1496T_fm_i-xxvi 1/6/06 02:56 Page iii
Materials Science and Engineering
An Introduction
1496T_fm_i-xxvi 1/6/06 22:25 Page v
SEVENTH EDITION
Materials Science
and Engineering
An Introduction
William D. Callister, Jr.
Department of Metallurgical Engineering
The University of Utah
with special contributions by
David G. Rethwisch
The University of Iowa
John Wiley & Sons, Inc.
1496T_fm_i-xxvi 1/11/06 23:05 Page vi
Front Cover: A unit cell for diamond (blue-gray spheres represent carbon atoms), which is positioned
above the temperature-versus-logarithm pressure phase diagram for carbon; highlighted in blue is the
region for which diamond is the stable phase.
Back Cover: Atomic structure for graphite; here the gray spheres depict carbon atoms. The region of
graphite stability is highlighted in orange on the pressure-temperature phase diagram for carbon,
which is situated behind this graphite structure.
ACQUISITIONS EDITOR
MARKETING DIRECTOR
SENIOR PRODUCTION EDITOR
SENIOR DESIGNER
COVER ART
TEXT DESIGN
SENIOR ILLUSTRATION EDITOR
COMPOSITOR
ILLUSTRATION STUDIO
Joseph Hayton
Frank Lyman
Ken Santor
Kevin Murphy
Roy Wiemann
Michael Jung
Anna Melhorn
Techbooks/GTS, York, PA
Techbooks/GTS, York, PA
This book was set in 10/12 Times Ten by Techbooks/GTS, York, PA and printed and bound by
Quebecor Versailles. The cover was printed by Quebecor.
This book is printed on acid free paper.
Copyright © 2007 John Wiley & Sons, Inc. All rights reserved.
No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form
or by any means, electronic, mechanical, photocopying, recording, scanning or otherwise, except as
permitted under Sections 107 or 108 of the 1976 United States Copyright Act, without either the prior
written permission of the Publisher, or authorization through payment of the appropriate per-copy fee
to the Copyright Clearance Center, 222 Rosewood Drive, Danvers, MA 01923, (508)750-8400, fax
(508)750-4470. Requests to the Publisher for permission should be addressed to the Permissions
Department, John Wiley & Sons, Inc., 605 Third Avenue, New York, NY 10158-0012, (212) 850-6011,
fax (212) 850-6008, E-Mail:
[email protected].
To order books or for customer service please call 1(800)225-5945.
Library of Congress Cataloging-in-Publication Data
Callister, William D., 1940Materials science and engineering : an introduction / William D. Callister, Jr.—7th ed.
p. cm.
Includes bibliographical references and index.
ISBN-13: 978-0-471-73696-7 (cloth)
ISBN-10: 0-471-73696-1 (cloth)
1. Materials. I. Title.
TA403.C23 2007
620.1’1—dc22
2005054228
Printed in the United States of America
10 9 8 7 6 5 4 3 2 1
1496T_fm_i-xxvi 1/6/06 02:56 Page vii
Dedicated to
my colleagues and friends in Brazil and Spain
1496T_fm_i-xxvi 1/6/06 02:56 Page viii
1496T_fm_i-xxvi 1/6/06 03:19 Page xv
Contents
LIST
OF
SYMBOLS xxiii
1. Introduction
1.1
1.2
1.3
1.4
1.5
1.6
1
Learning Objectives 2
Historical Perspective 2
Materials Science and Engineering 3
Why Study Materials Science and Engineering? 5
Classification of Materials 5
Advanced Materials 11
Modern Materials’ Needs 12
References 13
2. Atomic Structure and Interatomic Bonding
2.1
15
Learning Objectives 16
Introduction 16
ATOMIC STRUCTURE
16
2.2
2.3
2.4
Fundamental Concepts 16
Electrons in Atoms 17
The Periodic Table 23
2.5
2.6
2.7
2.8
Bonding Forces and Energies 24
Primary Interatomic Bonds 26
Secondary Bonding or van der Waals Bonding 30
Molecules 32
ATOMIC BONDING
IN
SOLIDS
24
Summary 34
Important Terms and Concepts 34
References 35
Questions and Problems 35
3. The Structure of Crystalline Solids
3.1
Learning Objectives 39
Introduction 39
3.2
3.3
3.4
3.5
3.6
Fundamental Concepts 39
Unit Cells 40
Metallic Crystal Structures 41
Density Computations 45
Polymorphism and Allotropy 46
CRYSTAL STRUCTURES
38
39
• xv
1496T_fm_i-xxvi 1/6/06 02:56 Page xvi
xvi • Contents
3.7
Crystal Systems 46
CRYSTALLOGRAPHIC POINTS, DIRECTIONS,
PLANES 49
3.8
3.9
3.10
3.11
3.12
Point Coordinates 49
Crystallographic Directions 51
Crystallographic Planes 55
Linear and Planar Densities 60
Close-Packed Crystal Structures
AND
3.17
61
Single Crystals 63
Polycrystalline Materials 64
Anisotropy 64
X-Ray Diffraction: Determination of
Crystal Structures 66
Noncrystalline Solids 71
Summary 72
Important Terms and Concepts 73
References 73
Questions and Problems 74
4. Imperfections in Solids 80
6. Mechanical Properties of Metals
6.1
6.2
6.3
6.4
6.5
Learning Objectives 81
Introduction 81
6.9
POINT DEFECTS
6.10
4.2
4.3
4.4
Vacancies and Self-Interstitials 81
Impurities in Solids 83
Specification of Composition 85
81
88
4.5
4.6
4.7
4.8
Dislocations–Linear Defects 88
Interfacial Defects 92
Bulk or Volume Defects 96
Atomic Vibrations 96
4.9
4.10
4.11
General 97
Microscopic Techniques 98
Grain Size Determination 102
MICROSCOPIC EXAMINATION
5.1
5.2
5.3
AND
DESIGN/SAFETY
Variability of Material Properties 161
Design/Safety Factors 163
Summary 165
Important Terms and Concepts 166
References 166
Questions and Problems 166
Design Problems 172
97
Summary 104
Important Terms and Concepts 105
References 105
Questions and Problems 106
Design Problems 108
5. Diffusion
143
Tensile Properties 144
True Stress and Strain 151
Elastic Recovery after Plastic
Deformation 154
Compressive, Shear, and Torsional
Deformation 154
Hardness 155
PROPERTY VARIABILITY
FACTORS 161
6.11
6.12
137
Stress-Strain Behavior 137
Anelasticity 140
Elastic Properties of Materials 141
PLASTIC DEFORMATION
6.6
6.7
6.8
131
Learning Objectives 132
Introduction 132
Concepts of Stress and Strain 133
ELASTIC DEFORMATION
4.1
MISCELLANEOUS IMPERFECTIONS
Nonsteady-State Diffusion 114
Factors That Influence Diffusion 118
Other Diffusion Paths 125
Summary 125
Important Terms and Concepts 126
References 126
Questions and Problems 126
Design Problems 129
CRYSTALLINE AND NONCRYSTALLINE
MATERIALS 63
3.13
3.14
3.15
3.16
5.4
5.5
5.6
109
Learning Objectives 110
Introduction 110
Diffusion Mechanisms 111
Steady-State Diffusion 112
7. Dislocations and Strengthening
Mechanisms 174
7.1
Learning Objectives 175
Introduction 175
DISLOCATIONS
DEFORMATION
7.2
7.3
7.4
7.5
7.6
7.7
PLASTIC
175
AND
Basic Concepts 175
Characteristics of Dislocations 178
Slip Systems 179
Slip in Single Crystals 181
Plastic Deformation of Polycrystalline
Materials 185
Deformation by Twinning 185
1496T_fm_i-xxvi 01/10/06 22:13 Page xvii
Contents • xvii
MECHANISMS OF STRENGTHENING
METALS 188
IN
7.8
7.9
7.10
Strengthening by Grain Size
Reduction 188
Solid-Solution Strengthening 190
Strain Hardening 191
RECOVERY, RECRYSTALLIZATION,
GROWTH 194
7.11
7.12
7.13
AND
9.2
9.3
9.4
9.5
9.6
BINARY PHASE DIAGRAMS
GRAIN
9.7
9.8
9.9
Recovery 195
Recrystallization 195
Grain Growth 200
Summary 201
Important Terms and Concepts 202
References 202
Questions and Problems 202
Design Problems 206
9.10
9.11
9.12
9.13
8. Failure
8.1
207
Learning Objectives 208
Introduction 208
FRACTURE
8.2
8.3
8.4
8.5
8.6
9.14
9.15
208
Fundamentals of Fracture 208
Ductile Fracture 209
Brittle Fracture 211
Principles of Fracture Mechanics 215
Impact Fracture Testing 223
FATIGUE
227
8.7
8.8
8.9
8.10
8.11
Cyclic Stresses 228
The S–N Curve 229
Crack Initiation and Propagation 232
Factors That Affect Fatigue Life 234
Environmental Effects 237
8.12
8.13
8.14
8.15
Generalized Creep Behavior 238
Stress and Temperature Effects 239
Data Extrapolation Methods 241
Alloys for High-Temperature
Use 242
CREEP
9.16
9.17
9.18
9.19
9.20
Summary 302
Important Terms and Concepts 303
References 303
Questions and Problems 304
10. Phase Transformations in Metals:
Development of Microstructure
and Alteration of Mechanical
Properties 311
10.1
9.1
10.2
10.3
252
Learning Objectives 253
Introduction 253
DEFINITIONS
AND
BASIC CONCEPTS
10.4
253
290
The Iron–Iron Carbide (Fe–Fe3C) Phase
Diagram 290
Development of Microstructure in
Iron–Carbon Alloys 293
The Influence of Other Alloying
Elements 301
Learning Objectives 312
Introduction 312
PHASE TRANSFORMATIONS
9. Phase Diagrams
258
Binary Isomorphous Systems 258
Interpretation of Phase Diagrams 260
Development of Microstructure in
Isomorphous Alloys 264
Mechanical Properties of Isomorphous
Alloys 268
Binary Eutectic Systems 269
Development of Microstructure in
Eutectic Alloys 276
Equilibrium Diagrams Having
Intermediate Phases or
Compounds 282
Eutectic and Peritectic Reactions 284
Congruent Phase
Transformations 286
Ceramic and Ternary Phase
Diagrams 287
The Gibbs Phase Rule 287
THE IRON–CARBON SYSTEM
238
Summary 243
Important Terms and Concepts 245
References 246
Questions and Problems 246
Design Problems 250
Solubility Limit 254
Phases 254
Microstructure 255
Phase Equilibria 255
One-Component (or Unary) Phase
Diagrams 256
312
Basic Concepts 312
The Kinetics of Phase
Transformations 313
Metastable versus Equilibrium
States 324
1496T_fm_i-xxvi 01/10/06 22:13 Page xviii
xviii • Contents
MICROSTRUCTURAL AND PROPERTY CHANGES
IRON–CARBON ALLOYS 324
10.5
10.6
10.7
10.8
10.9
Isothermal Transformation Diagrams 325
Continuous Cooling Transformation
Diagrams 335
Mechanical Behavior of Iron–Carbon
Alloys 339
Tempered Martensite 343
Review of Phase Transformations and
Mechanical Properties for Iron–Carbon
Alloys 346
Summary 350
Important Terms and Concepts 351
References 352
Questions and Problems 352
Design Problems 356
11. Applications and Processing of
Metal Alloys 358
11.1
OF
METAL ALLOYS
OF
METALS
382
11.4
11.5
11.6
Forming Operations 383
Casting 384
Miscellaneous Techniques 386
11.7
11.8
11.9
Annealing Processes 388
Heat Treatment of Steels 390
Precipitation Hardening 402
THERMAL PROCESSING
OF
METALS
Summary 453
Important Terms and Concepts 454
References 454
Questions and Problems 455
Design Problems 459
13. Applications and Processing of
Ceramics 460
13.1
12.2
12.3
12.4
12.5
12.6
Learning Objectives 461
Introduction 461
Glasses 461
Glass–Ceramics 462
Clay Products 463
Refractories 464
Abrasives 466
Cements 467
Advanced Ceramics 468
OF
13.9
Fabrication and Processing of Glasses
and Glass–Ceramics 471
13.10 Fabrication and Processing of Clay
Products 476
13.11 Powder Pressing 481
13.12 Tape Casting 484
Summary 484
Important Terms and Concepts 486
References 486
Questions and Problems 486
Design Problem 488
Learning Objectives 415
Introduction 415
14. Polymer Structures
CERAMIC STRUCTURES
14.1
14.2
14.3
14.4
Crystal Structures 415
Silicate Ceramics 426
Carbon 430
Imperfections in Ceramics 434
Diffusion in Ionic Materials 438
OF
FABRICATION AND PROCESSING
CERAMICS 471
12. Structures and Properties of
Ceramics 414
415
442
12.8 Brittle Fracture of Ceramics 442
12.9 Stress–Strain Behavior 447
12.10 Mechanisms of Plastic
Deformation 449
12.11 Miscellaneous Mechanical
Considerations 451
387
Summary 407
Important Terms and Concepts 409
References 409
Questions and Problems 410
Design Problems 411
12.1
Ceramic Phase Diagrams 439
MECHANICAL PROPERTIES
13.2
13.3
13.4
13.5
13.6
13.7
13.8
359
Ferrous Alloys 359
Nonferrous Alloys 372
FABRICATION
12.7
TYPES AND APPLICATIONS
CERAMICS 461
Learning Objectives 359
Introduction 359
TYPES
11.2
11.3
IN
14.5
489
Learning Objectives 490
Introduction 490
Hydrocarbon Molecules 490
Polymer Molecules 492
The Chemistry of Polymer
Molecules 493
Molecular Weight 497
1496T_fm_i-xxvi 01/10/06 22:13 Page xix
Contents • xix
14.6
14.7
14.8
14.9
14.10
14.11
14.12
14.13
14.14
Molecular Shape 500
Molecular Structure 501
Molecular Configurations 503
Thermoplastic and Thermosetting
Polymers 506
Copolymers 507
Polymer Crystallinity 508
Polymer Crystals 512
Defects in Polymers 514
Diffusion in Polymeric Materials 515
Summary 517
Important Terms and Concepts 519
References 519
Questions and Problems 519
15.2
15.3
15.4
15.5
15.6
OF
POLYMERS
MECHANISMS OF DEFORMATION AND FOR
STRENGTHENING OF POLYMERS 535
15.7
15.8
15.9
Deformation of Semicrystalline
Polymers 535
Factors That Influence the Mechanical
Properties of Semicrystalline
Polymers 538
Deformation of Elastomers 541
CRYSTALLIZATION, MELTING, AND GLASS
TRANSITION PHENOMENA IN POLYMERS 544
15.10 Crystallization 544
15.11 Melting 545
15.12 The Glass Transition 545
15.13 Melting and Glass Transition
Temperatures 546
15.14 Factors That Influence Melting and Glass
Transition Temperatures 547
POLYMER TYPES
15.15
15.16
15.17
15.18
15.19
16.1
560
Polymerization 561
Polymer Additives 563
Forming Techniques for Plastics 565
Fabrication of Elastomers 567
Fabrication of Fibers and Films 568
577
Learning Objectives 578
Introduction 578
16.4
16.5
16.6
16.7
16.8
16.9
16.10
16.11
16.12
16.13
580
Large-Particle Composites 580
Dispersion-Strengthened
Composites 584
FIBER-REINFORCED COMPOSITES
585
Influence of Fiber Length 585
Influence of Fiber Orientation and
Concentration 586
The Fiber Phase 595
The Matrix Phase 596
Polymer-Matrix Composites 597
Metal-Matrix Composites 603
Ceramic-Matrix Composites 605
Carbon–Carbon Composites 606
Hybrid Composites 607
Processing of Fiber-Reinforced
Composites 607
STRUCTURAL COMPOSITES
610
16.14 Laminar Composites 610
16.15 Sandwich Panels 611
Summary 613
Important Terms and Concepts 615
References 616
Questions and Problems 616
Design Problems 619
17. Corrosion and Degradation of
Materials 621
549
Plastics 549
Elastomers 552
Fibers 554
Miscellaneous Applications 555
Advanced Polymeric Materials 556
PROCESSING
PARTICLE-REINFORCED COMPOSITES
524
Stress–Strain Behavior 524
Macroscopic Deformation 527
Viscoelastic Deformation 527
Fracture of Polymers 532
Miscellaneous Mechanical
Characteristics 533
AND
Summary 569
Important Terms and Concepts 571
References 571
Questions and Problems 572
Design Questions 576
16.2
16.3
Learning Objectives 524
Introduction 524
MECHANICAL BEHAVIOR
15.20
15.21
15.22
15.23
15.24
16. Composites
15. Characteristics, Applications, and
Processing of Polymers 523
15.1
POLYMER SYNTHESIS
17.1
Learning Objectives 622
Introduction 622
17.2
17.3
Electrochemical Considerations 623
Corrosion Rates 630
CORROSION
OF
METALS
622
1496T_fm_i-xxvi 01/10/06 22:13 Page xx
xx • Contents
17.4 Prediction of Corrosion Rates 631
17.5 Passivity 638
17.6 Environmental Effects 640
17.7 Forms of Corrosion 640
17.8 Corrosion Environments 648
17.9 Corrosion Prevention 649
17.10 Oxidation 651
CORROSION
OF
DEGRADATION
CERAMIC MATERIALS
OF
POLYMERS
654
655
18.20 Types of Polarization 708
18.21 Frequency Dependence of the Dielectric
Constant 709
18.22 Dielectric Strength 711
18.23 Dielectric Materials 711
OTHER ELECTRICAL CHARACTERISTICS
MATERIALS 711
18.24 Ferroelectricity 711
18.25 Piezoelectricity 712
Summary 713
Important Terms and Concepts 715
References 715
Questions and Problems 716
Design Problems 720
17.11 Swelling and Dissolution 655
17.12 Bond Rupture 657
17.13 Weathering 658
Summary 659
Important Terms and Concepts 660
References 661
Questions and Problems 661
Design Problems 644
18. Electrical Properties
18.1
665
Learning Objectives 666
Introduction 666
ELECTRICAL CONDUCTION
18.2
18.3
18.4
18.5
18.6
18.7
18.8
18.9
666
Ohm’s Law 666
Electrical Conductivity 667
Electronic and Ionic Conduction 668
Energy Band Structures in
Solids 668
Conduction in Terms of Band and
Atomic Bonding Models 671
Electron Mobility 673
Electrical Resistivity of Metals 674
Electrical Characteristics of Commercial
Alloys 677
SEMICONDUCTIVITY
679
18.10 Intrinsic Semiconduction 679
18.11 Extrinsic Semiconduction 682
18.12 The Temperature Dependence of Carrier
Concentration 686
18.13 Factors That Affect Carrier Mobility 688
18.14 The Hall Effect 692
18.15 Semiconductor Devices 694
ELECTRICAL CONDUCTION
POLYMERS 700
IN
IONIC CERAMICS
AND IN
18.16 Conduction in Ionic Materials 701
18.17 Electrical Properties of Polymers 701
DIELECTRIC BEHAVIOR
702
18.18 Capacitance 703
18.19 Field Vectors and Polarization 704
OF
19. Thermal Properties
19.1
19.2
19.3
19.4
19.5
W1
Learning Objectives W2
Introduction W2
Heat Capacity W2
Thermal Expansion W4
Thermal Conductivity W7
Thermal Stresses W12
Summary W14
Important Terms and Concepts W15
References W15
Questions and Problems W15
Design Problems W17
20. Magnetic Properties
W19
Learning Objectives W20
Introduction W20
Basic Concepts W20
Diamagnetism and
Paramagnetism W24
20.4 Ferromagnetism W26
20.5 Antiferromagnetism and
Ferrimagnetism W28
20.6 The Influence of Temperature on
Magnetic Behavior W32
20.7 Domains and Hysteresis W33
20.8 Magnetic Anisotropy W37
20.9 Soft Magnetic Materials W38
20.10 Hard Magnetic Materials W41
20.11 Magnetic Storage W44
20.12 Superconductivity W47
20.1
20.2
20.3
Summary W50
Important Terms and Concepts W52
References W52
Questions and Problems W53
Design Problems W56
1496T_fm_i-xxvi 01/10/06 22:13 Page xxi
Contents • xxi
21.
