was
used. The container was balanced by a counterweight. A preheated alumina short
fiber preform was set on the bottom of the container which was rotated. When the
rotational speed reached the target value, molten pure aluminum was poured into a
pouring device concentric with the rotation shaft and a composite was produced.
The experimentally obtained relationship between the rotational speed and the
infiltrated distance for different volume fractions of fiber and initial preform
thickness is shown in Fig. 5.13. The diameter for the preform was 24 mm, and
the initial thicknesses of the preforms were 20 mm for Vf ¼ 6 %, 19 mm for
Vf ¼ 9 %, and 18 mm for Vf ¼ 13 %. The infiltrated distances for Vf ¼ 6 % and
5.1 Infiltration of Molten Metal Using Centrifugal Force 107

Fig. 5.14 Macrostructures of alumina/silica short fiber reinforced aluminum composite samples
fabricated using centrifugal force

9 % were either zero or 100 %. That is, in the case of Vf ¼ 6 %, infiltration did not
advance at 65 kPa (N ¼ 10 rev/s) but fully advanced at 70 kPa (N ¼ 11 rev/s). For
Vf ¼ 9 %, infiltration did not take place at 70 kPa (N ¼ 16 rev/s), but occurred fully
at 85 kPa (N ¼ 17 rev/s). For Vf ¼ 13 %, the molten metal advanced a little and
then stopped at 160 kPa (N ¼ 26 rev/s), but advanced much further at 230 kPa
(N ¼ 32 rev/s). Since the preheat temperatures of the preform and graphite con-
tainer were 703 and 673 K, respectively, the infiltration velocity was slow in this
case and the metal solidified during infiltration. Pressure higher than 300 kPa was
needed for full infiltration for the Vf ¼ 13 % preform.
The structural characteristics of samples are presented in Figs. 5.14 and 5.15.
A partially infiltrated sample with Vf ¼ 13 % is also shown in Fig. 5.14. The
preform also contained many particles, because the quality of the alumina/silica
fibers was not high. The minimum pressures for the start of infiltration were
obtained from this experiment and are listed in Table 5.1, along with the threshold
pressure values calculated theoretically using Eq. (5.11). Both sets of values are in
agreement.
108 5 Centrifugal Casting of Metal Matrix Composites

Fig. 5.15 Microstructure of alumina/silica short fiber reinforced aluminum composites fabricated
using centrifugal force

Table 5.1 Comparison of threshold pressures


Vf (%) Pc (kPa) (theoretical) Pc (kPa) (experimental)
6 57.9 66
9 89.7 81
13 135.6 156

5.2 Centrifugal Casting of Particle Dispersed Molten Metal

When molten metal which contains ceramic particles is cast in a centrifugal force
field, the particles tend to segregate because of the density difference between the
molten metal and the ceramic particles. Here, we discuss the movement of the
particles and the forces which act on the particles. To simplify the model, solidifi-
cation will be neglected.
A particle having density rp and radius Rp is present in molten metal having
density rm. The centrifugal force, which is proportional to the rotating radius r, is
acting on the particle. The equation for the force, Fc, which acts on the particle as a
result of the acceleration, o2r is similar to Eq. (5.1):

4
Fc ¼ pR3p ðrp  rm Þo2 r: (5.40)
3

If the particle moves with very low velocity in the molten metal, as shown in
Fig. 5.16, Stokes’ drag formula will be applicable [9]. The viscous drag is given by:

dr
Fv ¼ 6pmm Rp ; (5.41)
dt

where Fv acts on the particle in the opposite direction to the centrifugal force, and
has a negative sign. mm is the viscosity of the molten metal.
The fundamental expression of the movement of an object is: [force] ¼ [mass]
 [acceleration]. For the particle, the total force is Fc + Fv. Thus, we obtain the
following differential equation:
5.2 Centrifugal Casting of Particle Dispersed Molten Metal 109

Fig. 5.16 Viscous drag, Fv,


and centrifugal force, Fc, particle
which act on a particle during
centrifugal casting of particle
reinforced molten metal
Fv
Fc
Rp

4 3 dr 4 d2 r
pRp ðrp  rm Þo2 r  6pmm Rp ¼ pR3p rp 2 : (5.42)
3 dt 3 dt
Gravity is neglected, because the gravity is much smaller than the centrifugal force.
The solution of Eq. (5.42), i.e., the location of the particle as a function of time, is:
28 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi9 3
<9mm þ 81m2m þ 16ðrp  rm Þrp R4p o2 =
r ¼ r0 exp4 t5; (5.43)
: 4Rp rp ;

where r0 is the initial location of the particle. The solution is a simple exponential
function. If the particle goes far away from the rotating center, the velocity of the
particle will increase and Stokes’ drag formula (equation (5.41)) will be invalid,
because the formula is only valid when the Reynolds number around the particle is
less than 0.5.
Examples of calculations using Eq. (5.43) are shown in Figs. 5.17 and 5.18. In these
calculations, the particle is a spherical SiC particle with density rp ¼ 3.15  103
kg/m3; the density rm of the molten aluminum is 2.38  103 kg/m3 and the viscosity
of the molten aluminum mm ¼ 0.948  103 Pa s. Figure 5.16 shows the movement
with time of particles with different radii, Rp, when the revolution number N is 10 rev/s.
Equation (5.43) gives the dimensionless displacement ratio of the location of the particle
to its initial location r/r0, which represents the movement of the particle. When the
particle radius is 1 mm (i.e., the diameter is 2 mm), the particle quickly moves away
from the rotating center. However, when the particle radius is 0.1 mm, the particle will
move slowly. Figure 5.17 shows the displacement ratio, r/r0 of a particle of radius
0.1 mm, for different revolution numbers N. When N ¼ 5 rev/s, the particle essentially
remains at its initial location for the entire 300 s. However, when N is larger than 15 rev/s
(900 rpm), the particle moves quickly.
However, Eq. (5.43) is quite different to “Stokes’ law” which gives:

dz 2R2p ðrp  rL Þg
¼ ; (5.44)
dt 9mL
110 5 Centrifugal Casting of Metal Matrix Composites

Fig. 5.17 Curves of displacement ratio r/r0 with time for various particle radii; revolution number
N is 10 rev/s

Fig. 5.18 Curves of displacement ratio r/r0 with time for various values of revolution number N;
particle radius is 0.1 mm

where g is the gravitational acceleration, rL and mL are the density and viscosity of
the fluid, respectively. The z-axis is taken as the gravity direction. In this equation,
the acceleration is constant, although it is o2r in the centrifugal force field.
In actual centrifugal casting, solidification of metal occurs along with the
movement of particles. The solidification in this case is very complicated and it is
very difficult to analytically calculate the behavior, because solidification depends
References 111

on the shape of container (mold), the composition of the alloy and on other factors.
In addition, Eq. (5.43) is the model of the movement for a single particle only.
When the number of particles is large, we should consider the interaction between
particles. Then, the apparent viscosity mapp must be used instead of the real viscosity
of the molten metal; the apparent viscosity increases when the number of particles
increases, and is given by [10]:

5
mapp ¼ mm ð1 þ Vf þ 10:05Vf2 Þ: (5.45)
2

This mapp can then be used instead of mm in Eq. (5.43).


Centrifugal casting with solidification in the presence of solid particles was
discussed numerically by Panda et al. [11]. They showed that the segregation of
particles in the alumina particles/aluminum or silicon carbide particles/aluminum
systems depends on the size of particle, rotational speed, heat transfer coefficient at
the casting/mold interface and volume fraction of particles.
Examples of the application of centrifugal casting to composites include copper
alloy/graphite particle composites, which were developed by Kim et al. to improve
tribological properties of the inner surfaces of products [12]. Functionally graded
composites were also developed in TiB2 or TiC particles/aluminum alloy systems
by Kumar et al. [13].

References

1. Sugishita, J.: Fabrication and properties of centrifugally cast surface composites based on
Al-11%Si alloy. J. Jpn. Foundry Eng. Soc. 57, 102–107 (1985)
2. Tsunekawa, Y., Okumiya, M., Niimi, I., Yoneyama, K.: Centrifugally cast aluminum matrix
composites containing segregated alumina fibers. J. Mater. Sci. Lett. 7, 830–832 (1988)
3. Tsunekawa, Y., Okumiya, M., Niimi, I., Maeda, T.: Improvement of bending strength in
aluminum alloy matrix composites containing short alumina fibers by centrifugal force.
J. Jpn. Inst. Light Met. 40, 7–12 (1990)
4. Nishida, Y., Ohira, G.: Modelling of infiltration of molten metal in fibrous preform by
centrifugal force. Acta Mater. 47, 841–852 (1999)
5. Geiger, G.H., Poirier, D.R.: Transport Phenomena in Metallurgy, p. 48. Addison-Wesley,
Reading (1980)
6. Collins, R.E.: Flow of Fluids Through Porous Materials. Reinhold Publishing Corp, New York
(1961)
7. Jackson, G.W., James, D.F.: The permeability of fibrous porous media. Can. J. Chem. Eng. 64,
364–374 (1986)
8. Nishida, Y., Shirayanagi, I., Sakai, Y.: Infiltration of fibrous preform by molten aluminum in a
centrifugal force field. Metall. Mater. Trans. 27A, 4163–4169 (1996)
9. Geiger, G.H., Poirier, D.R.: Transport Phenomena in Metallurgy, p. 71. Addison-Wesley,
Reading (1980)
10. Schowalter, W.R.: Mechanics of Non-Newtonian Fluids, p. 288. Pergamon, London (1978)
112 5 Centrifugal Casting of Metal Matrix Composites

11. Panda, E., Mehrotra, S.P., Mazumdar, D.: Mathematical modeling of particle segregation
during centrifugal casting of metal matrix composites. Metall. Mater. Trans. 37A,
1675–1687 (2006)
12. Kim, J.K., Kestursatya, M., Rohatgi, P.K.: Tribological properties of centrifugally cast copper
alloy-graphite particle composite. Metall. Mater. Trans. 31A, 1283–1293 (2000)
13. Kumar, S., Sarma, V.S., Murty, B.S.: Functionally graded alloy matrix in situ composites.
Metall. Mater. Trans. 41A, 242–254 (2010)
Chapter 6
Properties of Composites

Abstract In this chapter the mechanical and physical properties of metal matrix
composites are discussed. Mechanical properties such as strength and elastic modu-
lus depend on the shape and properties of the reinforcements, their distribution and
volume fraction and the bonding strength at the reinforcement/matrix interface, as
well as the properties of the matrix itself. Basic ideas and proposed models are
introduced to understand the mechanical properties. Some physical properties such
as specific heat and density are determined by the intrinsic properties of reinforce-
ments and the matrix. However, other properties, such as thermal expansion coeffi-
cient, depend on the distribution, volume fraction, shape and state of the
reinforcements. Some proposed models for physical properties are also introduced.

Some properties of composites such as tensile strength are usually closely


associated with the interaction at the interface between constituents, rather than
the properties of the constituents (matrix and reinforcements). Other properties
such as density are given by a weighted sum of the intrinsic properties of
constituents. When a chemical reaction at the reinforcement/matrix interface
occurs, we can easily imagine that the chemical reaction greatly influences the
properties of the reinforcements and the composites. To obtain good bonding at
the interface, the constituents are usually heated to high temperature, where a
chemical reaction often occurs. Most of these chemical reactions result in the
degradation of the reinforcement properties so that the desired property is not
obtained. Therefore, in fabricating composites, it is very important to achieve the
best possible bonding at the reinforcement/matrix interface without degradation.
Since it is difficult to quantitatively assess the state of the interface, the mechani-
cal properties of the composite are usually discussed on the basis of some
assumed shear strength between the reinforcement and matrix metal. In this
chapter, the fundamental knowledge needed to understand the properties of
MMCs is briefly introduced [1].

Y. Nishida, Introduction to Metal Matrix Composites: Fabrication and Recycling, 113


DOI 10.1007/978-4-431-54237-7_6, # Springer Japan 2013
114 6 Properties of Composites

6.1 Mechanical Properties

6.1.1 Elastic Modulus

The elastic modulus (Young’s modulus) of composites is discussed here. As shown


in Fig. 6.1a, we consider a continuous fiber reinforced composite to which tensile
stress sCL is applied parallel to the fiber direction. The composite is deformed
elastically. The subscripts CL and CT correspond to parallel (longitudinal) to the
fiber direction and normal (transverse) to the fiber direction, respectively. In this
case, it is assumed that the bonding between fiber and matrix metal is perfect and
the amount of elastic deformation at the interface is equal for both metal and fiber.
In addition, it is also assumed that the Poisson’s ratios of fiber and matrix metal are
equal. Then, the following equations hold:

sf ¼ e f E f ; sm ¼ e m E m ; and sCL ¼ eCL ECL ; (6.1)

where s, e and E are stress, elastic strain and elastic modulus, respectively. The
suffixes f, m and C correspond to fiber, matrix metal and composite, respectively.
As the elastic deformations of the fiber and the matrix in the composite are the same
in this case, we obtain:

eCL ¼ ef ¼ em : (6.2)

The area fractions of fiber and matrix in the top surface of the composite are Vf
and (1–Vf), respectively. So the fiber and the matrix share the stress sCL and the
stress on the composite is given by:

sCL ¼ Vf sf þ ð1  Vf Þsm : (6.3)

This equation is a direct application of the rule of mixtures discussed in Sect. 1.2
of Chap. 1. Substituting Eq. (6.1) into Eq. (6.3) gives:

a b
Fiber

Fig. 6.1 Applied stress


(a) parallel to fiber direction,
(b) normal to fiber direction
of continuous fiber reinforced
composite
6.1 Mechanical Properties 115

eCL ECL ¼ Vf ef Ef þ ð1  Vf Þem Em : (6.4)

As the elastic strains are equal, as shown in Eq. (6.2), division of Eq. (6.4) by
strain e yields:

ECL ¼ Vf Ef þ ð1  Vf ÞEm : (6.5)

This equation shows that the elastic modulus parallel to the fiber direction is
obtained by direct application of the rule of mixtures. This elastic modulus,
however, gives the lower limit of the real value, as pointed out by Hill [2] and
Hashin et al. [3], because the interaction between fiber and matrix has not been
taken into consideration.
Alternatively, as shown in Fig. 6.1b, we can also consider a composite to which
tensile stress sCT is applied transverse (i.e., normal) to the fiber direction and where
the composite is deformed elastically. Equation (6.1) also holds in this case.
However, it is obvious from the figure that the following equation holds instead
of Eq. (6.2):

sCT ¼ sf ¼ sm : (6.6)

The fraction of fibers in unit length of the composite along stress direction is Vf,
and the fraction of matrix is (1–Vf). Then, the strain eCT is related to the strains of
the fibers and the matrix by:

eCT ¼ Vf ef þ ð1  Vf Þem : (6.7)

However, we will assume that the elastic modulus of the fibers in the radial
direction is equal to that along the fiber. Substitution of Eq. (6.1) into Eq. (6.7)
yields:

sCT Vf sf ð1  Vf Þsm
¼ þ : (6.8)
ECT Ef Em

Therefore, because of Eq. (6.6), the elastic modulus of a composite normal to the
fiber direction is given by:

E f Em
ECT ¼ : (6.9)
Em Vf þ Ef ð1  Vf Þ

Thus, for continuous fiber reinforced composites, the elastic modulus parallel to
the fiber direction is obtained simply by applying the rule of mixtures [Eq. (1.5)]
considering the microscopic relationship of stress and strain in the composites.
However, the elastic modulus normal to the fiber direction is given by Eq. (6.9).
Figure 6.2 shows an example of the difference between the elastic moduli parallel
116 6 Properties of Composites

Fig. 6.2 Comparison


between elastic moduli
of a continuous SiC fiber
reinforced aluminum
composite parallel to the fiber
direction ECL and normal
to the fiber direction ECT

Fig. 6.3 (a) Short fiber a Short fiber b Particle


reinforced composite and
(b) particulate reinforced
composite

to and normal to the fiber direction, as a function of Vf, for a SiC continuous fiber
reinforced aluminum composite. The difference can be very significant, and there is
a maximum in the difference between the moduli.
Equations (6.5) and (6.8) are not applicable to short fiber reinforced composites
or particulate reinforced composites as shown in Fig. 6.3. If we apply the rule of
mixtures [Eq. (1.5)] to these composites to calculate elastic modulus, the calcula-
tion will only be a rough estimation, and it is difficult to get good agreement with
experimental results. However, if short fibers are aligned in one direction in a
composite, as shown in Fig. 6.3a, a “shear lag model”, which is proposed by Cox
[4], can be applied. We will discuss the shear lag model following Cox’s
explanation.
Usually, if stress is applied to a composite along the short fiber axis direction, the
strain of the fiber is small, and that of the matrix metal is large, because the elastic
modulus of ceramic fibers is usually larger than that of matrix metal. We can imagine
that the strain of matrix metal in a composite will be smaller than that of the matrix
itself (without fibers) under the same stress. If the bonding at the fiber/matrix
interface is perfect, the strain of the matrix side at the interface is equal to that of
the fiber. The strain of the matrix increases with increasing radial distance from
the fiber surface until the strain approaches the average strain of the composite.
6.1 Mechanical Properties 117

Fig. 6.4 Schematic of a short fiber surrounded by matrix metal

before stress application

after stress application

c stress σ in the fiber and


shear stress τ on the
surface of the fiber

Fig. 6.5 Strain and stress distributions in the matrix surrounding a short fiber. (a) Before stress
application, (b) after stress application, (c) stress s in the fiber and shear stress t on the surface of
the fiber

Thus, it is obvious that shear stress acts on the matrix at any distance r normal to the
fiber axis (r ¼ 0).
A schematic description of a short fiber in a composite is shown in Fig. 6.4. A
fiber having diameter of 2r0 and length l is covered with matrix to produce a
cylindrical composite having diameter of 2R which is the average distance from
the axis of a fiber to the axis of the nearest neighboring fiber. In this case, the
deformation of the matrix is shown schematically in Fig. 6.5. Overall, stress sc acts
on the composite and the elastic strain is e. Defining x ¼ 0 at one end of the fiber
and force P as acting on the fiber at location x from the end of the fiber, the
relationship between P and x is given by:

dP
¼ Hðu  vÞ; (6.10)
dx
118 6 Properties of Composites

where u is the longitudinal displacement in the fiber, v is the corresponding


displacement of the matrix would undergo at the same point if the fiber were absent.
H is a constant. Further, P ¼ Asf, where A is the cross-sectional area of the fiber. sf
is the stress applied to the fiber. Hooke’s law holds for the relationship between the
displacement of the fiber and the force. Then, we obtain:
 
du
P ¼ Ef A ; (6.11)
dx

where Ef is the elastic modulus of the fiber in the axis direction. The relationship
between v and e is given by:

dv
¼ e ðconstantÞ: (6.12)
dx

Further differentiation of Eq. (6.10) with respect to x and substitution of


Eqs. (6.11) and (6.12) yields:
 
d2 P P
¼H e : (6.13)
dx2 Ef A

This equation is an ordinary differential equation between P and x. The general


solution is given by:

P ¼ Ef eA þ C1 sinh bx þ C2 cosh bx; (6.14)

where C1 and C2 are constants.


The stress at both ends of the fiber is zero. Then, C1 and C2 are obtained by using
the boundary conditions P ¼ 0 at x ¼ 0 and at x ¼ l. Finally, sf is given by:
 
cosh bðl=2  xÞ
sf ¼ Ef e 1  : ð0  x  l=2Þ; (6.15)
coshðbl=2Þ

where b ¼ (H/Ef)1/2.
Equation (6.15) gives the stress at x in the fiber. The average stress over the
whole fiber is given by:
ð l=2  
2 tanhðbl=2Þ
sf ¼ sf dx ¼ Ef e 1  : (6.16)
l 0 bl=2

The strain of the matrix bonded with the fiber at r ¼ r0 is the same as that of the
fiber. The strain of the matrix at any location r > r0 is larger than that of the fiber.
Then, a shear stress t (r) in the matrix acts on any point of the curved plane at
distance r from the fiber axis. The stress balance of t (r) yields:
6.1 Mechanical Properties 119

2prtðrÞ ¼ 2pr0 tðr0 Þ ¼ ðconstantÞ: (6.17)

The displacement w of the matrix between r ¼ r0 (the surface of fiber) and


r ¼ R takes the following form:

dw tðrÞ tðr0 Þr0


¼ ge ¼ ¼ ; (6.18)
dr Gm Gm r

where ge and Gm are the shear stress and the shear modulus of the matrix. Integra-
tion of Eq. (6.18) with respect to r from r0 to R yields:
ðR ðR  
tðr0 Þr0 tðr0 Þr0 R
dw ¼ dr ¼ ln : (6.19)
r0 r0 Gm r Gm r0

Since the integrated value of dw from r0 to R is (v–u), the following relationship


is obtained:
 
tðr0 Þr0 R
vu¼ ln : (6.20)
Gm r0

The shear stress acts on the surface of the fiber at r ¼ r0 and elastic deformation
of the fiber occurs. The force balance is given by:

dP
¼ 2pr0 tðr0 Þ ¼ Hðu  vÞ: (6.21)
dx

Therefore:

2pr0 tðr0 Þ 2pGm


H¼ ¼ : (6.22)
uv lnðR=r0 Þ

Substitution of this equation into Eq. (6.15) and rearranging for b yields:
 1=2
2pGm
b¼ : (6.23)
Ef lnðR=r0 Þ

Alternatively, by using Eqs. (6.21) and (6.15), t (r0) is:


 
r0 eEf b sinh bðl=2  xÞ
tðr0 Þ ¼ : ð0  x  l=2Þ (6.24)
2 coshðbl=2Þ

Equation (6.16) gives the apparent elastic modulus and the stress that one fiber
bears, when the total strain of the composite is e. Then, if the rule of mixtures of
120 6 Properties of Composites

Eq. (1.5) is applicable to the apparent elastic modulus of the fiber and the elastic
modulus of the matrix, the elastic modulus of the composite Ec is:
 
tanhðbl=2Þ
Ec ¼ E f Vf 1  þ Em ð1  Vf Þ: (6.25)
bl=2

The relationship between the stress sf in the fiber and the shear stress t (r0) on the
surface of the fiber is shown schematically in Fig. 6.5c. sf takes a maximum value in
the central part of the fiber span and decreases towards the ends of the fiber.
In contrast, the shear stress shows maximum magnitude at both ends of the fiber;
it is negative at one end and positive at the other.
This model is widely used because it is simple. This elastic modulus was
introduced on the assumption that the stresses at both ends of the fiber are
negligible. However, it is known that when the fiber length is very short, the
agreement between calculated and experimental values is poor. The shear lag
model has been explained plainly by Kelly and Macmillan [5]. The modified
model, including the stress on the both ends of fiber, has been discussed by Clyne
and Withers [6].
However, the shear lag model is not applicable to particulate reinforced metals
as shown in Fig. 6.3b. When the particles in a composite are spheroid, another
model, known as the “Eshelby method” or “equivalent inclusion method” is
applicable. This model applies when the particles are not in contact with each
other and are uniformly dispersed. The fundamental idea of this model was pro-
posed in 1957 by Eshelby [7] and further developed by Mura, Mori, Taya and other
researchers. This model has been omitted from this book, because the model is very
complicated, and beyond the level of this book. The author refers readers who are
interested in this model to the books by Clyne and Withers [6] and Taya and
Arsenault [8].