21.1
Optical Properties
Learning Objectives W58
Introduction W58
BASIC CONCEPTS
21.2
21.3
21.4
22.8
22.9
W57
ARTIFICIAL TOTAL HIP REPLACEMENT
W58
Electromagnetic Radiation W58
Light Interactions with Solids W60
Atomic and Electronic
Interactions W61
OPTICAL PROPERTIES
OF
METALS
OPTICAL PROPERTIES
OF
NONMETALS
W62
W63
21.5 Refraction W63
21.6 Reflection W65
21.7 Absorption W65
21.8 Transmission W68
21.9
Color W69
21.10 Opacity and Translucency in
Insulators W71
APPLICATIONS
OF
OPTICAL PHENOMENA
W72
Summary W82
Important Terms and Concepts W83
References W84
Questions and Problems W84
Design Problem W85
22. Materials Selection and Design
Considerations W86
Learning Objectives W87
Introduction W87
MATERIALS SELECTION FOR A TORSIONALLY
STRESSED CYLINDRICAL SHAFT W87
22.2
22.3
Strength Considerations–Torsionally
Stressed Shaft W88
Other Property Considerations and the
Final Decision W93
AUTOMOTIVE VALVE SPRING
22.4
22.5
22.6
22.7
W108
22.10 Anatomy of the Hip Joint W108
22.11 Material Requirements W111
22.12 Materials Employed W112
CHEMICAL PROTECTIVE CLOTHING
W115
22.13 Introduction W115
22.14 Assessment of CPC Glove Materials to
Protect Against Exposure to Methylene
Chloride W115
MATERIALS FOR INTEGRATED CIRCUIT
PACKAGES W119
21.11 Luminescence W72
21.12 Photoconductivity W72
21.13 Lasers W75
21.14 Optical Fibers in Communications W79
22.1
Testing Procedure and Results W102
Discussion W108
22.15
22.16
22.17
22.18
22.19
22.20
Introduction W119
Leadframe Design and Materials W120
Die Bonding W121
Wire Bonding W124
Package Encapsulation W125
Tape Automated Bonding W127
Summary W129
References W130
Design Questions and Problems W131
23. Economic, Environmental, and
Societal Issues in Materials Science
and Engineering W135
23.1
Learning Objectives W136
Introduction W136
ECONOMIC CONSIDERATIONS
23.2
23.3
23.4
W136
Component Design W137
Materials W137
Manufacturing Techniques W137
ENVIRONMENTAL AND SOCIETAL
CONSIDERATIONS W137
23.5
Recycling Issues in Materials Science and
Engineering W140
Summary W143
References W143
Design Question W144
W94
Mechanics of Spring Deformation W94
Valve Spring Design and Material
Requirements W95
One Commonly Employed Steel
Alloy W98
Appendix A The International System of
Units A1
FAILURE OF AN AUTOMOBILE REAR
AXLE W101
B.1
B.2
B.3
Introduction W101
Appendix B Properties of Selected
Engineering Materials A3
Density A3
Modulus of Elasticity A6
Poisson’s Ratio A10
1496T_fm_i-xxvi 01/10/06 22:13 Page xxii
xxii • Contents
B.4
B.5
B.6
B.7
B.8
B.9
B.10
Strength and Ductility A11
Plane Strain Fracture Toughness A16
Linear Coefficient of Thermal
Expansion A17
Thermal Conductivity A21
Specific Heat A24
Electrical Resistivity A26
Metal Alloy Compositions A29
Appendix E Glass Transition and Melting
Temperatures for Common Polymeric
Materials A41
Glossary
Answers to Selected Problems
Index
Appendix C Costs and Relative Costs for
Selected Engineering Materials A31
Appendix D Repeat Unit Structures for
Common Polymers A37
G0
I1
S1
1496T_fm_i-xxvi 1/6/06 02:56 Page ix
Preface
Imaterials
n this Seventh Edition I have retained the objectives and approaches for teaching
science and engineering that were presented in previous editions. The first,
and primary, objective is to present the basic fundamentals on a level appropriate for
university/college students who have completed their freshmen calculus, chemistry, and
physics courses. In order to achieve this goal, I have endeavored to use terminology
that is familiar to the student who is encountering the discipline of materials science
and engineering for the first time, and also to define and explain all unfamiliar terms.
The second objective is to present the subject matter in a logical order, from the
simple to the more complex. Each chapter builds on the content of previous ones.
The third objective, or philosophy, that I strive to maintain throughout the text
is that if a topic or concept is worth treating, then it is worth treating in sufficient
detail and to the extent that students have the opportunity to fully understand it
without having to consult other sources; also, in most cases, some practical relevance
is provided. Discussions are intended to be clear and concise and to begin at
appropriate levels of understanding.
The fourth objective is to include features in the book that will expedite the
learning process. These learning aids include:
• Numerous illustrations, now presented in full color, and photographs to
help visualize what is being presented;
• Learning objectives;
• “Why Study . . .” and “Materials of Importance” items that provide relevance to topic discussions;
• Key terms and descriptions of key equations highlighted in the margins for
quick reference;
• End-of-chapter questions and problems;
• Answers to selected problems;
• A glossary, list of symbols, and references to facilitate understanding the
subject matter.
The fifth objective is to enhance the teaching and learning process by using
the newer technologies that are available to most instructors and students of
engineering today.
FEATURES THAT ARE NEW
TO
THIS EDITION
New/Revised Content
Several important changes have been made with this Seventh Edition. One of
the most significant is the incorporation of a number of new sections, as well
• ix
1496T_fm_i-xxvi 01/10/06 22:13 Page x
x • Preface
as revisions/amplifications of other sections. New sections/discussions are as
follows:
• One-component (or unary) phase diagrams (Section 9.6)
• Compacted graphite iron (in Section 11.2, “Ferrous Alloys”)
• Lost foam casting (in Section 11.5, “Casting”)
• Temperature dependence of Frenkel and Schottky defects (in Section 12.5,
“Imperfections in Ceramics”)
• Fractography of ceramics (in Section 12.8, “Brittle Fracture of Ceramics”)
• Crystallization of glass-ceramics, in terms of isothermal transformation
and continuous cooling transformation diagrams (in Section 13.3,
“Glass-Ceramics”)
• Permeability in polymers (in Section 14.14, “Diffusion in Polymeric
Materials”)
• Magnetic anisotropy (Section 20.8)
• A new case study on chemical protective clothing (Sections 22.13 and
22.14).
Those sections that have been revised/amplified, include the following:
• Treatments in Chapter 1 (“Introduction”) on the several material types
have been enlarged to include comparisons of various property values (as
bar charts).
• Expanded discussions on crystallographic directions and planes in hexagonal
crystals (Sections 3.9 and 3.10); also some new related homework problems.
• Comparisons of (1) dimensional size ranges for various structural elements,
and (2) resolution ranges for the several microscopic examination techniques (in Section 4.10, “Microscopic Techniques”).
• Updates on hardness testing techniques (Section 6.10).
• Revised discussion on the Burgers vector (Section 7.4).
• New discussion on why recrystallization temperature depends on the purity
of a metal (Section 7.12).
• Eliminated some detailed discussion on fracture mechanics—i.e., used
“Concise Version” from sixth edition (Section 8.5).
• Expanded discussion on nondestructive testing (Section 8.5).
• Used Concise Version (from sixth edition) of discussion on crack initiation
and propagation (for fatigue, Section 8.9), and eliminated section on crack
propagation rate.
• Refined terminology and representations of polymer structures (Sections
14.3 through 14.8).
• Eliminated discussion on fringed-micelle model (found in Section 14.12 of
the sixth edition).
• Enhanced discussion on defects in polymers (Section 14.13).
• Revised the following sections in Chapter 15 (“Characteristics, Applications, and Processing of Polymers”): fracture of polymers (Section 15.5),
deformation of semicrystalline polymers (Section 15.7), adhesives (in
Section 15.18), polymerization (Section 15.20), and fabrication of fibers and
films (Section 15.24).
• Revised treatment of polymer degradation (Section 17.12).
1496T_fm_i-xxvi 01/10/06 22:13 Page xi
Preface • xi
Materials of Importance
One new feature that has been incorporated in this edition is “Materials of Importance” pieces; in these we discuss familiar and interesting materials/applications
of materials. These pieces lend some relevance to topical coverage, are found in
most chapters in the book, and include the following:
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
Carbonated Beverage Containers
Water (Its Volume Expansion Upon Freezing)
Tin (Its Allotropic Transformation)
Catalysts (and Surface Defects)
Aluminum for Integrated Circuit Interconnects
Lead-Free Solders
Shape-Memory Alloys
Metal Alloys Used for Euro Coins
Carbon Nanotubes
Piezoelectric Ceramics
Shrink-Wrap Polymer Films
Phenolic Billiard Balls
Nanocomposites in Tennis Balls
Aluminum Electrical Wires
Invar and Other Low-Expansion Alloys
An Iron-Silicon Alloy That is Used in Transformer Cores
Light-Emitting Diodes
Concept Check
Another new feature included in this seventh edition is what we call a “Concept
Check,” a question that tests whether or not a student understands the subject matter on a conceptual level. Concept check questions are found within most chapters; many of them appeared in the end-of-chapter Questions and Problems sections
of the previous edition. Answers to these questions are on the book’s Web site,
www.wiley.com/college/callister (Student Companion Site).
And, finally, for each chapter, both the Summary and the Questions and Problems are organized by section; section titles precede their summaries and questions/problems.
Format Changes
There are several other major changes from the format of the sixth edition. First of
all, no CD-ROM is packaged with the in-print text; all electronic components are
found on the book’s Web site (www.wiley.com/college/callister). This includes the
last five chapters in the book—viz. Chapter 19, “Thermal Properties;” Chapter 20,
“Magnetic Properties;” Chapter 21, “Optical Properties;” Chapter 22, “Materials
Selection and Design Considerations;” and Chapter 23, “Economic, Environmental,
and Societal Issues in Materials Science and Engineering.” These chapters are in
Adobe Acrobat® pdf format and may be downloaded.
Furthermore, only complete chapters appear on the Web site (rather than selected sections for some chapters per the sixth edition). And, in addition, for all sections of the book there is only one version—for the two-version sections of the sixth
edition, in most instances, the detailed ones have been retained.
1496T_fm_i-xxvi 01/10/06 22:13 Page xii
xii • Preface
Six case studies have been relegated to Chapter 22, “Materials Selection and
Design Considerations,” which are as follows:
•
•
•
•
•
•
Materials Selection for a Torsionally Stressed Cylindrical Shaft
Automobile Valve Spring
Failure of an Automobile Rear Axle
Artificial Total Hip Replacement
Chemical Protective Clothing
Materials for Integrated Circuit Packages
References to these case studies are made in the left-page margins at appropriate
locations in the other chapters. All but “Chemical Protective Clothing” appeared in
the sixth edition; it replaces the “Thermal Protection System on the Space Shuttle
Orbiter” case study.
STUDENT LEARNING RESOURCES
(WWW.WILEY.COM/COLLEGE/CALLISTER)
Also found on the book’s Web site (under “Student Companion Site”) are several
important instructional elements for the student that complement the text; these
include the following:
1. VMSE: Virtual Materials Science and Engineering. This is essentially the
same software program that accompanied the previous edition, but now browserbased for easier use on a wider variety of computer platforms. It consists of interactive simulations and animations that enhance the learning of key concepts
in materials science and engineering, and, in addition, a materials properties/cost
database. Students can access VMSE via the registration code included with all
new copies.
Throughout the book, whenever there is some text or a problem that is supplemented by VMSE, a small “icon” that denotes the associated module is included in one of the margins. These modules and their corresponding icons are as
follows:
Metallic Crystal Structures
and Crystallography
Phase Diagrams
Ceramic Crystal Structures
Diffusion
Repeat Unit and Polymer
Structures
Tensile Tests
Dislocations
Solid-Solution Strengthening
2. Answers to the Concept Check questions.
3. Direct access to online self-assessment exercises. This is a Web-based assessment program that contains questions and problems similar to those found in the
text; these problems/questions are organized and labeled according to textbook
sections. An answer/solution that is entered by the user in response to a question/problem is graded immediately, and comments are offered for incorrect responses. The student may use this electronic resource to review course material, and
to assess his/her mastery and understanding of topics covered in the text.
1496T_fm_i-xxvi 1/6/06 02:56 Page xiii
Preface • xiii
4. Additional Web resources, which include the following:
• Index of Learning Styles. Upon answering a 44-item questionnaire, a user’s
learning style preference (i.e., the manner in which information is assimilated and processed) is assessed.
• Extended Learning Objectives. A more extensive list of learning objectives
than is provided at the beginning of each chapter.
• Links to Other Web Resources. These links are categorized according to
general Internet, software, teaching, specific course content/activities, and
materials databases.
INSTRUCTORS’ RESOURCES
The “Instructor Companion Site” (www.wiley.com/college/callister) is available for
instructors who have adopted this text. Resources that are available include the
following:
1. Detailed solutions of all end-of-chapter questions and problems (in both
Microsoft Word® and Adobe Acrobat® PDF formats).
2. Photographs, illustrations, and tables that appear in the book (in PDF and
JPEG formats); an instructor can print them for handouts or prepare transparencies in his/her desired format.
3. A set of PowerPoint® lecture slides developed by Peter M. Anderson (The
Ohio State University) and David G. Rethwisch (The University of Iowa). These
slides follow the flow of topics in the text, and include materials from the text and
other sources as well as illustrations and animations. Instructors may use the slides
as is or edit them to fit their teaching needs.
4. A list of classroom demonstrations and laboratory experiments that portray
phenomena and/or illustrate principles that are discussed in the book; references
are also provided that give more detailed accounts of these demonstrations.
5. Suggested course syllabi for the various engineering disciplines.
WileyPLUS
WileyPLUS gives you, the instructor, the technology to create an environment where
students reach their full potential and experience academic success that will last a
lifetime! With WileyPLUS, students will come to class better prepared for your lectures, get immediate feedback and context-sensitive help on assignments and quizzes,
and have access to a full range of interactive learning resources including a complete online version of their text. WileyPLUS gives you a wealth of presentation and
preparation tools, easy-to-navigate assessment tools including an online gradebook,
and a complete system to administer and manage your course exactly as you wish.
Contact your local Wiley representative for details on how to set up your WileyPLUS
course, or visit the website at www.wiley.com/college/wileyplus.
FEEDBACK
I have a sincere interest in meeting the needs of educators and students in the
materials science and engineering community, and therefore would like to solicit
feedback on this seventh edition. Comments, suggestions, and criticisms may be
submitted to me via e-mail at the following address:
[email protected].
ACKNOWLEDGMENTS
Appreciation is expressed to those who have made contributions to this edition.
I am especially indebted to David G. Rethwisch, who, as a special contributor,
1496T_fm_i-xxvi 01/10/06 22:13 Page xiv
xiv • Preface
provided invaluable assistance in updating and upgrading important material in a
number of chapters. In addition, I sincerely appreciate Grant E. Head’s expert programming skills, which he used in developing the Virtual Materials Science and Engineering software. Important input was also furnished by Carl Wood of Utah State
University and W. Roger Cannon of Rutgers University, to whom I also give thanks.
In addition, helpful ideas and suggestions have been provided by the following:
Tarek Abdelsalam, East Carolina University
Keyvan Ahdut, University of the District of
Columbia
Mark Aindow, University of Connecticut (Storrs)
Pranesh Aswath, University of Texas at Arlington
Mir Atiqullah, St. Louis University
Sayavur Bakhtiyarov, Auburn University
Kristen Constant, Iowa State University
Raymond Cutler, University of Utah
Janet Degrazia, University of Colorado
Mark DeGuire, Case Western Reserve University
Timothy Dewhurst, Cedarville University
Amelito Enriquez, Canada College
Jeffrey Fergus, Auburn University
Victor Forsnes, Brigham Young University (Idaho)
Paul Funkenbusch, University of Rochester
Randall German, Pennsylvania State University
Scott Giese, University of Northern Iowa
Brian P. Grady, University of Oklahoma
Theodore Greene, Wentworth Institute of
Technology
Todd Gross, University of New Hampshire
Jamie Grunlan, Texas A & M University
Masanori Hara, Rutgers University
Russell Herlache, Saginaw Valley State University
Susan Holl, California State University
(Sacramento)
Zhong Hu, South Dakota State University
Duane Jardine, University of New Orleans
Jun Jin, Texas A & M University at Galveston
Paul Johnson, Grand Valley State University
Robert Johnson, University of Texas at Arlington
Robert Jones, University of Texas (Pan American)
Maureen Julian, Virginia Tech
James Kawamoto, Mission College
Edward Kolesar, Texas Christian University
Stephen Krause, Arizona State University (Tempe)
Robert McCoy, Youngstown State University
Scott Miller, University of Missouri (Rolla)
Devesh Misra, University of Louisiana at
Lafayette
Angela L. Moran, U.S. Naval Academy
James Newell, Rowan University
Toby Padilla, Colorado School of Mines
Timothy Raymond, Bucknell University
Alessandro Rengan, Central State University
Bengt Selling, Royal Institute of Technology
(Stockholm, Sweden)
Ismat Shah, University of Delaware
Patricia Shamamy, Lawrence Technological
University
Adel Sharif, California State University at
Los Angeles
Susan Sinnott, University of Florida
Andrey Soukhojak, Lehigh University
Erik Spjut, Harvey Mudd College
David Stienstra, Rose-Hulman Institute of
Technology
Alexey Sverdlin, Bradley University
Dugan Um, Texas State University
Raj Vaidyanatha, University of Central Florida
Kant Vajpayee, University of Southern Mississippi
Kumar Virwani, University of Arkansas
(Fayetteville)
Mark Weaver, University of Alabama (Tuscaloosa)
Jason Weiss, Purdue University (West Lafayette)
I am also indebted to Joseph P. Hayton, Sponsoring Editor, and to Kenneth Santor,
Senior Production Editor at Wiley for their assistance and guidance on this revision.
Since I undertook the task of writing my first text on this subject in the early80’s, instructors and students too numerous to mention have shared their input and
contributions on how to make this work more effective as a teaching and learning
tool. To all those who have helped, I express my sincere “Thanks!”
Last, but certainly not least, the continual encouragement and support of my
family and friends is deeply and sincerely appreciated.
WILLIAM D. CALLISTER, JR.
Salt Lake City, Utah
January 2006
1496T_fm_i-xxvi 01/10/06 22:13 Page xxiii
List of Symbols
T
he number of the section in which a symbol is introduced or explained is given
in parentheses.
A area
Å angstrom unit
Ai atomic weight of element i (2.2)
APF atomic packing factor (3.4)
a lattice parameter: unit cell
x-axial length (3.4)
a crack length of a surface crack (8.5)
at% atom percent (4.4)
B magnetic flux density
(induction) (20.2)
Br magnetic remanence (20.7)
BCC body-centered cubic crystal
structure (3.4)
b lattice parameter: unit cell
y-axial length (3.7)
b Burgers vector (4.5)
C capacitance (18.18)
Ci concentration (composition) of
component i in wt% (4.4)
Ci concentration (composition) of
component i in at% (4.4)
Cv, Cp heat capacity at constant volume,
pressure (19.2)
CPR corrosion penetration rate (17.3)
CVN Charpy V-notch (8.6)
%CW percent cold work (7.10)
c lattice parameter: unit cell
z-axial length (3.7)
c velocity of electromagnetic radiation in a vacuum (21.2)
D diffusion coefficient (5.3)
D dielectric displacement (18.19)
DP degree of polymerization (14.5)
d diameter
d average grain diameter (7.8)
dhkl interplanar spacing for planes of
Miller indices h, k, and l (3.16)
E energy (2.5)
E modulus of elasticity or Young’s
modulus (6.3)
e electric field intensity (18.3)
Ef Fermi energy (18.5)
Eg band gap energy (18.6)
Er(t) relaxation modulus (15.4)
%EL ductility, in percent
elongation (6.6)
e electric charge per electron (18.7)
e electron (17.2)
erf Gaussian error function (5.4)
exp e, the base for natural logarithms
F force, interatomic or mechanical
(2.5, 6.3)
f Faraday constant (17.2)
FCC face-centered cubic crystal
structure (3.4)
G shear modulus (6.3)
H magnetic field strength (20.2)
Hc magnetic coercivity (20.7)
HB Brinell hardness (6.10)
HCP hexagonal close-packed crystal
structure (3.4)
HK Knoop hardness (6.10)
HRB, HRF Rockwell hardness: B and F
scales (6.10)
• xxiii
1496T_fm_i-xxvi 01/11/06 0:06 Page xxiv
xxiv • List of Symbols
HR15N, HR45W superficial Rockwell
hardness: 15N and 45W
scales (6.10)
HV Vickers hardness (6.10)
h Planck’s constant (21.2)
(hkl) Miller indices for a
crystallographic plane (3.10)
I electric current (18.2)
I intensity of electromagnetic
radiation (21.3)
i current density (17.3)
iC corrosion current density
(17.4)
J diffusion flux (5.3)
J electric current density (18.3)
Kc fracture toughness (8.5)
KIc plane strain fracture
toughness for mode I
crack surface displacement
(8.5)
k Boltzmann’s constant (4.2)
k thermal conductivity (19.4)
l length
lc critical fiber length (16.4)
ln natural logarithm
log logarithm taken to base 10
M magnetization (20.2)
Mn polymer number-average
molecular weight (14.5)
Mw polymer weight-average
molecular weight (14.5)
mol% mole percent
N number of fatigue cycles (8.8)
NA Avogadro’s number (3.5)
Nf fatigue life (8.8)
n principal quantum number
(2.3)
n number of atoms per unit
cell (3.5)
n strain-hardening exponent
(6.7)
n number of electrons in
an electrochemical
reaction (17.2)
n number of conducting
electrons per cubic
meter (18.7)
n index of refraction (21.5)
n for ceramics, the number
of formula units per unit
cell (12.2)
ni intrinsic carrier (electron and
hole) concentration (18.10)
P dielectric polarization (18.19)
P–B ratio Pilling–Bedworth ratio (17.10)
p number of holes per cubic
meter (18.10)
Q activation energy
Q magnitude of charge stored
(18.18)
R atomic radius (3.4)
R gas constant
%RA ductility, in percent reduction
in area (6.6)
r interatomic distance (2.5)
r reaction rate (17.3)
rA, rC anion and cation ionic radii
(12.2)
S fatigue stress amplitude (8.8)
SEM scanning electron
microscopy or microscope
T temperature
Tc Curie temperature (20.6)
TC superconducting critical temperature (20.12)
Tg glass transition temperature
(13.9, 15.12)
Tm melting temperature
TEM transmission electron
microscopy or microscope
TS tensile strength (6.6)
t time
tr rupture lifetime (8.12)
Ur modulus of resilience (6.6)
[uvw] indices for a crystallographic
direction (3.9)
V electrical potential difference
(voltage) (17.2, 18.2)
1496T_fm_i-xxvi 01/10/06 22:13 Page xxv
List of Symbols • xxv
VC unit cell volume (3.4)
VC corrosion potential (17.4)
VH Hall voltage (18.14)
Vi volume fraction of phase i (9.8)
v velocity
vol% volume percent
Wi mass fraction of phase i (9.8)
wt% weight percent (4.4)
x length
x space coordinate
Y dimensionless parameter or function in
fracture toughness expression (8.5)
y space coordinate
z space coordinate
lattice parameter: unit cell y–z
interaxial angle (3.7)
, , phase designations
l linear coefficient of thermal expansion
(19.3)
lattice parameter: unit cell x–z
interaxial angle (3.7)
lattice parameter: unit cell x–y
interaxial angle (3.7)
shear strain (6.2)
precedes the symbol of a parameter to
denote finite change
engineering strain (6.2)
dielectric permittivity (18.18)
r dielectric constant or relative
permittivity (18.18)
s steady-state creep rate (8.12)
T true strain (6.7)
viscosity (12.10)
overvoltage (17.4)
Bragg diffraction angle (3.16)
D Debye temperature (19.2)
wavelength of electromagnetic
radiation (3.16)
magnetic permeability (20.2)
B Bohr magneton (20.2)
r relative magnetic permeability (20.2)
e electron mobility (18.7)
h hole mobility (18.10)
n Poisson’s ratio (6.5)
n frequency of electromagnetic
radiation (21.2)
density (3.5)
electrical resistivity (18.2)
t radius of curvature at the tip of a
crack (8.5)
engineering stress, tensile or
compressive (6.2)
electrical conductivity (18.3)
* longitudinal strength
(composite) (16.5)
c critical stress for crack propagation
(8.5)
fs flexural strength (12.9)
m maximum stress (8.5)
m mean stress (8.7)
m stress in matrix at composite
failure (16.5)
T true stress (6.7)
w safe or working stress (6.12)
y yield strength (6.6)
shear stress (6.2)
c fiber–matrix bond strength/matrix shear
yield strength (16.4)
crss critical resolved shear stress (7.5)
m magnetic susceptibility (20.2)
SUBSCRIPTS
c composite
cd discontinuous fibrous composite
cl longitudinal direction (aligned fibrous
composite)
ct transverse direction (aligned fibrous
composite)
f final
f at fracture
f fiber
i instantaneous
m matrix
m, max maximum
min minimum
0 original
0 at equilibrium
0 in a vacuum
1496T_fm_i-xxvi 1/6/06 02:56 Page xxvi
1496T_c01_01-14 12/20/05 7:11 Page 1
2nd REVISE PAGES
Chapter
A
1
Introduction
familiar item that is fabricated from three different material types is the beverage
container. Beverages are marketed in aluminum (metal) cans (top), glass (ceramic)
bottles (center), and plastic (polymer) bottles (bottom). (Permission to use these
photographs was granted by the Coca-Cola Company. Coca-Cola, Coca-Cola Classic,
the Contour Bottle design and the Dynamic Ribbon are registered trademarks of The
Coca-Cola Company and used with its express permission.)
• 1
1496T_c01_01-14 12/20/05 7:11 Page 2
2nd REVISE PAGES
Learning Objectives
After careful study of this chapter you should be able to do the following:
1. List six different property classifications of
4. (a) List the three primary classifications of solid
materials that determine their applicability.
materials, and then cite the distinctive
2. Cite the four components that are involved in
chemical feature of each.
the design, production, and utilization of
(b) Note the two types of advanced materials
materials, and briefly describe the interrelationand, for each, its distinctive feature(s).
ships between these components.
5. (a) Briefly define “smart material/system.”
3. Cite three criteria that are important in the ma(b) Briefly explain the concept of “nanotechterials selection process.
nology” as it applies to materials.
1.1 HISTORICAL PERSPECTIVE
Materials are probably more deep-seated in our culture than most of us realize.
Transportation, housing, clothing, communication, recreation, and food production—
virtually every segment of our everyday lives is influenced to one degree or another
by materials. Historically, the development and advancement of societies have been
intimately tied to the members’ ability to produce and manipulate materials to fill
their needs. In fact, early civilizations have been designated by the level of their
materials development (Stone Age, Bronze Age, Iron Age).1
The earliest humans had access to only a very limited number of materials,
those that occur naturally: stone, wood, clay, skins, and so on. With time they discovered techniques for producing materials that had properties superior to those
of the natural ones; these new materials included pottery and various metals. Furthermore, it was discovered that the properties of a material could be altered by
heat treatments and by the addition of other substances. At this point, materials utilization was totally a selection process that involved deciding from a given, rather
limited set of materials the one best suited for an application by virtue of its characteristics. It was not until relatively recent times that scientists came to understand
the relationships between the structural elements of materials and their properties.
This knowledge, acquired over approximately the past 100 years, has empowered
them to fashion, to a large degree, the characteristics of materials. Thus, tens of thousands of different materials have evolved with rather specialized characteristics that
meet the needs of our modern and complex society; these include metals, plastics,
glasses, and fibers.
The development of many technologies that make our existence so comfortable has been intimately associated with the accessibility of suitable materials.
An advancement in the understanding of a material type is often the forerunner to the stepwise progression of a technology. For example, automobiles
would not have been possible without the availability of inexpensive steel or
some other comparable substitute. In our contemporary era, sophisticated electronic devices rely on components that are made from what are called semiconducting materials.
1
The approximate dates for the beginnings of Stone, Bronze, and Iron Ages were 2.5 million
3500 BC and 1000 BC, respectively.
BC,
1496T_c01_01-14 11/9/05 17:02 Page 3
REVISED PAGES
1.2 Materials Science and Engineering • 3
1.2 MATERIALS SCIENCE AND ENGINEERING
Sometimes it is useful to subdivide the discipline of materials science and engineering into materials science and materials engineering subdisciplines. Strictly
speaking, “materials science” involves investigating the relationships that exist
between the structures and properties of materials. In contrast, “materials engineering” is, on the basis of these structure–property correlations, designing or engineering the structure of a material to produce a predetermined set of properties.2
From a functional perspective, the role of a materials scientist is to develop or synthesize new materials, whereas a materials engineer is called upon to create new
products or systems using existing materials, and/or to develop techniques for processing materials. Most graduates in materials programs are trained to be both
materials scientists and materials engineers.
“Structure” is at this point a nebulous term that deserves some explanation. In
brief, the structure of a material usually relates to the arrangement of its internal
components. Subatomic structure involves electrons within the individual atoms and
interactions with their nuclei. On an atomic level, structure encompasses the organization of atoms or molecules relative to one another. The next larger structural
realm, which contains large groups of atoms that are normally agglomerated together, is termed “microscopic,” meaning that which is subject to direct observation
using some type of microscope. Finally, structural elements that may be viewed with
the naked eye are termed “macroscopic.”
The notion of “property” deserves elaboration. While in service use, all materials are exposed to external stimuli that evoke some type of response. For example, a specimen subjected to forces will experience deformation, or a polished metal
surface will reflect light. A property is a material trait in terms of the kind and magnitude of response to a specific imposed stimulus. Generally, definitions of properties are made independent of material shape and size.
Virtually all important properties of solid materials may be grouped into six different categories: mechanical, electrical, thermal, magnetic, optical, and deteriorative.
For each there is a characteristic type of stimulus capable of provoking different responses. Mechanical properties relate deformation to an applied load or force; examples include elastic modulus and strength. For electrical properties, such as electrical
conductivity and dielectric constant, the stimulus is an electric field. The thermal behavior of solids can be represented in terms of heat capacity and thermal conductivity. Magnetic properties demonstrate the response of a material to the application of
a magnetic field. For optical properties, the stimulus is electromagnetic or light radiation; index of refraction and reflectivity are representative optical properties. Finally,
deteriorative characteristics relate to the chemical reactivity of materials. The chapters
that follow discuss properties that fall within each of these six classifications.
In addition to structure and properties, two other important components are
involved in the science and engineering of materials—namely, “processing” and
“performance.” With regard to the relationships of these four components, the structure of a material will depend on how it is processed. Furthermore, a material’s performance will be a function of its properties. Thus, the interrelationship between
processing, structure, properties, and performance is as depicted in the schematic
illustration shown in Figure 1.1. Throughout this text we draw attention to the
2
Throughout this text we draw attention to the relationships between material properties
and structural elements.
1496T_c01_01-14 12/20/05 7:11 Page 4
2nd REVISE PAGES
4 • Chapter 1 / Introduction
Processing
Structure
Properties
Performance
Figure 1.1 The four components of the discipline of materials science and
engineering and their interrelationship.
relationships among these four components in terms of the design, production, and
utilization of materials.
We now present an example of these processing-structure-properties-performance
principles with Figure 1.2, a photograph showing three thin disk specimens placed
over some printed matter. It is obvious that the optical properties (i.e., the light
transmittance) of each of the three materials are different; the one on the left is transparent (i.e., virtually all of the reflected light passes through it), whereas the disks in
the center and on the right are, respectively, translucent and opaque. All of these specimens are of the same material, aluminum oxide, but the leftmost one is what we call
a single crystal—that is, it is highly perfect—which gives rise to its transparency. The
center one is composed of numerous and very small single crystals that are all connected; the boundaries between these small crystals scatter a portion of the light reflected from the printed page, which makes this material optically translucent. Finally,
the specimen on the right is composed not only of many small, interconnected crystals, but also of a large number of very small pores or void spaces. These pores also
effectively scatter the reflected light and render this material opaque.
Thus, the structures of these three specimens are different in terms of crystal
boundaries and pores, which affect the optical transmittance properties. Furthermore, each material was produced using a different processing technique. And, of
course, if optical transmittance is an important parameter relative to the ultimate
in-service application, the performance of each material will be different.
Figure 1.2 Photograph of three thin disk specimens of aluminum oxide, which have been
placed over a printed page in order to demonstrate their differences in light-transmittance
characteristics. The disk on the left is transparent (that is, virtually all light that is reflected
from the page passes through it), whereas the one in the center is translucent (meaning that
some of this reflected light is transmitted through the disk). And, the disk on the right is
opaque—i.e., none of the light passes through it. These differences in optical properties are a
consequence of differences in structure of these materials, which have resulted from the way
the materials were processed. (Specimen preparation, P. A. Lessing; photography by S. Tanner.)
1496T_c01_01-14 11/9/05 17:02 Page 5
REVISED PAGES
1.4 Classification of Materials • 5
1.3 WHY STUDY MATERIALS SCIENCE
AND ENGINEERING?
Why do we study materials? Many an applied scientist or engineer, whether mechanical, civil, chemical, or electrical, will at one time or another be exposed to a
design problem involving materials. Examples might include a transmission gear,
the superstructure for a building, an oil refinery component, or an integrated circuit
chip. Of course, materials scientists and engineers are specialists who are totally
involved in the investigation and design of materials.
Many times, a materials problem is one of selecting the right material from the
many thousands that are available. There are several criteria on which the final
decision is normally based. First of all, the in-service conditions must be characterized, for these will dictate the properties required of the material. On only rare
occasions does a material possess the maximum or ideal combination of properties.
Thus, it may be necessary to trade off one characteristic for another. The classic example involves strength and ductility; normally, a material having a high strength
will have only a limited ductility. In such cases a reasonable compromise between
two or more properties may be necessary.
A second selection consideration is any deterioration of material properties that
may occur during service operation. For example, significant reductions in mechanical
strength may result from exposure to elevated temperatures or corrosive environments.
Finally, probably the overriding consideration is that of economics: What will
the finished product cost? A material may be found that has the ideal set of properties but is prohibitively expensive. Here again, some compromise is inevitable.
The cost of a finished piece also includes any expense incurred during fabrication
to produce the desired shape.
The more familiar an engineer or scientist is with the various characteristics
and structure–property relationships, as well as processing techniques of materials,
the more proficient and confident he or she will be to make judicious materials
choices based on these criteria.
1.4 CLASSIFICATION OF MATERIALS
Solid materials have been conveniently grouped into three basic classifications: metals, ceramics, and polymers. This scheme is based primarily on chemical makeup and
atomic structure, and most materials fall into one distinct grouping or another,
although there are some intermediates. In addition, there are the composites, combinations of two or more of the above three basic material classes. A brief explanation of these material types and representative characteristics is offered next.Another
classification is advanced materials—those used in high-technology applications—
viz. semiconductors, biomaterials, smart materials, and nanoengineered materials;
these are discussed in Section 1.5.
Metals
Materials in this group are composed of one or more metallic elements (such as iron,
aluminum, copper, titanium, gold, and nickel), and often also nonmetallic elements (for
example, carbon, nitrogen, and oxygen) in relatively small amounts.3 Atoms in metals
and their alloys are arranged in a very orderly manner (as discussed in Chapter 3),
and in comparison to the ceramics and polymers, are relatively dense (Figure 1.3).With
3
The term metal alloy is used in reference to a metallic substance that is composed of two
or more elements.
1496T_c01_01-14 11/9/05 17:02 Page 6
REVISED PAGES
Figure 1.3
Bar-chart of roomtemperature density
values for various
metals, ceramics,
polymers, and
composite materials.
Density (g/cm3) (logarithmic scale)
6 • Chapter 1 / Introduction
40
Metals
20
Platinum
Silver
10
8
6
Copper
Iron/Steel
Titanium
4
Aluminum
2
Magnesium
Ceramics
ZrO2
Al2O3
SiC,Si3N4
Glass
Concrete
Polymers
PTFE
Composites
GFRC
CFRC
PVC
PS
PE
Rubber
1.0
0.8
0.6
Woods
0.4
0.2
0.1
regard to mechanical characteristics, these materials are relatively stiff (Figure 1.4)
and strong (Figure 1.5), yet are ductile (i.e., capable of large amounts of deformation
without fracture), and are resistant to fracture (Figure 1.6), which accounts for their
widespread use in structural applications. Metallic materials have large numbers of
nonlocalized electrons; that is, these electrons are not bound to particular atoms. Many
properties of metals are directly attributable to these electrons. For example, metals
are extremely good conductors of electricity (Figure 1.7) and heat, and are not transparent to visible light; a polished metal surface has a lustrous appearance. In addition, some of the metals (viz., Fe, Co, and Ni) have desirable magnetic properties.
Figure 1.8 is a photograph that shows several common and familiar objects that
are made of metallic materials. Furthermore, the types and applications of metals
and their alloys are discussed in Chapter 11.
Ceramics
Figure 1.4
Bar-chart of roomtemperature stiffness
(i.e., elastic modulus)
values for various
metals, ceramics,
polymers, and
composite materials.
Stiffness [Elastic (or Young’s) Modulus (in units of
gigapascals)] (logarithmic scale)
Ceramics are compounds between metallic and nonmetallic elements; they are most
frequently oxides, nitrides, and carbides. For example, some of the common ceramic
1000
100
10
1.0
Metals
Tungsten
Iron/Steel
Titanium
Aluminum
Magnesium
Ceramics
Composites
SiC
AI2O3
Si3N4
ZrO2
Glass
Concrete
CFRC
GFRC
Polymers
PVC
PS, Nylon
PTFE
PE
0.1
Rubbers
0.01
0.001
Woods
1496T_c01_01-14 12/20/05 7:11 Page 7
2nd REVISE PAGES
1.4 Classification of Materials • 7
Metals
Strength (Tensile Strength, in units of
megapascals) (logarithmic scale)
Composites
Ceramics
1000
Steel
alloys
Cu,Ti
alloys
100
Aluminum
alloys
Gold
CFRC
Si3N4
Al2O3
GFRC
SiC
Polymers
Glass
Nylon
PVC
PS
PE
Woods
PTFE
10
materials include aluminum oxide (or alumina,Al2O3), silicon dioxide (or silica, SiO2),
silicon carbide (SiC), silicon nitride (Si3N4), and, in addition, what some refer to as
the traditional ceramics—those composed of clay minerals (i.e., porcelain), as well as
cement, and glass. With regard to mechanical behavior, ceramic materials are relatively stiff and strong—stiffnesses and strengths are comparable to those of the metals (Figures 1.4 and 1.5). In addition, ceramics are typically very hard. On the other
hand, they are extremely brittle (lack ductility), and are highly susceptible to fracture
(Figure 1.6). These materials are typically insulative to the passage of heat and electricity (i.e., have low electrical conductivities, Figure 1.7), and are more resistant to
high temperatures and harsh environments than metals and polymers. With regard to
optical characteristics, ceramics may be transparent, translucent, or opaque (Figure
1.2), and some of the oxide ceramics (e.g., Fe3O4) exhibit magnetic behavior.
Metals
Resistance to Fracture (Fracture Toughness,
in units of MPa m) (logarithmic scale)
Figure 1.5
Bar-chart of roomtemperature strength
(i.e., tensile strength)
values for various
metals, ceramics,
polymers, and
composite materials.
100
Steel
alloys
Composites
Titanium
alloys
Aluminum
alloys
10
CFRC
GFRC
Ceramics
Polymers
Si3N4
Al2O3
SiC
1.0
Nylon
Polystyrene
Polyethylene
Wood
Glass
Polyester
Concrete
0.1
Figure 1.6 Bar-chart of room-temperature resistance to fracture (i.e., fracture toughness)
for various metals, ceramics, polymers, and composite materials. (Reprinted from Engineering
Materials 1: An Introduction to Properties, Applications and Design, third edition, M. F. Ashby
and D. R. H. Jones, pages 177 and 178, Copyright 2005, with permission from Elsevier.)
1496T_c01_01-14 12/20/05 7:11 Page 8
2nd REVISE PAGES
8 • Chapter 1 / Introduction
Metals
108
Semiconductors
Electrical Conductivity (in units of reciprocal
ohm-meters) (logarithmic scale)
Figure 1.7
Bar-chart of roomtemperature
electrical
conductivity ranges
for metals, ceramics,
polymers, and
semiconducting
materials.
104
1
10–4
10–8
Ceramics
Polymers
10–12
10–16
10–20
Several common ceramic objects are shown in the photograph of Figure 1.9.
The characteristics, types, and applications of this class of materials are discussed
in Chapters 12 and 13.
Polymers
Polymers include the familiar plastic and rubber materials. Many of them are organic
compounds that are chemically based on carbon, hydrogen, and other nonmetallic
elements (viz. O, N, and Si). Furthermore, they have very large molecular structures,
often chain-like in nature that have a backbone of carbon atoms. Some of the common and familiar polymers are polyethylene (PE), nylon, poly(vinyl chloride)
(PVC), polycarbonate (PC), polystyrene (PS), and silicone rubber. These materials
typically have low densities (Figure 1.3), whereas their mechanical characteristics
are generally dissimilar to the metallic and ceramic materials—they are not as stiff
nor as strong as these other material types (Figures 1.4 and 1.5). However, on the
basis of their low densities, many times their stiffnesses and strengths on a per mass
Figure 1.8 Familiar
objects that are
made of metals and
metal alloys: (from
left to right)
silverware (fork and
knife), scissors, coins,
a gear, a wedding
ring, and a nut and
bolt. (Photograpy by
S. Tanner.)
1496T_c01_01-14 12/20/05 7:11 Page 9
2nd REVISE PAGES
1.4 Classification of Materials • 9
Figure 1.9
Common objects
that are made of
ceramic materials:
scissors, a china tea
cup, a building brick,
a floor tile, and a
glass vase.
(Photography by
S. Tanner.)
basis are comparable to the metals and ceramics. In addition, many of the polymers
are extremely ductile and pliable (i.e., plastic), which means they are easily formed
into complex shapes. In general, they are relatively inert chemically and unreactive
in a large number of environments. One major drawback to the polymers is their
tendency to soften and/or decompose at modest temperatures, which, in some instances, limits their use. Furthermore, they have low electrical conductivities (Figure 1.7) and are nonmagnetic.
The photograph in Figure 1.10 shows several articles made of polymers that are
familiar to the reader. Chapters 14 and 15 are devoted to discussions of the structures, properties, applications, and processing of polymeric materials.
Figure 1.10 Several
common objects
that are made of
polymeric materials:
plastic tableware
(spoon, fork, and
knife), billiard balls,
a bicycle helmet, two
dice, a lawnmower
wheel (plastic hub
and rubber tire), and
a plastic milk carton.