6.1.2 Strength of Composites

6.1.2.1 Factors Which Influence Strength of Composites

High strength is typically a target of composite materials development. However,


there are many factors which influence the strength of composites, and it is very
difficult to estimate the strength of a composite precisely. In a fiber reinforced
composite, the strength of fibers has a distribution and is not uniform. Many factors
including fiber arrangement and distribution, chemical reaction products at the
fiber/matrix interface and crack propagation after crack formation influence the
strength of a composite.
6.1 Mechanical Properties 121

Fig. 6.6 Probability density function f (s) and cumulative distribution function F (s)

6.1.2.2 Distribution of Fiber Strength

High strength fibers such as SiC, alumina, boron, carbon or glass fibers are used to
strengthen metals. When a stress is applied to these fibers and the stress reaches the
limit of the strength of a fiber, the fiber fractures in a brittle manner. It is known that
the fracture strength of brittle materials shows a Weibull distribution [9]. When the
number of tests carried out on the fibers N is sufficiently large, the number of fibers
fractured at any particular stress s divided by N shows a distribution curve similar
to that shown in Fig. 6.6a. This curve is the “probability density function” f (s), and
is expressed by:
ð1
f ðsÞds ¼ 1: (6.26)
0

Therefore, the hatched area between “a” and “b” in Fig. 6.6a corresponds to the
proportion of fibers which fractured at stresses between a and b. Then, N  this area
is the total number of fibers which fractured at stresses between a and b. The
hatched area in Fig. 6.6b gives the proportion of fibers which fractured at or
below stress sa. This is expressed by F (s) which is called the “cumulative
distribution function”. By using the expression s instead of sa, the relationship
between f (s) and F (s) becomes:
ðs
FðsÞ ¼ f ðsÞds: (6.27)
0

F (s) is expressed by:

FðsÞ ¼ 1  e’ðsÞ : (6.28)

By choosing a suitable function for ’(s), most distribution functions can be


expressed by Eq. (6.28) [10–12]. For example, the Weibull distribution is expressed
as:
m
FðsÞ ¼ 1  eðss1 Þ =s0
: (6.29)
122 6 Properties of Composites

Fig. 6.7 Weibull


distributions for different
values of m parameter

Therefore, the corresponding probability density function is:

mðs  s1 Þm1 ðss1 Þm =s0


f ðsÞ ¼ e ; (6.30)
s0

where m is the shape parameter, s0 the scale parameter and s1 the location parame-
ter. In this case, when the stress s is lower than s1, the fiber does not fracture. m is the
most important parameter, and is usually called the “Weibull parameter”.
The shape of the distribution depends on the m value, as shown in Fig. 6.7. When
m ¼ 1, the distribution is an ordinary exponential distribution. When m ¼ 2, a peak
appears in the distribution curve. The peak becomes sharper and higher with
increasing values of m. This means that as the value of m increases, the probability
of fracture becomes high near the peak value and the peak approaches s ¼ 2 (in this
case, since s0 ¼ s1 ¼ 1). Further, the probability of fracture becomes very small at
places far away from the peak value of s. In summary, the width of the distribution
becomes narrow and the fiber strength value becomes more reliable.
Most similar phenomena (for example the strength distributions of metals)
usually show a normal distribution, which is symmetric about the peak value.
In contrast, the Weibull distribution is asymmetric. The mean m and the standard
deviation s for the Weibull distribution are:
ð1  
1=m 1
m¼ sf ðsÞds ¼ s0 G 1þ ; (6.31)
0 m
6.1 Mechanical Properties 123

Fig. 6.8 Relationship


between fracture stress s and
F (s) of sintered silicon
nitride. (Circle) 2 mm deep
scratches parallel to the
tension direction; (square)
2 mm deep scratches normal
to the tension direction; and
(cross) 6 mm deep scratches
parallel to the tension
direction

and
    1=2
1=m 2 1
s ¼ s0 G 1þ  G2 1 þ ; (6.32)
m m

where GðxÞ is the gamma function:


ð1
GðxÞ ¼ px1 ep dp: (6.33)
0

The probability of non-breakage of a particular fiber at a particular stress is 1–F


(s), which is, from Eq. (6.29):
m
1  FðsÞ ¼ eðss1 Þ =s0
: (6.34)

Taking the natural logarithm of Eq. (6.34) twice and rearranging gives:

1
ln ln ¼ m lnðs  s1 Þ  ln s0 : (6.35)
1  FðsÞ

This means that if we plot a graph with the independent variable s as the abscissa
and the left-hand side of Eq. (6.35) as the ordinate, we expect to obtain a straight
line with gradient m. The linearity of this graph means that the distribution is in fact
a Weibull distribution. Examples of Weibull distributions are shown in Fig. 6.8
(S. Ito, private communication, 1995). This graph shows the relationship between
the fracture stress s and F (s) for bending tests on sintered silicon nitride samples
with differences in surface roughness. In this graph, ○: 2 mm deep scratches parallel
124 6 Properties of Composites

to the tension direction; △: 2 mm deep scratches normal to the tension direction;


and : 6 mm deep scratches parallel to the tension direction. The Weibull
parameters for samples ○, △ and  are m ¼ 19, 33 and 34, respectively. Although
the Weibull parameter of sample  is large, the strength of this sample is low.
The m value of fine ceramics like sintered silicon nitride is usually about 20.
By comparison, if we were to apply a Weibull distribution to the strength of
metal matrices (without reinforcements), the m value would be 30–40.
Alternatively, supposing that the diameter of a fiber is uniform, the probability of
fracture will depend on the length of the fiber. Therefore, the following probability
density function which includes the fiber length L has been proposed [13]:
m
f ðsÞ ¼ kLmsm1 ekLs ; (6.36)

where k is a constant and s1 ¼ 0 in this case. The mean m and standard deviation
s of this distribution are given by:
 1=m  
1 1
m¼ G 1þ ; (6.37)
kL m

and
 1=m     1=2
1 2 1
s¼ G 1þ  G2 1 þ : (6.38)
kL m m

It is known that the m value for glass fibers is about 11, and those of boron and
SiC fibers are 2.7–5.8 [13]. A comparison with Fig. 6.7 shows that the distribution
of the strength of fibers such as SiC fiber is quite wide.
As discussed above, since the strength distribution of fibers has a finite width, the
strength distribution of composites must also have a finite width.

6.1.2.3 Continuous Fiber Reinforced Composites

In this section, to make the discussion simple, we assume that the fracture strength
of fibers is uniform and does not have a distribution, although in reality there is
some distribution in the fracture strength. In addition, the bonding between the fiber
and the matrix is assumed to be perfect. When a continuous fiber reinforced
composite in which fibers are aligned in one direction is subjected to a tensile test
parallel to the fiber direction, the rule of mixtures [Eq. (1.5)] is applicable and the
strength is given by:

sc ¼ sff Vf þ s0m ð1  Vf Þ; (6.39)


6.1 Mechanical Properties 125

Fig. 6.9 Relationship


between tensile strength sc
and Vf of continuous fiber
reinforced metal

where sc is the strength of the composite, sff is the fracture strength of the fiber, Vf is
the volume fraction of the fiber, and s0 m is the stress which the matrix shares when
the stress of the fiber reaches its fracture stress. Usually, ceramic fibers deform
elastically and their strain is small, because they are brittle materials. However,
matrix metals are ductile materials and their elongation is large. This means that the
fiber reaches its fracture stress in the composite before the matrix metal reaches its
fracture stress, as shown in Fig. 6.9. su in the figure is the fracture stress of the
matrix metal and s0 m is usually lower than su. If the fracture strain of the fiber is
equal to that of the matrix metal, s0 m coincides with su, and in Eq. (6.39) s0 m should
be replaced with su. When Vf is smaller than some value Vmin, the composite does
not fracture even after the fibers have fractured. The composite deforms continu-
ously with work hardening and finally fractures at a higher stress than that obtained
using Eq. (6.39). In this case, the fracture strength of the composite is given by:

sc ¼ su ð1  Vf Þ: (6.40)

This equation shows that the fracture strength of the composite is given only by
the matrix strength, when Vf is smaller than Vmin. Therefore, this result shows that Vf
should be larger than the minimum fiber volume fraction Vmin to get a strengthening
effect. Vmin is given by the intersection of the two lines for sc, as shown in Fig. 6.9,
and can be calculated as:

su  s0m
Vmin ¼ : (6.41)
sff þ su  s0m

In addition, it is obvious from Fig. 6.9 that to obtain higher strength than that of
the matrix metal, Vf should be a higher volume fraction than Vcrt, which is given by:
126 6 Properties of Composites

su  s0m
Vcrt ¼ : (6.42)
sff  s0m

The strength of continuous fiber reinforced metal matrix composites has been
discussed comprehensively by Kagawa [14].

6.1.2.4 Discontinuous Fiber Reinforced Composites

The fracture strength of a composite reinforced with short fibers which are aligned
uniformly as shown in Fig. 6.3a is given by:

Vf þ s0m ð1  Vf Þ;
sc ¼ s (6.43)

where it is assumed that bonding between the fiber and the matrix is perfect. The
average strength s of the fiber will be obtained in the following discussion, which is
based on reference [5].
As mentioned in our discussion of the elastic modulus of short fiber reinforced
composites, and shown in Fig. 6.5c, the shear stress t (r0) applied to the surface of
the fiber by the matrix has maximum magnitude at both ends of a short fiber. t (r0)
at the end of the fiber increases as the outside load increases. However, t (r0) never
exceeds the yield shear stress ty of the matrix metal. If t (r0) reaches ty, the matrix
metal deforms plastically holding t (r0) ¼ ty. In this case, it is assumed that the
shear strength (bonding strength) between the fiber and the matrix is higher than ty.
Then, even if the shear stress reaches ty, the matrix will not debond at the interface
between the fiber and the matrix. The shear stress t (r0) curves are schematically
shown in Fig. 6.10a. One curve is for the shear stress when yielding has occurred
between 0 and a, and the other curve is for the shear stress before yielding. The
stress distribution of the fiber is also shown schematically in Fig. 6.10b. Since the
matrix metal has yielded between x ¼ 0 and x ¼ a, a constant shear stress ty is
given to the fiber between 0 and a. From Eq. (6.21), the stress in the fiber is given
by:

2ty x
s¼ ð0<x<aÞ (6.44)
r0

As the load applied to the fiber increases, the value a of the fiber expands and the
stress s of the fiber in Fig. 6.10b becomes high. If s exceeds sff before a reaches l/2,
the fiber will break at some place in the central area (a < x < l–a). When a reaches
l/2 and s is lower than sff, the fiber will not break and the matrix metal will fracture
after plastic deformation. Therefore, there is a minimum length lc at which the stress
s reaches sff and the fiber fractures. If the fiber length is shorter than lc, the fiber will
not break. We assume that the stress s in the fiber in Fig. 6.10b reaches sff when a is
equal to l/2, namely x ¼ l/2. Then, by replacing l with lc in Eq. (6.44), we have:
6.1 Mechanical Properties 127

Fig. 6.10 Stress distributions


under applied tensile stress
(a) distribution of shear stress
at fiber/matrix interface and
(b) distribution of stress in
short fiber

r0 sff
lc ¼ : (6.45)
ty

This lc is called the “critical length” by Kelly and Macmillan [5]. Further, the
critical aspect ratio is given by:

lc sff
¼ ; (6.46)
d 2ty

where d is the diameter of the fiber. This result means that if the fiber length is
longer than lc, the fiber shares the stress in the composite and the composite
fractures when the fiber breaks. If the fiber is shorter than lc, the fiber will not
break until the composite fractures.
Alternatively, when the stress applied to the central part of a fiber reaches the
fracture stress sff, the average stress which the fiber shares is lower than sff . When
the composite fractures, the fiber shares this average stress s. The following method
was used by Kelly and Macmillan [5] to obtain s  . As shown in Fig. 6.11, as the
matrix metal is yielded from x ¼ 0 to x ¼ lc/2, the stress of the fiber s increases
linearly and reaches sff . It is assumed that the stress s of the fiber remains constant
at sff from x ¼ lc/2 to x ¼ l  lc/2. Therefore, we obtain a trapezium-shaped stress
curve as shown in Fig. 6.11. s  corresponds to the area under the trapezium, divided
by the fiber length, and is given by:
 
lc
 ¼ sff 1 
s : (6.47)
2l
128 6 Properties of Composites

Fig. 6.11 Assumed stress


distribution in a short fiber

Fig. 6.12 Tensile strength of


composite as a function of Vf
for different values of fiber
length

By substituting Eq. (6.47) into Eq. (6.43), we obtain:


 
lc
sc ¼ sff 1  Vf þ s0m ð1  Vf Þ; (6.48)
2l

where this equation is applicable when Vf > Vmin. When the fiber length l becomes
infinitely large, Eq. (6.48) will coincide with Eq. (6.39) which gives the strength of
continuous fiber reinforced composites. Then, the relationship between the strength of
the composite sc and Vf is shown in Fig. 6.12, for different values of fiber length l. It is
obvious from Eq. (6.48) that when l ¼ 10lc, the strength of the short fiber reinforced
composite becomes about 95 % of that of the continuous fiber reinforced composite.

6.1.2.5 Particulate Reinforced Composites

The concepts used to discuss the strengthening mechanisms for continuous or


discontinuous fiber reinforced composites do not apply to particulate reinforced
6.1 Mechanical Properties 129

composites where the effect of strengthening is based on the shape of reinforcement.


If we were to apply the strengthening mechanisms suggested for discontinuous fiber
reinforced composites to composites containing particles we would find that appar-
ently no strengthening can be expected from particles.
Alternatively, if hard, fine particles are dispersed uniformly in a metal, the
metal will be “dispersion strengthened”. The strengthening mechanism, proposed
by Orowan, is based on restricting the movement of dislocations and is known as
“Orowan looping” or the “Orowan bypass mechanism” [5]. If a particle is
sufficiently hard, it cannot be sheared easily by a dislocation. Instead, the dislo-
cation often bows around the particle. The dislocation is able to bow out between
particles, but the bowing process is resisted by the dislocation tension. Atoms at
dislocations have insufficient bonding (like atoms at grain boundaries as
discussed in Sect. 1.3 of Chap. 1). This means that atoms at a dislocation line
are at a higher energy level than other atoms. The bowing of the dislocation
creates more dislocation line and increases the energy of the system. Eventually,
the dislocation is able to continue gliding, leaving behind a circle of dislocation
around the particle. These processes resist the dislocation movement and contrib-
ute to strengthening of the metal.
The yield strength sy of the metal is approximately:

Gm b
s y ¼ s0 þ ; (6.49)
l

where s0 is the yield strength of the matrix metal, Gm is the shear modulus of the
matrix metal, b is the Burgers vector, and l is the average interparticle spacing [15].
This equation shows that the yield strength sy of the metal increases with decreas-
ing l. We can determine the relationship between l and Vf, using the following
model. Assume that the distribution of particles is a simple cubic system with one
lattice point on each corner of the cube. Each corner is occupied by a particle. In this
case, the volume fraction Vf is given by:

 3
p d
Vf ¼ ; (6.50)
6 dþl

where d is the diameter of a particle. The relationship between 1/l and Vf for
various values of d is shown in Fig. 6.13 [16]. In the case of d ¼ 0.1 mm, when Vf is
0.02, 1/l is approximately 5. Although the volume fraction of particles is very low,
we can obtain a high strengthening effect. However, when the particle diameter is
larger than 1 mm, the strengthening effect 1/l increases gradually with increasing
Vf. When d is larger than 3 mm, the strengthening effect 1/l is very small up to
around Vf. ¼ 0.3. Therefore, to attain a high strengthening effect, the particles must
be very fine and hard.
130 6 Properties of Composites

Fig. 6.13 Relationship


between Vf and average inter-
particle spacing l for
different values of particle
diameter d

Fig. 6.14 Three typical true


stress–true strain curves:
(a) brittle material,
(b) composite, (c) ductile
material

6.1.3 Fracture of Composites

Figure 6.14 compares typical examples of the relationship between true stress and
true strain for three common types of materials. Curve “a” is that of a brittle
material such as a ceramic; curve “c” is that of a ductile material like aluminum,
and curve “b” is the stress–strain curve of a material with the properties intermedi-
ate between brittle and ductile. A metal matrix composite with a very high Vf shows
a stress–strain curve similar to “a”, because most of the load is taken by the
reinforcements such as ceramic fibers or carbon fibers. When Vf is relatively low,
the composite shows a curve similar to “b”.
The area under the curves corresponds to the energy absorbed up until the
material fractures (shown as the hatched area for curve “c”). For curve “c”, the
absorbed energy is large. In contrast, the absorbed energy for curve “a” is very
small, although the fracture strength is high. Therefore, if a large impact is applied
to a structure made of material “a”, the structure cannot absorb the impact energy
and will break instantly. However, if the structure is made of material “c”, the
6.1 Mechanical Properties 131

structure will be deformed plastically by the impact, absorbing most of the energy,
although the fracture strength of the material “c” is not high. In this latter case,
rupture of the structure will not propagate to other parts of the system.
The hatched area shown for curve “c”, the absorbed energy E, can be expressed by:

E ¼ gA þ W; (6.51)

where A is the total surface area created by deformation and fracture, g is the surface
energy of this material, W is the energy consumed through plastic deformation,
dislocation multiplication, internal stress and other mechanisms. The difference in g
values between brittle materials and ductile materials is small, as shown in
Table 1.2. This means that if we use the same shaped test pieces, the difference
in gA between material “a” and material “c” will not be large. Therefore, the major
source of the difference in the absorbed energy between material “a” and material
“c” is the value of W.
The science which tackles fracture phenomena of materials is “fracture mechan-
ics”. It is thought that the real science of fracture mechanics started with Griffith
[17]. Griffith thought that the term gA in Eq. (6.51) was dominant in the energy
absorbed during the fracture of brittle materials. However, as is clearly shown by
comparing “a” with “c” in Fig. 6.14, the term W is dominant in the energy absorbed
during fracture of ductile materials.
Alternatively, the stress intensity factor (K value) was proposed by Irwin et al.
to explain crack initiation and propagation at the edge of the crack [18, 19].
In addition, the fracture of ductile materials is discussed in nonlinear fracture
mechanics using the “J-integral”. As these fracture mechanics are beyond the
scope of this book, monographs by Fujii and Zako [20] or Yokobori [21] are
recommended to readers who are interested in fracture mechanics.
The fracture of composites is a more complicated phenomenon than the fracture
of homogeneous materials. Fracture of composites during testing can be roughly
classified into static failure, failure by impact and failure by fatigue.

6.1.3.1 Static Failure

This category includes fracture of composites at low strain rate (during tensile
tests). Interesting experiments have been performed by Ochiai et al. [22]. A copper
sample reinforced by a single tungsten fiber was subjected to tensile tests and the
results are shown in Fig. 6.15. This sample corresponds to the case Vf < Vmin, as
discussed in Sect. 6.1.2.4. First, the fiber broke in one place during a tensile test, and
the stress dropped suddenly. But, the stress recovered quickly, elongation
continued, and the matrix metal work hardened without fracture of the composite.
Fiber breakage occurred many more times, and each breakage produced a sharp
drop in the stress as seen in the stress–strain curves. Finally, fracture of the
composite occurred. The appearance of the sample is shown in the inset to the
graph, where many breakages of the fiber can be seen. This is called “Multiple
132 6 Properties of Composites

Fig. 6.15 Load–strain curve


of copper reinforced with a
single tungsten fiber, showing
multiple fracture [26]
(courtesy of Prof. S. Ochiai)

Fig. 6.16 Schematic of fiber


pullout near the crack front of
a continuous fiber reinforced
brittle matrix composite

Fiber

fracture”. To relate their experiments to the fracture of composites, Ochiai et al.


clarified the relationship between the fracture behavior, the aspect ratio of fiber and
Vf using Monte Carlo simulations and found good agreement with experimental
results [22].
Alternatively, a crack in a continuous fiber reinforced brittle matrix composite is
schematically drawn in Fig. 6.16. The figure shows the pullout of broken fibers at
the crack. As a result of the pullout of broken fibers, the fiber/matrix interface has
been turned into newly formed fiber surface and matrix surface. This means that A
in Eq. (6.51) increases and the absorbed energy also increases during fracture of the
composite. The fracture toughness of brittle matrix composites can be improved by
increasing A in Eq. (6.51). Furthermore, the friction between fiber and matrix
during fiber pullout also improves the toughness of the composites.
As it is very difficult to discuss theoretically the fracture of composites
reinforced with many fibers, the fracture of those composites has been investigated
by Ochiai et al. using Monte Carlo simulation [23–25].
In the case of particulate reinforced metal composites, the fracture mode is
different. Williams et al. showed that if the strength of a matrix is very high
(as found for peak-aged Al–Cu alloy reinforced with SiC particles) the main
damage mode is SiC particle fracture [26].
6.1 Mechanical Properties 133

6.1.3.2 Failure by Impact

The fracture of composites at high strain rates belongs in this category. Impact
fracture is usually studied using either a Charpy impact tester or an Izod impact
tester. Few papers on the fracture of metal matrix composites by impact have been
published, though the energy absorbed during fracture is an important measure of
the reliability and the toughness of composites. The fracture phenomena of alumi-
num alloys reinforced with SiC whiskers or particles were reported by Kobayashi
et al. [27, 28]. The amount of energy absorbed during fracture decreased steeply
with increasing Vf of SiC whiskers or particles in the aluminum alloy. Especially,
for the case of A6061 matrix, the absorbed energies of the composites containing 10
or 20 wt.% SiC particles decreased to 1/4 and 1/5 of that of the matrix metal,
respectively. Furthermore, it was reported by Kagawa that the absorbed energy of
continuous boron fiber reinforced aluminum composites decreased to 1/5 of that of
the matrix metal [14]. Therefore, these results show that an increase in the term gA
in Eq. (6.51) in an aluminum alloy matrix is not expected for any shape of
reinforcement (particle, whisker or continuous fiber). W in Eq. (6.51) should be
increased to increase the absorbed energy. However, because reinforcements act as
obstacles to dislocation motion, and dislocation motion is needed for plastic
deformation of the matrix, the plasticity of the composite will decrease when
reinforcements are added, and W will decrease as well. Therefore, whenever we
use ceramic reinforcements in ductile metal matrices, a decrease in the toughness is
inevitable.