(Photography by
S. Tanner.)
1496T_c01_01-14 11/9/05 17:02 Page 10
REVISED PAGES
10 • Chapter 1 / Introduction
MATERIALS OF IMPORTANCE
Carbonated Beverage Containers
O
ne common item that presents some interesting material property requirements is the
container for carbonated beverages. The material
used for this application must satisfy the following constraints: (1) provide a barrier to the passage of carbon dioxide, which is under pressure in
the container; (2) be nontoxic, unreactive with the
beverage, and, preferably be recyclable; (3) be relatively strong, and capable of surviving a drop
from a height of several feet when containing the
beverage; (4) be inexpensive and the cost to fabricate the final shape should be relatively low;
(5) if optically transparent, retain its optical clarity; and (6) capable of being produced having
different colors and/or able to be adorned with
decorative labels.
All three of the basic material types—metal
(aluminum), ceramic (glass), and polymer (polyester plastic)—are used for carbonated beverage
containers (per the chapter-opening photographs
for this chapter).All of these materials are nontoxic
and unreactive with beverages. In addition, each
material has its pros and cons. For example, the
aluminum alloy is relatively strong (but easily
dented), is a very good barrier to the diffusion of
carbon dioxide, is easily recycled, beverages are
cooled rapidly, and labels may be painted onto its
surface. On the other hand, the cans are optically
opaque, and relatively expensive to produce. Glass
is impervious to the passage of carbon dioxide, is
a relatively inexpensive material, may be recycled,
but it cracks and fractures easily, and glass bottles
are relatively heavy. Whereas the plastic is relatively strong, may be made optically transparent,
is inexpensive and lightweight, and is recyclable, it
is not as impervious to the passage of carbon dioxide as the aluminum and glass. For example, you
may have noticed that beverages in aluminum and
glass containers retain their carbonization (i.e.,
“fizz”) for several years, whereas those in two-liter
plastic bottles “go flat” within a few months.
Composites
A composite is composed of two (or more) individual materials, which come from
the categories discussed above—viz., metals, ceramics, and polymers.The design goal
of a composite is to achieve a combination of properties that is not displayed by
any single material, and also to incorporate the best characteristics of each of the
component materials. A large number of composite types exist that are represented
by different combinations of metals, ceramics, and polymers. Furthermore, some
naturally-occurring materials are also considered to be composites—for example,
wood and bone. However, most of those we consider in our discussions are synthetic (or man-made) composites.
One of the most common and familiar composites is fiberglass, in which small
glass fibers are embedded within a polymeric material (normally an epoxy or
polyester).4 The glass fibers are relatively strong and stiff (but also brittle), whereas
the polymer is ductile (but also weak and flexible). Thus, the resulting fiberglass is
relatively stiff, strong, (Figures 1.4 and 1.5) flexible, and ductile. In addition, it has
a low density (Figure 1.3).
Another of these technologically important materials is the “carbon fiberreinforced polymer” (or “CFRP”) composite—carbon fibers that are embedded within
a polymer. These materials are stiffer and stronger than the glass fiber-reinforced
materials (Figures 1.4 and 1.5), yet they are more expensive. The CFRP composites
4
Fiberglass is sometimes also termed a “glass fiber-reinforced polymer” composite, abbreviated “GFRP.”
1496T_c01_01-14 11/9/05 17:02 Page 11
REVISED PAGES
1.5 Advanced Materials • 11
are used in some aircraft and aerospace applications, as well as high-tech sporting
equipment (e.g., bicycles, golf clubs, tennis rackets, and skis/snowboards). Chapter 16
is devoted to a discussion of these interesting materials.
1.5 ADVANCED MATERIALS
Materials that are utilized in high-technology (or high-tech) applications are sometimes termed advanced materials. By high technology we mean a device or product
that operates or functions using relatively intricate and sophisticated principles; examples include electronic equipment (camcorders, CD/DVD players, etc.), computers, fiber-optic systems, spacecraft, aircraft, and military rocketry.These advanced
materials are typically traditional materials whose properties have been enhanced,
and, also newly developed, high-performance materials. Furthermore, they may be
of all material types (e.g., metals, ceramics, polymers), and are normally expensive.
Advanced materials include semiconductors, biomaterials, and what we may term
“materials of the future” (that is, smart materials and nanoengineered materials),
which we discuss below. The properties and applications of a number of these
advanced materials—for example, materials that are used for lasers, integrated
circuits, magnetic information storage, liquid crystal displays (LCDs), and fiber
optics—are also discussed in subsequent chapters.
Semiconductors
Semiconductors have electrical properties that are intermediate between the electrical conductors (viz. metals and metal alloys) and insulators (viz. ceramics and
polymers)—Figure 1.7. Furthermore, the electrical characteristics of these materials are extremely sensitive to the presence of minute concentrations of impurity
atoms, for which the concentrations may be controlled over very small spatial regions. Semiconductors have made possible the advent of integrated circuitry that
has totally revolutionized the electronics and computer industries (not to mention
our lives) over the past three decades.
Biomaterials
Biomaterials are employed in components implanted into the human body for
replacement of diseased or damaged body parts. These materials must not produce
toxic substances and must be compatible with body tissues (i.e., must not cause
adverse biological reactions). All of the above materials—metals, ceramics, polymers, composites, and semiconductors—may be used as biomaterials. For example,
some of the biomaterials that are utilized in artificial hip replacements are discussed in Section 22.12.
Materials of the Future
Smart Materials
Smart (or intelligent) materials are a group of new and state-of-the-art materials
now being developed that will have a significant influence on many of our technologies.The adjective “smart” implies that these materials are able to sense changes
in their environments and then respond to these changes in predetermined manners—
traits that are also found in living organisms. In addition, this “smart” concept is being extended to rather sophisticated systems that consist of both smart and traditional materials.
Components of a smart material (or system) include some type of sensor (that
detects an input signal), and an actuator (that performs a responsive and adaptive
1496T_c01_01-14 12/20/05 7:11 Page 12
2nd REVISE PAGES
12 • Chapter 1 / Introduction
function). Actuators may be called upon to change shape, position, natural
frequency, or mechanical characteristics in response to changes in temperature,
electric fields, and/or magnetic fields.
Four types of materials are commonly used for actuators: shape memory alloys,
piezoelectric ceramics, magnetostrictive materials, and electrorheological/magnetorheological fluids. Shape memory alloys are metals that, after having been deformed,
revert back to their original shapes when temperature is changed (see the Materials of Importance piece following Section 10.9). Piezoelectric ceramics expand and
contract in response to an applied electric field (or voltage); conversely, they also
generate an electric field when their dimensions are altered (see Section 18.25). The
behavior of magnetostrictive materials is analogous to that of the piezoelectrics, except that they are responsive to magnetic fields. Also, electrorheological and magnetorheological fluids are liquids that experience dramatic changes in viscosity upon
the application of electric and magnetic fields, respectively.
Materials/devices employed as sensors include optical fibers (Section 21.14),
piezoelectric materials (including some polymers), and microelectromechanical
devices (MEMS, Section 13.8).
For example, one type of smart system is used in helicopters to reduce aerodynamic cockpit noise that is created by the rotating rotor blades. Piezoelectric
sensors inserted into the blades monitor blade stresses and deformations; feedback
signals from these sensors are fed into a computer-controlled adaptive device, which
generates noise-canceling antinoise.
Nanoengineered Materials
Until very recent times the general procedure utilized by scientists to understand
the chemistry and physics of materials has been to begin by studying large and complex structures, and then to investigate the fundamental building blocks of these
structures that are smaller and simpler. This approach is sometimes termed “topdown” science. However, with the advent of scanning probe microscopes (Section 4.10), which permit observation of individual atoms and molecules, it has become possible to manipulate and move atoms and molecules to form new structures
and, thus, design new materials that are built from simple atomic-level constituents
(i.e., “materials by design”). This ability to carefully arrange atoms provides opportunities to develop mechanical, electrical, magnetic, and other properties that
are not otherwise possible. We call this the “bottom-up” approach, and the study of
the properties of these materials is termed “nanotechnology”; the “nano” prefix denotes that the dimensions of these structural entities are on the order of a nanometer (109 m)—as a rule, less than 100 nanometers (equivalent to approximately 500
atom diameters).5 One example of a material of this type is the carbon nanotube,
discussed in Section 12.4. In the future we will undoubtedly find that increasingly
more of our technological advances will utilize these nanoengineered materials.
1.6 MODERN MATERIALS’ NEEDS
In spite of the tremendous progress that has been made in the discipline of materials
science and engineering within the past few years, there still remain technological
challenges, including the development of even more sophisticated and specialized
5
One legendary and prophetic suggestion as to the possibility of nanoengineering materials
was offered by Richard Feynman in his 1960 American Physical Society lecture that was
entitled “There is Plenty of Room at the Bottom.”
1496T_c01_01-14 11/9/05 17:02 Page 13
REVISED PAGES
References • 13
materials, as well as consideration of the environmental impact of materials production. Some comment is appropriate relative to these issues so as to round out
this perspective.
Nuclear energy holds some promise, but the solutions to the many problems
that remain will necessarily involve materials, from fuels to containment structures
to facilities for the disposal of radioactive waste.
Significant quantities of energy are involved in transportation. Reducing the
weight of transportation vehicles (automobiles, aircraft, trains, etc.), as well as
increasing engine operating temperatures, will enhance fuel efficiency. New highstrength, low-density structural materials remain to be developed, as well as materials that have higher-temperature capabilities, for use in engine components.
Furthermore, there is a recognized need to find new, economical sources of energy and to use present resources more efficiently. Materials will undoubtedly play
a significant role in these developments. For example, the direct conversion of solar into electrical energy has been demonstrated. Solar cells employ some rather
complex and expensive materials. To ensure a viable technology, materials that are
highly efficient in this conversion process yet less costly must be developed.
The hydrogen fuel cell is another very attractive and feasible energy-conversion
technology that has the advantage of being non-polluting. It is just beginning to be
implemented in batteries for electronic devices, and holds promise as the power
plant for automobiles. New materials still need to be developed for more efficient
fuel cells, and also for better catalysts to be used in the production of hydrogen.
Furthermore, environmental quality depends on our ability to control air and
water pollution. Pollution control techniques employ various materials. In addition,
materials processing and refinement methods need to be improved so that they produce less environmental degradation—that is, less pollution and less despoilage of
the landscape from the mining of raw materials. Also, in some materials manufacturing processes, toxic substances are produced, and the ecological impact of their
disposal must be considered.
Many materials that we use are derived from resources that are nonrenewable—
that is, not capable of being regenerated. These include polymers, for which the
prime raw material is oil, and some metals. These nonrenewable resources are gradually becoming depleted, which necessitates: (1) the discovery of additional reserves,
(2) the development of new materials having comparable properties with less adverse environmental impact, and/or (3) increased recycling efforts and the development of new recycling technologies. As a consequence of the economics of not
only production but also environmental impact and ecological factors, it is becoming
increasingly important to consider the “cradle-to-grave” life cycle of materials relative to the overall manufacturing process.
The roles that materials scientists and engineers play relative to these, as well as
other environmental and societal issues, are discussed in more detail in Chapter 23.
REFERENCES
Ashby, M. F. and D. R. H. Jones, Engineering Materials 1, An Introduction to Their Properties
and Applications, 3rd edition, ButterworthHeinemann, Woburn, UK, 2005.
Ashby, M. F. and D. R. H. Jones, Engineering Materials 2, An Introduction to Microstructures, Pro-
cessing and Design, 3rd edition, ButterworthHeinemann, Woburn, UK, 2005.
Askeland, D. R. and P. P. Phulé, The Science and
Engineering of Materials, 5th edition, Nelson
(a division of Thomson Canada), Toronto,
2006.
1496T_c01_01-14 11/9/05 17:02 Page 14
REVISED PAGES
14 • Chapter 1 / Introduction
Baillie, C. and L. Vanasupa, Navigating the Materials
World, Academic Press, San Diego, CA, 2003.
Flinn, R. A. and P. K. Trojan, Engineering Materials and Their Applications, 4th edition, John
Wiley & Sons, New York, 1994.
Jacobs, J. A. and T. F. Kilduff, Engineering Materials Technology, 5th edition, Prentice Hall PTR,
Paramus, NJ, 2005.
Mangonon, P. L., The Principles of Materials Selection for Engineering Design, Prentice Hall
PTR, Paramus, NJ, 1999.
McMahon, C. J., Jr., Structural Materials, Merion
Books, Philadelphia, 2004.
Murray, G. T., Introduction to Engineering Materials—Behavior, Properties, and Selection, Marcel
Dekker, Inc., New York, 1993.
Ralls, K. M., T. H. Courtney, and J. Wulff, Introduction to Materials Science and Engineering,
John Wiley & Sons, New York, 1976.
Schaffer, J. P., A. Saxena, S. D. Antolovich, T. H.
Sanders, Jr., and S. B. Warner, The Science and
Design of Engineering Materials, 2nd edition,
WCB/McGraw-Hill, New York, 1999.
Shackelford, J. F., Introduction to Materials Science
for Engineers, 6th edition, Prentice Hall PTR,
Paramus, NJ, 2005.
Smith, W. F. and J. Hashemi, Principles of Materials Science and Engineering, 4th edition,
McGraw-Hill Book Company, New York, 2006.
Van Vlack, L. H., Elements of Materials Science
and Engineering, 6th edition, Addison-Wesley
Longman, Boston, MA, 1989.
White, M. A., Properties of Materials, Oxford
University Press, New York, 1999.
1496T_c02_15-37 12/20/05 7:19 Page 15
Chapter
T
2nd REVISE PAGES
2
Atomic Structure and
Interatomic Bonding
his photograph shows the underside of a gecko.
Geckos, harmless tropical lizards, are extremely fascinating and extraordinary animals. They
have very sticky feet that cling to virtually any surface. This characteristic makes it possible for
them to rapidly run up vertical walls and along the undersides of horizontal surfaces. In fact, a
gecko can support its body mass with a single toe! The secret to this remarkable ability is the presence of an extremely large number of microscopically small hairs on each of their toe pads. When
these hairs come in contact with a surface, weak forces of attraction (i.e., van der Waals forces)
are established between hair molecules and molecules on the surface. The fact that these hairs are
so small and so numerous explains why the gecko grips surfaces so tightly. To release its grip, the
gecko simply curls up its toes, and peels the hairs away from the surface.
Another interesting feature of these toe pads is that they are self-cleaning—that is, dirt particles don’t stick to them. Scientists are just beginning to understand the mechanism of adhesion for
these tiny hairs, which may lead to the development of synthetic self-cleaning adhesives. Can you
image duct tape that never looses its stickiness, or bandages that never leave a sticky residue?
(Photograph courtesy of Professor Kellar Autumn, Lewis & Clark College, Portland, Oregon.)
WHY STUDY Atomic Structure and Interatomic Bonding?
An important reason to have an understanding of interatomic bonding in solids is that, in some instances,
the type of bond allows us to explain a material’s
properties. For example, consider carbon, which may
exist as both graphite and diamond. Whereas graphite
is relatively soft and has a “greasy” feel to it, diamond
is the hardest known material. This dramatic disparity
in properties is directly attributable to a type of interatomic bonding found in graphite that does not exist
in diamond (see Section 12.4).
• 15
1496T_c02_15-37 12/20/05 7:19 Page 16
2nd REVISE PAGES
Learning Objectives
After careful study of this chapter you should be able to do the following:
1. Name the two atomic models cited, and note
(b) Note on this plot the equilibrium separation
the differences between them.
and the bonding energy.
2. Describe the important quantum-mechanical
4. (a) Briefly describe ionic, covalent, metallic,
principle that relates to electron energies.
hydrogen, and van der Waals bonds.
3. (a) Schematically plot attractive, repulsive, and
(b) Note which materials exhibit each of these
net energies versus interatomic separation
bonding types.
for two atoms or ions.
2.1 INTRODUCTION
Some of the important properties of solid materials depend on geometrical atomic
arrangements, and also the interactions that exist among constituent atoms or molecules. This chapter, by way of preparation for subsequent discussions, considers
several fundamental and important concepts—namely, atomic structure, electron
configurations in atoms and the periodic table, and the various types of primary
and secondary interatomic bonds that hold together the atoms comprising a solid.
These topics are reviewed briefly, under the assumption that some of the material
is familiar to the reader.
A t o m i c St r u c t u r e
2.2 FUNDAMENTAL CONCEPTS
atomic number
isotope
atomic weight
atomic mass unit
Each atom consists of a very small nucleus composed of protons and neutrons, which
is encircled by moving electrons. Both electrons and protons are electrically charged,
the charge magnitude being 1.60 1019 C, which is negative in sign for electrons
and positive for protons; neutrons are electrically neutral. Masses for these subatomic particles are infinitesimally small; protons and neutrons have approximately
the same mass, 1.67 1027 kg, which is significantly larger than that of an electron, 9.11 1031 kg.
Each chemical element is characterized by the number of protons in the nucleus, or the atomic number (Z).1 For an electrically neutral or complete atom, the
atomic number also equals the number of electrons. This atomic number ranges in
integral units from 1 for hydrogen to 92 for uranium, the highest of the naturally
occurring elements.
The atomic mass (A) of a specific atom may be expressed as the sum of the
masses of protons and neutrons within the nucleus. Although the number of protons
is the same for all atoms of a given element, the number of neutrons (N) may be
variable. Thus atoms of some elements have two or more different atomic masses,
which are called isotopes. The atomic weight of an element corresponds to the
weighted average of the atomic masses of the atom’s naturally occurring isotopes.2
The atomic mass unit (amu) may be used for computations of atomic weight. A
1
Terms appearing in boldface type are defined in the Glossary, which follows Appendix E.
The term “atomic mass” is really more accurate than “atomic weight” inasmuch as, in this
context, we are dealing with masses and not weights. However, atomic weight is, by convention, the preferred terminology and will be used throughout this book. The reader should
note that it is not necessary to divide molecular weight by the gravitational constant.
2
1496T_c02_15-37 11/10/05 10:42 Page 17
REVISED PAGES
2.3 Electrons in Atoms • 17
scale has been established whereby 1 amu is defined as 121 of the atomic mass of the
most common isotope of carbon, carbon 12 1 12C2 1A 12.000002. Within this
scheme, the masses of protons and neutrons are slightly greater than unity, and
AZN
mole
(2.1)
The atomic weight of an element or the molecular weight of a compound may be
specified on the basis of amu per atom (molecule) or mass per mole of material.
In one mole of a substance there are 6.023 1023 (Avogadro’s number) atoms
or molecules. These two atomic weight schemes are related through the following
equation:
1 amu/atom 1or molecule2 1 g/mol
For example, the atomic weight of iron is 55.85 amu/atom, or 55.85 g/mol. Sometimes
use of amu per atom or molecule is convenient; on other occasions g (or kg)/mol
is preferred. The latter is used in this book.
Concept Check 2.1
Why are the atomic weights of the elements generally not integers? Cite two
reasons.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
2.3 ELECTRONS IN ATOMS
Atomic Models
quantum mechanics
Bohr atomic model
During the latter part of the nineteenth century it was realized that many phenomena involving electrons in solids could not be explained in terms of classical
mechanics. What followed was the establishment of a set of principles and laws that
govern systems of atomic and subatomic entities that came to be known as quantum
mechanics. An understanding of the behavior of electrons in atoms and crystalline
solids necessarily involves the discussion of quantum-mechanical concepts. However, a detailed exploration of these principles is beyond the scope of this book,
and only a very superficial and simplified treatment is given.
One early outgrowth of quantum mechanics was the simplified Bohr atomic
model, in which electrons are assumed to revolve around the atomic nucleus
in discrete orbitals, and the position of any particular electron is more or
less well defined in terms of its orbital. This model of the atom is represented in
Figure 2.1.
Another important quantum-mechanical principle stipulates that the energies
of electrons are quantized; that is, electrons are permitted to have only specific values of energy. An electron may change energy, but in doing so it must make a quantum jump either to an allowed higher energy (with absorption of energy) or to a
lower energy (with emission of energy). Often, it is convenient to think of these allowed electron energies as being associated with energy levels or states. These states
do not vary continuously with energy; that is, adjacent states are separated by finite
energies. For example, allowed states for the Bohr hydrogen atom are represented
in Figure 2.2a. These energies are taken to be negative, whereas the zero reference
is the unbound or free electron. Of course, the single electron associated with the
hydrogen atom will fill only one of these states.
1496T_c02_15-37 12/20/05 13:51 Page 18
2nd REVISE PAGES
18 • Chapter 2 / Atomic Structure and Interatomic Bonding
Figure 2.1 Schematic representation of the Bohr
atom.
Orbital electron
Nucleus
0
0
–1.5
n=3
–3.4
n=2
3d
3p
3s
2p
2s
–1 × 10–18
–10
–2 × 10–18
n=1
–13.6
–15
(a)
1s
(b)
Energy (J)
–5
Energy (eV)
wave-mechanical
model
Thus, the Bohr model represents an early attempt to describe electrons in
atoms, in terms of both position (electron orbitals) and energy (quantized energy
levels).
This Bohr model was eventually found to have some significant limitations
because of its inability to explain several phenomena involving electrons. A
resolution was reached with a wave-mechanical model, in which the electron is
considered to exhibit both wave-like and particle-like characteristics. With this
model, an electron is no longer treated as a particle moving in a discrete orbital; rather, position is considered to be the probability of an electron’s being at
various locations around the nucleus. In other words, position is described by a
probability distribution or electron cloud. Figure 2.3 compares Bohr and wavemechanical models for the hydrogen atom. Both these models are used throughout the course of this book; the choice depends on which model allows the more
simple explanation.
Figure 2.2 (a) The
first three electron
energy states for the
Bohr hydrogen atom.
(b) Electron energy
states for the first
three shells of the
wave-mechanical
hydrogen atom.
(Adapted from W. G.
Moffatt, G. W.
Pearsall, and J. Wulff,
The Structure and
Properties of
Materials, Vol. I,
Structure, p. 10.
Copyright © 1964 by
John Wiley & Sons,
New York. Reprinted
by permission of John
Wiley & Sons, Inc.)
1496T_c02_15-37 12/20/05 7:19 Page 19
2nd REVISE PAGES
2.3 Electrons in Atoms • 19
Figure 2.3 Comparison of
the (a) Bohr and (b) wavemechanical atom models
in terms of electron
distribution. (Adapted from
Z. D. Jastrzebski, The
Nature and Properties of
Engineering Materials, 3rd
edition, p. 4. Copyright ©
1987 by John Wiley & Sons,
New York. Reprinted by
permission of John Wiley &
Sons, Inc.)
Probability
1.0
0
Distance from nucleus
Orbital electron
Nucleus
(a)
(b)
Quantum Numbers
quantum number
Using wave mechanics, every electron in an atom is characterized by four parameters called quantum numbers. The size, shape, and spatial orientation of an electron’s
probability density are specified by three of these quantum numbers. Furthermore,
Bohr energy levels separate into electron subshells, and quantum numbers dictate the
number of states within each subshell. Shells are specified by a principal quantum
number n, which may take on integral values beginning with unity; sometimes these
shells are designated by the letters K, L, M, N, O, and so on, which correspond,
respectively, to n 1, 2, 3, 4, 5, . . . , as indicated in Table 2.1. Note also that this quantum number, and it only, is also associated with the Bohr model. This quantum number is related to the distance of an electron from the nucleus, or its position.
The second quantum number, l, signifies the subshell, which is denoted by a
lowercase letter—an s, p, d, or f; it is related to the shape of the electron subshell.
In addition, the number of these subshells is restricted by the magnitude of n.
Allowable subshells for the several n values are also presented in Table 2.1. The
number of energy states for each subshell is determined by the third quantum number, ml. For an s subshell, there is a single energy state, whereas for p, d, and f subshells, three, five, and seven states exist, respectively (Table 2.1). In the absence of
an external magnetic field, the states within each subshell are identical. However,
when a magnetic field is applied these subshell states split, each state assuming a
slightly different energy.
1496T_c02_15-37 11/10/05 10:42 Page 20
REVISED PAGES
20 • Chapter 2 / Atomic Structure and Interatomic Bonding
Table 2.1 The Number of Available Electron States in Some of the Electron
Shells and Subshells
Principal
Quantum
Number n
Shell
Designation
1
Number of Electrons
Subshells
Number
of States
Per Subshell
Per Shell
K
s
1
2
2
2
L
s
p
1
3
2
6
8
3
M
s
p
d
1
3
5
2
6
10
18
N
s
p
d
f
1
3
5
7
2
6
10
14
32
4
Energy
Associated with each electron is a spin moment, which must be oriented either
up or down. Related to this spin moment is the fourth quantum number, ms, for
which two values are possible ( 12 and 12), one for each of the spin orientations.
Thus, the Bohr model was further refined by wave mechanics, in which the introduction of three new quantum numbers gives rise to electron subshells within
each shell. A comparison of these two models on this basis is illustrated, for the
hydrogen atom, in Figures 2.2a and 2.2b.
A complete energy level diagram for the various shells and subshells using the
wave-mechanical model is shown in Figure 2.4. Several features of the diagram are
worth noting. First, the smaller the principal quantum number, the lower the energy
level; for example, the energy of a 1s state is less than that of a 2s state, which in
turn is lower than the 3s. Second, within each shell, the energy of a subshell level increases with the value of the l quantum number. For example, the energy of a 3d
state is greater than a 3p, which is larger than 3s. Finally, there may be overlap in
f
d
f
d
p
s
f
d
p
s
d
p
s
Figure 2.4 Schematic
representation of the relative
energies of the electrons for the
various shells and subshells. (From
K. M. Ralls, T. H. Courtney, and
J. Wulff, Introduction to Materials
Science and Engineering, p. 22.
Copyright © 1976 by John Wiley &
Sons, New York. Reprinted by
permission of John Wiley & Sons,
Inc.)
p
s
d
p
s
p
s
s
1
2
3
4
5
Principal quantum number, n
6
7
1496T_c02_15-37 11/10/05 13:32 Page 21
REVISED PAGES
2.3 Electrons in Atoms • 21
energy of a state in one shell with states in an adjacent shell, which is especially true
of d and f states; for example, the energy of a 3d state is greater than that for a 4s.
Electron Configurations
Pauli exclusion
principle
ground state
electron
configuration
valence electron
The preceding discussion has dealt primarily with electron states—values of energy
that are permitted for electrons.To determine the manner in which these states are filled
with electrons, we use the Pauli exclusion principle, another quantum-mechanical
concept. This principle stipulates that each electron state can hold no more than two
electrons, which must have opposite spins. Thus, s, p, d, and f subshells may each accommodate, respectively, a total of 2, 6, 10, and 14 electrons; Table 2.1 summarizes the
maximum number of electrons that may occupy each of the first four shells.
Of course, not all possible states in an atom are filled with electrons. For most
atoms, the electrons fill up the lowest possible energy states in the electron shells and
subshells, two electrons (having opposite spins) per state. The energy structure for a
sodium atom is represented schematically in Figure 2.5. When all the electrons occupy the lowest possible energies in accord with the foregoing restrictions, an atom
is said to be in its ground state. However, electron transitions to higher energy states
are possible, as discussed in Chapters 18 and 21. The electron configuration or structure of an atom represents the manner in which these states are occupied. In the
conventional notation the number of electrons in each subshell is indicated by a superscript after the shell–subshell designation. For example, the electron configurations
for hydrogen, helium, and sodium are, respectively, 1s1, 1s2, and 1s22s22p63s1. Electron
configurations for some of the more common elements are listed in Table 2.2.
At this point, comments regarding these electron configurations are necessary.
First, the valence electrons are those that occupy the outermost shell. These electrons are extremely important; as will be seen, they participate in the bonding between atoms to form atomic and molecular aggregates. Furthermore, many of the
physical and chemical properties of solids are based on these valence electrons.
In addition, some atoms have what are termed “stable electron configurations”;
that is, the states within the outermost or valence electron shell are completely
filled. Normally this corresponds to the occupation of just the s and p states for
the outermost shell by a total of eight electrons, as in neon, argon, and krypton;
one exception is helium, which contains only two 1s electrons. These elements
(Ne, Ar, Kr, and He) are the inert, or noble, gases, which are virtually unreactive
chemically. Some atoms of the elements that have unfilled valence shells assume
stable electron configurations by gaining or losing electrons to form charged ions,
3p
3s
Increasing energy
electron state
2p
2s
1s
Figure 2.5 Schematic representation of the
filled and lowest unfilled energy states for a
sodium atom.
1496T_c02_15-37 11/10/05 10:42 Page 22
REVISED PAGES
22 • Chapter 2 / Atomic Structure and Interatomic Bonding
Table 2.2 A Listing of the Expected Electron Configurations for
Some of the Common Elementsa
Element
Hydrogen
Helium
Lithium
Beryllium
Boron
Carbon
Nitrogen
Oxygen
Fluorine
Neon
Sodium
Magnesium
Aluminum
Silicon
Phosphorus
Sulfur
Chlorine
Argon
Potassium
Calcium
Scandium
Titanium
Vanadium
Chromium
Manganese
Iron
Cobalt
Nickel
Copper
Zinc
Gallium
Germanium
Arsenic
Selenium
Bromine
Krypton
Symbol
Atomic
Number
H
He
Li
Be
B
C
N
O
F
Ne
Na
Mg
Al
Si
P
S
Cl
Ar
K
Ca
Sc
Ti
V
Cr
Mn
Fe
Co
Ni
Cu
Zn
Ga
Ge
As
Se
Br
Kr
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Electron Configuration
1s1
1s2
1s22s1
1s22s2
1s22s22p1
1s22s22p2
1s22s22p3
1s22s22p4
1s22s22p5
1s22s22p6
1s22s22p63s1
1s22s22p63s2
1s22s22p63s23p1
1s22s22p63s23p2
1s22s22p63s23p3
1s22s22p63s23p4
1s22s22p63s23p5
1s22s22p63s23p6
1s22s22p63s23p64s1
1s22s22p63s23p64s2
1s22s22p63s23p63d14s2
1s22s22p63s23p63d24s2
1s22s22p63s23p63d34s2
1s22s22p63s23p63d54s1
1s22s22p63s23p63d54s2
1s22s22p63s23p63d64s2
1s22s22p63s23p63d74s2
1s22s22p63s23p63d84s2
1s22s22p63s23p63d104s1
1s22s22p63s23p63d104s2
1s22s22p63s23p63d104s24p1
1s22s22p63s23p63d104s24p2
1s22s22p63s23p63d104s24p3
1s22s22p63s23p63d104s24p4
1s22s22p63s23p63d104s24p5
1s22s22p63s23p63d104s24p6
a
When some elements covalently bond, they form sp hybrid bonds. This is especially true for C, Si, and Ge.
or by sharing electrons with other atoms. This is the basis for some chemical
reactions, and also for atomic bonding in solids, as explained in Section 2.6.
Under special circumstances, the s and p orbitals combine to form hybrid
spn orbitals, where n indicates the number of p orbitals involved, which may have
a value of 1, 2, or 3. The 3A, 4A, and 5A group elements of the periodic table
(Figure 2.6) are those that most often form these hybrids. The driving force for the
formation of hybrid orbitals is a lower energy state for the valence electrons. For
carbon the sp3 hybrid is of primary importance in organic and polymer chemistries.
The shape of the sp3 hybrid is what determines the 109 (or tetrahedral) angle found
in polymer chains (Chapter 14).
1496T_c02_15-37 11/10/05 10:42 Page 23
REVISED PAGES
2.4 The Periodic Table • 23
Concept Check 2.2
Give electron configurations for the Fe3 and S2 ions.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
2.4 THE PERIODIC TABLE
All the elements have been classified according to electron configuration in the
periodic table (Figure 2.6). Here, the elements are situated, with increasing atomic
number, in seven horizontal rows called periods. The arrangement is such that all
elements arrayed in a given column or group have similar valence electron structures, as well as chemical and physical properties. These properties change gradually, moving horizontally across each period and vertically down each column.
The elements positioned in Group 0, the rightmost group, are the inert gases,
which have filled electron shells and stable electron configurations. Group VIIA and
VIA elements are one and two electrons deficient, respectively, from having stable
structures. The Group VIIA elements (F, Cl, Br, I, and At) are sometimes termed
the halogens. The alkali and the alkaline earth metals (Li, Na, K, Be, Mg, Ca, etc.)
are labeled as Groups IA and IIA, having, respectively, one and two electrons in excess of stable structures. The elements in the three long periods, Groups IIIB through
IIB, are termed the transition metals, which have partially filled d electron states and
in some cases one or two electrons in the next higher energy shell. Groups IIIA,
IVA, and VA (B, Si, Ge, As, etc.) display characteristics that are intermediate between the metals and nonmetals by virtue of their valence electron structures.
periodic table
Metal
IA
Key
1
29
Atomic number
H
Cu
Symbol
1.0080
3
IIA
63.54
4
Li
Be
6.941
11
9.0122
12
0
Nonmetal
2
He
Atomic weight
Intermediate
VIII
IIIA
IVA
VA
VIA
VIIA
5
6
7
8
9
4.0026
10
B
C
N
O
F
Ne
10.811
13
12.011
14
14.007
15
15.999
16
18.998
17
20.180
18
Na
Mg
22.990
19
24.305
20
IIIB
IVB
VB
VIB
21
22
23
24
25
26
27
28
29
30
K
Ca
Sc
Ti
V
Cr
Mn
Fe
Co
Ni
Cu
Zn
Ga
Ge
As
Se
Br
Kr
39.098
40.08
44.956
47.87
50.942
51.996
54.938
55.845
58.933
58.69
63.54
65.41
69.72
72.64
74.922
78.96
79.904
83.80
VIIB
IB
IIB
Al
Si
P
S
Cl
Ar
26.982
31
28.086
32
30.974
33
32.064
34
35.453
35
39.948
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
Rb
Sr
Y
Zr
Nb
Mo
Tc
Ru
Rh
Pd
Ag
Cd
In
Sn
Sb
Te
I
Xe
85.47
55
87.62
56
88.91
91.22
72
92.91
73
95.94
74
(98)
75
101.07
76
102.91
77
106.4
78
107.87
79
112.41
80
114.82
81
118.71
82
121.76
83
127.60
84
126.90
85
131.30
86
Cs
Ba
132.91
87
137.34
88
Fr
Ra
(223)
(226)
Rare
earth
series
Actinide
series
Rare earth series
Actinide series
Hf
Ta
W
Re
Os
Ir
Pt
Au
Hg
Tl
Pb
Bi
Po
At
Rn
178.49
104
180.95
105
183.84
106
186.2
107
190.23
108
192.2
109
195.08
110
196.97
200.59
204.38
207.19
208.98
(209)
(210)
(222)
Rf
Db
Sg
Bh
Hs
Mt
Ds
(261)
(262)
(266)
(264)
(277)
(268)
(281)
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
La
Ce
Pr
Nd
Pm
Sm
Eu
Gd
Tb
Dy
Ho
Er
Tm
Yb
Lu
138.91
140.12
140.91
144.24
(145)
150.35
151.96
157.25
158.92
162.50
164.93
167.26
168.93
173.04
174.97
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
Ac
Th
Pa
U
Np
Pu
Am
Cm
Bk
Cf
Es
Fm
Md
No
Lr
(227)
232.04
231.04
238.03
(237)
(244)
(243)
(247)
(247)
(251)
(252)
(257)
(258)
(259)
(262)
Figure 2.6 The periodic table of the elements. The numbers in parentheses are the atomic
weights of the most stable or common isotopes.
1496T_c02_15-37 11/10/05 10:42 Page 24
REVISED PAGES
24 • Chapter 2 / Atomic Structure and Interatomic Bonding
0
IA
1
2
H
He
2.1
3
IIA
IIIA
IVA
VA
VIA
VIIA
4
5
6
7
8
9
–
10
Li
Be
B
C
N
O
F
Ne
1.0
11
1.5
12
2.0
13
2.5
14
3.0
15
3.5
16
4.0
17
–
18
Na
Mg
0.9
19
1.2
20
VIII
IIIB
IVB
VB
VIB
VIIB
21
22
23
24
25
26
27
28
IB
IIB
29
30
Al
Si
P
S
Cl
Ar
1.5
31
1.8
32
2.1
33
2.5
34
3.0
35
–
36
Kr
K
Ca
Sc
Ti
V
Cr
Mn
Fe
Co
Ni
Cu
Zn
Ga
Ge
As
Se
Br
0.8
1.0
1.3
1.5
1.6
1.6
1.5
1.8
1.8
1.8
1.9
1.6
1.6
1.8
2.0
2.4
2.8
–
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
Rb
Sr
Y
Zr
Nb
Mo
Tc
Ru
Rh
Pd
Ag
Cd
In
Sn
Sb
Te
I
Xe
0.8
55
1.0
56
1.2
57–71
1.4
72
1.6
73
1.8
74
1.9
75
2.2
76
2.2
77
2.2
78
1.9
79
1.7
80
1.7
81
1.8
82
1.9
83
2.1
84
2.5
85
–
86
Cs
Ba
La–Lu
Hf
Ta
W
Re
Os
Ir
Pt
Au
Hg
Tl
Pb
Bi
Po
At
Rn
0.7
87
0.9
88
1.1–1.2
89–102
1.3
1.5
1.7
1.9
2.2
2.2
2.2
2.4
1.9
1.8
1.8
1.9
2.0
2.2
–
Fr
Ra
Ac–No
0.7
0.9
1.1–1.7
Figure 2.7 The electronegativity values for the elements. (Adapted from Linus Pauling, The
Nature of the Chemical Bond, 3rd edition. Copyright 1939 and 1940, 3rd edition copyright ©
1960, by Cornell University. Used by permission of the publisher, Cornell University Press.)
electropositive
electronegative
As may be noted from the periodic table, most of the elements really come under the metal classification. These are sometimes termed electropositive elements,
indicating that they are capable of giving up their few valence electrons to become
positively charged ions. Furthermore, the elements situated on the right-hand side
of the table are electronegative; that is, they readily accept electrons to form negatively charged ions, or sometimes they share electrons with other atoms. Figure 2.7
displays electronegativity values that have been assigned to the various elements
arranged in the periodic table. As a general rule, electronegativity increases in moving from left to right and from bottom to top. Atoms are more likely to accept electrons if their outer shells are almost full, and if they are less “shielded” from (i.e.,
closer to) the nucleus.
Atomic Bonding in Solids
2.5 BONDING FORCES AND ENERGIES
An understanding of many of the physical properties of materials is predicated on
a knowledge of the interatomic forces that bind the atoms together. Perhaps the
principles of atomic bonding are best illustrated by considering the interaction
between two isolated atoms as they are brought into close proximity from an infinite separation. At large distances, the interactions are negligible, but as the atoms
approach, each exerts forces on the other. These forces are of two types, attractive
and repulsive, and the magnitude of each is a function of the separation or interatomic distance. The origin of an attractive force FA depends on the particular type
of bonding that exists between the two atoms. The magnitude of the attractive force
varies with the distance, as represented schematically in Figure 2.8a. Ultimately,
the outer electron shells of the two atoms begin to overlap, and a strong repulsive
force FR comes into play. The net force FN between the two atoms is just the sum
of both attractive and repulsive components; that is,
FN FA FR
(2.2)
1496T_c02_15-37 11/10/05 10:42 Page 25
REVISED PAGES
2.5 Bonding Forces and Energies • 25
+
Figure 2.8 (a) The
dependence of repulsive,
attractive, and net forces on
interatomic separation for
two isolated atoms. (b) The
dependence of repulsive,
attractive, and net potential
energies on interatomic
separation for two isolated
atoms.
Force F
Attraction
Attractive force FA
0
Repulsion
Interatomic separation r
r0
Repulsive force FR
Net force FN
–
(a)
Repulsion
Repulsive energy ER
Interatomic separation r
0
Net energy EN
Attraction
Potential energy E
+
E0
Attractive energy EA
–
(b)
which is also a function of the interatomic separation, as also plotted in Figure 2.8a.
When FA and FR balance, or become equal, there is no net force; that is,
FA FR 0
(2.3)
Then a state of equilibrium exists. The centers of the two atoms will remain separated by the equilibrium spacing r0, as indicated in Figure 2.8a. For many atoms, r0
is approximately 0.3 nm. Once in this position, the two atoms will counteract any
attempt to separate them by an attractive force, or to push them together by a
repulsive action.
Sometimes it is more convenient to work with the potential energies between
two atoms instead of forces. Mathematically, energy (E) and force (F) are related as
Force-potential
energy relationship
for two atoms
E
F dr
(2.4)
dr
(2.5)
or, for atomic systems,
r
EN
F
N
q
r
FA dr
q
EA ER
r
F
R
dr
(2.6)
q
(2.7)
1496T_c02_15-37 11/10/05 10:42 Page 26
REVISED PAGES
26 • Chapter 2 / Atomic Structure and Interatomic Bonding
bonding energy
primary bond
in which EN, EA, and ER are respectively the net, attractive, and repulsive energies
for two isolated and adjacent atoms.
Figure 2.8b plots attractive, repulsive, and net potential energies as a function of
interatomic separation for two atoms. The net curve, which is again the sum of the
other two, has a potential energy trough or well around its minimum. Here, the same
equilibrium spacing, r0, corresponds to the separation distance at the minimum of the
potential energy curve. The bonding energy for these two atoms, E0, corresponds to
the energy at this minimum point (also shown in Figure 2.8b); it represents the energy that would be required to separate these two atoms to an infinite separation.
Although the preceding treatment has dealt with an ideal situation involving
only two atoms, a similar yet more complex condition exists for solid materials because force and energy interactions among many atoms must be considered. Nevertheless, a bonding energy, analogous to E0 above, may be associated with each
atom. The magnitude of this bonding energy and the shape of the energy-versusinteratomic separation curve vary from material to material, and they both depend
on the type of atomic bonding. Furthermore, a number of material properties depend on E0, the curve shape, and bonding type. For example, materials having large
bonding energies typically also have high melting temperatures; at room temperature, solid substances are formed for large bonding energies, whereas for small
energies the gaseous state is favored; liquids prevail when the energies are of intermediate magnitude. In addition, as discussed in Section 6.3, the mechanical stiffness
(or modulus of elasticity) of a material is dependent on the shape of its force-versusinteratomic separation curve (Figure 6.7). The slope for a relatively stiff material at
the r r0 position on the curve will be quite steep; slopes are shallower for more
flexible materials. Furthermore, how much a material expands upon heating or contracts upon cooling (that is, its linear coefficient of thermal expansion) is related to
the shape of its E0-versus-r0 curve (see Section 19.3). A deep and narrow “trough,”
which typically occurs for materials having large bonding energies, normally correlates with a low coefficient of thermal expansion and relatively small dimensional
alterations for changes in temperature.