6.1.3.3 Failure by Fatigue

When a material is subjected to cyclic loading such as bending, twisting,


compressing or stretching, fatigue of the material occurs and the material can
eventually fracture. Ferrous alloys such as steel and cast iron do not fracture
below a certain stress amplitude, even after many cycles. This distinct stress
amplitude is called the “fatigue limit”. Other metals, such as aluminum and copper,
do not have such a distinct stress amplitude and will fracture even under small stress
amplitude after many cycles. For such metals, usually 107 cycles is chosen as the
limiting number of cycles, and the highest stress amplitude below which the metal
does not fracture even after 107 cycles is chosen as the fatigue limit. The highest
stress S is plotted against the logarithm of the number of cycles until failure N. This
relationship is called the “S–N curve”.
When a stress is applied to an aluminum composite reinforced with ceramic
fibers, the strain of the aluminum (which has a low Young’s modulus) is large and
the strain of the fiber (which has a high Young’s modulus) is small. Young’s moduli
of some metals and ceramics are shown in Table 6.1; Young’s moduli of ceramics
are usually about five times larger than those of metals. Thus, when the same
magnitude of stress is applied to ceramic fibers and matrix metal, the metal deforms
134 6 Properties of Composites

Table 6.1 Comparison of Young’s moduli of metals and ceramics


Material Young’s modulus (GPa) Shear modulus (GPa)
Metals Al 72 26
Mg 44 17
Cu 136 48
Fe 190 80
Ti 114 40
Ceramics Al2O3 460 147
SiC 560 –
C (fiber) 200–600 –
B (fiber) 380–400 –
9Al2O32B2O3 392 –
(whisker)
Si3N4 372 –
MgO 245 122
SiO2 94 29
TiC 450 –
TiN 600 –

much more, and the matrix might become detached from the fibers at the fiber/
matrix interface. In addition, if there is a fine defect at the interface, the defect may
be an origin for fatigue crack initiation; the difference in Young’s modulus between
the matrix and the fiber might then promote fatigue crack propagation. This implies
that composites might be vulnerable to cyclic loading. However, if Vf of fibers is
relatively high and the fibers have a rigid structure in the composite such that the
matrix metal is compressed, the composite may resist cyclic loading.
Figure 6.17 shows S–N curves of composite plate type specimens subjected to
cyclic bending stress. These specimens are alumina short fiber reinforced Al–11 wt.%
Si–1 wt.% Cu–1 wt.% Mg–1 wt.% Ni alloy (JIS AC8A) composites which were
produced by squeeze casting [29]. When the applied stress S is high, the number of
cycles to failure N is small. When S is low, N is large. The S–N curves plotted on a
logarithmic scale for N are almost linear. As the Vf of alumina short fibers increases,
the S–N curve is shifted toward higher stresses, and the fatigue limit increases. The
fatigue limit of the composite with Vf ¼ 20 % is about 60 % higher than that of the
matrix metal (i.e., the Vf ¼ 0 % curve plotted on the graph). However, it has been
reported that when Vf in an aluminum alloy composite was lower than 8 %, the fatigue
limit became lower than that of its matrix [30]. The increase in fatigue limit for the
Vf ¼ 20 % composite is associated with an increase in the elastic modulus of
the material.
Generally, alumina short fiber products include some non-fibrous “shot”. The
chemical composition of the non-fibrous material is the same as that of the fiber and
the material is formed during the production of the fiber. An example of the shot is
shown in Fig. 6.18. This kind of shot affects the fatigue properties of the
composites. The content of shot having a diameter over 100 mm is less than
0.1 % of the fiber volume in the composites of Fig. 6.17. The fracture surfaces of
6.1 Mechanical Properties 135

Fig. 6.17 Influence of Vf on S–N curves of alumina short fiber reinforced JIS AC8A alloy
(Al–12 wt.% Si–1 wt.% Cu–1 wt.% Mg–1 wt.% Ni)

Fig. 6.18 Optical microstructures of alumina short fiber reinforced JIS AC8A alloy (Al–12 wt.%
Si–1 wt.% Cu–1 wt.% Mg–1 wt.% Ni).

these test pieces are shown in Fig. 6.19. Since some large pieces of shot are
observed in these fracture surfaces, some of these pieces may be origins of crack
initiation. In addition, as many fibers parallel to the fracture surface are observed,
fatigue cracks tend to propagate along the interface between the fibers and the
matrix.
The fatigue crack growth rate of the same composites was measured and is
shown in Fig. 6.20. CT (compact type) specimens, compliant with ASTM E647-83,
Fig. 6.19 Fracture surfaces of alumina short fiber reinforced JIS AC8A alloy after fatigue tests.
Samples are the same as in Figs. 6.17 and 6.18

Fig. 6.20 Relationship between fatigue crack growth rate and stress intensity range DK of
alumina short fiber reinforced JIS AC8A alloy during fatigue tests
6.1 Mechanical Properties 137

were used for the measurements. The ordinate is the crack growth rate da/dN and
the abscissa is the stress intensity range DK. a is the length of the crack. The
relationship between DK and the applied load is given by:

DK ¼ f ðaÞðPmax  Pmin Þ; (6.52)

where f (a) is determined by the shape of the test piece and the length a of the crack.
Pmax is the applied maximum load, and Pmin is the minimum load. Paris et al.
derived the following relationship:

da
¼ CðDKÞm ; (6.53)
dN

where C and m are constants [31]. If we compare the crack growth rate da/dN of
the composites at a particular value of DK, the da/dN of the matrix metal is the
smallest and da/dN increases with increasing Vf of alumina short fiber. The m
value of the composites is almost the same as that of the matrix metal. The C value
of the composite with Vf ¼ 20 % is about 100 times larger than that of the matrix
metal, which means that a crack in the composite grows about 100 times faster
than in the matrix. It is obvious that if there is a crack in the aluminum alloy
composite, the crack will grow quickly and the composite will fracture in a short
time. In contrast, Masuda et al. showed that the crack growth rate for SiC whisker/
A2024 alloy and SiC particle/A357 alloy composites was lower than that of the
matrix metal [32]. These composites were prepared by a powder metallurgical
process. The crack growth rate seems to depend on the shape and size of rein-
forcement, the matrix metal, the production process etc. [33]. The results in
Fig. 6.20 show that Paris’ relationship is not only applicable to composites but
also to the aluminum alloy matrix.
Alternatively, the path of a crack observed by optical microscopy is shown in
Fig. 6.21. Figure 6.21a shows a composite with Vf ¼ 20 % and (b) shows the matrix
metal. When the crack reaches a fiber at an angle of 45–90 to the crack, the crack
seems to intersect the fiber, thus breaking it. When the angle of approach is lower
than 45 , the crack tends to advance along the surface of the fiber. This is consistent
with the observation in Fig. 6.19 of many fibers parallel to the fracture surface.
Fatigue of particulate reinforced composites has been researched by Chawla
et al. [34], who showed that the cyclic strength of a SiC particle/A2080 aluminum
alloy composite was significantly affected by the microstructure of the matrix.
In addition, it was reported that, during cyclic loading, many very fine cracks
(micro-cracks) were formed in front of the main crack [27, 35]. The fatigue crack
seems to grow by the coalescence of micro-cracks with the main crack. Toda et al.
have used computer simulations to clarify the propagation of the fatigue crack,
taking into account the formation mechanism via these micro-cracks, with the aim
of improving the toughness of composites [36, 37].
138 6 Properties of Composites

Fig. 6.21 Crack paths observed by optical microscopy. (a) 20 vol.% alumina short fiber
reinforced JIS AC8A alloy, and (b) matrix alloy only

6.2 Physical Properties

6.2.1 Density

Here we discuss the density of a composite with mass mc, volume vc and density rc.
The mass, the volume and the density of the matrix metal are mm, vm and rm,
respectively, while those of the reinforcements of the composite are mi, vi and ri
(i ¼ 1, 2, 3, . . .). Then, we have the following relationships:

mc ¼ mm þ m1 þ m2 þ       (6.54)

and

vc ¼ vm þ v1 þ v2 þ       (6.55)

By dividing Eq. (6.54) by mc and Eq. (6.55) by vc, we have:

1 ¼ Mm þ M1 þ M2 þ       (6.56)

and

1 ¼ Vm þ V1 þ V2 þ     ; (6.57)

where Mi is mass fraction and Vi is volume fraction. The density of the composite is
given by:

mc mm þ m1 þ m2 þ     
rc ¼ ¼ ¼ rm Vm þ r1 V1 þ r2 V2 þ       (6.58)
vc vc

When there is only one type of reinforcement, the density is given by:

rc ¼ rm Vm þ r1 V1 : (6.59)
6.2 Physical Properties 139

This equation is the same as that obtained by direct application of the rule of
mixtures [Eq. (1.5)]. If m2 is voids or pores, then m2 ¼ r2 ¼ 0 but V2 ¼ 6 0 if the
composite includes voids. In this case, from Eq. (6.57), the void fraction V2 is given by:

V2 ¼ 1  ðVm þ V1 Þ; (6.60)

where vm ¼ mm/rm, v1 ¼ m1/r1, Vm ¼ vm/vc and V1 ¼ v1/vc. We obtain the following


equation:
   
vm v1 1 mm m1
V2 ¼ 1  þ ¼1 þ : (6.61)
vc vc vc rm r1

The void fraction in the composite can be calculated using Eq. (6.61).
When the precise volume fraction of reinforcement is unknown, the volume
fraction can be determined from the measured density of the composite using
Eq. (6.58). Archimedes’ principle is often used to measure the density of
composites. To do this, the weight Wa in air and the weight Wl in a liquid of density
rl are measured. Neglecting the buoyant force of air, the density of the composite is
given by:

Wa rl
rc ¼ : (6.62)
Wa  Wl

Usually, water is used as the liquid. Care needs to be taken because the density of
water changes significantly with temperature.

6.2.2 Specific Heat

The constituents of a composite are distributed in the composite whilst retaining


their own microstructures. If the constituents are homogeneously distributed, the
specific heat of the composite can be calculated from the amounts of each of the
constituents included in unit volume of the composite. We consider unit volume of
a fiber reinforced composite with density rc. The composite absorbs heat qc and the
temperature rises DT. The temperature rise of the fibers and the matrix will also be
DT. If we denote the heats absorbed by the fiber and the matrix as qf and qm,
respectively, the following relationships hold:

qc ¼ rc Cc DT; qf ¼ Vf rf Cf DT; qm ¼ ð1  Vf Þrm Cm DT; and


qc ¼ qf þ qm ; (6.63)
140 6 Properties of Composites

where Cc, Cf and Cm are the specific heats of the composite, fiber and matrix,
respectively, and rf and rm are the densities of the fiber and the matrix. The specific
heat of the composite is given by:

1n o
Cc ¼ Vf rf Cf þ ð1  Vf Þrm Cm : (6.64)
rc

Although very small gas voids are formed during fabrication of the composite,
the specific heat of the gas will be small and have little effect. However, if some
chemical reaction occurs between the fiber and matrix, we cannot use Eq. (6.64),
because the specific heat of the reaction products is usually quite different from that
of the constituents.

6.2.3 Thermal Conductivity

Consider one-dimensional heat flow in a composite bar with uniform rectangular


cross-section, where the sides of the bar are thermally insulated. The heat flow qc at
a specific point x of the bar (the x axis is taken parallel to the heat flow direction) is
proportional to the temperature gradient dT/dx at the point. This is given by
Fourier’s law:

dT
qc ¼ lc ; (6.65)
dx

where the proportionality constant lc is the thermal conductivity of the composite,


and T is the temperature at the specific point x. If the length of the bar is L and
the temperatures of both ends of the bar are kept at T0 (x ¼ 0) and T1 (T0 > T1) for
a long time, a steady-state temperature distribution will be attained when a
constant heat flow through the bar maintains the temperature difference (T0–T1).
Then, the thermal conductivity can be calculated by measuring the heat flow qc
through the bar.
First, we consider the thermal conductivity lcL when heat flows along the
direction of the fiber axis in a continuous fiber reinforced composite as shown in
Fig. 6.22a. Using Eq. (6.65), the heat flows qf in the fiber, qm in the matrix and qc in
the composite are given by:

lf ðT1  T0 Þ lm ðT1  T0 Þ lcL ðT1  T0 Þ


qf ¼  ; qm ¼  ; qc ¼  (6.66)
L L L

where lf and lm are the thermal conductivities of the fiber and the matrix. When the
heat flow is in a steady state, the temperature gradient is linear. Then, in this model,
the temperature gradient in the fiber is the same as that of the matrix, and no heat
flows from the matrix to the fiber. The proportion of the top surface of the
6.2 Physical Properties 141

Fig. 6.22 Heat flows in a a qc b qc


continuous fiber reinforced Fiber
Fiber
composite. (a) Parallel to T0 T0
the fiber axis direction and
(b) normal to the fiber axis
direction; assuming that heat
flows linearly

L
X
T1 T1
qc qc

composite occupied by fiber is Vf as shown in Fig. 6.22a, and the matrix occupies
(1–Vf.). The heat which flows into or out of the composite per unit area and unit time
is given by:

qc ¼ qf Vf þ qm ð1  Vf Þ: (6.67)

By substituting Eq. (6.66) into Eq. (6.67), we obtain:

lcL ¼ lf Vf þ lm ð1  Vf Þ: (6.68)

This equation shows that when heat flows parallel to the fiber axis direction, the
thermal conductivity of the composite can be obtained by direct application of
the rule of mixtures. If the fiber has a higher thermal conductivity than that of the
matrix, it is easy to hold the matrix at a particular temperature T0. However, it is not
easy to maintain the temperature of the fiber at T0, because heat flows well in the
fiber, and the temperature of the fiber tends to fall below T0. Therefore, experimen-
tal results may vary slightly from that of Eq. (6.68).
Next, we consider the thermal conductivity lcT when heat flows normal to the
direction of the fiber axis in the composite as shown in Fig. 6.22b. In this case, we
assume that at a steady state, heat flows linearly in the composite as indicated by
the heat flow lines in Fig. 6.22b, although, in actual fact, heat tends to flow towards
the material having higher thermal conductivity. If we consider one heat flow line
which goes linearly through the matrix and the fiber, the total lengths of the heat
flow line in the fiber and in the matrix will be LVf and L(1–Vf), respectively. If DTf is
the temperature drop over the total lengths of the fiber in the heat flow line and DTm
is the total temperature drop across the matrix in the heat flow line, because the
amount of heat qc in the fiber is the same as in the matrix, we have:

DTf DTm
qc ¼ lf ; and qc ¼ lm : (6.69)
LVf Lð1  Vf Þ
142 6 Properties of Composites

Alternatively, for the overall composite, we have:

DTf þ DTm
qc ¼ lcT : (6.70)
L

Substitution of Eq. (6.69) into Eq. (6.70) gives the thermal conductivity when
heat flows normal to the fiber axis direction:

lf lm
lcT ¼ : (6.71)
lf ð1  Vf Þ þ lm Vf

The idea behind this equation is also applicable to particulate reinforced or short
fiber reinforced composites. However, as mentioned above, since heat tends to flow
towards the material having higher thermal conductivity, the real thermal conduc-
tivity may be slightly higher than the values obtained from Eqs. (6.68) and (6.71).
In addition, there is thermal resistance at the fiber/matrix interface. The heat flow
between the fiber and the matrix is greatly affected by this thermal resistance, which
has been discussed by Hasselman et al. [38]. Non-steady-state heat conduction in
composites has been discussed by Ohta et al. [39, 40].

6.2.4 Thermal Expansion Coefficient

Atoms in a solid material vibrate at their positions. When the material is heated, the
amplitude of the vibration will increase, and the distance between any two atoms
will also increase. This is how thermal expansion occurs. If the temperature of the
material rises dT and the length of the material increases from L to L + dL, the
fractional increase in the length per unit change in temperature, a, is:
 
1 dL
a¼ : (6.72)
L dT

a is the thermal expansion coefficient (or the coefficient of linear thermal expan-
sion). The thermal expansion of materials is not only one-dimensional, but also
volumetric. Therefore, for the expansion of volume, the volumetric coefficient of
thermal expansion b is defined as:
 
1 dV
b¼ : (6.73)
V dT

a and b depend slightly on temperature. For particulate reinforced composites, if the


particles are distributed homogeneously, the composite expands almost uniformly
6.2 Physical Properties 143

in three dimensions, which means that within a limited temperature range, we have
the following relationship:

b ¼ 3a: (6.74)

Obviously, the thermal expansion coefficient of continuous fiber reinforced


composites depends on the fiber axis direction. Therefore, the equations discussed
above are not applicable to continuous fiber reinforced composites.
Alternatively, if we apply the rule of mixtures [Eq. (1.5)] directly to particulate
reinforced composites in which the particles are distributed homogeneously, the
thermal expansion coefficient ac is:

ac ¼ ap Vp þ am ð1  Vp Þ; (6.75)

where the subscripts c, p and m correspond to composite, particle and matrix,


respectively, and Vp is the volume fraction of particles.
Several theoretical models to obtain the thermal expansion coefficient of partic-
ulate reinforced composites have been proposed. Some of them: Turner’s equation
[41], Kerner’s equation [42] and Schapery’s equation [43] are introduced in this
section. Turner’s model is based on the idea that the internal stress sp arising from
the temperature change of the particles is in equilibrium with the internal stress sm
of the matrix, and that shear deformation can be neglected. The internal stress of the
particles and the internal stress of matrix are given by:

sp ¼ ðbc  bp ÞKp DT; and sm ¼ ðbc  bm ÞKm DT; (6.76)

where K is the bulk modulus. The Young’s modulus E, K, and the shear modulus G
have the following relationship:

E
K¼ : (6.77)
3ð3  E=GÞ

The balance between the internal stress of the particles (volume fraction Vp) and
that of the matrix (volume fraction (1–Vp)) is given by:

sp Vp þ sm ð1  Vp Þ ¼ 0: (6.78)

By substituting Eq. (6.76) into Eq. (6.78) and using bc ¼ 3ac, we have:

ap Vp Kp þ am ð1  Vp ÞKm
ac ¼ : (6.79)
Vp Kp þ ð1  Vp ÞKm

This is Turner’s equation [41].


144 6 Properties of Composites

Alternatively, the thermal volumetric expansion of a spherical particle covered


with a matrix shell was discussed by Kerner [42] and is given by:

4Gm ðKc  Kp Þðbm  bp Þ


bc ¼ bp Vp þ bm ð1  Vp Þ þ ; (6.80)
Kc ð4Gm þ 3Kp Þ

where Kc is given by:

Km ð1  Vp Þð3Kp þ 4Gm Þ þ Kp Vp ð3Km þ 4Gm Þ


Kc ¼ : (6.80a)
ð1  Vp Þð3Kp þ 4Gm Þ þ Vp ð3Km þ 4Gm Þ

Because it is possible to regard particulate reinforced composites as three-


dimensionally uniform, we can obtain ac by replacing bc with 3ac. This means
that the equation for ac can be obtained just by replacing bc with ac in Eq. (6.80)
to give:

4Gm ðKc  Kp Þðam  ap Þ


ac ¼ ap Vp þ am ð1  Vp Þ þ : (6.80b)
Kc ð4Gm þ 3Kp Þ

Further, Schapery [43] proposed that an equilibrium of the internal stress arising
from the difference in thermal expansion between the particles and the matrix is
attained when the free energy of the system is minimized. Schapery’s equation is
given by:

1Vp V
bc Km þ Kpp  K1c
¼ ap Vp þ am ð1  Vp Þ  ðam  ap Þ: (6.81)
3 1
Km  K1p

Schapery used the definitions proposed by Hashin et al. for Kc [44]. Kc in this
case is given by:

Vp
Kc;upp ¼ Km þ 1V
; (6.82)
1
Kp Km þ Km þ4Gpm =3

and

1  Vp
Kc;low ¼ Kp þ V
: (6.82a)
1
Km Kp þ Kp þ4G
p
p =3

Equation (6.82) gives the upper limit of the volumetric coefficient of thermal
expansion and Eq. (6.82a) gives the lower limit. The upper limit of Schapery’s
equation coincides with Kerner’s equation.
6.2 Physical Properties 145

Fig. 6.23 Values of thermal


expansion coefficient of SiC
particle reinforced aluminum
alloy—comparison of various
models with experimental
values from Lemieux et al.
[45]

The results calculated using Eqs. (6.79), (6.80b) and (6.81) are plotted in
Fig. 6.23, and compared with experimental results obtained at 473 K by Lemieux
et al. [45]. The particles and matrix are SiC and aluminum alloy. The data used
in the calculation were: Kp ¼ 227 GPa, Gp ¼ 192 GPa, ap ¼ 4.5  106 K1,
Km ¼ 67.2 GPa, Gm ¼ 25.1 GPa, and am ¼ 23.2  106 K1. The lowest curve
in Fig. 6.23 is the result given by Turner’s equation, which lies well below the
experimental data. The curve from Kerner’s equation (also the upper limit of
Schapery’s equation) is close to the curve obtained using the rule of mixtures.
The theoretical models mentioned above are based on the assumption that each
particle is isolated and that the particles are distributed homogeneously in the
matrix. However, the reinforcements for composites produced by pressure infiltra-
tion methods are usually a soft “skeleton-like” structure, because slightly sintered
preforms containing some binder are used. Therefore, the thermal expansion
coefficient of these composites, especially at Vf > 50 % is greatly influenced by
the skeleton and will be closer to the value of the thermal expansion coefficient of
the reinforcement than to any of the values predicted using the rule of mixtures or
the models discussed above.
Alternatively, for fiber reinforced composites, it is intuitive that the thermal
expansion coefficient parallel to the fiber axis direction will be very different from
that normal to the fiber axis direction. Furthermore, the coefficients of continuous
fiber reinforced composites will be different from those of short fiber reinforced
composites. Theoretical models for the thermal expansion coefficient of fiber
reinforced composites have been discussed by Schapery [43], Ishikawa et al. [46]
and Rosen et al. [47]. In addition, the thermal expansion coefficient when slipping
at the fiber/matrix interface occurs has been discussed by Jasiuk et al. [48]. This
model has been omitted in this section.
146 6 Properties of Composites

Fig. 6.24 Thermal expansion


and contraction of alumina
short fiber/aluminum
composites during heating
and cooling

In the case of continuous fiber reinforced composites, plastic expansion of the


matrix metal parallel to the fiber axis direction has been observed upon repeated
heating and cooling (i.e., thermal cycling). This phenomenon is known as a
“thermal ratchet” [49]. The mechanism can be explained as follows: when the
temperature of a composite rises, shear stress will be produced at the fiber/matrix
interface, because the thermal expansion of the ceramic fibers is relatively small. If
the shear stress increases beyond the bonding strength of the fiber/matrix interface,
slipping and shear deformation of the matrix will occur at the fiber/matrix
interface, because the matrix metal is usually weaker than the ceramic fiber. After
the composite has cooled down, the slipped and plastically deformed interface area
does not recover, but remains slipped and deformed. This deformed state will be the
starting point for the next heating cycle, and further slipping and plastic deforma-
tion of the matrix metal will occur on each subsequent heating and cooling cycle.
The accumulated slipping and plastic deformation will cause plastic expansion of
the matrix metal parallel to the fiber axis.
A similar phenomenon has been observed for short fiber reinforced composites
upon thermal cycling. Heating and cooling curves for alumina short fiber reinforced
aluminum composites are shown in Fig. 6.24 [50]. The composites contain
10–20 vol.% alumina short fibers with diameter of about 3 mm and length of
about 1 mm. The heating curve does not coincide with the cooling curve, and the
cooling curve does not return to the starting point. Thus, plastic deformation occurs
even in short fiber reinforced composites when the temperature amplitude of the
thermal cycling is large. This plastic deformation is significant when the fiber
orientation is uniform and only slight when the orientation is random.
6.3 Interface 147

Fig. 6.25 Experimentally


obtained dimension changes
of an alumina short fiber
reinforced aluminum
specimen during thermal
cycling between room
temperature and 743 K

The alumina short fibers in the composites shown in Fig. 6.24 are distributed
randomly in a plane, with few fibers normal to the plane. These composites were
subjected to thermal cycling between room temperature and 743 K. The dimension
changes of the specimens were measured after thermal cycling and are shown in
Fig. 6.25. The plane in which the fibers are distributed randomly is the “horizontal
plane” in Fig. 6.25; the plane normal to the horizontal plane is the “vertical plane”. The
dimensional change of the vertical plane upon thermal cycling is positive (i.e., the
sample expands in this direction). In the vertical direction, the aluminum/10 vol.%
alumina short fiber composite expands by about 1.5 %. Larger expansions are seen in
samples with lower Vf . The horizontal plane contracts upon thermal cycling, with the
aluminum/10 vol.% alumina short fiber composite contracting by about 0.4 %. As
before, the lower the Vf, the bigger the contraction. When the fiber content is 20 vol.%,
the dimensional changes are small, and only occur at the beginning of the thermal
cycling; after about 2,000 cycles, the dimensions become stable. Then, when we use
short fiber reinforced composites in industry, we must pay attention to the fiber axis
orientation. From the industrial viewpoint, the fiber content of around 20 vol.% will
have the most stable dimensions, and is therefore more suitable. More detailed
research has been reported in the instructive book by Chawla [1].

6.3 Interface

In Sects. 1.3–1.5 of Chap. 1, we discussed the fundamentals of the formation and


the state of the reinforcement/matrix interface. We then discussed in Sect. 2.1 of
Chap. 2 that the interface formation energy corresponds to the minimum fabrication
148 6 Properties of Composites

energy of the composite, and that it is possible to convert mechanical energy into
interfacial energy to enable wetting when the contact angle is high. Such energy
conversion will be realized by pressure infiltration methods. The interfaces
obtained by pressure infiltration methods seem to be bonded very well. Usually
pressure infiltration is conducted by squeeze casting, and the molten metal solidifies
rapidly in the mold. However, if the molten metal solidifies gradually, chemical
reactions tend to occur at the interface, and the reinforcement may be degraded.
The fiber/matrix interface has been intensively researched because the properties
of composites are greatly influenced by the bonding at the interface. For example,
the strength of composites depends upon force transfer across the interface. Heat
transfer through the interface greatly influences the thermal conductivity of
composites. Many coating techniques have been developed to promote wetting
between fiber and molten metal. However, good wetting often induces some
chemical reaction and thus can result in fiber degradation. This means that coating
techniques have had to be developed to control chemical reaction as well as
promote wetting at the interface [51].
For example, in the case of the Al/C system, aluminum carbide (Al4C3) is formed
readily over 800  C. To suppress this chemical reaction, CVD processes for applying
TiB2, TiC and TiN coatings to carbon fiber and SiC fiber were developed [6].
Cu/C composite is used as a heat sink material because of its high thermal
conductivity. Carbon is stable in molten copper. However, it is not easy to produce
Cu/C composites by pressure infiltration methods, because of the high melting point
of copper and poor wetting. An example of a commercial 25 vol.% Cu/75 vol.% C
composite heat sink for a base station of a mobile phone was shown in Fig. 1.8; this
composite was produced by squeeze casting. Since this product has high thermal
conductivity, the Cu/C interface seems to be bonded very well. Cu/SiC composites
also have high thermal conductivity. However, SiC is not stable in molten copper
and copper silicide (Cu5Si) will be formed if the copper solidifies slowly.
In the case of Ti alloy/SiC fiber composites, TiB2-coated SiC fibers or TiB2/C
bilayer coated SiC fibers prepared by ion plating are used to prevent reaction
between the SiC fiber and the molten Ti [52].
The pressure infiltration method without coatings is probably the most suitable for
development of materials with high thermal conductivity, because coating materials
affect the heat transfer through the reinforcement/matrix interface, and usually add to
the thermal resistance. In contrast, fiber coating is widely used to develop high strength
composites, because coating materials play an important role in enhancing force
transfer, by promotion of wetting and reduction of deleterious reactions at the interface.

6.4 New Trends in MMCs

6.4.1 Carbon/Metal Composites with High Thermal Conductivity

Industrial demand stimulates the development of MMCs. The demand for high
thermal conductivity materials for heat sinks such as substrates, packaging and
6.4 New Trends in MMCs 149

cases for integrated circuit (IC) products has increased with progress in ICs.
To meet this demand, metal composites with high thermal conductivity such as
carbon/Al or carbon/Cu composites are being actively developed. It is well known
that graphite fibers, carbon nanotubes and carbon nanofibers have very high thermal
conductivities. For example, pitch-based carbon fiber with a thermal conductivity
of 900 W m1 K1 [53], carbon nanotubes with conductivities about
3,500 W m1 K1 [54] and carbon nanofibers of about 1,200 W m1 K1 [55]
along their axis directions have been developed recently and are commercially
available. New composites are being actively developed using those fibers, and
composites with very high thermal conductivity have recently reported [56, 57].
However, the thermal conductivity in the radial direction of these fibers is very low,
although these fibers have very high thermal conductivities along their axis
directions. This means that carbon fiber/metal composites only have high thermal
conductivity along the fiber axis direction, when the fiber alignment is
unidirectional.
Alternatively, graphite has a layered, planar structure, in which the carbon atoms
are arranged in a hexagonal lattice (alpha graphite). The distance between planes is
0.335 nm. The thermal conductivity within the planes is very high and two-
dimensional. Recently, scale-like graphite particles having high thermal conductiv-
ity in two dimensions were developed. Scale-like graphite particle/aluminum
composites were fabricated by SPS and also showed high thermal conductivity in
two dimensions [58]. A crucial issue for these composites is how to distribute
the scale-like graphite particles horizontally within the metal powder prior to SPS.
This scale-like graphite particle/aluminum composite is a more promising compos-
ite than carbon fiber/aluminum composites for heat sink applications, because the
carbon fiber/aluminum composite only has high thermal conductivity in one
dimension.

6.4.2 Smart Composites

For more than a decade, some researchers dreamt of developing a material which
works like human skin—sensing hot, cold or pain and recovering from damage by
itself (self-healing). They called it “Intelligent material” [59, 60]. However, they
could not make much progress, because they did not have a concrete idea to develop
smart composites. Later, some other researchers started to use the concept of an
“intelligent material” to perform some practical functions [61, 62].
If we embed a temperature sensor or a piezoelectric ceramic sensor in a metal
structure, we can monitor the temperature or the stress of the metal structure.
Research into this type of concept began in around 1990. If we embed an optical
fiber in a metal structure and watch the light through the fiber, we can monitor the
stress condition of the metal structure and detect damage prior to fracture because
the intensity and waveform of the light through the fiber changes with the stress
applied to the metal structure. This optical fiber technique can be applied not only to
150 6 Properties of Composites

Table 6.2 Classification Sensor Actuator Controller Processor


of smart materials
suggested by Boller [63] Intelligent material ○ ○ ○ ○
Adaptive material ○ ○ ○
Active material ○ ○
Sensory material ○
Passive material

metal structures, but to structures such as buildings or bridges that are made of other
materials. Materials with sensor capabilities are classed as “Sensory materials”
[63], while ordinary materials are regarded as “Passive materials”. These ideas and
techniques are under investigation for monitoring the “health” of aircraft, and other
vehicles. The classification of smart materials has been proposed by Boller and is
shown in Table 6.2 [63]. The term “Active material” can be used if the material
includes actuators. If sensors and actuators are linked via a controller allowing the
material to adapt itself to various prescribed conditions, the material can be called
an “adaptive material”. When the adaptive material also includes a processor
allowing itself to adapt to various conditions by self-learning, the material becomes
an “Intelligent material” or “Smart material”. Most “smart materials” are actually
smart composites.
We can make a shape memory alloy fiber/metal composite which changes its
shape with stress and environmental temperature. Such a material has potential as
an actuator and is the subject of fundamental ongoing research [64].

References

1. Chawla, K.K.: Composite Materials, Science and Engineering, 2nd edn. Springer, New York
(1998)
2. Hill, R.: Theory of mechanical properties of fibre-strengthened materials: I. Elastic behavior.
J. Mech. Phys. Solids 12, 199–212 (1964)
3. Hashin, Z., Rosen, B.W.: The elastic moduli of fiber-reinforced materials. J. Appl. Mech. 31,
223–232 (1964)
4. Cox, H.L.: The elasticity and strength of paper and other fibrous materials. Br. J. Appl. Phys. 3,
72–78 (1952)
5. Kelly, A., Macmillan, N.H.: Strong Solids, 3rd edn. Oxford University Press, Oxford (1986)
6. Clyne, T.W., Withers, P.J.: An Introduction to Metal Matrix Composites, p. 12. Cambridge
University Press, Cambridge (1993)
7. Eshelby, J.D.: The determination of the elastic field of an ellipsoidal inclusion, and related
problems. Proc. R. Soc. Lond. A241, 376–396 (1957)
8. Taya, M., Arsenault, R.J.: Metal Matrix Composites—Thermo-Mechanical Behavior.
Pergamon, Oxford (1989)
9. Coleman, B.D.: On the strength of classical fibres and fibre bundles. J. Mech. Phys. Solids 7,
60–70 (1958)
10. Chawla, K.K.: Fibrous Materials, p. 258. Cambridge University Press, Cambridge (1998)
11. Weibull, W.: A statistical distribution function of wide applicability. J. Appl. Mech. Trans.
ASME 18, 293–297 (1951)
References 151

12. The Ceramics Society of Japan: Ceramic Advanced Materials, p. 1. Ohmsha, Ltd., Tokyo
(1991)
13. Rosen, B.W.: Mechanics of Composite Materials: Recent Advances, p. 105. Pergamon, Oxford
(1983)
14. Kagawa, Y.: Fiber Reinforced Metal Composites, p. 253. CMC Publisher, Tokyo (1985)
(in Japanese)
15. Kelly, A., Macmillan, N.H.: Strong Solids, 3rd edn, p. 201. Oxford University Press, Oxford
(1986)
16. Imai, T., Nishida, Y., Yamada, M., Shirayanagi, I., Matsubara, H.: Effect of cold rolling on the
mechanical properties of Al/alumina short fiber composite. J. Jpn. Inst. Light Met. 37, 179–184
(1987)
17. Griffith, A.A.: The phenomena of rupture and flow in solid. Philos. Trans. R. Soc. Lond. A221,
163–198 (1920)
18. Irwin, G.R.: Analysis of stresses and strains near the end of a crack traversing a plate. J. Appl.
Mech. 24, 361–364 (1957)
19. Irwin, G.R., Kies, J.A., Smith, H.L.: Fracture strengths relative to onset and arrest of crack
propagation. Proc. Am. Soc. Test Mater. 58, 640–657 (1958)
20. Fujii, T., Zako, M.: Fracture and Mechanics of Composite Materials, p. 57. Jikkyo Shuppan
Co. Ltd., Tokyo (1978) (in Japanese)
21. Yokobori, T.: An Interdisciplinary Approach to Fracture and Strength of Solids (trans and edit:
Crisp, J.D.C.). Walter Nordhoff, Pub., Groningen (1968)
22. Ochiai, S., Osamura, K.: A study of multiple fracture phenomenon of a coating film on a metal
fibre by means of computer simulation. J. Mater. Sci. 21, 2735–2743 (1986)
23. Ochiai, S., Osamura, K.: Influence of matrix ductility, interfacial bonding strength, and fiber
volume fraction on tensile strength of unidirectional metal matrix composite. Metall. Trans.
21A, 971–977 (1990)
24. Ochiai, S., Hayashi, K., Osamura, K.: Estimation of interfacial shear strength between
superconducting oxides and silver sheath from multiple-fracture phenomenon of the oxide.
Metall. Trans. 25A, 349–356 (1994)
25. Ochiai, S., Hojo, M.: Application of Monte Carlo simulation to mesomechanics of fiber-
reinforced composite materials. Mater. Jpn. 33, 1397–1406 (1994)
26. Williams, J.J., Chapman, N.C., Jakkali, V., Tanna, V.A., Chawla, N.: Characterization of
damage evolution in SiC particle reinforced Al alloy matrix composites by in-situ X-ray
synchrotoron tomography. Metall. Mater. Trans. 42A, 2999–3005 (2011)
27. Kobayashi, T., Iwanari, H., Kim, H.-J., Yoon, E.-P., Watanabe, S.: Fracture toughness of SiCp/
6061-T6 composite. J. Jpn. Inst. Light Met. 41, 89–94 (1991)
28. Yoshino, M., Iwanari, H., Niinomi, M., Kobayashi, T.: Mechanical properties of SiC whisker
reinforced aluminum alloys. J. Jpn. Inst. Light Met. 38, 593–599 (1988)
29. Shirayanagi, I., Nishida, Y., Matsubara, H., Nakanishi, M., Kato, E.: Fatigue of alumina short
fiber reinforced AC8A aluminum alloy composites. J. Jpn. Inst. Light Met. 41, 471–476 (1991)
30. Donomoto, T., Miura, N., Funatani, K., Miyake, N.: Ceramic fiber reinforced piston for high
performance diesel engines. SAE Paper No. 830252 (1983)
31. Paris, P.C., Erdogan, F.: A critical analysis of crack propagation laws. Trans. ASME D 85,
528–534 (1963)
32. Masuda, C., Tanaka, Y.: Fatigue crack propagation mechanisms of SiC whiskers or SiC
particulates reinforced aluminum composites. In: Proceedings of the Fifth Japan–U.S. Confer-
ence on Composite Materials, Tama-City, Tokyo, June 1990, pp. 321–328 (1990)
33. Shang, J.K., Yu, W., Ritchie, R.O.: Role of silicon carbide particles in fatigue crack growth in
SiC-particulate-reinforced aluminum alloy composites. Mater. Sci. Eng. A 102, 181–192
(1988)
34. Chawla, N., Habel, U., Shen, Y.-L., Andres, C., Jones, J.W.: The effect of matrix microstruc-
ture on the tensile and fatigue behavior of SiC particle-reinforced 2080 Al matrix composites.
Metall. Mater. Trans. 31A, 531–540 (2000)
152 6 Properties of Composites

35. Lewandowski, J.J., Liu, C., Hunt Jr., W.H.: Effects of matrix microstructure and particle
distribution on fracture of an aluminum metal matrix composite. Mater. Sci. Eng. A107,
241–255 (1989)
36. Toda, H., Kobayashi, T., Wada, Y.: Fracture mechanical simulation of a crack propagating in
discontinuously-reinforced metal matrix composites. J. Jpn. Inst. Met. 59, 94–102 (1995)
37. Toda, H., Kobayashi, T., Wada, Y., Inoue, N.: Evaluation of fracture toughness and proposal of
microstructurally-controlled composites by fracture-mechanics simulation. J. Jpn. Inst. Met.
59, 198–205 (1995)
38. Hasselman, D.P., Johnson, L.F.: Effective thermal conductivity of composites with interfacial
thermal barrier resistance. J. Compos. Mater. 21, 508–515 (1987)
39. Ota, H., Tomota, Y.: Estimation of the effective thermal conductivity and the analysis of
temperature response in the transient state for composite materials. Bull. Jpn. Inst. Met. 29,
147–154 (1990)
40. Ota, H., Tomota, Y.: A list of the thermal conductivity of composite materials. Bull. Jpn. Inst.
Met. 29, 155–158 (1990)
41. Turner, P.S.: Thermal-expansion stresses in reinforced plastics. J. Res. Natl. Bur. Stand. 37,
239–250 (1946)
42. Kerner, E.H.: The elastic and thermo-elastic properties of composite media. Proc. Phys. Soc.
(Lond.) 69B, 808–813 (1956)
43. Schapery, R.A.: Thermal expansion coefficients of composite materials based on energy
principles. J. Compos. Mater. 2, 380–404 (1968)
44. Hashin, Z., Shtrikman, S.: A variational approach to the theory of the elastic behavior of
multiphase materials. J. Mech. Phys. Solids 11, 127–140 (1963)
45. Lemieux, S., Elomari, S., Nemes, J.A., Skibo, M.D.: Thermal expansion of isotropic Duralcan
metal-matrix composites. J. Mater. Sci. 33, 4381–4387 (1998)
46. Ishikawa, T., Koyama, K., Kobayashi, S.: Thermal expansion coefficients of unidirectional
composites. J. Compos. Mater. 12, 153–168 (1978)
47. Rosen, B.W., Hashin, Z.: Effective thermal expansion coefficients and specific heats of
composite materials. Int. J. Eng. Sci. 8, 157–173 (1970)
48. Jasiuk, I., Mura, T., Tsuchida, E.: Thermal stresses and thermal expansion coefficients of short
fiber composites with sliding interfaces. Trans. ASME 110, 96–100 (1988)
49. Yoda, S., Kurihara, N., Wakashima, K., Umekawa, S.: Thermal cycling-induced deformation
of fibrous composites with particular reference to the tungsten-copper system. Metall. Trans.
9A, 1229–1236 (1978)
50. Nakanishi, M., Nishida, Y., Sakai, Y.: Effect of thermal cycling on the properties of alumina
short fiber-reinforced aluminum. In: Proceedings of the Fifth Japan–U.S. Conference on
Composite Materials, Tokyo, pp. 301–308 (1990)
51. Rocher, J.P., Quenisset, J.M., Naslain, R.: Wetting improvement of carbon or silicon carbide
by aluminium alloys based on a K2ZrF6 surface treatment: application to composite material
casting. J. Mater. Sci. 24, 2697–2703 (1989)
52. Choy, K.-L., Derby, B.: Potential coating systems for inhibiting SiC/Ti interfacial reactions.
In: Vincenzini, P. (ed.) Advanced Structural Fiber Composites, pp. 179–184. National
Research Council/Techna, Italy (1995)
53. Catalog. Nippon Graphite Fiber Co., Tokyo. http://www.ngfworld.com
54. Pop, E., Mann, D., Wang, Q., Goodson, K., Dai, H.: Thermal conductance of an individual
single-wall carbon nanotube above room temperature. Nano Lett. 6(1), 96–100 (2006)
55. Showa Denko. http://www.sdk.co.jp/products
56. Imanishi, T., Sasaki, K., Katagiri, K., Kakitsuji, A.: Thermal and mechanical properties of
VGCF-containing aluminum. Trans. Jpn. Soc. Mech. Eng. A 74(5), 23–29 (2008)
57. Imanishi, T., Sasaki, K., Katagiri, K., Kakitsuji, A.: Effect of CNT addition on thermal
properties of VGCF/aluminum composites. Trans. Jpn. Soc. Mech. Eng. A 75(1), 27–33 (2009)
58. Ueno, T., Yoshioka, H.: Japanese Patent JP 4441768
59. Shimamura, S.: Kikai no kenkyu. 30, 99–105 (1978) (in Japanese)
References 153

60. Shimamura, S. (ed.): Miraio Hiraku Sentanzairyo (Advanced Materials with Bright Future),
pp. 207–213. Kogyo Chyosakai, Tokyo (1982) (in Japanese)
61. Takagi, T., et al.: The Concept of Intelligent Materials and the Guidelines on R & D
Promotion. Science and Technology Agency, Tokyo (1989)
62. Takeuchi, E., Matsuoka, S., Miyahara, K., Hirukawa, H., Ikeda, Y.: Proceedings of the 3rd
National Intelligent Materials Symposium, pp. 31–33 (1994) (in Japanese)
63. Boller, C.: General Introduction. AGARD (Advisory Group for Aerospace Research &
Development), Smart Structure and Materials. Implications for Military Aircraft of New
Generation. North Atlantic Treaty Organization, I-1 to I-7 (1996)
64. Coughlin, J.P., Williams, J.J., Crawford, G.A., Chawla, N.: Interfacial reactions in model NiTi
shape memory alloy fiber-reinforced Sn matrix “Smart” composites. Metall. Mater. Trans.
40A, 176–184 (2009)
Chapter 7
Superplasticity of Composites

Abstract The superplasticity of metal matrix composites is introduced along with


production methods for superplastic composites. Composites are strengthened by
particles or fibers and usually have poor ductility, so superplastic composites
(which were discovered in 1984) are definitely unusual. Superplasticity in MMCs
occurs at high strain rates and at high temperatures near the solidus line of the
matrix alloys. The strain rate is 100–1,000 times faster than that required to produce
superplasticity in alloys. In this chapter, the mechanism of superplasticity is
discussed using constitutive equations, where the shapes of the reinforcements
are limited to particles or short, fine fibers. Equal channel angular pressing is
introduced as one production method for superplastic MMCs.

7.1 Background for Superplasticity

It is well known that heated glass can be extended easily and exhibits large elonga-
tion in a manner reminiscent of chewing gum. A similar phenomenon occurs in
metals produced by a special process. This phenomenon is called “superplasticity”,
but it is different from the large elongation of glass. Superplasticity is defined as very
large and uniform elongation of a polycrystalline material.
Very large elongation of fine grained alloys was observed by Rosenhain et al.
[1], Jenkins [2] and Pearson [3]. Their studies seem to be the first reported
investigations of superplasticity. In 1945, Soviet researchers Bochvar et al.
named such very large elongation “Superplasticity” and carried out systematic
research [4]. Since Underwood published a review paper on superplasticity in
1962 [5], many researchers, especially American researchers, have taken an interest
in superplasticity, and the research area has become very active.
Superplasticity of a metal matrix composite was discovered by Nieh et al. in
1984 [6]. Their sample was 20 vol.% SiC whisker reinforced 2124 aluminum alloy,
which was produced by a powder metallurgical route, hot extruded and rolled.
The sample was subjected to a tensile test at a strain rate of 0.3 s1, and the largest

Y. Nishida, Introduction to Metal Matrix Composites: Fabrication and Recycling, 155


DOI 10.1007/978-4-431-54237-7_7, # Springer Japan 2013
156 7 Superplasticity of Composites

Fig. 7.1 Appearance of tensile test specimen of 20 vol.% b-Si3N4 whisker reinforced 2124
aluminum alloy tested at 525  C (courtesy of Dr. T. Imai): (a) before test, (b) after test at strain
rate of 0.05 s1, (c) 0.17 s1, (d) 0.50 s1

elongation they obtained was 300 %. As this strain rate is about 100 times higher
than the conventional strain rate at which alloys become superplastic;
this superplasticity is known as “High strain rate superplasticity (HSRS)”. The
superplasticity of ceramics was also discovered by Wakai et al. in 1986 [7]. Since
then, superplasticity of many kinds of metal matrix composites has been observed.
This indicates that not only special materials exhibit superplasticity; superplasticity
is a common phenomenon of very fine grained normal materials. Figure 7.1 shows
an example of a superplastic composite (b-Si3N4 whisker reinforced 2124 alumi-
num alloy), which exhibited a maximum elongation of 250 % [8].
Figure 7.2 shows a component with a more complicated shape made from 25 vol.%
SiC particle reinforced 2124 aluminum alloy, which was formed by gas pressure at a
strain rate of 0.1 s1. This is interesting because, as metal matrix composites usually
contain ceramic reinforcements such as SiC particles or alumina fibers, it is very
difficult to machine them. It is, however, possible to create complex shapes using a
superplastic forming process.