Three different types of primary or chemical bond are found in solids—ionic,
covalent, and metallic. For each type, the bonding necessarily involves the valence
electrons; furthermore, the nature of the bond depends on the electron structures
of the constituent atoms. In general, each of these three types of bonding arises
from the tendency of the atoms to assume stable electron structures, like those of
the inert gases, by completely filling the outermost electron shell.
Secondary or physical forces and energies are also found in many solid materials; they are weaker than the primary ones, but nonetheless influence the physical properties of some materials. The sections that follow explain the several kinds
of primary and secondary interatomic bonds.
2.6 PRIMARY INTERATOMIC BONDS
Ionic Bonding
ionic bonding
Ionic bonding is perhaps the easiest to describe and visualize. It is always found in
compounds that are composed of both metallic and nonmetallic elements, elements
that are situated at the horizontal extremities of the periodic table. Atoms of a metallic element easily give up their valence electrons to the nonmetallic atoms. In the
process all the atoms acquire stable or inert gas configurations and, in addition, an
electrical charge; that is, they become ions. Sodium chloride (NaCl) is the classic ionic
material. A sodium atom can assume the electron structure of neon (and a net single
1496T_c02_15-37 11/10/05 10:42 Page 27
REVISED PAGES
2.6 Primary Interatomic Bonds • 27
Figure 2.9 Schematic representation of ionic
bonding in sodium chloride (NaCl).
Coulombic bonding force
coulombic force
Na+
Cl–
Na+
Cl–
Na+
Cl–
Na+
Cl–
Na+
Cl–
Na+
Cl–
Na+
Cl–
Na+
Cl–
Na+
Cl–
Na+
Cl–
positive charge) by a transfer of its one valence 3s electron to a chlorine atom. After such a transfer, the chlorine ion has a net negative charge and an electron configuration identical to that of argon. In sodium chloride, all the sodium and chlorine
exist as ions. This type of bonding is illustrated schematically in Figure 2.9.
The attractive bonding forces are coulombic; that is, positive and negative ions,
by virtue of their net electrical charge, attract one another. For two isolated ions,
the attractive energy EA is a function of the interatomic distance according to3
Attractive energyinteratomic
separation
relationship
EA
A
r
(2.8)
An analogous equation for the repulsive energy is
Repulsive energyinteratomic
separation
relationship
ER
B
rn
(2.9)
In these expressions, A, B, and n are constants whose values depend on the particular ionic system. The value of n is approximately 8.
Ionic bonding is termed nondirectional; that is, the magnitude of the bond is equal
in all directions around an ion. It follows that for ionic materials to be stable, all
positive ions must have as nearest neighbors negatively charged ions in a threedimensional scheme, and vice versa. The predominant bonding in ceramic materials is
ionic. Some of the ion arrangements for these materials are discussed in Chapter 12.
Bonding energies, which generally range between 600 and 1500 kJ/mol (3 and
8 eV/atom), are relatively large, as reflected in high melting temperatures.4 Table
2.3 contains bonding energies and melting temperatures for several ionic materials.
3
The constant A in Equation 2.8 is equal to
1
1Z e21Z2e2
4p0 1
where 0 is the permittivity of a vacuum 18.85 1012 F/m2, Z1 and Z2 are the valences of
the two ion types, and e is the electronic charge 11.602 1019 C2.
4
Sometimes bonding energies are expressed per atom or per ion. Under these circumstances the electron volt (eV) is a conveniently small unit of energy. It is, by definition, the
energy imparted to an electron as it falls through an electric potential of one volt. The
joule equivalent of the electron volt is as follows: 1.602 1019 J 1 eV.
1496T_c02_15-37 11/10/05 10:42 Page 28
REVISED PAGES
28 • Chapter 2 / Atomic Structure and Interatomic Bonding
Table 2.3 Bonding Energies and Melting Temperatures for
Various Substances
Bonding Energy
Bonding Type
Substance
Melting
Temperature
(°C)
kJ/mol
eV/Atom,
Ion, Molecule
640
1000
3.3
5.2
801
2800
Ionic
NaCl
MgO
Covalent
Si
C (diamond)
450
713
4.7
7.4
1410
3550
Metallic
Hg
Al
Fe
W
68
324
406
849
0.7
3.4
4.2
8.8
39
660
1538
3410
van der Waals
Ar
Cl2
7.7
31
0.08
0.32
189
101
Hydrogen
NH3
H2O
35
51
0.36
0.52
78
0
Ionic materials are characteristically hard and brittle and, furthermore, electrically and
thermally insulative. As discussed in subsequent chapters, these properties are a direct
consequence of electron configurations and/or the nature of the ionic bond.
Covalent Bonding
covalent bonding
In covalent bonding, stable electron configurations are assumed by the sharing of
electrons between adjacent atoms. Two atoms that are covalently bonded will each
contribute at least one electron to the bond, and the shared electrons may be considered to belong to both atoms. Covalent bonding is schematically illustrated in
Figure 2.10 for a molecule of methane 1CH4 2 . The carbon atom has four valence
electrons, whereas each of the four hydrogen atoms has a single valence electron.
Each hydrogen atom can acquire a helium electron configuration (two 1s valence
electrons) when the carbon atom shares with it one electron. The carbon now has
four additional shared electrons, one from each hydrogen, for a total of eight valence
electrons, and the electron structure of neon. The covalent bond is directional; that
is, it is between specific atoms and may exist only in the direction between one atom
and another that participates in the electron sharing.
H
Shared electron
from carbon
Shared electron
from hydrogen
H
C
H
H
Figure 2.10 Schematic representation of
covalent bonding in a molecule of methane
1CH4 2 .
1496T_c02_15-37 11/10/05 10:42 Page 29
REVISED PAGES
2.6 Primary Interatomic Bonds • 29
Many nonmetallic elemental molecules 1H2, Cl2, F2, etc.2 as well as molecules
containing dissimilar atoms, such as CH4, H2O, HNO3, and HF, are covalently
bonded. Furthermore, this type of bonding is found in elemental solids such as diamond (carbon), silicon, and germanium and other solid compounds composed of
elements that are located on the right-hand side of the periodic table, such as gallium arsenide (GaAs), indium antimonide (InSb), and silicon carbide (SiC).
The number of covalent bonds that is possible for a particular atom is determined by the number of valence electrons. For N¿ valence electrons, an atom can
covalently bond with at most 8 N¿ other atoms. For example, N¿ 7 for chlorine,
and 8 N¿ 1, which means that one Cl atom can bond to only one other atom,
as in Cl2. Similarly, for carbon, N¿ 4, and each carbon atom has 8 4, or four,
electrons to share. Diamond is simply the three-dimensional interconnecting structure wherein each carbon atom covalently bonds with four other carbon atoms. This
arrangement is represented in Figure 12.15.
Covalent bonds may be very strong, as in diamond, which is very hard and has
a very high melting temperature, 7 3550C 16400F2, or they may be very weak, as
with bismuth, which melts at about 270C 1518F2 . Bonding energies and melting
temperatures for a few covalently bonded materials are presented in Table 2.3.
Polymeric materials typify this bond, the basic molecular structure being a long
chain of carbon atoms that are covalently bonded together with two of their available four bonds per atom. The remaining two bonds normally are shared with other
atoms, which also covalently bond. Polymeric molecular structures are discussed in
detail in Chapter 14.
It is possible to have interatomic bonds that are partially ionic and partially covalent, and, in fact, very few compounds exhibit pure ionic or covalent bonding. For
a compound, the degree of either bond type depends on the relative positions of
the constituent atoms in the periodic table (Figure 2.6) or the difference in their
electronegativities (Figure 2.7).The wider the separation (both horizontally—relative
to Group IVA—and vertically) from the lower left to the upper-right-hand corner
(i.e., the greater the difference in electronegativity), the more ionic the bond. Conversely, the closer the atoms are together (i.e., the smaller the difference in electronegativity), the greater the degree of covalency. The percentage ionic character
of a bond between elements A and B (A being the most electronegative) may be
approximated by the expression
% ionic character 51 exp310.2521XA XB 2 2 4 6 100
(2.10)
where XA and XB are the electronegativities for the respective elements.
Metallic Bonding
metallic bonding
Metallic bonding, the final primary bonding type, is found in metals and their alloys. A relatively simple model has been proposed that very nearly approximates
the bonding scheme. Metallic materials have one, two, or at most, three valence electrons. With this model, these valence electrons are not bound to any particular atom
in the solid and are more or less free to drift throughout the entire metal. They may
be thought of as belonging to the metal as a whole, or forming a “sea of electrons”
or an “electron cloud.” The remaining nonvalence electrons and atomic nuclei form
what are called ion cores, which possess a net positive charge equal in magnitude
to the total valence electron charge per atom. Figure 2.11 is a schematic illustration
of metallic bonding. The free electrons shield the positively charged ion cores from
mutually repulsive electrostatic forces, which they would otherwise exert upon one
another; consequently the metallic bond is nondirectional in character. In addition,
these free electrons act as a “glue” to hold the ion cores together. Bonding energies
1496T_c02_15-37 12/20/05 7:19 Page 30
2nd REVISE PAGES
30 • Chapter 2 / Atomic Structure and Interatomic Bonding
Ion cores
+
+
–
+
–
+
–
+
+
+
+
–
+
–
+
+
–
–
–
+
+
Figure 2.11 Schematic illustration of metallic
bonding.
+
–
+
+
Sea of valence electrons
and melting temperatures for several metals are listed in Table 2.3. Bonding may
be weak or strong; energies range from 68 kJ/mol (0.7 eV/atom) for mercury to
850 kJ/mol (8.8 eV/atom) for tungsten. Their respective melting temperatures are
39 and 3410C 138 and 6170F2 .
Metallic bonding is found in the periodic table for Group IA and IIA elements
and, in fact, for all elemental metals.
Some general behaviors of the various material types (i.e., metals, ceramics,
polymers) may be explained by bonding type. For example, metals are good conductors of both electricity and heat, as a consequence of their free electrons (see
Sections 18.5, 18.6 and 19.4). By way of contrast, ionically and covalently bonded
materials are typically electrical and thermal insulators, due to the absence of large
numbers of free electrons.
Furthermore, in Section 7.4 we note that at room temperature, most metals and
their alloys fail in a ductile manner; that is, fracture occurs after the materials have
experienced significant degrees of permanent deformation. This behavior is explained in terms of deformation mechanism (Section 7.2), which is implicitly related
to the characteristics of the metallic bond. Conversely, at room temperature ionically bonded materials are intrinsically brittle as a consequence of the electrically
charged nature of their component ions (see Section 12.10).
Concept Check 2.3
Offer an explanation as to why covalently bonded materials are generally less dense
than ionically or metallically bonded ones.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
2.7 SECONDARY BONDING OR
VAN DER WAALS BONDING
secondary bond
van der Waals bond
Secondary, van der Waals, or physical bonds are weak in comparison to the primary
or chemical ones; bonding energies are typically on the order of only 10 kJ/mol
(0.1 eV/atom). Secondary bonding exists between virtually all atoms or molecules,
1496T_c02_15-37 12/20/05 7:19 Page 31
2nd REVISE PAGES
2.7 Secondary Bonding or van der Waals Bonding • 31
+
–
+
–
Figure 2.12 Schematic illustration of van der Waals
bonding between two dipoles.
Atomic or molecular dipoles
dipole
hydrogen bonding
but its presence may be obscured if any of the three primary bonding types is present. Secondary bonding is evidenced for the inert gases, which have stable electron
structures, and, in addition, between molecules in molecular structures that are
covalently bonded.
Secondary bonding forces arise from atomic or molecular dipoles. In essence,
an electric dipole exists whenever there is some separation of positive and negative portions of an atom or molecule. The bonding results from the coulombic attraction between the positive end of one dipole and the negative region of an
adjacent one, as indicated in Figure 2.12. Dipole interactions occur between induced
dipoles, between induced dipoles and polar molecules (which have permanent
dipoles), and between polar molecules. Hydrogen bonding, a special type of secondary bonding, is found to exist between some molecules that have hydrogen as
one of the constituents. These bonding mechanisms are now discussed briefly.
Fluctuating Induced Dipole Bonds
A dipole may be created or induced in an atom or molecule that is normally electrically symmetric; that is, the overall spatial distribution of the electrons is symmetric with respect to the positively charged nucleus, as shown in Figure 2.13a. All
atoms are experiencing constant vibrational motion that can cause instantaneous
and short-lived distortions of this electrical symmetry for some of the atoms or molecules, and the creation of small electric dipoles, as represented in Figure 2.13b. One
of these dipoles can in turn produce a displacement of the electron distribution of
an adjacent molecule or atom, which induces the second one also to become a dipole that is then weakly attracted or bonded to the first; this is one type of van der
Waals bonding. These attractive forces may exist between large numbers of atoms
or molecules, which forces are temporary and fluctuate with time.
The liquefaction and, in some cases, the solidification of the inert gases and
other electrically neutral and symmetric molecules such as H2 and Cl2 are realized
because of this type of bonding. Melting and boiling temperatures are extremely
low in materials for which induced dipole bonding predominates; of all possible
intermolecular bonds, these are the weakest. Bonding energies and melting temperatures for argon and chlorine are also tabulated in Table 2.3.
Polar Molecule-Induced Dipole Bonds
Permanent dipole moments exist in some molecules by virtue of an asymmetrical
arrangement of positively and negatively charged regions; such molecules are
Atomic nucleus
Atomic nucleus
Electron cloud
Electron cloud
+
(a)
–
(b)
Figure 2.13 Schematic
representations of (a) an
electrically symmetric
atom and (b) an induced
atomic dipole.
1496T_c02_15-37 11/10/05 10:42 Page 32
REVISED PAGES
32 • Chapter 2 / Atomic Structure and Interatomic Bonding
H
Cl
+
polar molecule
Figure 2.14 Schematic representation of a polar hydrogen chloride (HCl)
molecule.
–
termed polar molecules. Figure 2.14 is a schematic representation of a hydrogen
chloride molecule; a permanent dipole moment arises from net positive and negative charges that are respectively associated with the hydrogen and chlorine ends
of the HCl molecule.
Polar molecules can also induce dipoles in adjacent nonpolar molecules, and
a bond will form as a result of attractive forces between the two molecules. Furthermore, the magnitude of this bond will be greater than for fluctuating induced dipoles.
Permanent Dipole Bonds
Van der Waals forces will also exist between adjacent polar molecules. The associated
bonding energies are significantly greater than for bonds involving induced dipoles.
The strongest secondary bonding type, the hydrogen bond, is a special case of polar molecule bonding. It occurs between molecules in which hydrogen is covalently
bonded to fluorine (as in HF), oxygen (as in H2O), and nitrogen (as in NH3). For
each H—F, H—O, or H—N bond, the single hydrogen electron is shared with the
other atom. Thus, the hydrogen end of the bond is essentially a positively charged
bare proton that is unscreened by any electrons. This highly positively charged end
of the molecule is capable of a strong attractive force with the negative end of an
adjacent molecule, as demonstrated in Figure 2.15 for HF. In essence, this single proton forms a bridge between two negatively charged atoms. The magnitude of the hydrogen bond is generally greater than that of the other types of secondary bonds
and may be as high as 51 kJ/mol (0.52 eV/molecule), as shown in Table 2.3. Melting
and boiling temperatures for hydrogen fluoride and water are abnormally high in
light of their low molecular weights, as a consequence of hydrogen bonding.
2.8 MOLECULES
Many of the common molecules are composed of groups of atoms that are bound
together by strong covalent bonds; these include elemental diatomic molecules
1F2, O2, H2, etc.2 as well as a host of compounds 1H2O, CO2, HNO3, C6H6, CH4, etc.2.
In the condensed liquid and solid states, bonds between molecules are weak secondary ones. Consequently, molecular materials have relatively low melting and
boiling temperatures. Most of those that have small molecules composed of a few
atoms are gases at ordinary, or ambient, temperatures and pressures. On the other
hand, many of the modern polymers, being molecular materials composed of
extremely large molecules, exist as solids; some of their properties are strongly
dependent on the presence of van der Waals and hydrogen secondary bonds.
H
F
H
Hydrogen
bond
F
Figure 2.15 Schematic representation of hydrogen
bonding in hydrogen fluoride (HF).
1496T_c02_15-37 11/10/05 10:43 Page 33
REVISED PAGES
2.8 Molecules • 33
MATERIAL OF IMPORTANCE
Water (Its Volume Expansion Upon Freezing)
U
pon freezing (i.e., transforming from a liquid to a solid upon cooling), most substances
experience an increase in density (or, correspondingly, a decrease in volume). One exception
is water, which exhibits the anomalous and familiar expansion upon freezing—approximately
9 volume percent expansion. This behavior may
be explained on the basis of hydrogen bonding.
Each H2O molecule has two hydrogen atoms that
can bond to oxygen atoms; in addition, its single
O atom can bond to two hydrogen atoms of other
H2O molecules. Thus, for solid ice, each water
molecule participates in four hydrogen bonds as
shown in the three-dimensional schematic of
Figure 2.16a; here hydrogen bonds are denoted
by dashed lines, and each water molecule has 4
nearest-neighbor molecules. This is a relatively
open structure—i.e., the molecules are not
closely packed together—and, as a result, the
density is comparatively low. Upon melting, this
structure is partially destroyed, such that the
water molecules become more closely packed together (Figure 2.16b)—at room temperature the
average number of nearest-neighbor water molecules has increased to approximately 4.5; this
leads to an increase in density.
Consequences of this anomalous freezing
phenomenon are familiar. This phenomenon explains why icebergs float, why, in cold climates, it
is necessary to add antifreeze to an automobile’s
cooling system (to keep the engine block from
cracking), and why freeze-thaw cycles break up
the pavement in streets and cause potholes to
form.
H
H
O
Hydrogen bond
H
H
O
O
H
H
O
H
O
H
H
(a)
H
H
H
O
H
H
H
H
H
H
H
O
H
O
O
H
H
H
O
O
O
H
H
H
H
H
O
O
A watering can that ruptured along a side panelbottom panel seam. Water that was left in the can
during a cold late-autumn night expanded as it
froze and caused the rupture. (Photography
by S. Tanner.)
O
H
(b)
Figure 2.16 The arrangement of water 1H2O2
molecules in (a) solid ice, and (b) liquid water.
1496T_c02_15-37 11/10/05 10:43 Page 34
REVISED PAGES
34 • Chapter 2 / Atomic Structure and Interatomic Bonding
SUMMARY
Electrons in Atoms
The Periodic Table
This chapter began with a survey of the fundamentals of atomic structure, presenting the Bohr and wave-mechanical models of electrons in atoms. Whereas the Bohr
model assumes electrons to be particles orbiting the nucleus in discrete paths, in
wave mechanics we consider them to be wave-like and treat electron position in
terms of a probability distribution.
Electron energy states are specified in terms of quantum numbers that give rise
to electron shells and subshells. The electron configuration of an atom corresponds
to the manner in which these shells and subshells are filled with electrons in compliance with the Pauli exclusion principle.The periodic table of the elements is generated
by arrangement of the various elements according to valence electron configuration.
Bonding Forces and Energies
Primary Interatomic Bonds
Atomic bonding in solids may be considered in terms of attractive and repulsive
forces and energies. The three types of primary bond in solids are ionic, covalent,
and metallic. For ionic bonds, electrically charged ions are formed by the transference of valence electrons from one atom type to another; forces are coulombic.There
is a sharing of valence electrons between adjacent atoms when bonding is covalent.
With metallic bonding, the valence electrons form a “sea of electrons” that is uniformly dispersed around the metal ion cores and acts as a form of glue for them.
Secondary Bonding or van der Waals Bonding
Both van der Waals and hydrogen bonds are termed secondary, being weak in comparison to the primary ones. They result from attractive forces between electric
dipoles, of which there are two types—induced and permanent. For the hydrogen
bond, highly polar molecules form when hydrogen covalently bonds to a nonmetallic
element such as fluorine.
I M P O R TA N T T E R M S A N D C O N C E P T S
Atomic mass unit (amu)
Atomic number
Atomic weight
Bohr atomic model
Bonding energy
Coulombic force
Covalent bond
Dipole (electric)
Electron configuration
Electron state
Electronegative
Electropositive
Ground state
Hydrogen bond
Ionic bond
Isotope
Metallic bond
Mole
Pauli exclusion principle
Periodic table
Polar molecule
Primary bonding
Quantum mechanics
Quantum number
Secondary bonding
Valence electron
van der Waals bond
Wave-mechanical model
Note: In each chapter, most of the terms listed in the “Important Terms and Concepts”
section are defined in the Glossary, which follows Appendix E. The others are important
enough to warrant treatment in a full section of the text and can be referenced from the
table of contents or the index.