7.2 Mechanism of Superplastic Forming

The plastic deformation of metals is explained by the movement of dislocations


(which are types of lattice defects). This explanation is supported by the fact that
dislocations are relatively easy to move in fcc or bcc structured metals which have
7.2 Mechanism of Superplastic Forming 157

Fig. 7.2 A component of


25 vol.% SiC particle
reinforced 2124 aluminum
alloy superplastically formed
by gas pressure at a strain rate
of 101 s1 (courtesy of
Kawasaki Heavy Industries,
Ltd)

many slip planes and good plasticity. In fact, fcc structured metals such as alumi-
num or copper are easy to deform plastically, while plastic deformation of hcp
structured metals such as titanium or magnesium is not easy. Let us consider a grain
in a metal specimen which is subjected to tensile stress. If a dislocation moves in the
stress direction, the grain will be deformed, but the elongation limit of the grain will
be 200–300 %. However, over 1,000 % elongation has been obtained for some
superplastically deformed metals. Such large elongations cannot be explained by
the movement of dislocations. In addition, superplasticity of ceramics has been
found, although dislocations barely move in ceramics even at high temperature.
Therefore, these results suggest that dislocations are not the dominant mechanism
in superplastic deformation.
Many researchers investigated the superplastic deformation mechanism experi-
mentally, and realized that sliding between the grains was occurring at the grain
boundaries. In addition, it also became clear that each individual grain is only
deformed slightly [9–11]. When sliding at grain boundaries occurs, stress
concentrations will form at the triple junctions of grain boundaries and at some
other complicated grain boundaries. This means that stress relaxation is required for
grain-boundary sliding at these boundaries. Some accommodation process is also
required to prevent cavity formation at grain boundaries. Many models for super-
plastic deformation have been proposed [12].

7.2.1 Superplastic Deformation Mechanism of Alloys

The strain rate e_ of a material during high temperature creep is given by:
    
AGb b p s n Q
e_ ¼ D0 exp  ; (7.1)
kT d G RT
158 7 Superplasticity of Composites

where A is a dimensionless constant, G is the shear modulus, b is the Burgers


vector, k is the Boltzmann constant, T is the absolute temperature, d is the grain size
(i.e., average grain diameter), s is the stress, D0 is the diffusion coefficient, Q is the
activation energy, R is the gas constant, p is the grain size exponent and n is the
stress exponent. Equation (7.1) is known as the “constitutive equation” [13].
When intergranular lattice diffusion plays an important role in the movement of
atoms during high temperature creep, the exponents of Eq. (7.1) take the values
p ¼ 2 and n ¼ 1. This type of creep is known as “Nabarro-Herring creep”, and in
this case, the value of the activation energy from Eq. (7.1) seems to agree well with
the activation energy for lattice diffusion. When the temperature is slightly lower
than in the above case, and grain-boundary diffusion dominates the movement of
atoms during creep deformation, the exponents of Eq. (7.1) take the values p ¼ 3
and n ¼ 1. This type of creep is called “Coble creep”. The diffusion coefficient in
this case agrees well with the diffusion coefficients for grain-boundary diffusion
obtained by other processes. In addition, when dislocations play the major role in
creep deformation, the exponents of Eq. (7.1) are p ¼ 0 and n > 3.
As described above, when the amount of creep deformation is small, it is
possible to explain the phenomenon by the movement of atoms which corresponds
with the deformation.
Alternatively, for superplasticity, results from many researchers suggest that the
exponents of Eq. (7.1) are p ¼ 2 and n ¼ 2 [14]. Then, Eq. (7.1) becomes:
    
AGb b 2 s  s0 2 Q
e_ ¼ D0 exp  ; (7.2)
kT d G RT

where s0 is the threshold stress. In many cases, particularly for metal matrix
composites, the stress dependence has been found to be a function of (s  s0),
rather than just s. It has been confirmed that superplastic deformation of fine-
grained materials occurs by grain-boundary sliding. It is thought that the stress
relaxation and the prevention of cavity formation at the triple junctions of grain
boundaries can be explained by grain-boundary sliding accommodated by grain-
boundary diffusion. In addition, it was shown that the apparent activation energy
obtained by fitting Eq. (7.2) to experimental data is similar to the activation energy
of grain-boundary diffusion [15].

7.2.2 Superplastic Deformation Mechanism of Composites

Conventional superplasticity occurs at strain rates of around 104 s1. However,


metal matrix composites exhibit superplasticity at approximately 101 s1, which is
100–1,000 times greater. Superplasticity in MMCs is limited to composites
containing fine particles, short fibers or whiskers as reinforcement. Further, the
largest elongation of metal matrix composites is usually obtained near their solidus,
7.3 Production Methods of Superplastic Materials 159

which is much higher than the temperature required for conventional superplasticity
(which occurs at intermediate temperatures). This means that stress relaxation and
the process of accommodation at the triple junctions of grain boundaries must occur
very quickly in superplasticity at high temperature and high strain rate. The
accommodation process which takes place during superplastic deformation of
composites has not yet been clarified. However, the following possibilities are
being discussed for the accommodation process:
1. Grain-boundary diffusion, as occurs in superplasticity of conventional alloys.
2. Lattice diffusion.
3. Liquid phases arising from included impurities.
4. Dislocations.
The apparent activation energy of diffusion is an important way of assessing the
above ideas. The apparent activation energies obtained for metal matrix composites are
usually three to five times higher than that for grain-boundary diffusion in the matrix
alloy. In addition, as superplasticity of metal matrix composites occurs near the solidus,
it is reasonable to think that processes (1), (2) and (4) can all be taking place at the same
time, and that the main question is the relative importance of each process. Nieh et al.
proposed process (3) [16, 17]. However, if the liquid phase plays an important role, it
should be very easy to provide atoms at the triple junctions of grain boundaries and the
apparent activation energy should be lower. In fact, for the superplasticity of SiO2
doped zirconia, it was observed that above the glass transition temperature of amor-
phous SiO2 (where it forms a more liquid-like phase), the apparent activation energy
decreased [18]. In contrast, very large apparent activation energies have been obtained
for metal matrix composites. The interactions between reinforcement and matrix
grains might contribute to the large apparent activation energy.

7.3 Production Methods of Superplastic Materials

Some materials with very fine-grained microstructure exhibit superplasticity. Such


superplasticity is known as “fine-structure superplasticity” (FSS) because it
requires a very fine-grained microstructure [2]. However, if grain growth of a
material occurs during deformation, superplastic deformation will stop and the
material will fracture. When the material contains a second phase such as the
reinforcement in a composite, the second phase contributes to the suppression of
grain growth. Therefore, the prerequisites for FSS are:
1. Fine-grained polycrystalline materials:
For metals: grain size < 10 m.
For ceramics: grain size < 1 m.
2. At least two phases present in the material.
It is difficult to produce a very fine-grained metal directly by casting. Instead,
after casting, some type of plastic deformation process followed by recrystallization
160 7 Superplasticity of Composites

Fig. 7.3 Schematic


description of typical ECAP
during which the sample
undergoes straining by shear

is usually applied to the metal to produce fine grains. In industrial processes, plastic
deformation with recrystallization of metals is induced by hot extrusion, hot rolling
and/or hot forging. Hot rolling after hot extrusion is an effective way to produce
very fine microstructure. Lim et al. succeeded in producing a superplastic fine-
grained TiC particulate reinforced magnesium alloy composite by effectively using
hot rolling after hot extrusion [19]. However, if a large strain is applied to the metal
during these plastic deformation processes, the metal becomes a thin bar or a plate
which means that there is a limit to the amount of strain that can be applied.
Therefore, it is very difficult to produce very fine-grained metals by conventional
plastic deformation processes. To overcome this, equal channel angular pressing
(ECAP) [20–22] and accumulative roll bonding (ARB) processes [23, 24] have
been proposed as methods for maintaining dimensions close to the initial
dimensions of metals during and after plastic deformation.
ECAP is a new deformation process that can produce ultrafine-grained bulk
metals without the limitations of dimension. As shown in Fig. 7.3, a billet (the
“sample” in Fig. 7.3) is repeatedly passed through two channels of equal cross-
section intersecting at an angle. This technique was first developed by Segal et al. as
a means to introduce intense plastic strains into a metal without changing the cross-
sectional area [25–27]. It has also been studied as a method for producing
submicron-size grains in alloys.
Despite its unique capability, ECAP has not been widely used, primarily because
the strain per pass through the die channel is not very large. To obtain a desired
microstructure by using ECAP, about eight to ten passes through the die are
typically required. For each pass, the pressed billet must be removed from the die
and re-inserted into the die for the next pass, often after reheating the billet in a
7.3 Production Methods of Superplastic Materials 161

Fig. 7.4 Die configuration


for RD-ECAP

separate furnace. This makes the current ECAP process inefficient as well as
difficult to control. To circumvent these shortcomings, a more efficient ECAP
process that uses a rotary die has been developed [28]. This new ECAP is termed
rotary-die ECAP (RD-ECAP).
The die configuration for RD-ECAP is shown schematically in Fig. 7.4, and the
process sequence is shown in Fig. 7.5. Two channels of equal cross-section
intersecting at 90 are provided in the die. First, three punches of equal length are
inserted in the channel; one in the vertical channel and the other two in the
horizontal channels. The side and bottom plates confine the motion of the right
and bottom punches, respectively. The left horizontal punch can move freely. The
metal billet is inserted into the vertical channel with its bottom resting on the
bottom vertical punch. Then, the upper punch is inserted from the top so that it
can be pressed by the plunger.
The metal billet is extruded through the left horizontal channel as shown in
Fig. 7.5b. The plunger stops when it comes level with the upper surface of the die,
completing the first pressing cycle. Then, the die is rotated 90 . This rotation brings
the ECAP process back to its initial configuration and permits a new pressing cycle
to follow without billet removal and reinsertion. Prior to each pressing cycle, the
entire die assembly is placed in a movable furnace kept at the pressing temperature.
Just before pressing, the furnace is removed and the die assembly is placed on the
hydraulic press. After pressing, the die assembly is put back in the furnace to bring
the die temperature back to the set pressing temperature.
162 7 Superplasticity of Composites

Fig. 7.5 RD-ECAP process sequence: (a) initial state, (b) after one pass and (c) after 90 die
rotation

The macrostructural changes in an Al–7 wt.% Si–0.35 wt.% Mg alloy processed


by RD-ECAP at 300  C are demonstrated in Fig. 7.6 [29]. The images are of the
perpendicular cross-section along the billet center axis in the pressing direction
(i.e., to the right in these images). After one pressing, Fig. 7.6a, a strong downward
convex metal flow is observed in the middle of the billet. This is the characteristic
deformation mode in RD-ECAP. Deformation in the middle of the billet is quite
different from that at its perimeter. After the second pressing, Fig. 7.6b, the strongly
deformed area has expanded throughout the billet. Since the pressing direction in
the second pressing is the opposite to that of the first pressing, the billet has received
almost the same deformation as in the first pressing, but in the opposite direction.
After three presses, Fig. 7.6c, metal flow similar to strong agitation is seen in the
billet. The metal flow became much more complicated after four presses, as seen in
Fig. 7.6d. By eight presses, Fig. 7.6e, strong metal flow lines had become very
complicated and much agitation had been generated. The strong metal flow seen
after 20 presses, Fig. 7.6f, is similar to that after eight presses, showing that the
agitation has persisted. The overall structure, however, has become much finer and
more homogenized. This agitation of bulk solid metal should be effective in
producing fine grains with high angle grain boundaries, because the high disloca-
tion density and local non-uniform deformation necessary for nucleation of new
grains are created. The average grain size of this alloy was about 2 mm after ten
presses, as observed by transmission electron microscopy. The RD-ECAP has also
been successfully applied to produce a superplastic SiC whisker reinforced 7075
alloy composite [30, 31].
In ECAP (one of the severe plastic deformation processes), including RD-
ECAP, a hard second phase (such as the reinforcement in composites) has an
important role in producing very fine-grained metals. The reinforcements behave
as many miniature rollers in a rolling process inside the matrix metal during the
deformation processes. Therefore, we can produce very fine-grained metal matrix
composites efficiently by ECAP. However, plastic deformation during the ECAP
needs to be performed at intermediate temperatures, and the samples should have
some ductility. Those are some disadvantages of this process.
References 163

Fig. 7.6 Macrostructures of Al–7 wt.% Si–0.35 wt.% Mg alloy after (a) 1, (b) 2, (c) 3, (d) 4, (e)
8 and (f) 20 pressing cycles (extruded from left to right) at 573 K. Agitation with vortexes is clearly
observed in (d), (e) and (f)

References

1. Rosenhain, W., Ewen, D.: Intercrystalline cohesion in metals. J. Inst. Met. 8, 149–185 (1912)
2. Jenkins, C.H.M.: Strength of Cd–Zn and Sn–Pb alloy solder. J. Inst. Met. 40, 21–32 (1928)
3. Pearson, C.E.: Viscous properties of extruded eutectic alloys of Pb–Sn. J. Inst. Met. 54,
111–123 (1934)
4. Bochvar, A.A., Sviderskaya, Z.A.: Superplasticity in zinc–aluminum alloys. Izvest. Akad.
Nauk SSSR Otdel. Tekh. Nauk. 9, 821–827 (1945)
5. Underwood, E.E.: A review of superplasticity and related phenomenon. J. Met. 14, 914–919
(1962)
6. Nieh, T.G., Henshall, C.A., Wadsworth, J.: Superplasticity at high strain rate in SiC-2124 Al
composite. Scripta Metall. 18, 1405–1408 (1984)
7. Wakai, F., Sakaguchi, S., Matsuno, Y.: Superplasticity of yttria-stabilized tetragonal ZrO2
polycrystals. Adv. Ceram. Mater. 1, 259–263 (1986)
8. Imai, T., Mabuchi, M., Tozawa, Y., Yamada, M.: Superplasticity in b-silicon nitride whisker-
reinforced 2124 aluminum composite. J. Mater. Sci. Lett. 9, 255–257 (1990)
9. Lin, Z.-R., Chokshi, A.H., Langdon, T.G.: An investigation of grain boundary sliding in
superplasticity at high elongations. J. Mater. Sci. 23, 2712–2722 (1988)
164 7 Superplasticity of Composites

10. Matsuki, K., Morita, H., Yamada, M., Murakami, Y.: Relative motion of grains during
superplastic flow in an Al–9Zn–1 wt.% Mg alloy. Met. Sci 11, 156–163 (1977)
11. Matsuki, K.: Development and property of superplastic aluminum alloys. Bull. Jpn. Inst. Met.
26, 263–271 (1987)
12. Nieh, T.G., Wadsworth, J., Sherby, O.D.: Superplasticity in Metals and Ceramics. Cambridge
University Press, Cambridge (1996)
13. Maruyama, K., Nakajima, H.: High Temperature Strength of Materials, p. 15. Uchida
Rokakuho Publishing Co., Ltd., Tokyo (1997) (in Japanese)
14. Langdon, T.G.: A unified approach to grain boundary sliding in creep and superplasticity. Acta
Metall. Mater. 42, 2437–2443 (1994)
15. Mishra, R.S., Bieler, T.R., Mukherjee, A.K.: Superplasticity in powder metallurgy aluminum
alloys and composites. Acta Metall. Mater. 43, 877–891 (1995)
16. Nieh, T.G., Wadsworth, J.: High-strain-rate superplasticity in aluminum matrix composites.
Mater. Sci. Eng. A147, 129–142 (1991)
17. Nieh, T.G., Wadsworth, J., Imai, T.: A rheological view of high-strain-rate superplasticity in
alloys and metal-matrix composites. Scripta Metall. Mater. 26, 703–708 (1992)
18. Kajihara, K., Yoshizawa, Y., Sakuma, T.: The enhancement of superplastic flow in tetragonal
zirconia polycrystals with SiO2-doping. Acta Metall. Mater. 43, 1235–1242 (1995)
19. Lim, S.-W., Imai, T., Nishida, Y., Choh, T.: High strain rate superplasticity of TiC particulate
reinforced magnesium alloy composite by vortex method. Scripta Metall. Mater. 32,
1713–1717 (1995)
20. Segal, V.M., Goforth, R.E., Hartwig, K.T.: The application of equal channel angular extrusion
to produce extraordinary properties in advanced metallic materials. In: Henein, H., Oki, T.
(eds.) Processing Materials for Properties, pp. 971–974. Warrendale, TMS (1991)
21. Langdon, T.G., Furukawa, M., Nemoto, M., Horita, Z.: Using equal-channel angular pressing
for refining grain size. JOM 52(4), 30–33 (2000)
22. Lowe, T.C., Valiev, R.Z.R.Z.: Producing nanoscale microstructures through severe plastic
deformation. JOM 52(4), 27–29 (2000)
23. Tsuji, N., Shiotsuki, K., Saito, Y.: Superplasticity of ultra-fine grained Al–Mg alloy produced
by accumulative roll-bonding. Mater. Trans. JIM 40, 765–771 (1999)
24. Saito, Y., Utsunomiya, H., Tsuji, N., Sakai, T.: Novel ultra-high straining process for bulk
materials – development of the accumulative roll-bonding (ARB) process. Acta Mater. 47,
579–583 (1999)
25. Segal, V.M., Reznikov, V.I., Drobyshevskiy, A.E., Kopylov, V.I.: Plastic working of metals by
simple shear. Russ. Metall. (Metally) 1(99–115) (1981)
26. Segal, V.M.: Materials processing by simple shear. Mater. Sci. Eng. A197, 157–164 (1995)
27. Valiev, R.Z., Korznikov, A.V., Mulyukov, R.R.: Structure and properties of ultrafine-grained
materials produced by severe plastic deformation. Mater. Sci. Eng. A168, 141–148 (1993)
28. Nishida, Y., Arima, H., Kim, J.-C., Ando, T.: Rotary-die equal-channel angular pressing of an
Al–7 mass% Si–0.35 mass% Mg alloy. Scripta Mater. 45, 261–266 (2001)
29. Ma, A., Nishida, Y., Suzuki, K., Shigematsu, I., Saito, N.: Characteristics of plastic deforma-
tion by rotary-die equal-channel angular pressing. Scripta Mater. 52, 433–437 (2005)
30. Nishida, Y., Shigematsu, I., Arima, H., Kim, J.-C., Ando, T.: Superplasticity of SiC whisker
reinforced 7075 composite processed by rotary-die equal-channel angular pressing. J. Mater.
Sci. Lett. 21, 465–468 (2002)
31. Ma, A., Suzuki, K., Nishida, Y., Saito, N., Shigematsu, I., Takagi, M., Iwata, H., Watazu, A.,
Imura, T.: Impact toughness of an ultrafine-grained Al-11 mass% Si alloy processed by rotary-
die equal-channel angular pressing. Acta Mater. 53, 211–220 (2005)
Chapter 8
Materials for the Fabrication of Composites

Abstract Properties of the major reinforcement materials used for metal matrix
composites are introduced in this chapter and their characteristic features are
discussed and compared. The properties of a composite are partly determined by
the properties of each constituent material, because each constituent retains its own
microstructure and properties within the composite. Production methods for
reinforcements are briefly described, along with their influence on the properties
of the reinforcements including their compatibility with matrix metal.
Reinforcements may be particles or fibers, including whiskers, nanofibers and
nanotubes. They are made from ceramics or carbon. Carbon fibers with high thermal
conductivity are introduced, because they are likely to be useful in carbon/metal
composite heat sink materials.

Composites have been developed to pursue properties beyond those of conventional


materials. Many studies have combined different materials in anticipation of
acquiring properties intermediate between the properties of the constituents, or
even quite different properties. In some cases, however, the expected properties
have not been realized because of mismatch or degradation of constituents by
chemical reactions at the interface between them. This has led to extensive work
to improve the properties of constituents and the interfaces between them.
To solve problems in the fabrication of composites, researchers and engineers
need to have some knowledge about the properties of the constituents. The purpose
of this chapter is to provide that knowledge. The production processes will only be
described for the major types of reinforcements.

Y. Nishida, Introduction to Metal Matrix Composites: Fabrication and Recycling, 165


DOI 10.1007/978-4-431-54237-7_8, # Springer Japan 2013
166 8 Materials for the Fabrication of Composites

8.1 Characteristics of Reinforcements and Matrix Metals

The constituents of metal matrix composites are matrix metal and reinforcements
(fibers or particles which are added to the matrix metal). From an industrial
viewpoint, the most useful matrix metals (considering cost of materials, commer-
cial availability and recycling) are aluminum, magnesium, titanium, iron, copper
and their alloys. However, other matrix metals can be used if needed to develop
functional composite materials. The major properties of the most practical metals
are listed in Table 8.1. The properties of major ceramics (including a carbon fiber)
are listed in Table 8.2. The characteristic properties of ceramics are high elastic
modulus (Young’s modulus) (about ten times higher than that of metals), low
thermal expansion coefficient (about 1/5 of that of metals), low thermal conductiv-
ity (except for AlN, SiC and C) and very high melting point or high decomposition
temperature.
The main reasons to fabricate composites using the above constituents are to
develop composites with:
1. High strength
2. High elastic modulus
3. High strength and high toughness
4. Low thermal expansion coefficient
5. High thermal conductivity and low thermal expansion coefficient
6. High wear resistance
Furthermore, if we need lightweight materials, we can add “and light weight” to
each of the above properties. As most ceramics have high strength, high elastic
modulus, high hardness, low thermal expansion coefficient and relatively low
density, their use as reinforcement helps meet the above requirements.
However, as discussed in Sect. 6.1.3 of Chap. 6, when metal matrix composites
are fractured, most of the energy is absorbed by plastic deformation of the matrix
metal. We must consider that if the volume fraction of ceramic reinforcement
increases, the fracture toughness inevitably decreases.

Table 8.1 Properties of major matrix metals


Thermal
expansion Specific
Thermal coefficient heat Melting Vickers Young’s Shear
Density conductivity a (kJ kg1 point hardness modulus modulus
Material (kg m3) (W m1 K1) (106 K1) K1) (K) Hv (GPa) (GPa)
Al 2,698 238 23.9 0.9 933 18 75.7 26
Cu 8,930 416 17.1 0.39 1,356 35 136 40–50
Mg 1,740 171 26.1 1.02 932 37 44.3 16.6
Ti 4,500 15 8.8 0.52 1,904 60 114 39.8
Fe 7,870 71 12.1 0.44 1,809 150–370 190 80
Table 8.2 Properties of major ceramic and carbon reinforcements
Thermal
expansion
Thermal coefficient Melting Vickers Young’s Shear
Density conductivity a Specific point hardness modulus modulus
Material (kg m3) (W m1 K1) (106 K1) heat (kJ kg1 K1) (K) Hv (GPa) (GPa)
Al2O3 3,800–3,900 29 8.5 0.78 2,323 1,500–2,000 460 147
AlN 3,250 100–260 2.5 0.71 2,673 1,600 275 –
TiO2 4,175 8.4 – 0.69 2,113 – 170 –
TiN 4,780 25 6.3 0.6 3,223 – 600 –
TiC 4,770 31.8 7.6 0.56 3,413 2,270 450 –
SiO2 2,200 1.38 0.5 0.69 1,983 635 65–71 29
Si3N4 3,150 31 3.3–3.6 0.71 2,273a 1,400–1,550 372 –
SiC 3,146 270 5.1–5.8 0.67 3,103a 2,800 560 –
cBN 2,070 28 0.2–2.9 0.8 3,273a 3,500 400–900 –
WC 15,700 29 5–6 0.18 3,143 1,400 245 –
MgO 3,560 42 13.5 0.92 3,073 190 290 122
ZrO2 5,684 1.9–3.8 9.2–10.0 0.45 2,943 1,470–1,800 150–260 –
8.1 Characteristics of Reinforcements and Matrix Metals

Carbon 1,800–2,200 60–500 1.678 0.7 3,915b – 200–600 –


fiber
a
Decomposition
b
Sublimation point
167
168 8 Materials for the Fabrication of Composites

8.2 Production Processes for Reinforcements

8.2.1 Ceramic Particles

8.2.1.1 Silicon Carbide Particles

Silicon carbide (SiC) is one of the most suitable reinforcement materials to make
composites with high thermal conductivity and a low thermal expansion coefficient,
because it has a high thermal conductivity and a low thermal expansion coefficient.
In addition, because SiC has high hardness, it is also an excellent material to
produce highly wear resistant composites.
Industrially available SiC is produced by mixing SiO2 (silica sand) and C
(carbon, coke) and heating them to over 1,500  C in an electric furnace to induce
the following chemical reaction: SiO2 þ 3C ! SiC þ 2CO.
The main product of this reaction is b-SiC [1], although the reaction products also
include residual SiO2 and C. To obtain purer b-SiC, the products are then purified by
decarburizing and pickling, before being milled. a-SiC can then be produced by
heating b-SiC to over 2,000  C. Hexagonal close-packed a-SiC is stable at higher
temperatures than cubic b-SiC (which tends to be oxidized at high temperature).

8.2.1.2 Alumina Particles

Alumina (Al2O3) is roughly classified into regular alumina, low soda alumina,
reactive alumina and high-purity alumina. Each type of alumina is produced by a
different process. Regular alumina is produced from bauxite by the Bayer process
(invented by Bayer) [2]. If we dissolve bauxite in sodium hydroxide, the content of
alumina in bauxite changes into sodium aluminate and impurities such as Fe2O3,
TiO2 and SiO2 deposit as “red mud”. By hydrolyzing the sodium aluminate, gibbsite
(Al(OH)3) precipitates. a-Al2O3 is obtained by heating gibbsite over 1,100  C. The
alumina produced in this way is usually secondary particles, 30–50 mm in diameter,
which are agglomerates of smaller primary particles, 2–3 mm in diameter. We can
obtain particles close to the diameter of the primary particles by milling the
secondary particles. Low soda alumina and reactive alumina can be produced by
improved Bayer processes. High-purity alumina is produced by hydrolyzing high-
purity organoaluminum compounds or by pyrolysis of aluminum chloride.

8.2.1.3 Titanium Carbide Particles

Titanium carbide (TiC) is a widely used material for cutting tools, because its
properties are similar to those of cermet. Titanium carbide is produced by a solid
state reaction between titanium dioxide (TiO2) and carbon (C) at reduced pressure,
or by a solid state reaction between titanium hydride (TiH2) and carbon [3].
8.2 Production Processes for Reinforcements 169

8.2.1.4 Aluminum Nitride Particles

Aluminum nitride (AlN) is one of the most suitable reinforcing materials to produce
composites with high heat conductivity and low thermal expansion, because it has
high heat conductivity and an extremely small thermal expansion coefficient.
However, aluminum nitride is easily oxidized, which makes it more difficult to
use as a reinforcement in composite production. In addition, as AlN tends to react
with moisture in air and release ammonia gas, it must be handled carefully. There
are many production processes for aluminum nitride. The commercialized pro-
cesses are direct nitriding of molten aluminum under a nitrogen gas or ammonia gas
atmosphere, or reduction of alumina with carbon black under nitrogen atmosphere.

8.2.1.5 Silicon Nitride Particles

Silicon nitride (Si3N4) has strong covalent bonding and a low thermal expansion
coefficient. There are two crystal structures: a-type and b-type silicon nitrides. b-
type silicon nitride includes many impurities, so a-type silicon nitride is more
widely used. Silicon nitride is produced by “nitriding”, which involves heating
metallic silicon powder under a nitrogen or ammonia gas atmosphere at 1,400  C.
The reaction is controlled by the particle size of the silicon raw material, to suppress
the formation of b-type silicon nitrides and grain growth. In addition, residual
silicon and impurities that have become mingled during milling of the powder
must be removed [4].
Alternatively, silicon nitride powder can also be produced by decomposition by
heating silicon diimide (Si(NH)2) under a non-oxidizing atmosphere. Silicon
diimide is produced from silicon tetrachloride (SiCl4) and NH3.

8.2.2 Ceramic Fibers and Carbon Fibers

Ceramic fibers are roughly classified into two categories: melt-spun fibers and
chemical vapor deposited (CVD) fibers. The major fibers and whiskers are listed
in Table 8.3. The diameters of melt-spun fibers are 3–15 mm, and the diameters of
CVD fibers range from 100 to 150 mm. The large diameters of CVD fibers are
because these fibers are produced by depositing boron (B) or SiC on the surface of
tungsten or carbon fibers.

8.2.2.1 Alumina Fibers

Most of the alumina fibers used for metal matrix composites are alumina short
fibers. The properties of alumina fibers are listed in Table 8.3, although the
170 8 Materials for the Fabrication of Composites

Table 8.3 Fibers available for reinforcement of metals


Diameter Tensile strength Young’s Density
Fibers Shape (mm) (MPa) modulus (GPa) (g cm3)
Boron fiber (CVD) Continuous 100–150 3,400 390 2.4
SiC/boron fiber (CVD) Continuous 100–150 3,000 390 2.6
SiC fiber (CVD) Continuous 100–150 3,000 390 3.1
SiC (Nippon carbon, Continuous 10–15 2,500 176 2.6
precursor)
Si–M–C–O fiber Continuous 10–12 3,000 180 2.3
(M ¼ Ti,
Zr, Al) (Ube,
precursor)
C fiber (pitch) Continuous 5–10 2,100 390 2
C fiber (PAN) Continuous 7–9 3,000 235–264 1.7
Alumina Continuous 9 2,500 245 3.2
(Sumitomo
Chemical)
SiC Whisker 0.3–0.6 2,100–13,800 551 3.19
Si3N4 Whisker 0.1–1.6 13,800 379 3.18
Potassium Whisker 0.2–0.5 >6,900 >274 3.3
titanate
(K2O6TiO2)
Aluminum borate Whisker 0.5–1.0 8,000 400 2.93
(9Al2O32B2O3)
Alumina (e.g. Saffil) Short fiber 3 1,000 – 3.4
Alumina/silica (e.g. Short fiber 3 2,500 200 3.3
Denka)

properties differ slightly between fibers from different manufacturers.


Many studies have been performed on alumina fiber reinforced aluminum alloys,
and some of these composites have been commercialized [5, 6]. As is clear from
Table 8.3, the tensile strength of alumina fibers is not particularly high and
thus their strengthening effect is not large. However, the advantages of alumina
fiber reinforced composites are that the tensile strength only decreases gradually
with increasing temperature, and that these composites have excellent wear
properties.
It is difficult to directly melt-spin alumina (as is done to produce organic fibers or
glass fibers), because the melting point of alumina is 2,050  C, and the viscosity of
molten alumina is low. This means that a solution including aluminum chloride and
polymers must first be made as a starting material, and then green fibers (precursor
fibers) can be produced by spinning that solution. Finally the precursor fibers are
subjected to controlled pyrolysis to make alumina fibers. This process is called
“precursor spinning”, and a flow diagram of the process is shown in Fig. 8.1 [7, 8].
An SEM image of 80 wt.% Al2O3/20 wt.% SiO2 fibers is shown in Fig. 8.2. Both
continuous alumina fibers and alumina short fibers are commercially available, and are
produced by similar processes.
8.2 Production Processes for Reinforcements 171

Fig. 8.1 Flow diagram for


Alumina raw material Other materials
the production of alumina
fiber

Mixing

Polymeric

Spinning

Precursor fiber

Controlled pyrolysis

Alumina fiber

Fig. 8.2 SEM image of


alumina short fibers (80 wt.%
Al2O3/20 wt.% SiO2)
(courtesy of DENKA)

8.2.2.2 Silicon Carbide and Related Fibers

The specific strength and specific elastic modulus of silicon carbide fibers (SiC) are
very high and their strength decreases only slightly with increasing temperature.
Silicon carbide fibers have good compatibility with metals such as aluminum alloys
and are excellent reinforcements for composites. Silicon carbide fibers are produced
by either CVD processes or precursor spinning using organosilicon polymers.
172 8 Materials for the Fabrication of Composites

During the CVD process, tungsten fibers, about 10 mm in diameter, or carbon


fibers, are heated up to 1,200  C, and the fibers are passed through a gas mixture
containing methyldichlorosilane and hydrogen. SiC deposits on the surface of the
fibers, and the fiber diameter increases to 100–150 mm. The large diameter of these
fibers means that it is difficult to weave cloth with them.
Alternatively, precursor spinning using organosilicon polymers was invented in
the late 1970s by Prof. Seishi Yajima’s research group at Tohoku University in
Japan [9, 10]. They demonstrated that the SiC system fiber produced by precursor
spinning was so strong that its strength did not decrease even in the flame of a gas
burner. Since then, the fiber has become well-known for its excellent properties; the
fiber technology has been transferred to private companies and further developed
by them. At present, two kinds of precursor spun fibers are commercially available.
(a) SiC fiber
Polycarbosilane as a precursor is used, and green fibers (polycarbosilane fibers)
are produced by precursor spinning. The green fibers are subjected to heating and
oxidation curing at 190  C, followed by heating at 600–800  C for pyrolysis, and
then further heating up to 800–1,200  C for sintering, to produce the final SiC fiber.
As the sintering temperature increases, the structure of the fiber changes from an
amorphous to a fine polycrystalline structure, which increases the strength and
elastic modulus of the fiber [9, 10].
(b) Si–M–C–O fiber (M ¼ Ti, Zr, Al etc.)
The flow diagram for the production process of this fiber is shown in Fig. 8.3 [11].
The precursor is polymetallocarbosilane, which contains metal atoms (one of Ti, Zr,
Al, among others) within the main carbosilane structure. The green fibers are produced
by precursor spinning and subjected to oxidation curing or electron beam irradiation
curing. There are several curing processes which enable the polymerization of the
molecules within the green fibers without degradation of the fiber shape.
Curing by heating fibers in air at intermediate temperatures is employed for
industrial production. The cured polymer fibers are heated again under nitrogen at
temperatures over 1,000  C to sinter them. This fiber was developed by Ube
Industry, Ltd. and is known as “Tyranno® fiber”. Amorphous fibers and sintered
fine polycrystalline fibers have been developed. The fine polycrystalline fiber
contains Al as the M to improve the high temperature strength. The appearance
and the fracture surfaces of those fibers are shown in Fig. 8.4.

8.2.2.3 Carbon Fibers

Carbon fibers have high tensile strength, high elastic modulus, a low thermal
expansion coefficient and relatively low density. These fibers would appear to be
excellent for the reinforcement of metal matrix composites. However, carbon fibers
are barely wet by molten aluminum, and a reaction product (Al4C3) between carbon
fiber and aluminum tends to be formed during fabrication of composites.
8.2 Production Processes for Reinforcements 173

Polydimethylsilane

Polycarbosilane
Ti alkoxide
Zr complex compound, Al complex compound
Polymetallocarbosilane

Melt-spinning

Spun Fiber

Radiation Curing Oxidation Curing Oxidation Curing

Cured Fiber Cured Fiber Cured Fiber

Pyrolysis Pyrolysis Pyrolysis

Si-M-C-O Fiber (M: Ti, Zr) Si-M-C-O Fiber (M: Ti, Zr) Si-Al-C-O Fiber

Sintering

Si-Al-C Fiber

Fig. 8.3 Flow diagram for the production of Si–M–C–O fibers

Fig. 8.4 Appearance and fracture surfaces of Si–M–C–O fibers (Tyranno® fiber) observed using
SEM (courtesy of UBE) (a) fine crystalline fiber, (b) fracture surface of fine crystalline fiber, (c)
fracture surface of amorphous fiber

Carbon fibers available for composites are polyacrylonitrile (PAN)-based carbon


fibers and pitch-based carbon fibers. Both fibers are produced by the carbonization
of organic fibers which have been spun from solution. The flow diagram for
production is shown in Fig. 8.5 [12]. In addition, vapor grown carbon fibers and
carbon nanotubes have been developed.
174 8 Materials for the Fabrication of Composites

Fig. 8.5 Flow diagram for the production of carbon fiber

(a) PAN-based carbon fiber


The precursor fiber becomes stable by forming a pyridine ladder structure upon
dehydration and curing in the presence of oxygen. When the stabilized fiber is
heated up to 1,000  C, the precursor fiber becomes carbon fiber and eventually is
transformed into graphite fiber by further heating over 2,800  C. During carboniza-
tion and graphitization, the fiber is extended to improve the crystallinity of the
polymer and the orientation of the crystallites with respect to its fiber axis. These
procedures greatly improve the strength of the fiber.
(b) Pitch-based carbon fiber
Pitch has high carbon content and can be directly melt-spun as shown in Fig. 8.5.
After melt-spinning, the pitch fiber is transformed into carbon fiber by curing,
followed by carbonization.
Recently, the high thermal conductivity of carbon fibers has attracted special
interest. Carbon fibers (for example: GRANOC (commercial name)) having ther-
mal conductivity of 900 W m1 K1 along their axes have been developed [13].
These fibers are suitable for making composites for heat sinks, because the diameter
of the fibers is about 10 mm, and composite fabrication is not so difficult.
8.2 Production Processes for Reinforcements 175

(c) Carbon nanotubes


Cylindrical nanosize carbon crystals were discovered in the carbon soot of graphite
electrodes during an arc discharge, and analyzed by TEM in 1991 [14]. These crystals
are called “carbon nanotubes” (CNTs), and are further classified as single-walled
nanotubes (SWNTs) or multi-walled nanotubes (MWNTs). CNTs exhibit extraordi-
nary strength, electrical properties and thermal conductivity. The diameter of CNTs
ranges from 0.4 to 50 nm, while some CNTs grow up to several mm long. The Young’s
modulus of CNTs has been reported as 1.8 TPa [15], while their tensile strength has
been reported as 63 GPa [16]. The thermal conductivity of SWNTs along their axis has
been obtained by Pop et al. as about 3,500 W m1 K1 [17].
Arc discharge is one method used for synthesis of CNTs. Chemical vapor
deposition (CVD) using a catalyst such as nickel, cobalt or iron nanoparticles is
also widely used for large-scale synthesis [18].
However, as these nanotubes are very fine, the fabrication technique for
composites is very difficult, although some researchers are tackling the problem
of fabricating CNT/light metal composites [19]. Usually, the thermal conductivity
of the carbon wall of the nanotube is reported. If we take into account the hollow
part inside the nanotube, the effective thermal conductivity will be much lower than
the value reported by Pop et al. [17].
In addition, when we align fine fibers such as carbon nanotubes in a matrix metal
in one direction to develop a heat sink, heat flows from the fiber into the matrix, and
then into the next fiber. In this case, there is an interface thermal resistance between
the fiber and the matrix. If the length of the fibers is short, the total thermal
resistance will be very large, although the fibers themselves have very high thermal
conductivity. This means that it is difficult to produce composites with high thermal
conductivity using short fibers, and long fibers are better for this purpose.
(d) Vapor grown carbon fiber
Vapor grown carbon fibers (VGCFs) are produced by CVD using a catalyst such
as nickel, iron or cobalt particles [20]. Gas phase molecules are decomposed at high
temperature, carbon is deposited, and carbon fibers grow around the catalyst
particles. VGCFs are also hollow fibers similar to carbon nanotubes. Aluminum
composites with thermal conductivity of about 700 W m1 K1 have been reported
by Imanishi et al. [21, 22]. These composites contain 60 vol.% VGCFs aligned in
one dimension.

8.2.3 Ceramic Whiskers

Ceramic whiskers are whisker-like fibers. Typically, their length is a couple


of 100 mm and their diameter is very fine (<1 mm) as shown in Table 8.3.
A characteristic feature of whiskers is that they are single crystals with very few
crystal defects, which means that their strengths are close to theoretical values.
176 8 Materials for the Fabrication of Composites

Fig. 8.6 SEM image of SiC


whiskers

Processes to grow whiskers are classified into vapor phase, liquid phase and solid
phase routes. The vapor phase route is suitable for high quality, industrial produc-
tion. SiC, Si3N4 and Al2O3 whiskers are produced commercially by the vapor phase
route. In this process, the whisker is grown from vapor at a supersaturated vapor
pressure below the vapor/solid line of a phase diagram. This means that a large
amount of vapor is needed to grow a small number of whiskers. Compared with the
vapor phase route, the liquid phase route is an economical process to grow whiskers
from supersaturated solution. However, the disadvantage of the liquid phase route is
that the whiskers produced are lower in purity and therefore less perfect crystals.
It is very important to note that whiskers may be harmful to health, because both
ends of whiskers are fine and sharp (similar to asbestos fibers), and dry whiskers can
drift and fly about in the air easily. Whiskers require very careful handling, and their
use is prohibited in some countries.

8.2.3.1 Silicon Carbide Whiskers

The vapor phase reaction of CO gas and SiO gas produces SiC whiskers. CO gas is
produced by oxidation of carbon. SiO gas is produced by reduction of SiO2 raw
material. The simplified chemical equation for this reaction is:

SiO2 þ 3C ! SiC þ 2CO

The Gibbs standard free energy change of this reaction is positive. This means
that, if the partial pressure of CO gas at 1,350  C is continuously maintained below
the equilibrium pressure, the above chemical reaction will proceed to the right-hand
side and SiC whiskers can be grown [23]. The reason why SiC grows in the shape of
whiskers is explained by the VLS (vapor–liquid–solid) mechanism. In this mecha-
nism, a metal impurity is put on a substrate, and a droplet containing the impurity
then forms on the substrate. The impurity acts as a catalyst to grow whiskers. C and
Si atoms dissolve in the droplet, before eventually the droplet becomes supersatu-
rated with respect to SiC, and precipitation of SiC occurs in the shape of the
whisker. SiC whiskers observed under SEM are shown in Fig. 8.6.
8.2 Production Processes for Reinforcements 177

8.2.3.2 Silicon Nitride Whiskers

A mixture of silica sand (silicon source) and carbon (or oxidized chaff after acid
treatment) is used as the raw material for silicon nitride whisker (Si3N4) production.
Silicon nitride whiskers are grown by holding the raw materials under a nitrogen
atmosphere at temperatures between 1,250  C and 1,450  C in the presence of either
Fe or fluoride as a catalyst [24]. The simplified chemical equation for this reaction is:

3SiO2 þ 6C þ 2N2 ! Si3 N4 þ 6CO

The Gibbs standard free energy change of this reaction is also positive, which
means that if we maintain the CO gas pressure below the equilibrium pressure, the
chemical reaction will proceed to the right-hand side and Si3N4 whiskers can be
grown.
There are two kinds of Si3N4 whiskers, a-type and b-type whiskers. When the
partial pressure of SiO is higher and the partial pressure of N2 is lower than the
equilibrium pressure, a-type Si3N4 whiskers grow, and when the conditions are
reversed, b-type whiskers grow. The a-type whiskers are a mixture of whiskers with
square and triangular cross-sections [25].

8.2.3.3 Potassium Titanate Whiskers

Potassium titanate (K2O6TiO2) whisker grows from a liquid, i.e., this whisker is
produced by a liquid phase route. The crystal structure of this whisker is a tunnel
structure made from chains of octahedra, with potassium ions occupying positions
inside the tunnel. Potassium titanate whiskers contain more crystal defects such as
vacancies and dislocations than the other whiskers that are produced by the vapor
phase route. This means that the strength and the toughness of potassium titanate
whiskers are lower than those of other whiskers. This whisker is good for wear and
abrasion applications, because the hardness of this whisker is not so high. However,
since potassium titanate whiskers react with molten aluminum at high temperatures,
it is not easy to fabricate composites [26].

8.2.3.4 Aluminum Borate Whiskers

Aluminum borate whiskers (9Al2O32B2O3) are produced by heating raw materials


from which Al2O3 and B2O3 are formed between 1,000  C and 1,200  C, using a flux
which contains alkali metal chloride and sulfate or carbonate. This process is called
the “external flux process”. If the heating temperature is 800–1,000  C,
2Al2O32B2O3 whiskers will be obtained. This process was developed jointly by
Shikoku Chemicals Co. and the National Institute of AIST (Shikoku), Japan [27, 28].
178 8 Materials for the Fabrication of Composites

Fig. 8.7 SEM images of aluminum borate whiskers (courtesy of SUZUKI) (a) as received, (b)
granulated whiskers (diameter: about 50 mm)

The suitability of this whisker for reinforcing MMC was shown by Suganuma et al.
who prepared an aluminum borate whisker/AC8A alloy (Al–12 wt.% Si–1 wt.%
Cu–1 wt.% Mg–1 wt.% Ni) composite which showed high tensile strength and high
modulus [29].
The problems with aluminum borate whiskers are that the whiskers react readily
with molten aluminum and that the compressive deformation of the preform during
pressure infiltration is large [30]. However, the problem of chemical reactions
between molten aluminum and the whiskers was solved by protecting the whiskers
with an inorganic binder that covered their entire surface. This approach enabled
the fabrication of aluminum alloy/aluminum borate whisker composites without
unwanted reactions at high temperature. The problem of compressive deformation
has also been solved by using granulated whiskers, and some products have been
commercialized [31, 32]. The granulated whiskers are shown in Fig. 8.7.