1496T_c02_15-37 11/10/05 10:43 Page 35
REVISED PAGES
Questions and Problems • 35
REFERENCES
Most of the material in this chapter is covered in
college-level chemistry textbooks.Two are listed
here as references.
Brady, J. E., and F. Senese, Chemistry: Matter and
Its Changes, 4th edition, John Wiley & Sons,
Inc., Hoboken, NJ, 2004.
Ebbing, D. D., S. D. Gammon, and R. O. Ragsdale,
Essentials of General Chemistry, 2nd edition,
Houghton Mifflin Company, Boston, 2006.
QUESTIONS AND PROBLEMS
Fundamental Concepts
Electrons in Atoms
2.1 (a) Cite the difference between atomic mass
and atomic weight.
2.2 Silicon has three naturally-occurring isotopes: 92.23% of 28Si, with an atomic weight
of 27.9769 amu, 4.68% of 29Si, with an atomic
weight of 28.9765 amu, and 3.09% of 30Si,
with an atomic weight of 29.9738 amu. On the
basis of these data, confirm that the average
atomic weight of Si is 28.0854 amu.
2.3 (a) How many grams are there in one amu of
a material?
(b) Mole, in the context of this book, is taken
in units of gram-mole. On this basis, how
many atoms are there in a pound-mole of a
substance?
2.4 (a) Cite two important quantum-mechanical
concepts associated with the Bohr model of
the atom.
(b) Cite two important additional refinements
that resulted from the wave-mechanical atomic
model.
2.5 Relative to electrons and electron states,
what does each of the four quantum numbers
specify?
2.6 Allowed values for the quantum numbers of
electrons are as follows:
n 1, 2, 3, . . .
l 0, 1, 2, 3, . . . , n 1
ml 0, 1, 2, 3, . . . , l
ms 12
The relationships between n and the shell designations are noted in Table 2.1. Relative to
the subshells,
l
l
l
l
0
1
2
3
corresponds
corresponds
corresponds
corresponds
to
to
to
to
an s subshell
a p subshell
a d subshell
an f subshell
For the K shell, the four quantum numbers
for each of the two electrons in the 1s state,
in the order of nlmlms, are 1001 12 2 and
100112 2. Write the four quantum numbers
for all of the electrons in the L and M shells,
and note which correspond to the s, p, and
d subshells.
2.7 Give the electron configurations for the
following ions: P5, P3, Sn4, Se2, I, and
Ni2.
2.8 Potassium iodide (KI) exhibits predominantly
ionic bonding. The K and I ions have electron structures that are identical to which two
inert gases?
The Periodic Table
2.9 With regard to electron configuration, what
do all the elements in Group IIA of the periodic table have in common?
2.10 To what group in the periodic table would an
element with atomic number 112 belong?
2.11 Without consulting Figure 2.6 or Table 2.2, determine whether each of the electron configurations given below is an inert gas, a halogen,
an alkali metal, an alkaline earth metal, or a
transition metal. Justify your choices.
1496T_c02_15-37 11/10/05 10:43 Page 36
REVISED PAGES
36 • Chapter 2 / Atomic Structure and Interatomic Bonding
(a) 1s22s22p63s23p5
(b) 1s22s22p63s23p63d74s2
(c) 1s22s22p63s23p63d104s24p6
(d) 1s22s22p63s23p64s1
(e) 1s22s22p63s23p63d104s24p64d55s2
(f) 1s22s22p63s2
2.12 (a) What electron subshell is being filled for
the rare earth series of elements on the periodic table?
(b) What electron subshell is being filled for
the actinide series?
Bonding Forces and Energies
2.13 Calculate the force of attraction between a
Ca2 and an O2 ion the centers of which are
separated by a distance of 1.25 nm.
2.14 The net potential energy between two adjacent ions, EN, may be represented by the sum
of Equations 2.8 and 2.9; that is,
EN
B
A
n
r
r
(2.11)
Calculate the bonding energy E0 in terms of
the parameters A, B, and n using the following procedure:
1. Differentiate EN with respect to r, and then
set the resulting expression equal to zero,
since the curve of EN versus r is a minimum
at E0.
2. Solve for r in terms of A, B, and n, which
yields r0, the equilibrium interionic spacing.
3. Determine the expression for E0 by substitution of r0 into Equation 2.11.
2.15 For an NaCl ion pair, attractive and repulsive energies EA and ER, respectively,
depend on the distance between the ions r,
according to
EA
ER
1.436
r
7.32 106
r8
For these expressions, energies are expressed
in electron volts per Na Cl pair, and r is
the distance in nanometers. The net energy
EN is just the sum of the two expressions
above.
(a) Superimpose on a single plot EN , ER, and
EA versus r up to 1.0 nm.
(b) On the basis of this plot, determine (i) the
equilibrium spacing r0 between the Na and
Cl ions, and (ii) the magnitude of the bonding energy E0 between the two ions.
(c) Mathematically determine the r0 and E0
values using the solutions to Problem 2.14
and compare these with the graphical results
from part (b).
2.16 Consider a hypothetical X Y ion pair
for which the equilibrium interionic spacing
and bonding energy values are 0.38 nm and
5.37 eV, respectively. If it is known that
n in Equation 2.11 has a value of 8, using
the results of Problem 2.14, determine
explicit expressions for attractive and repulsive energies EA and ER of Equations 2.8
and 2.9.
2.17 The net potential energy EN between two adjacent ions is sometimes represented by the
expression
EN
r
C
D exp a b
r
r
(2.12)
in which r is the interionic separation and C,
D, and r are constants whose values depend
on the specific material.
(a) Derive an expression for the bonding energy E0 in terms of the equilibrium interionic
separation r0 and the constants D and r using
the following procedure:
1. Differentiate EN with respect to r and
set the resulting expression equal to zero.
2. Solve for C in terms of D, r, and r0.
3. Determine the expression for E0 by
substitution for C in Equation 2.12.
(b) Derive another expression for E0 in terms
of r0, C, and r using a procedure analogous to
the one outlined in part (a).
Primary Interatomic Bonds
2.18 (a) Briefly cite the main differences between
ionic, covalent, and metallic bonding.
(b) State the Pauli exclusion principle.
2.19 Compute the percentage ionic character of
the interatomic bond for each of the following compounds: MgO, GaP, CsF, CdS, and
FeO.
1496T_c02_15-37 11/10/05 10:43 Page 37
REVISED PAGES
Questions and Problems • 37
2.20 Make a plot of bonding energy versus melting temperature for the metals listed in Table
2.3. Using this plot, approximate the bonding
energy for molybdenum, which has a melting
temperature of 2617C.
2.21 Using Table 2.2, determine the number of covalent bonds that are possible for atoms of the
following elements: silicon, bromine, nitrogen,
sulfur, and neon.
2.22 What type(s) of bonding would be expected
for each of the following materials: solid xenon,
calcium fluoride (CaF2), bronze, cadmium telluride (CdTe), rubber, and tungsten?
Secondary Bonding or van der Waals Bonding
2.23 Explain why hydrogen fluoride (HF) has a
higher boiling temperature than hydrogen
chloride (HCl) (19.4 vs. 85C), even though
HF has a lower molecular weight.
1496T_c03_38-79 12/20/05 7:38 Page 38
Chapter
3
2nd REVISE PAGES
The Structure of
Crystalline Solids
Diffracted
beams
Incident
beam
Single crystal
X-ray source
Lead screen
Photographic plate
(a)
(b)
(a) X-ray diffraction photograph [or Laue photograph (Section 3.16)] for a single crystal of magnesium. (b) Schematic diagram illustrating how the spots (i.e., the diffraction pattern) in (a) are produced. The lead screen blocks out all beams generated from the x-ray source, except for a narrow beam traveling in a single direction. This incident beam is diffracted by individual crystallographic planes in the single crystal (having different orientations), which gives rise to the various diffracted
beams that impinge on the photographic plate. Intersections of these beams with the plate appear as spots when the film is
developed.
The large spot in the center of (a) is from the incident beam, which is parallel to a [0001] crystallographic direction. It
should be noted that the hexagonal symmetry of magnesium’s hexagonal close-packed crystal structure is indicated by the
diffraction spot pattern that was generated.
[Figure (a) courtesy of J. G. Byrne, Department of Metallurgical Engineering, University of Utah. Figure (b) from
J. E. Brady and F. Senese, Chemistry: Matter and Its Changes, 4th edition. Copyright © 2004 by John Wiley & Sons,
Hoboken, NJ. Reprinted by permission of John Wiley & Sons, Inc.]
WHY STUDY The Structure of Crystalline Solids?
The properties of some materials are directly related
to their crystal structures. For example, pure and undeformed magnesium and beryllium, having one crystal structure, are much more brittle (i.e., fracture at
lower degrees of deformation) than are pure and undeformed metals such as gold and silver that have yet
another crystal structure (see Section 7.4).
38 •
Furthermore, significant property differences
exist between crystalline and noncrystalline materials
having the same composition. For example, noncrystalline ceramics and polymers normally are optically
transparent; the same materials in crystalline (or semicrystalline) form tend to be opaque or, at best,
translucent.
1496T_c03_38-79 12/20/05 7:38 Page 39
2nd REVISE PAGES
Learning Objectives
After studying this chapter you should be able to do the following:
1. Describe the difference in atomic/molecular
5. Given three direction index integers, sketch the
structure between crystalline and noncrystalline
direction corresponding to these indices within a
materials.
unit cell.
2. Draw unit cells for face-centered cubic, body6. Specify the Miller indices for a plane that has
centered cubic, and hexagonal close-packed
been drawn within a unit cell.
crystal structures.
7. Describe how face-centered cubic and hexagonal
3. Derive the relationships between unit cell edge
close-packed crystal structures may be generated
length and atomic radius for face-centered cubic
by the stacking of close-packed planes of atoms.
and body-centered cubic crystal structures.
8. Distinguish between single crystals and polycrys4. Compute the densities for metals having facetalline materials.
centered cubic and body-centered cubic crystal
9. Define isotropy and anisotropy with respect to
structures given their unit cell dimensions.
material properties.
3.1 INTRODUCTION
Chapter 2 was concerned primarily with the various types of atomic bonding, which
are determined by the electron structures of the individual atoms. The present discussion is devoted to the next level of the structure of materials, specifically, to some
of the arrangements that may be assumed by atoms in the solid state. Within this
framework, concepts of crystallinity and noncrystallinity are introduced. For crystalline solids the notion of crystal structure is presented, specified in terms of a unit
cell. The three common crystal structures found in metals are then detailed, along
with the scheme by which crystallographic points, directions, and planes are expressed. Single crystals, polycrystalline, and noncrystalline materials are considered.
The final section of this chapter briefly describes how crystal structures are determined experimentally using x-ray diffraction techniques.
C r ys t a l St r u c t u r e s
3.2 FUNDAMENTAL CONCEPTS
crystalline
crystal structure
Solid materials may be classified according to the regularity with which atoms or
ions are arranged with respect to one another. A crystalline material is one in which
the atoms are situated in a repeating or periodic array over large atomic distances;
that is, long-range order exists, such that upon solidification, the atoms will position
themselves in a repetitive three-dimensional pattern, in which each atom is bonded
to its nearest-neighbor atoms. All metals, many ceramic materials, and certain polymers form crystalline structures under normal solidification conditions. For those
that do not crystallize, this long-range atomic order is absent; these noncrystalline
or amorphous materials are discussed briefly at the end of this chapter.
Some of the properties of crystalline solids depend on the crystal structure of
the material, the manner in which atoms, ions, or molecules are spatially arranged.
There is an extremely large number of different crystal structures all having longrange atomic order; these vary from relatively simple structures for metals to
exceedingly complex ones, as displayed by some of the ceramic and polymeric
materials. The present discussion deals with several common metallic crystal structures. Chapters 12 and 14 are devoted to crystal structures for ceramics and polymers, respectively.
1496T_c03_38-79 11/11/05 19:05 Page 40
REVISED PAGES
40 • Chapter 3 / The Structure of Crystalline Solids
(a)
( b)
Figure 3.1 For the facecentered cubic crystal
structure, (a) a hard
sphere unit cell
representation, (b) a
reduced-sphere unit cell,
and (c) an aggregate of
many atoms. [Figure
(c) adapted from W. G.
Moffatt, G. W. Pearsall,
and J. Wulff, The Structure
and Properties of
Materials, Vol. I, Structure,
p. 51. Copyright © 1964 by
John Wiley & Sons, New
York. Reprinted by
permission of John Wiley
& Sons, Inc.]
(c)
lattice
When describing crystalline structures, atoms (or ions) are thought of as being
solid spheres having well-defined diameters. This is termed the atomic hard sphere
model in which spheres representing nearest-neighbor atoms touch one another.
An example of the hard sphere model for the atomic arrangement found in some
of the common elemental metals is displayed in Figure 3.1c. In this particular case
all the atoms are identical. Sometimes the term lattice is used in the context of crystal structures; in this sense “lattice” means a three-dimensional array of points
coinciding with atom positions (or sphere centers).
3.3 UNIT CELLS
unit cell
The atomic order in crystalline solids indicates that small groups of atoms form a
repetitive pattern. Thus, in describing crystal structures, it is often convenient to subdivide the structure into small repeat entities called unit cells. Unit cells for most
crystal structures are parallelepipeds or prisms having three sets of parallel faces;
one is drawn within the aggregate of spheres (Figure 3.1c), which in this case happens to be a cube. A unit cell is chosen to represent the symmetry of the crystal
structure, wherein all the atom positions in the crystal may be generated by translations of the unit cell integral distances along each of its edges. Thus, the unit cell
is the basic structural unit or building block of the crystal structure and defines
the crystal structure by virtue of its geometry and the atom positions within.
1496T_c03_38-79 11/11/05 19:05 Page 41
REVISED PAGES
3.4 Metallic Crystal Structures • 41
Convenience usually dictates that parallelepiped corners coincide with centers of
the hard sphere atoms. Furthermore, more than a single unit cell may be chosen for
a particular crystal structure; however, we generally use the unit cell having the
highest level of geometrical symmetry.
3.4 METALLIC CRYSTAL STRUCTURES
The atomic bonding in this group of materials is metallic and thus nondirectional
in nature. Consequently, there are minimal restrictions as to the number and position of nearest-neighbor atoms; this leads to relatively large numbers of nearest
neighbors and dense atomic packings for most metallic crystal structures. Also, for
metals, using the hard sphere model for the crystal structure, each sphere represents an ion core. Table 3.1 presents the atomic radii for a number of metals. Three
relatively simple crystal structures are found for most of the common metals: facecentered cubic, body-centered cubic, and hexagonal close-packed.
The Face-Centered Cubic Cr ystal Structure
face-centered cubic
(FCC)
The crystal structure found for many metals has a unit cell of cubic geometry, with
atoms located at each of the corners and the centers of all the cube faces. It is aptly
called the face-centered cubic (FCC) crystal structure. Some of the familiar metals having this crystal structure are copper, aluminum, silver, and gold (see also Table 3.1).
Figure 3.1a shows a hard sphere model for the FCC unit cell, whereas in Figure 3.1b
the atom centers are represented by small circles to provide a better perspective of
atom positions. The aggregate of atoms in Figure 3.1c represents a section of crystal
consisting of many FCC unit cells. These spheres or ion cores touch one another across
a face diagonal; the cube edge length a and the atomic radius R are related through
Unit cell edge length
for face-centered
cubic
Crystal Systems and
Unit Cells for Metals
a 2R12
(3.1)
This result is obtained in Example Problem 3.1.
For the FCC crystal structure, each corner atom is shared among eight unit cells,
whereas a face-centered atom belongs to only two. Therefore, one-eighth of each of
the eight corner atoms and one-half of each of the six face atoms, or a total of four
whole atoms, may be assigned to a given unit cell. This is depicted in Figure 3.1a,
where only sphere portions are represented within the confines of the cube. The
Table 3.1 Atomic Radii and Crystal Structures for 16 Metals
Metal
Aluminum
Cadmium
Chromium
Cobalt
Copper
Gold
Iron ()
Lead
Crystal
Structurea
Atomic
Radiusb
(nm)
FCC
HCP
BCC
HCP
FCC
FCC
BCC
FCC
0.1431
0.1490
0.1249
0.1253
0.1278
0.1442
0.1241
0.1750
Metal
Crystal
Structure
Atomic
Radius
(nm)
Molybdenum
Nickel
Platinum
Silver
Tantalum
Titanium ()
Tungsten
Zinc
BCC
FCC
FCC
FCC
BCC
HCP
BCC
HCP
0.1363
0.1246
0.1387
0.1445
0.1430
0.1445
0.1371
0.1332
FCC face-centered cubic; HCP hexagonal close-packed; BCC body-centered cubic.
A nanometer (nm) equals 109 m; to convert from nanometers to angstrom units (Å),
multiply the nanometer value by 10.
a
b
1496T_c03_38-79 11/11/05 19:05 Page 42
REVISED PAGES
42 • Chapter 3 / The Structure of Crystalline Solids
coordination number
atomic packing
factor (APF)
Definition of atomic
packing factor
cell comprises the volume of the cube, which is generated from the centers of the
corner atoms as shown in the figure.
Corner and face positions are really equivalent; that is, translation of the cube
corner from an original corner atom to the center of a face atom will not alter the
cell structure.
Two other important characteristics of a crystal structure are the coordination
number and the atomic packing factor (APF). For metals, each atom has the same
number of nearest-neighbor or touching atoms, which is the coordination number.
For face-centered cubics, the coordination number is 12. This may be confirmed by
examination of Figure 3.1a; the front face atom has four corner nearest-neighbor
atoms surrounding it, four face atoms that are in contact from behind, and four other
equivalent face atoms residing in the next unit cell to the front, which is not shown.
The APF is the sum of the sphere volumes of all atoms within a unit cell (assuming the atomic hard sphere model) divided by the unit cell volume—that is
APF
volume of atoms in a unit cell
total unit cell volume
(3.2)
For the FCC structure, the atomic packing factor is 0.74, which is the maximum
packing possible for spheres all having the same diameter. Computation of this APF
is also included as an example problem. Metals typically have relatively large atomic
packing factors to maximize the shielding provided by the free electron cloud.
The Body-Centered Cubic Cr ystal Structure
body-centered cubic
(BCC)
Another common metallic crystal structure also has a cubic unit cell with atoms
located at all eight corners and a single atom at the cube center. This is called a
body-centered cubic (BCC) crystal structure. A collection of spheres depicting this
crystal structure is shown in Figure 3.2c, whereas Figures 3.2a and 3.2b are diagrams
of BCC unit cells with the atoms represented by hard sphere and reduced-sphere
models, respectively. Center and corner atoms touch one another along cube diagonals, and unit cell length a and atomic radius R are related through
Unit cell edge length
for body-centered
cubic
(a)
a
( b)
4R
13
(3.3)
(c)
Figure 3.2 For the body-centered cubic crystal structure, (a) a hard sphere unit cell
representation, (b) a reduced-sphere unit cell, and (c) an aggregate of many atoms. [Figure
(c) from W. G. Moffatt, G. W. Pearsall, and J. Wulff, The Structure and Properties of
Materials, Vol. I, Structure, p. 51. Copyright © 1964 by John Wiley & Sons, New York.
Reprinted by permission of John Wiley & Sons, Inc.]
1496T_c03_38-79 11/11/05 21:57 Page 43
REVISED PAGES
3.4 Metallic Crystal Structures • 43
H
F
G
E
J
c
D
B
C
A
a
(a)
(b)
Figure 3.3 For the hexagonal close-packed crystal structure, (a) a reduced-sphere unit
cell (a and c represent the short and long edge lengths, respectively), and (b) an aggregate
of many atoms. [Figure (b) from W. G. Moffatt, G. W. Pearsall, and J. Wulff, The Structure
and Properties of Materials, Vol. I, Structure, p. 51. Copyright © 1964 by John Wiley &
Sons, New York. Reprinted by permission of John Wiley & Sons, Inc.]
Crystal Systems and
Unit Cells for Metals
Chromium, iron, tungsten, as well as several other metals listed in Table 3.1 exhibit
a BCC structure.
Two atoms are associated with each BCC unit cell: the equivalent of one atom
from the eight corners, each of which is shared among eight unit cells, and the single center atom, which is wholly contained within its cell. In addition, corner and
center atom positions are equivalent. The coordination number for the BCC crystal structure is 8; each center atom has as nearest neighbors its eight corner atoms.
Since the coordination number is less for BCC than FCC, so also is the atomic packing factor for BCC lower—0.68 versus 0.74.