References

1. Weimer, A.W. (ed.): Carbide. Nitride and Boride Materials Synthesis and Processing. Chap-
man and Hall, New York (1997)
2. Altenpohl, D.G.: Aluminum: Technology, Applications, and Environment, a Profile of a
Modern Metal, 6th edn, p. 7. The Aluminum Association, Inc. and TMS, Warrendale (1998)
3. Yanagida, H.: Fine Ceramics, p. 46. Ohm Publishing Co., Ltd, Tokyo (1982) (in Japanese)
4. Somiya, S.: Ceramics Science and Technology at the Present and in the Future, p. 35. Uchida
Rokakuho Publishing Co., Ltd, Tokyo (1982) (in Japanese)
5. Donomoto, T., Miura, N., Funatani, K., Miyake, N.: Ceramic fiber reinforced piston for high
performance diesel engines. SAE Paper No. 830252 (1983)
6. Ushio, H., Hayashi, N., Shibata, K., Ebisawa, Y.: Wear properties of hybrid fiber reinforced
aluminum matrix composites and application to an automotive engine block. J. Jpn Inst. Light
Met. 40, 787–792 (1990)
7. Takada, M.: Preprint of 125th Reinforced Plastic Committee. Society of Materials Science,
Japan (1997)
8. Tsuchida, T., Akimoto, H.: Review of alumina fiber. J. Jpn Inst. Light Met. 37, 762–766 (1987)
References 179

9. Yajima, S., Okamoto, K., Hayashi, J., Omori, M.: Synthesis of continuous SiC fibers with high
tensile strength. J. Am. Ceram. Soc. 59, 324–327 (1976)
10. Yajima, S.: Special heat-resisting materials from organometallic polymers. Am. Ceram. Soc.
Bull. 62, 893–898 (1983)
11. Satoh, M.: Development of tyranno-fiber and application to ceramic matrix composites. J. Jpn
Soc. Compos. Mater. 25, 9–15 (1999)
12. Chawla, K.K.: Fibrous Materials, p. 211. Cambridge University Press, Cambridge (1998)
13. Catalog. Nippon Graphite Fiber Co., Tokyo. http://www.ngfworld.com
14. Iijima, S.: Helical microtubes of graphite carbon. Nature 354, 56–58 (1991)
15. Treacy, M.M.J., Ebbesen, T.W., Gibson, J.M.: Exceptionally high young’s modulus observed
for individual carbon nanotubes. Nature 381, 678–680 (1996)
16. Yu, M.-F., Lourie, O., Dyer, M.J., Moloni, K., Kelly, T.F., Ruoff, R.S.: Strength and breaking
mechanism of multiwalled carbon nanotubes under tensile load. Science 287, 637–640 (2000)
17. Pop, E., Mann, D., Wang, Q., Goodson, K., Dai, H.: Thermal conductance of an individual
single-wall carbon nanotube above room temperature. Nano Lett. 6(1), 96–100 (2006)
18. Ebbesen, T.W., Ajayan, P.M.: Large-scale synthesis of carbon nanotubes. Nature 358,
220–222 (1992)
19. Uozumi, H., Kobayashi, K., Nakanishi, K., Matsunaga, T., Shinozaki, K., Sakamoto, H.,
Tsukada, T., Masuda, C., Yoshida, M.: Fabrication process of carbon nanotube/light metal
matrix composites by squeeze casting. Mater. Sci. Eng. A 495, 282–287 (2008)
20. Hatano, M., Ohsaki, T., Arakawa, K.: Graphite whiskers by new process and their composites,
advancing technology in materials and processes. Sci. Adv. Mater. Process. Natl. SAMPE
Symp. 30, 1467–1476 (1985)
21. Imanishi, T., Sasaki, K., Katagiri, K., Kakitsuji, A.: Thermal and mechanical properties of
VGCF-containing aluminum. Trans. Jpn. Soc. Mech. Eng. A 74, 655–661 (2008)
22. Imanishi, T., Sasaki, K., Katagiri, K., Kakitsuji, A.: Effect of CNT addition on thermal
properties of VGCF/aluminum composites. Trans. Jpn. Soc. Mech. Eng. A 75, 27–33 (2009)
23. Saito, H., Dei, T.: Vapor phase growth of b-SiC whiskers with fluorosilicate melt. Yogyo-
Kyokai-Shi (Ceram. Soc. Jpn.) 88, 265–270 (1980)
24. Saito, H., Hayashi, T., Miura, K.: Vapor growth of Si3N4 whiskers by nitriding of Si2-C-
Na3AlF6 system. Nippon Kagaku kaishi (Chem. Soc. Jpn.) 1981, 1371–1377 (1981)
25. Sasaki, K., Kuroda, K., Imura, T., Saka, K.: Microstructural investigation of a–Si3N4 whiskers
by transmission electron microscopy. Yogyo-Kyokai-Shi (Ceram. Soc. Jpn.) 94, 773–778
(1986)
26. Nishida, Y., Imai, T., Yamada, M., Matsubara, H., Shirayanagi, I.: Fabrication of potassium
titanate whisker/aluminum composites and some their properties. J. Jpn. Inst. Light Met. 38,
515–521 (1988)
27. Wada, H., Sakane, K., Kitamura, T., Hata, H.: The reaction sequence in the synthesis of
aluminium borate whiskers. J. Mater. Sci. 31, 537–544 (1996)
28. Wada, H., Sakane, K., Kitamura, T., Hata, H., Kambara, H.: Synthesis of aluminium borate
whiskers in potassium sulphate flux. J. Mater. Sci. Lett. 10, 1076–1077 (1991)
29. Suganuma, K., Fujita, T., Sasaki, G., Suzuki, N.: Evaluation of strength and heat-resistance for
aluminum-borate whisker reinforced AC8A aluminum alloy composite. J. Jpn. Inst. Light Met.
41, 270–275 (1991)
30. Saito, N., Nakanishi, M., Nishida, Y.: Effect of heat treatment on the mechanical properties of
aluminum-borate whisker reinforced 6061 aluminum alloy. J. Jpn. Inst. Light Met. 44, 86–90
(1994)
31. Yamauchi, T., Nishida, Y., Nakae, H.: Pressure infiltration of molten aluminum into preform
of granulated whiskers. J. Jpn. Inst. Met. 61, 158–165 (1997)
32. Yamauchi, T., Suzuki, M., Takahashi, M., Takada, I., Toda, M.: Trans. Soc. Automot. Eng.
Jpn. 30, 31 (1999)
Chapter 9
Recycling of Composites

Abstract The fundamental ideas for the recycling of metal matrix composites are
introduced in this chapter. The state of reinforcements in the composite is discussed
from a free energy viewpoint. Reinforcements are stable within the composite.
However, it is shown thermodynamically that if we change the energy at the
reinforcement/matrix metal interface, we can separate the reinforcements from
the matrix. Reinforcements can be separated by either mechanical or chemical
methods; the theoretical basis for both types of methods is discussed, and experi-
mental examples are presented. Reinforcements can be screened or separated from
the molten matrix metal using pressure (an example of a mechanical method). Some
molten salts with low surface energy can be very effective at separating
reinforcements from molten matrix metal. They do this by replacing the molten
matrix metal at the interface (an example of a chemical method). The entropy
increase upon addition of particles is also discussed, and the interface entropy is
calculated. Finally, one approach for assessing the viability and value of a metal
matrix composite is proposed.

The human population was small in the era before the industrial revolution, and the
influence of human activity on the environment was negligible. However, the
consumption of energy, especially fossil fuels, and natural resources per person
increased rapidly because of a desire for a more convenient life; in combination
with a large increase in population, this meant that human activity began to
seriously influence the environment. Therefore, recycling and reclamation of
materials have become important issues to preserve the natural environment, save
energy and reduce waste matter [1–4].
However, the drive to increase the efficiency of machines and thus save energy
places new demands on the properties of metallic materials. MMCs have been
developed to meet this demand for properties over and above those of ordinary
metallic materials. Since the amount of MMC products is still small, recycling and
reclamation of MMCs has not been a serious problem until now. Recycling,
however, will become an important issue in the near future, because the number

Y. Nishida, Introduction to Metal Matrix Composites: Fabrication and Recycling, 181


DOI 10.1007/978-4-431-54237-7_9, # Springer Japan 2013
182 9 Recycling of Composites

of aluminum matrix composite products has been steadily increasing. Therefore,


technologies for separating the reinforcements from composites, to enable
recycling, will be discussed in this chapter.

9.1 Composite Ratio in Products and Re-melting

Major examples of MMC products commercialized today are pistons for diesel
engines [5], pistons for outboard motors [6] and cylinder blocks for automobiles
[7]. The volume of the composite part of these products is very small. For example,
in the case of pistons for diesel engines, only the part surrounding the piston ring is
made from composite. The reinforcement of this part is alumina/silica short fibers
(diameter of about 3 mm), and the fiber volume fraction is 7–8 %. The composite
part of the cylinder blocks for automobiles is the inner surface of the cylinder; it is
3 mm thick, and contains about 20 vol.% fibers. The fibers used for the cylinder
block are alumina short fibers (diameter of about 3 mm) and carbon fibers (diameter
of about 7 mm). Aluminum alloy pistons reinforced with SiC whiskers or aluminum
borate whiskers (diameter of 0.5–1 mm) have also been commercialized. These
composites occupy only a small volume in the pistons. The volume fraction of
whiskers in the composite is about 20 %, indicating that the composite does not
seriously affect the recycling of metals.
When scraps of the composite products are melted, the composite part sinks to
the bottom of the crucible, keeping its original shape, because the reinforcing fibers
usually have higher specific density than the matrix metal, and each fiber does not
drift separately in the melt. Because the preforms are usually made of reinforce-
ment containing inorganic binders, the composite parts can be easily separated
from the remainder of the MMC products. This means that after melting the
composite products, it is possible to take out the composite part by simply
decanting the molten metal, or by filtering the melt. However, the matrix metal
within the composite part remains, without separating from the reinforcement. It
would be desirable that the composite parts be recycled without separating
reinforcements. But this is difficult, because the composite part is small and usually
has a specific shape. Therefore, a crucial requirement for the recycling and
reclamation of MMCs is to develop technologies for the separation of matrix
metal from reinforcements [8].
Alternatively, ceramic particulate reinforced aluminum alloys are expected to be
used for the disc brakes of the Shinkansen (Japanese high speed trains) and some
MMC disc brakes have been commercialized for automobiles. In this case,
the whole product is made from MMC. When these composite products are melted,
the reinforcement particles drift in the melt, although some sink, float or agglomer-
ate slightly. Therefore, it is possible to recycle these products as composites.
However, technologies for the separation of reinforcement particles from matrix
metal will be important for recovering the matrix metal.
9.2 Separation of Fibers or Particles 183

9.2 Separation of Fibers or Particles

9.2.1 State of Composites

Metal matrix composites consist of matrix metal and reinforcements, which each
have their own microstructure (for example, crystal structure), and interfaces
between them. The entropy of mixing does not increase much upon the fabrication
of a composite, because composites are not atomic-level mixtures. A comparison
between a composite and an atomic-level mixture (such as an alloy or a chemical
compound) is shown schematically in Fig. 9.1. Alloys and chemical compounds
have very high entropies of mixing, and require energy-intensive processes, such as
electrolysis or chemical methods to separate atoms in the atomic-level mixture.
However, the separation of reinforcements from composites only requires separa-
tion of the interfaces between the reinforcements and the matrix metal.
First, we will discuss the fabrication process. As shown in Fig. 2.1 in Chap. 2,
before fabrication of a composite, the surfaces of fibers having a surface energy of
gfa are surrounded by air. After fabrication of the composite, the fibers are in the
matrix metal, and interface energy is gfm. When fibers are made of ceramics such as
alumina, which do not wet well with molten aluminum, gfm is much larger than gfa.
This means that to fabricate the composite, work equivalent to the energy difference
between gfa and gfm must be done to the interface. As discussed in Chap. 2, the
fabrication energy is given by:

W ¼ ðgfm  gfa ÞA; (9.1)

a
Composite Interface

Matrix metal

Reinforcement

b
Alloy or compound

Fig. 9.1 (a) Reinforcement/


matrix metal interface of
composite, (b) atomic-level
mixture as found in alloys or
chemical compounds
184 9 Recycling of Composites

Fig. 9.2 Interface free


energy difference before and composite free energy
after fabrication of hill
composite. Gm is the free
energy of the matrix/
reinforcement interface, Ga is
the surface free energy of the
reinforcement before
fabrication
fabrication separation

before fabrication

(arbitrary)

where A is the total area of interface between fiber and matrix metal in unit volume
of the composite [9]. If the diameter of the fibers becomes smaller for a given
volume fraction of fibers, the total interface area per unit volume of the composite
increases, and, from Eq. (9.1), more work must be done to fabricate the same
amount of composite.
Alternatively, supposing that gfm and gfa are constants, independent of the
location on the fiber surface, the interface free energy change of the fiber per unit
area before and after fabrication of the composite is given by:

DG ¼ gfm  gfa : (9.2)

Equation (9.2) shows that the interface free energy of the fiber in the matrix
metal is larger than the surface free energy of the fiber in air. The free energy
difference is shown schematically in Fig. 9.2. This means that the fibers having gfa
in air are essentially more stable than those having gfm in the matrix metal. If some
stimulus is given to the interface, the separation of the fibers might occur, because
there is a driving force. However, as shown in Fig. 9.2, there is a free energy barrier
or hill to be overcome, which means that the fibers in the matrix are stable. If the
matrix/fiber interface area decreases by a small amount dA, although these free
energies are in equilibrium, matrix metal surface area dA and fiber surface area dA
will be newly formed, and the total surface free energy will change. As the free
energy change for this case is always positive, the separation of fibers does not
occur spontaneously in the composite.
However, if the separation of a fiber occurs, the separation will start from an
exposed fiber in the surface of the composite. Usually, it is difficult for fibers to
separate from matrix inside the composite. In this case, if metal/fiber interface area
dA separates, the metal/fiber interface area decreases by dA and metal/air interface
area dA and fiber/air interface area dA are newly formed as shown in Fig. 9.3. The
free energy change DG associated with this process is:
9.2 Separation of Fibers or Particles 185

Fig. 9.3 Separation of an


infinitely small area, dA, at gma
the triple junction of molten
matrix metal, fiber and air
Matrix metal
Air
dA
gfm gfa
Fiber

Fig. 9.4 Schematic diagram


of the interfacial energy Molten metal gma
balance of a sessile drop

q
gfm
gfa

Reinforcement

DG ¼ gfm dA þ gfa dA þ gma dA ¼ ðgfm þ gfa þ gma ÞdA; (9.3)

where gma is the surface energy of the molten metal. The reported values for surface
energy of molten metal are measured in saturated vapor of the metal and are slightly
different from the surface energy in air. In addition, as molten aluminum oxidizes
easily in air, the surface energy of aluminum is not the same as gma. For the sake of
convenience, we will suppose that the surface energy of aluminum is the same as
gma to make the discussion simple. If DG in the above equation is negative, the
separation of fibers will begin. Therefore, the requirement for separation is:

gfm >gma þ gfa : (9.4)

A molten metal droplet on a ceramic plate is shown in Fig. 9.4. In this figure, if
the contact angle y is 180 , the left-hand side and right-hand side of Eq. (9.4) are
equal. So even if y is 180 , the criterion of Eq. (9.4) is not satisfied. Therefore, the
fiber and molten matrix metal do not separate spontaneously. However, if the
surface area of the molten metal does not increase as a result of the separation, as
explained in Fig. 2.2, the separation will occur spontaneously. In general, some
thermodynamic assistance is needed to separate fibers. The separation of fibers
from matrix metal can be achieved either mechanically or chemically (in a similar
way that mechanical or chemical means can be used to enable wetting of fibers by
molten metal).
186 9 Recycling of Composites

Fig. 9.5 Schematic diagram Punch


of the mechanical separation
method: (a) composite held
above the melting point of the Composite
matrix metal, (b) separation
of the matrix metal from the
composite by compression Base

Fiber

Matrix metal

Fig. 9.6 Microscopic view of Fiber


mechanical separation: (a)
molten matrix metal between
fibers before compression, (b) Fiber
molten metal pushed out from Molten metal
between fibers by
compression

Molten matrix metal


Fiber

9.2.2 Mechanical Separation of Reinforcements

Above the melting temperature, molten matrix metal can be easily squeezed out
from a composite to form a pool of molten metal alongside the compressed
composite, as shown schematically in Fig. 9.5. This is a macroscopic separation,
and some matrix metal remains in the compressed composite. Another mechanical
method is the filtering of molten matrix metal from particle reinforced composites.
These macroscopic phenomena coincide with our experience of other separations.
In these macroscopic methods, the interface energy required for separation is not
necessarily provided by the mechanical means; instead some matrix metal may
remain in contact with the fibers.
An example of microscopic separation is shown in Fig. 9.6. Before compressing
a composite, fibers are surrounded by molten matrix metal. After compressing,
some fibers come into contact with each other, and the molten metal between them
9.2 Separation of Fibers or Particles 187

Fig. 9.7 Schematic diagram


of chemical separation: a gmc
candidate material infiltrates
area dA of the matrix/
reinforcement interface
Matrix metal Candidate
material C
dA

gfm gfc

Fiber

is squeezed out. This separation occurs partly on the surface of the fibers. But this
result also suggests the possibility of microscopic mechanical separation of
reinforcements. Microscopic separation can be achieved if enough energy can be
supplied by mechanical means to overcome the surface free energy barrier for the
separation of the molten matrix metal from reinforcement as shown in Fig. 9.2.
In addition, in Fig. 2.2 in Chap. 2, when the pressure of molten metal is DP
higher than that of the outside, the molten metal is in equilibrium. Therefore, if the
pressure of molten metal decreases and becomes lower than DP, the molten metal in
the ceramic pipe will recede. This means that mechanical separation can occur
microscopically. In free energy terms, this is the case where the free energy barrier
in Fig. 9.2 is not present.
Another example is the use of centrifugal force on molten metal to separate the
molten metal from reinforcements. This is the reverse of centrifugal casting,
which was discussed in Chap. 5. Therefore, macroscopic and microscopic sepa-
ration can be achieved by the expulsion of molten matrix metal using centrifugal
force. There are many mechanical means by which the free energy barrier can be
overcome.

9.2.3 Chemical Separation of Reinforcements

To separate matrix metal from a composite by chemical means, it is important to


find a material which makes an interface of lower interface energy with the
reinforcement than the matrix metal does. If this material comes into contact with
the reinforcement/matrix interface, the material will infiltrate into the interface and
replace the matrix metal at the interface. Furthermore, if this material has low
solubility in the matrix metal, separation of the reinforcement from the matrix will
commence.
If a small area dA of the reinforcement/matrix interface is replaced by a third
material, the candidate material C, a new matrix/candidate material interface of
area dA and a new reinforcement/candidate material interface of area dA will be
formed, as shown in Fig. 9.7. By replacing the air in Eq. (9.3) with the candidate
material C, the free energy change of the interface DG is:
188 9 Recycling of Composites

DG ¼ gfm dA þ gfc dA þ gmc dA ¼ ðgfm þ gfc þ gmc ÞdA; (9.5)

where gfc and gmc are the reinforcement/candidate material C interface energy and
the matrix/candidate material C interface energy, respectively. When the free
energy change of Eq. (9.5) is negative, the separation of reinforcement occurs.
The requirement of negative DG is:

gfm >gfc þ gmc : (9.6)

If the candidate material satisfies Eq. (9.6), the separation of the reinforcement
will occur.
The remaining issue involves finding a suitable candidate material. Such a
material [10] must have small gfc and gmc values in Eq. (9.6) when the material
contacts both reinforcement and matrix metal. When two different materials, which
have quite different structures (for example aluminum and SiC fibers) come into
contact with each other, there seems to be no theoretical method to quantitatively
estimate the interface energy, as discussed in Chap. 1. If gA, gB and gAB are the
surface energies of material A and material B and the energy of the material A/
material B interface, respectively, the relationship between them is:

gAB  gA þ gB : (9.7)

As is obvious from Fig. 9.4, if the contact angle y is close to 180 , gAB will be
close to the sum of gA and gB. Equation (9.7) shows that if the values of gA and gB are
very small, gAB will also be very small. Since we can regard gAB as gfc, if we choose
a candidate material with a low value of g, gfc will be much lower than gfm. We can
now apply relationship (9.7) to the matrix metal (material A) and the candidate
material (material B). When gB is much lower than the surface energy of the molten
matrix metal gma, gAB (which corresponds to gmc in this case) will be low. Because
we cannot change the surface energy of the reinforcement, nor of the matrix metal,
we need to choose a candidate material that forms interfaces of relatively low
energy with both the reinforcement and the matrix metal. As the entropy change
arising from the formation of the interface is negligibly small in general and
G ¼ H  TS, the interface free energy G per unit area is equal to g (the interface
energy). Examples of candidate materials are listed in Table 9.1. From the table it is
clear that as molten salts and alkali metals have low surface energy and low
solubility in aluminum, they are good candidate materials. Further, when the
contact angle between two materials is small, the interface energy can be estimated
by Girifalco and Good’s equation, as discussed in Chap. 1. The equation is:

gAB ¼ gA þ gB  2fðgA gB Þ1=2 ; (9.8)

where f is a constant [11, 12]. The experimentally obtained value of f for a molten
salt/molten aluminum system is 0.41 [13].
9.3 Separation of Fiber and Metal from Composites 189

Table 9.1 Surface energies of candidate materials


Surface energy Temperature Density
Material (mJ m2) (measured) (K) (103 kg m3)
Al 860–900 973 2.3
NaCl 190.8 1,083 1.55
KCl 155.2 1,063 1.54
BaCl2 162.6 1,254 3.12
Na 151 773 0.829
K 86 473 0.795
Bi 356 873 9.66
Ca 225 1,123 1.44
Hg 395 623 12.76
Li 430 453 0.508
Mg 526 1,023 1.56
Pb 414–426 873 10.3
Sb 383 973 6.45
Sr 165 1,043 2.48
Te 300 723 –

9.3 Separation of Fiber and Metal from Composites

9.3.1 Chemical Method and Ratio of Separation

As shown in Fig. 9.8, aluminum (the same material as the matrix metal) was melted
in a crucible. A rectangular block of the composite was dropped into the aluminum
and held in the crucible until the matrix metal had melted. After the matrix
aluminum had melted, the crucible was taken out from the furnace. Addition of a
flux to the molten system was carried out outside the furnace, with agitation. When
the composite sample contacted the flux, separation of the matrix metal began and
the sample floated on the molten aluminum surface. The composite sample was
removed from the melt before the melt solidified.
A composite sample solidified in a crucible without flux is shown in Fig. 9.9a.
The composite is pure aluminum reinforced with 10 vol.% alumina short fiber
(Al2O3: 85 wt.%, SiO2: 15 wt.%). The composite is the dark gray, rectangular part
in the casting near the bottom. Fig. 9.9b shows the appearance of the composite,
which has floated up to the melt surface and solidified. The flux used for this
experiment was a sodium-based flux (NaClKClNa2SiF6) and the amount of flux
added was 1 wt.%. The relationship between the remaining matrix metal (alumi-
num) and the amount of flux added into the molten aluminum is shown in Fig. 9.10,
for the pure aluminum/10 vol.% alumina short fiber composite. The proportion of
the matrix metal separated (“separation ratio”) increased with increasing amount of
Na-containing flux. However, the separation ratio reached a limit of 50–60 vol.% at
around 2 wt.% of flux when this method was applied.
190 9 Recycling of Composites

Agitation bar

Crucible

Flux

Molten aluminum

Composite

Fig. 9.8 Diagram of chemical separation using a flux

Fig. 9.9 Appearance of a composite sample during chemical separation: (a) aluminum/10 vol.%
alumina short fiber composite (gray part of the ingot) solidified without flux in the crucible, and
(b) composite that has floated and solidified on top of the molten metal

Fig. 9.10 Relationship between the amount of Na-containing flux added into the melt and the
proportion of matrix metal remaining for an aluminum composite containing 10 vol.% alumina
short fibers
9.3 Separation of Fiber and Metal from Composites 191

Fig. 9.11 Relationship


between the amount of flux
added into the melt and the
proportion of matrix metal
remaining for an aluminum
composite containing 20 vol.
% alumina short fibers.
(Filled circle) NaCl, (open
circle) Na-containing flux

Fig. 9.12 Relationship


between the amount of flux
added into the melt and the
proportion of matrix metal
remaining for 6,061
aluminum alloy reinforced
with 10 vol.% alumina short
fibers. (Open circle) flux with
Na, (open square): flux
without Na

Figure 9.11 shows the separation ratios of the matrix metal from a pure alumi-
num/20 vol.% alumina short fiber composite. To separate the matrix metal, two
different candidate additive materials were used: the Na-containing flux, and pure
NaCl powder. The separation limit was 50–60 vol.% when the flux was added with
agitation of the molten metal. When NaCl powder was used instead of the flux, the
separation ratio reached a limit of about 20 vol.% by addition of 3 wt.% of NaCl
powder. This result indicates that the effect of potassium and fluorine in the flux on
separation is large.
The separation ratios produced using fluxes containing Na and without Na
(KClAlF3K3AlF6) are compared in Fig. 9.12, for the case of a 6,061 aluminum
alloy/10 vol.% alumina short fiber composite. The limit of the separation ratio
produced using the flux containing Na was about 30 vol.%, but the limit using the
Na-free flux was 50–60 vol.%.
The separation ratios are shown in Fig. 9.13 for a 6,061/25 vol.% SiC whisker
composite. Fluxes with and without Na were used. The separation ratios obtained
using both fluxes were low (10–20 vol.%). Since the fiber diameter of SiC whiskers
192 9 Recycling of Composites

Fig. 9.13 Relationship


between the amount of flux
added into the melt and the
proportion of matrix metal
remaining for 6,061
aluminum alloy reinforced
with 25 vol.% SiC whiskers.
(Open circle) flux with Na,
(open square): flux without
Na

is small (about 0.6 mm) compared with that of the alumina short fibers (about 3 mm),
the total surface area of the whiskers is much larger than that of the alumina short
fibers at the same volume fraction. Therefore, a greater amount of flux should be
needed for separation of whiskers.
The separation ratios mentioned above were obtained by measuring the amount
of matrix metal which flowed out spontaneously when the melt around the floating
composite sample was agitated. However, if the composite sample is broken during
the agitation of the melt, the separation ratio will increase.

9.3.2 Phenomena Associated with Separation Using Fluxes

The separation of matrix metals from fibers using fluxes has been demonstrated
experimentally. Therefore, DG in Eq. (9.5) in these cases should be negative. To
calculate DG, we need the gfm, gfc and gmc values of the materials used. However,
the interface energies of these materials have not been obtained. Instead, the surface
energy of an Al2O3 single crystal is used to approximate that of the alumina short
fiber, and that of an equimolar mixture of NaCl and KCl is used to approximate the
surface energy of the Na-containing flux. The detailed calculations are as follows:
1. gfm: As the contact angle between Al2O3 and molten aluminum is about 0.9p rad
[14], the interface energy between them is obtained as gfm ¼ 1,823 mJ m2
using Young’s equation.
2. gfc: The surface energy of the molten salt is 106 mJ m2 [15, 16]. In addition, the
surface energy of Al2O3 is 1,015 mJ m2 at the experimental temperature of
1,023 K [17]. Then, the interface energy between the molten salt and Al2O3 is
obtained as 852 mJ m2 by Eq. (9.8) using f ¼ 0.41.
3. gmc: The experimental value, gmc, between molten aluminum and the NaClKCl
is 783 mJ m2 [16].
9.3 Separation of Fiber and Metal from Composites 193

Fig. 9.14 Droplet of matrix


aluminum pushed out from an
aluminum/10 vol.% alumina
short fiber composite upon
treatment with Na-containing
flux

The interface energy gfc between Al2O3 and NaClKCl is much smaller than gfm
between Al2O3 and aluminum. If these interface energies are used for the calcula-
tion of DG in Eq. (9.5), DG ¼ 233 dA (mJ), which is a negative value. This means
that the replacement of the matrix metal with the flux takes place. A droplet of
pushed-out molten aluminum forms, as shown in Fig. 9.14, and only fibers remain
in the composite.
When the fluxes are added onto the molten metal surface, the fluxes float
separately from the molten aluminum, because these fluxes have low solubility in
molten aluminum, and their densities are smaller than that of aluminum. However,
when the fluxes contact the composite as the molten metal is agitated, the fluxes
infiltrate the matrix metal/fiber interface, pushing the matrix metal. This explana-
tion involves simple replacement of the matrix metal with fluxes and does not mean
that molten matrix metal flows out from the composite leaving cavities behind.
However, it is difficult to explain the amount of molten matrix metal flowing out of
the composite using only the simple replacement mechanism mentioned above. It is
likely that if some gases are released from the flux after the flux infiltrates the
composite, these gases will also push the molten metal out from the composite.
Therefore, the separation mechanism probably consists of two steps:
1. A flux infiltrates the fiber/matrix metal interface, releasing chlorine and fluorine
gases.
2. Sodium and potassium formed by the decomposition of the flux continue to
infiltrate into the interface. Since these metals are very reactive with oxygen, the
air follows these metals and enters the composite.
The surface energies of sodium and potassium are about 1/6 and 1/10 of the
surface energy of aluminum, respectively. The interface energies of sodium/fiber
and potassium/fiber are much smaller than that of aluminum/fiber. As these metals
have very little solubility in aluminum even at the molten state, and also have
relatively low densities, these metals float on the surface of the aluminum melt. In
these experiments, the flux was added to the surface of the melt, although the
composite was on the bottom of the crucible; so initially the flux did not touch the
composite. By agitating the melt, the composite came into contact with the flux,
194 9 Recycling of Composites

Fig. 9.15 Schematic diagram Fiber Flux


of the infiltration of flux (Na, K)

(Na, K) into the interface


between molten aluminum
and a fiber
Molten matrix
metal

Flux
Fiber (Na, K)

Molten matrix
metal

floated up to the surface of the melt and separation began. Then, sodium and
potassium might cover the fibers. These phenomena are shown schematically in
Fig. 9.15.
Alternatively, low surface energy elements like sodium and potassium segre-
gate around the fibers instead of aluminum, because these elements are almost
insoluble in aluminum, and the interface free energy decreases upon their infiltra-
tion into the fiber/aluminum interface. In contrast, calcium and magnesium,
which also have low surface energy, dissolve quite well in aluminum. These
elements then make the aluminum surface energy lower. Therefore, we have to
note that the presence of calcium or magnesium decreases the interface energies
gfm and gmc, and that these elements therefore work against the separation of
reinforcements.
The contact angle of aluminum alloy on small SiC particles was observed by
Madarasz et al. using an emulsification technique [18]. Readers are referred to the
experimental results presented by Madarasz et al. for additional information about
the interface energies between SiC/flux, SiC/Al alloy and Al alloy/flux.

9.4 Entropy of Mixing by the Addition of Reinforcement


Particles

As discussed in Sect. 9.2, the entropy increase upon the addition of reinforcements
is only small when ordinary sized reinforcements are added into a matrix metal.
However, if the reinforcements are very fine particles or fibers such as carbon
nanotubes, the entropy might increase much more. In this section, we discuss the
entropy change associated with the addition of reinforcement particles.
9.4 Entropy of Mixing by the Addition of Reinforcement Particles 195

Fig. 9.16 Mixing of an ideal


solution

9.4.1 Entropy of Mixing

If we mix liquids A and B to form an ideal solution, the total volume will be the sum
of their volumes without heat release or heat absorption. As the heat release or
absorption Q is zero, it appears that DS ¼ Q/T ¼ 0, as shown in Fig. 9.16. How-
ever, the entropy change of the system is not zero, and the entropy increases due to
the “entropy of mixing”. In this case, the entropy of mixing is:

DSm ¼ R½XA ln XA þ XB ln XB ; (9.9)

where XA and XB are mole fractions of A and B, and R is the gas constant. Equation
(9.9) is independent of temperature and pressure. This equation is applicable to
ideal solutions and ideal gases, but is only approximately applicable to nonideal
solutions or solid solutions, because Eq. (9.9) does not include material properties.
Upon mixing, nonideal solutions usually release heat, and the total volume is not
the sum of their individual volumes. When more than two ideal liquids are mixed,
the entropy of mixing is given by:

DSm ¼ R½X1 ln X1 þ X2 ln X2 þ      þ Xi ln Xi þ      þ Xr ln Xr : (9.10)

This equation can be simplified to:

X
r
DSm ¼ kN ðNi =NÞ lnðNi =NÞ; (9.11)
i¼1

where i ¼ 1, 2, 3, 4, . . ..r, k is the Boltzmann constant, and N is the number of


atoms of a particular type. The mole fractions, Xi ¼ Ni =N, and:

N ¼ N1 þ N2 þ N3 þ       þNi þ      þ Nr : (9.11a)

If the total number of molecules (atoms) N is Avogadro’s number, kN ¼ R.

9.4.2 Entropy Increase upon the Addition of Particles

If particles become very small and eventually each molecule of the particles is
separately dispersed in the molten matrix metal without chemical reaction between
196 9 Recycling of Composites

Fig. 9.17 Entropy of mixing as a function of mole fraction of MgO for molecular-level mixing

Table 9.2 Energy of MgO Contact angle, y ( ) gSL (J m2) Temperature (K)
particle/molten aluminum
interface at different 130 1.75 1,173
temperatures, calculated 120 1.63 1,273
from contact angle 110 1.49 1,333

them, the entropy change will be calculated by Eq. (9.9). The relationship between
the entropy of mixing and the mole fraction of reinforcement (once the reinforce-
ment has been separated into constituent molecules) is shown in Fig. 9.17.
The entropy of mixing increases with increasing mole fraction of the reinforcement,
and a maximum appears at a mole fraction of 0.5.
We can calculate the total interface entropy of MgO particles introduced into the
molten aluminum on the condition that the total number of molecules (the sum of
Al atoms and MgO molecules) is constant, where chemical reactions between MgO
and aluminum are neglected. The relationship between free energy change DG and
entropy change DSS per unit area is obtained using Eq. (1.25):
   
@DG dg
DSS ¼  ¼ : (9.12)
@T P dT

This equation shows that it is possible to obtain DSS if the temperature dependence
of DG or g is known. The temperature dependence of the contact angle between
molten aluminum and MgO was shown in Table 1.3 in Chap. 1. The interface energy
gSL between MgO and molten aluminum can be obtained using Young’s equation
[Eq. (1.31) in Chap. 1] and is shown in Table 9.2. The data are plotted in Fig. 9.18, and
the gradient of the line obtained by the least squares method is 0.00158 J m2 K1.
The interface entropy per unit area is obtained from this gradient and Eq. (9.12). The
value of interface entropy is DSS ¼ 1.58  103 J m2 K1.
9.4 Entropy of Mixing by the Addition of Reinforcement Particles 197

Fig. 9.18 Temperature dependence of MgO/molten Al interface energy

Taking XAl and XMgO as the mole fractions of aluminum and MgO, the sum of
these is:

XAl þ XMgO ¼ 1: (9.13)

For this calculation, we will assume that the volume fraction Vf ¼ 0.1.
According to Eq. (9.13), the mole fractions are:

XAl ¼ 0:9054 and XMgO ¼ 0:0946:

Then, the total weight Wtotal and total volume Vtotal of 1 mol of composite are:
Wtotal ¼ 0.02824 kg and total volume Vtotal ¼ 0.00001044 m3.
The weight of MgO in 1 mol of composite is: WMgO ¼ 0.003812 kg.
As the surface area and the volume of a spherical particle (of diameter d) are
A ¼ 4p(d/2)2 and V ¼ 4p(d/2)3/3, the total surface area At and the total volume Vt
of particles in 1 mol of composite are given by:
 2
d
At ¼ 4np ; (9.14)
2

and
 3
d
4np
2
Vt ¼ ; (9.14a)
3
198 9 Recycling of Composites

Fig. 9.19 Total interface entropy as a function of particle diameter for a composite containing
MgO particles (Vf ¼ 0.1)

where n is the total number of particles in 1 mol of composite and d is the average
diameter of the particles. Therefore, n is given by:

2:495  107
n¼ d 3 : (9.15)
2

Substituting Eq. (9.15) into Eq. (9.14), the total surface area of MgO particles is:

6:267  106
At ¼ ðm2 Þ: (9.16)
d

Assuming that the interface entropy per unit area of particles is constant,
independent of particle diameter, the total interface entropy of 1 mol of composite
containing atomic-level MgO particles (Vf ¼ 0.1) is:

9:903  109
DSS ¼ ðJ K1 mol1 Þ: (9.17)
d

This relationship is shown in Fig. 9.19.


Alternatively, when Vf of MgO particles is 0.1, XAl ¼ 0.9054 and XMgO
¼ 0.0946. Therefore, the entropy of mixing can be obtained using Eq. (9.9), as:
DSm ¼ 2.603 J K1 mol1.
This result is for ideal solutions. The real entropy of mixing of MgO molecules
in molten aluminum might be slightly different from the above result. However,
when the particle diameter is 1 nm, the total interface entropy 9.903 J K1 mol1 of
1 mol of composite is the same order of magnitude as the calculated entropy of
mixing, 2.603 J K1 mol1.
9.5 Assessment of Metal Matrix Composites 199

The total interface entropy was calculated on the assumption that the macro-
scopic interface entropy per unit area is independent of MgO particle diameter. We
have obtained relatively good agreement, although this assumption may not be
applicable to very fine particles. If the particle diameter becomes very small, the
percentage of MgO molecules which contact aluminum atoms will increase until
finally all MgO molecules contact aluminum atoms. This state is the same as an
atomic-level mixture. Therefore, the above conclusion is reasonable.
This result shows that it will be very difficult to separate very fine reinforcements
from matrix metal. Such separations are likely to require very energy-intensive
processes such as electrolysis.

9.5 Assessment of Metal Matrix Composites

Additional energy is necessary to produce composite products compared with


ordinary metal products. In addition, after use, supplemental energy is required
for recycling or disposal of composite products. From the viewpoint of environ-
mental problems and saving energy, some people criticize the use of composites.
Alternatively, a compact engine block has been realized by using a composite
instead of a cast iron sleeve, as shown in Fig. 1.2 in Chap. 1. In addition, higher
engine efficiency has been attained by composite pistons. Therefore, we need a way
to estimate the overall value of composites. To that effect, we propose the following
approach. Consider the total energy consumed for production, recycling of the
composite and operation of a machine containing composite products for the
lifetime of the machine, and compare this with the total energy consumed in
producing, recycling and operating a similar machine that does not use composites.
If the total energy is decreased by using the composite, the composite will be a
useful material, because, overall we will be able to save energy. The total energy
Etotal is expressed by:

Etotal ¼ EP þ EU þ ER ; (9.18)

where EP is the energy for production of the composite, EU is the total operation
energy (fuel) used for a machine for the average lifetime of the machine, and ER is
the energy for recycling or disposal of the composite. EP and ER are clear and easy
to understand. A good example to help explain EU is the main wing of an airplane. If
the wings are made of a fiber reinforced polymer (FRP) such as is found in the
B787, the weight of the airplane will decrease and fuel consumption will be
significantly reduced compared to the consumption of an aircraft using aluminum
alloy wings. In other words, the EU of the airplane with the FRP wing will be
significantly smaller than that of an airplane with the aluminum alloy wing over the
average lifetime of the airplane. This difference in EU more than compensates for
differences in EP and ER, and means that the total energy, Etotal, will be smaller than
that of an airplane with an aluminum wing. This approach can be used to assess
200 9 Recycling of Composites

most applications of composites. However, Eq. (9.18) is not an appropriate way to


assess the value of special functional composites such as oil-less self-lubricating
materials or lightweight, thermally conductive materials, which have been realized
using carbon/metal composites. Equation (9.18) may need to be modified so that it
may be applied to these functional composites.

References

1. Ohnishi, T.: Social environment and problems in aluminum recycling. J. Jpn. Inst. Light Met.
46, 525–532 (1996)
2. Murata, F.: Research and development of technology to promote recycling of aluminum
materials. J. Jpn. Inst. Light Met. 46, 551–556 (1996)
3. Peterson, R.D.: Issues in the melting and reclamation of aluminum scrap. JOM. 47, 27–29
(1995)
4. Friesen, K.J., Utigard, T.A., Dupuis, C., Martin, J.P.: Coalescence behavior of aluminium
droplets under a molten salt flux cover. Light Metals, pp. 857–864. TMS, Warrendale (1997)
5. Donomoto, T., Miura, N., Funatani, K., Miyake, N.: Ceramic fiber reinforced piston for high
performance diesel engines. SAE Paper No. 830252 (1983)
6. Yamauchi, T.: Development of SiC whiskers reinforced piston. SAE Paper No. 911284 (1991)
7. Hayashi, T., Ushio, H., Ebisawa, M.: The properties of hybrid fiber reinforced metal and its
application for engine block. SAE Paper No. 890559 (1989)
8. Inoue, T., Inayoshi, H., Kanematsu, H., Kunieda, Y., Hayashi, S., Oki, T.: The recovery of
aluminum from aluminum matrix composites by a molten salt process. J. Jpn. Inst. Light Met.
46, 183–188 (1996)
9. Nishida, Y.: Development of pressure infiltration method for fabrication of metal matrix
composites. Mater. Jpn. 36, 40–46 (1997)
10. Nishida, Y., Izawa, N., Kuramasu, Y.: Recycling of aluminum matrix composites. Metall.
Mater. Trans. 30A, 839–844 (1999)
11. Girifalco, L.A., Good, R.J.: A theory for the estimation of surface and interfacial energy.
1. Derivation and application to interfacial tension. J. Phys. Chem. 61, 904–909 (1957)
12. Adamson, A.W.: Physical Chemistry of Surfaces, 4th edn, p. 107. Wiley, New York (1982)
13. Utigard, T.A., Toguri, J.M.: Interfacial tension of aluminum in cryolite melts. Metall. Trans.
16B, 333–338 (1985)
14. Brennan, J.J., Pask, J.A.: Effect of nature of surfaces on wetting of sapphire by liquid
aluminum. J. Am. Ceram. Soc. 51, 569–573 (1968)
15. Ho, F.K., Sahai, Y.: Interfacial tension in molten aluminum and salt systems. Light Metals,
pp. 717–720. TMS, Warrendale (1990)
16. Silny, A., Utigard, T.A.: Interfacial tension between aluminum and flux. Light Metals,
pp. 871–878. TMS, Warrendale (1997)
17. Kingery, W.D.: Metal-ceramic interactions: IV, absolute measurement of metal-ceramic
interfacial energy and the interfacial adsorption of silicon from iron-silicon alloys. J. Am.
Ceram. Soc. 37, 42–45 (1954)
18. Madarasz, D., Budai, I., Kaptay, G.: Fabrication of SiC-particles-shielded Al spheres upon
recycling Al/SiC composites. Metall. Mater. Trans. 42A, 1439–1443 (2011)
Index

A Cumulative distribution function, 123


Active material, 152 CVD. See Chemical vapor deposition (CVD)
Actuator, 152 CVD process, 148, 173, 174
Adaptive material, 152
Apparent viscosity, 43, 111
D
Darcy’s law, 23–25, 64, 67, 69, 71, 72, 95
B Diffusion bonding process, 39, 40
Bulk modulus, 145 Discontinuous fiber reinforced composites, 3,
128–130
Dislocation, 9, 131, 133, 135, 158–161, 179
C Dispersion strengthened metals, 3, 9, 38, 39,
Carbon fiber reinforced carbon (CC 49, 131
composite), 2, 3
Carbon nanofiber, 151
Carbon nanotube, 87, 88, 151, 175, 177 E
Centrifugal casting, 28, 91–111, 189 Equivalent inclusion method, 122
Centrifugal force, 91–110, 189 Eshelby method, 122
Charpy impact tester, 133
Chemical vapor deposition (CVD), 28, 34,
46–47, 150, 171, 173, 174, 177 F
Coefficient of linear thermal expansion, 18 Fatigue limit, 135, 136
Compact type specimen (CT specimen), 137 Fiber reinforced ceramics (FRC), 2, 3
Compo-casting, 28, 34, 43–44 Fiber reinforced glasses (FRG), 2, 3
Constitutive equation, 160 Fiber reinforced metals (FRM), 1, 4, 54,
Contact angle, 16, 19–22, 30–32, 41, 45, 127, 128
62, 63, 87, 95, 150, 187, 190, Fiber reinforced plastics (FRP), 2–4, 201
194, 196, 198 Fourier’s law, 142
Continuous fiber reinforced composites, 2, 3, Fracture mechanics, 133
34, 39, 40, 54, 116, 117, 126–128, FRG. See Fiber reinforced glasses (FRG)
130, 134, 142, 143, 145, 148 FRM. See Fiber reinforced metals (FRM)
Crack growth rate, 137–139 Front, 62, 68–83, 88, 92, 93, 96–100, 102–105,
Critical aspect ratio, 129 134, 139
Critical length, 129 FRP. See Fiber reinforced plastics (FRP)

Y. Nishida, Introduction to Metal Matrix Composites: Fabrication and Recycling, 201


DOI 10.1007/978-4-431-54237-7, # Springer Japan 2013
202 Index

G P
Gas atomization, 37 PAN system carbon fiber, 172, 175, 176
Gas state fabrication technique, 27, 28, 34 Particulate reinforced composites, 3, 7, 89,
Graphite fiber, 151, 176 118, 130–132, 139, 144, 146
Passive material, 152
Permeability, 62, 67, 68, 75, 77, 82–84,
H 86–88, 96, 100
Hagen–Poiseuille law, 63, 64 Permeability coefficient, 23
High-strain-rate superplasticity (HSRS), 158 Physical vapor deposition (PVD), 28, 34,
Hooke’s law, 120 46–48
Hot isostatic pressing (HIP), 36 Poiseuille equation, 63
Hot press (HP), 36, 40 Preform, 4, 54, 56, 67–88, 91, 147,
180, 184
Pressure infiltration process, 4, 7, 28, 34,
I 53–55, 59, 73, 87–89
Infiltration velocity model, 72–88 Pressureless infiltration process, 28,
In-situ fabrication technique, 34, 48–49 44–45
In-situ process, 27, 28 Probability density function, 123, 124
Intelligent material, 151, 152 Pullout, 134
Interface energy (interfacial energy), 11–16, PVD. See Physical vapor
19, 21, 28, 30, 32, 34, 64, 150, 185, deposition (PVD)
187–190, 194, 195, 198, 199
Intermetallic compound matrix composites
(IMC), 2, 3 R
Internal nitridation process, 49 Reynolds number, 25, 67, 71,
Internal oxidation process, 48–49 72, 109
Rheocasting process, 43
Rule of mixtures, 9–11, 116, 117, 121, 126,
K 143, 145, 147
Kerner’s equation, 145–147

S
L Scale-like graphite particle, 41, 151
Lamellar spacing, 49 Schapery’s equation, 145–147
Laminar flow, 23, 25, 67, 68 Sensory material, 152
Lanxide, 28, 44, 45 Shear lag model, 118, 122
Liquid state fabrication technique, 28, 30, Shear modulus, 121, 131, 136, 145,
34, 41–46 160, 168, 169
Shot, 136, 137
SiC whisker, 6, 54, 72, 74, 77–81, 84–87,
M 135, 139, 157, 164, 178, 184,
Matrix, 2 193, 194
Mechanical alloying (MA), 9, 28, 38–39 Sintered aluminum products (SAP), 9, 28,
Metal matrix composites (MMC), 2, 3 37, 38
Slurry, 43, 44
Smart composite, 151–152
N Smart material, 152
Navier–Stokes equations, 25 S–N curve, 135–137
Newtonian fluids, 23, 63 Solid state fabrication technique, 28, 29,
Nonlinear fracture mechanics, 133 34–41
Specific permeability, 23
Sputtering process, 47
O Standard deviation, 126
Orowan looping, 131 Stokes’ drag formula, 108, 109
Index 203

Stokes’ law, 109 Turbulent flow, 25


Stress intensity Turner’s equation, 145, 147
factor, 133
range, 138, 139
Substrate, 45–48, 150, 178 V
Superplasticity, 157–165 Vapor–liquid–solid (VLS) mechanism, 178
Surface energy, 11–19, 27–30, 32, 36, 95, Volumetric coefficient of thermal expansion,
185, 187, 190, 191, 194–196 144, 146
Vortex addition technique, 28, 41–42

T
Target, 47, 48, 106, 122 W
TD nickel, 9 Weibull distribution, 123–126
Thermal expansion coefficient (coefficient Wettability, 19–22, 46
of linear thermal expansion), 4, 10, Whisker reinforced composite, 85, 87
144–149, 168–171 Work of adhesion, 21
Thermal ratchet, 148 Work of cohesion, 22
Threshold
pressure, 45, 55–58, 62–64, 68, 69,
71–73, 75–77, 83, 87, 95–97, 99, Y
102, 103, 108 Young’s equation, 16, 19–21, 30–32, 56,
stress, 160 194, 198

You might also like