The Hexagonal Close-Packed Cr ystal Structure
hexagonal closepacked (HCP)
Crystal Systems and
Unit Cells for Metals
Not all metals have unit cells with cubic symmetry; the final common metallic
crystal structure to be discussed has a unit cell that is hexagonal. Figure 3.3a shows
a reduced-sphere unit cell for this structure, which is termed hexagonal closepacked (HCP); an assemblage of several HCP unit cells is presented in Figure
3.3b.1 The top and bottom faces of the unit cell consist of six atoms that form
regular hexagons and surround a single atom in the center. Another plane that
provides three additional atoms to the unit cell is situated between the top and
bottom planes. The atoms in this midplane have as nearest neighbors atoms in
both of the adjacent two planes. The equivalent of six atoms is contained in each
unit cell; one-sixth of each of the 12 top and bottom face corner atoms, one-half
of each of the 2 center face atoms, and all 3 midplane interior atoms. If a and c
represent, respectively, the short and long unit cell dimensions of Figure 3.3a, the
ca ratio should be 1.633; however, for some HCP metals this ratio deviates from
the ideal value.
1
Alternatively, the unit cell for HCP may be specified in terms of the parallelepiped
defined by atoms labeled A through H in Figure 3.3a. As such, the atom denoted J
lies within the unit cell interior.
1496T_c03_38-79 11/11/05 19:05 Page 44
REVISED PAGES
44 • Chapter 3 / The Structure of Crystalline Solids
The coordination number and the atomic packing factor for the HCP crystal
structure are the same as for FCC: 12 and 0.74, respectively. The HCP metals include
cadmium, magnesium, titanium, and zinc; some of these are listed in Table 3.1.
EXAMPLE PROBLEM 3.1
Determination of FCC Unit Cell Volume
Calculate the volume of an FCC unit cell in terms of the atomic radius R.
Solution
In the FCC unit cell illustrated,
R
a
4R
a
the atoms touch one another across a face-diagonal the length of which is 4R.
Since the unit cell is a cube, its volume is a3, where a is the cell edge length.
From the right triangle on the face,
a2 a2 14R2 2
or, solving for a,
a 2R 12
(3.1)
The FCC unit cell volume VC may be computed from
VC a3 12R122 3 16R3 12
(3.4)
EXAMPLE PROBLEM 3.2
Computation of the Atomic Packing Factor for FCC
Show that the atomic packing factor for the FCC crystal structure is 0.74.
Solution
The APF is defined as the fraction of solid sphere volume in a unit cell, or
APF
VS
volume of atoms in a unit cell
total unit cell volume
VC
Both the total atom and unit cell volumes may be calculated in terms of the
atomic radius R. The volume for a sphere is 43 pR3, and since there are four
1496T_c03_38-79 11/11/05 19:05 Page 45
REVISED PAGES
3.5 Density Computations • 45
atoms per FCC unit cell, the total FCC atom (or sphere) volume is
VS 142 43 pR3 163 pR3
From Example Problem 3.1, the total unit cell volume is
VC 16R3 12
Therefore, the atomic packing factor is
APF
1 163 2 pR3
VS
0.74
VC
16R3 12
3.5 DENSITY COMPUTATIONS
A knowledge of the crystal structure of a metallic solid permits computation of its
theoretical density r through the relationship
Theoretical density
for metals
r
nA
VC NA
(3.5)
where
n number of atoms associated with each unit cell
A atomic weight
VC volume of the unit cell
NA Avogadro’s number 16.023 1023 atoms/mol2
EXAMPLE PROBLEM 3.3
Theoretical Density Computation for Copper
Copper has an atomic radius of 0.128 nm, an FCC crystal structure, and an
atomic weight of 63.5 g/mol. Compute its theoretical density and compare the
answer with its measured density.
Solution
Equation 3.5 is employed in the solution of this problem. Since the crystal
structure is FCC, n, the number of atoms per unit cell, is 4. Furthermore, the
atomic weight ACu is given as 63.5 g/mol. The unit cell volume VC for FCC was
determined in Example Problem 3.1 as 16R3 12, where R, the atomic radius,
is 0.128 nm.
Substitution for the various parameters into Equation 3.5 yields
r
nACu
nACu
VCNA
116R3 122 NA
14 atoms/unit cell2163.5 g/mol2
3161211.28 108 cm2 3/unit cell4 16.023 1023 atoms/mol2
8.89 g/cm3
The literature value for the density of copper is 8.94 g/cm3, which is in very
close agreement with the foregoing result.
1496T_c03_38-79 11/11/05 19:05 Page 46
REVISED PAGES
46 • Chapter 3 / The Structure of Crystalline Solids
3.6 POLYMORPHISM AND ALLOTROPY
polymorphism
allotropy
Some metals, as well as nonmetals, may have more than one crystal structure, a phenomenon known as polymorphism. When found in elemental solids, the condition
is often termed allotropy. The prevailing crystal structure depends on both the temperature and the external pressure. One familiar example is found in carbon:
graphite is the stable polymorph at ambient conditions, whereas diamond is formed
at extremely high pressures. Also, pure iron has a BCC crystal structure at room
temperature, which changes to FCC iron at 912C (1674F). Most often a modification of the density and other physical properties accompanies a polymorphic
transformation.
3.7 CRYSTAL SYSTEMS
Crystal Systems and
Unit Cells for Metals
lattice parameters
crystal system
Since there are many different possible crystal structures, it is sometimes convenient to divide them into groups according to unit cell configurations and/or atomic
arrangements. One such scheme is based on the unit cell geometry, that is, the shape
of the appropriate unit cell parallelepiped without regard to the atomic positions
in the cell. Within this framework, an x, y, z coordinate system is established with
its origin at one of the unit cell corners; each of the x, y, and z axes coincides with
one of the three parallelepiped edges that extend from this corner, as illustrated in
Figure 3.4. The unit cell geometry is completely defined in terms of six parameters:
the three edge lengths a, b, and c, and the three interaxial angles a, b, and g. These
are indicated in Figure 3.4, and are sometimes termed the lattice parameters of a
crystal structure.
On this basis there are seven different possible combinations of a, b, and c, and
a, b, and g, each of which represents a distinct crystal system. These seven crystal
systems are cubic, tetragonal, hexagonal, orthorhombic, rhombohedral,2 monoclinic,
and triclinic. The lattice parameter relationships and unit cell sketches for each are
represented in Table 3.2.The cubic system, for which a b c and a b g 90,
has the greatest degree of symmetry. Least symmetry is displayed by the triclinic
system, since a b c and a b g.
From the discussion of metallic crystal structures, it should be apparent that
both FCC and BCC structures belong to the cubic crystal system, whereas HCP
falls within hexagonal. The conventional hexagonal unit cell really consists of three
parallelepipeds situated as shown in Table 3.2.
z
c
Figure 3.4 A unit cell with x, y, and z coordinate axes,
showing axial lengths (a, b, and c) and interaxial angles
1a, b, and g2 .
␣

y
␥
a
b
x
2
Also called trigonal.
1496T_c03_38-79 11/11/05 19:05 Page 47
REVISED PAGES
3.7 Crystal Systems • 47
Table 3.2 Lattice Parameter Relationships and Figures Showing Unit Cell
Geometries for the Seven Crystal Systems
Crystal System
Cubic
Axial
Relationships
abc
Interaxial Angles
Unit Cell Geometry
a b g 90
a a
Hexagonal
abc
a b 90, g 120
a
c
a
Tetragonal
abc
a b g 90
c
a
Rhombohedral
(Trigonal)
abc
Orthorhombic
abc
a b g 90
a b g 90
aa
a
␣
a
c
a
Monoclinic
abc
a
a
a g 90 b
b
c
a

b
Triclinic
abc
a b g 90
c
␣

␥
b
a
1496T_c03_38-79 11/11/05 19:05 Page 48
REVISED PAGES
48 • Chapter 3 / The Structure of Crystalline Solids
MATERIAL OF IMPORTANCE
Tin (Its Allotropic Transformation)
A
nother common metal that experiences an allotropic change is tin. White (or b) tin, having a body-centered tetragonal crystal structure at
room temperature, transforms, at 13.2C (55.8F),
to gray (or a) tin, which has a crystal structure similar to diamond (i.e., the diamond cubic crystal
structure); this transformation is represented
schematically as follows:
13.2°C
Cooling
White () tin
The rate at which this change takes place is extremely slow; however, the lower the temperature
(below 13.2C) the faster the rate. Accompanying
this white tin-to-gray tin transformation is an increase in volume (27 percent), and, accordingly, a
decrease in density (from 7.30 g/cm3 to 5.77 g/cm3).
Consequently, this volume expansion results in the
disintegration of the white tin metal into a coarse
powder of the gray allotrope. For normal subambient temperatures, there is no need to worry
about this disintegration process for tin products,
due to the very slow rate at which the transformation occurs.
This white-to-gray-tin transition produced
some rather dramatic results in 1850 in Russia.The
winter that year was particularly cold, and record
low temperatures persisted for extended periods
of time.The uniforms of some Russian soldiers had
tin buttons, many of which crumbled due to these
extreme cold conditions, as did also many of the
tin church organ pipes. This problem came to be
known as the “tin disease.”
Gray (␣) tin
Specimen of white tin (left). Another specimen
disintegrated upon transforming to gray tin (right)
after it was cooled to and held at a temperature
below 13.2C for an extended period of time.
(Photograph courtesy of Professor Bill Plumbridge,
Department of Materials Engineering, The Open
University, Milton Keynes, England.)
1496T_c03_38-79 11/11/05 19:05 Page 49
REVISED PAGES
3.8 Point Coordinates • 49
Concept Check 3.1
What is the difference between crystal structure and crystal system?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
It is important to note that many of the principles and concepts addressed
in previous discussions in this chapter also apply to crystalline ceramic and polymeric systems (Chapters 12 and 14). For example, crystal structures are most often described in terms of unit cells, which are normally more complex than those
for FCC, BCC, and HCP. In addition, for these other systems, we are often interested in determining atomic packing factors and densities, using modified
forms of Equations 3.2 and 3.5. Furthermore, according to unit cell geometry,
crystal structures of these other material types are grouped within the seven crystal systems.
C r ys t a l l o g r a p h i c Po i n t s ,
Directions, and Planes
When dealing with crystalline materials, it often becomes necessary to specify a particular point within a unit cell, a crystallographic direction, or some crystallographic
plane of atoms. Labeling conventions have been established in which three numbers or indices are used to designate point locations, directions, and planes. The basis
for determining index values is the unit cell, with a right-handed coordinate system
consisting of three (x, y, and z) axes situated at one of the corners and coinciding
with the unit cell edges, as shown in Figure 3.4. For some crystal systems—namely,
hexagonal, rhombohedral, monoclinic, and triclinic—the three axes are not mutually perpendicular, as in the familiar Cartesian coordinate scheme.
3.8 POINT COORDINATES
The position of any point located within a unit cell may be specified in terms of its
coordinates as fractional multiples of the unit cell edge lengths (i.e., in terms of a,
b, and c). To illustrate, consider the unit cell and the point P situated therein as shown
in Figure 3.5. We specify the position of P in terms of the generalized coordinates
z
Figure 3.5 The manner in which the q, r,
and s coordinates at point P within the
unit cell are determined. The q coordinate
(which is a fraction) corresponds to the
distance qa along the x axis, where a is the
unit cell edge length. The respective r and
s coordinates for the y and z axes are
determined similarly.
b
a
P
qrs
c
sc
y
qa
rb
x
1496T_c03_38-79 11/11/05 19:05 Page 50
REVISED PAGES
50 • Chapter 3 / The Structure of Crystalline Solids
q, r, and s where q is some fractional length of a along the x axis, r is some fractional
length of b along the y axis, and similarly for s. Thus, the position of P is designated
using coordinates q r s with values that are less than or equal to unity. Furthermore,
we have chosen not to separate these coordinates by commas or any other punctuation marks (which is the normal convention).
EXAMPLE PROBLEM 3.4
Location of Point Having Specified Coordinates
For the unit cell shown in the accompanying sketch (a), locate the point having coordinates 14 1 12.
z
z
0.46 nm
m
8n
0.4
1 1
1
4 2
P
0.20 nm
0.12 nm M
N
O
0.46 nm
0.40 nm
y
x
(a)
x
y
(b)
Solution
From sketch (a), edge lengths for this unit cell are as follows: a 0.48 nm,
b 0.46 nm, and c 0.40 nm. Furthermore, in light of the above discussion,
fractional lengths are q 14, r 1, and s 12. Therefore, first we move from
the origin of the unit cell (point M) qa 14 (0.48 nm) 0.12 nm units along
the x axis (to point N), as shown in the (b) sketch. Similarly, we proceed
rb (1)(0.46 nm) 0.46 nm parallel to the y axis, from point N to point O.
Finally, we move from this position, sc 12 (0.40 nm) 0.20 nm units parallel
to the z axis to point P as noted again in sketch (b). This point P then corresponds to the 14 1 12 point coordinates.
EXAMPLE PROBLEM 3.5
Specification of Point Coordinates
Specify point coordinates for all atom positions for a BCC unit cell.
Solution
For the BCC unit cell of Figure 3.2, atom position coordinates correspond to
the locations of the centers of all atoms in the unit cell—that is, the eight
corner atoms and single center atom. These positions are noted (and also numbered) in the following figure.
1496T_c03_38-79 11/11/05 19:05 Page 51
REVISED PAGES
3.9 Crystallographic Directions • 51
z
a
a
9
6
8
7
5
a
4
1
2
y
3
x
Point coordinates for position number 1 are 0 0 0; this position is located
at the origin of the coordinate system, and, as such, the fractional unit cell
edge lengths along the x, y, and z axes are, respectively, 0a, 0a, and 0a. Furthermore, for position number 2, since it lies one unit cell edge length along
the x axis, its fractional edge lengths are a, 0a, and 0a, respectively, which yield
point coordinates of 1 0 0. The following table presents fractional unit cell
lengths along the x, y, and z axes, and their corresponding point coordinates
for each of the nine points in the above figure.
Point
Number
1
2
3
4
5
6
7
8
9
Fractional Lengths
x axis
y axis
z axis
0
1
1
0
0
0
1
1
0
0
0
0
1
2
1
2
1
2
0
1
1
0
0
0
1
1
1
1
1
1
Point
Coordinates
0
1
1
0
0
0
1
1
0
0
0
0
111
222
0
1
1
0
0
0
1
1
1
1
1
1
3.9 CRYSTALLOGRAPHIC DIRECTIONS
A crystallographic direction is defined as a line between two points, or a vector.
The following steps are utilized in the determination of the three directional indices:
Crystallographic
Directions
1. A vector of convenient length is positioned such that it passes through the
origin of the coordinate system. Any vector may be translated throughout the
crystal lattice without alteration, if parallelism is maintained.
2. The length of the vector projection on each of the three axes is determined;
these are measured in terms of the unit cell dimensions a, b, and c.
3. These three numbers are multiplied or divided by a common factor to reduce
them to the smallest integer values.
4. The three indices, not separated by commas, are enclosed in square brackets,
thus: [uvw]. The u, v, and w integers correspond to the reduced projections
along the x, y, and z axes, respectively.
1496T_c03_38-79 11/11/05 19:05 Page 52
REVISED PAGES
52 • Chapter 3 / The Structure of Crystalline Solids
z
Figure 3.6 The [100], [110], and [111] directions within a
unit cell.
[111]
y
[110]
[100]
x
For each of the three axes, there will exist both positive and negative coordinates. Thus negative indices are also possible, which are represented by a bar over
the appropriate index. For example, the [111] direction would have a component in
the y direction. Also, changing the signs of all indices produces an antiparallel
direction; that is, [111] is directly opposite to [111]. If more than one direction (or
plane) is to be specified for a particular crystal structure, it is imperative for the
maintaining of consistency that a positive–negative convention, once established,
not be changed.
The [100], [110], and [111] directions are common ones; they are drawn in the
unit cell shown in Figure 3.6.
EXAMPLE PROBLEM 3.6
Determination of Directional Indices
Determine the indices for the direction shown in the accompanying figure.
z
Projection on
x axis (a/2)
Projection on
y axis (b)
y
c
a
b
x
Solution
The vector, as drawn, passes through the origin of the coordinate system, and
therefore no translation is necessary. Projections of this vector onto the x, y,
and z axes are, respectively, a 2, b, and 0c, which become 12, 1, and 0 in terms
of the unit cell parameters (i.e., when the a, b, and c are dropped). Reduction
of these numbers to the lowest set of integers is accompanied by multiplication of each by the factor 2. This yields the integers 1, 2, and 0, which are then
enclosed in brackets as [120].
1496T_c03_38-79 11/11/05 19:05 Page 53
REVISED PAGES
3.9 Crystallographic Directions • 53
This procedure may be summarized as follows:
Projections
Projections (in terms of a, b, and c)
Reduction
Enclosure
x
y
z
a2
b
1
2
[120]
0c
0
0
1
2
1
EXAMPLE PROBLEM 3.7
Construction of Specified Cr ystallographic Direction
Draw a [110] direction within a cubic unit cell.
Solution
First construct an appropriate unit cell and coordinate axes system. In the
accompanying figure the unit cell is cubic, and the origin of the coordinate
system, point O, is located at one of the cube corners.
z
a
O
–y
[110] Direction
P
–a
+y
a
a
a
x
This problem is solved by reversing the procedure of the preceding example.
For this [110] direction, the projections along the x, y, and z axes are a, a,
and 0a, respectively. This direction is defined by a vector passing from the origin to point P, which is located by first moving along the x axis a units, and
from this position, parallel to the y axis a units, as indicated in the figure.
There is no z component to the vector, since the z projection is zero.
For some crystal structures, several nonparallel directions with different indices
are actually equivalent; this means that the spacing of atoms along each direction
is the same. For example, in cubic crystals, all the directions represented by the following indices are equivalent: [100], [100], [010], [010], [001], and [001]. As a convenience, equivalent directions are grouped together into a family, which are enclosed
in angle brackets, thus: 81009. Furthermore, directions in cubic crystals having the
same indices without regard to order or sign, for example, [123] and [213], are equivalent. This is, in general, not true for other crystal systems. For example, for crystals
of tetragonal symmetry, [100] and [010] directions are equivalent, whereas [100] and
[001] are not.
1496T_c03_38-79 11/11/05 21:57 Page 54
REVISED PAGES
54 • Chapter 3 / The Structure of Crystalline Solids
z
Figure 3.7 Coordinate axis system for a hexagonal unit cell
(Miller–Bravais scheme).
a2
a3
120°
a1
Hexagonal Cr ystals
A problem arises for crystals having hexagonal symmetry in that some crystallographic equivalent directions will not have the same set of indices. This is circumvented by utilizing a four-axis, or Miller–Bravais, coordinate system as shown in
Figure 3.7. The three a1, a2, and a3 axes are all contained within a single plane (called
the basal plane) and are at 120 angles to one another. The z axis is perpendicular
to this basal plane. Directional indices, which are obtained as described above, will
be denoted by four indices, as [uvtw]; by convention, the first three indices pertain
to projections along the respective a1, a2, and a3 axes in the basal plane.
Conversion from the three-index system to the four-index system,
3u¿v¿w¿ 4 ¡ 3uvtw4
is accomplished by the following formulas:
u
1
12u¿ v¿2
3
(3.6a)
v
1
12v¿ u¿2
3
(3.6b)
t 1u v2
w w¿
(3.6c)
(3.6d)
where primed indices are associated with the three-index scheme and unprimed
with the new Miller–Bravais four-index system. (Of course, reduction to the lowest
set of integers may be necessary, as discussed above.) For example, the [010] direction becomes [1210]. Several different directions are indicated in the hexagonal unit
cell (Figure 3.8a).
z
z
[0001]
(1010)
(1011)
a2
[1120]
a3
a3
a1
[1100]
(a)
a1
(0001)
(b)
Figure 3.8 For the
hexagonal crystal system,
(a) [0001], [1100], and [1120]
directions, and (b) the
(0001), (1011), and (1010)
planes.
1496T_c03_38-79 11/11/05 19:05 Page 55
REVISED PAGES
3.10 Crystallographic Planes • 55
EXAMPLE PROBLEM 3.8
Determination of Directional Indices for a
Hexagonal Unit Cell
Determine the indices for the direction shown in the hexagonal unit cell of
sketch (a) below.
z
z
H
F
G
E
a2
a2
c
D
B
a3
a3
C
a
a1
(a)
a
A
a1
(b)
Solution
In sketch (b), one of the three parallelepipeds comprising the hexagonal cell
is delineated—its corners are labeled with letters A through H, with the origin
of the a1-a2-a3-z axes coordinate system located at the corner labeled C. We
use this unit cell as a reference for specifying the directional indices. It now becomes necessary to determine projections of the direction vector on the a1, a2,
and z axes. These respective projections are a (a1 axis), a (a2 axis) and c (z axis),
which become 1, 1, and 1 in terms of the unit cell parameters. Thus,
u¿ 1
v¿ 1
w¿ 1
Also, from Equations 3.6a, 3.6b, 3.6c, and 3.6d
u
1
1
1
12u¿ v¿2 3 122112 14
3
3
3
v
1
1
1
12v¿ u¿2 3 122112 14
3
3
3
1
1
2
t 1u v2 a b
3
3
3
w w¿ 1
Multiplication of the above indices by 3 reduces them to the lowest set, which
yields values for u, v, t, and w of 1, 1, 2 and 3, respectively. Hence, the direction shown in the figure is [1123].
3.10 CRYSTALLOGRAPHIC PLANES
Miller indices
The orientations of planes for a crystal structure are represented in a similar manner. Again, the unit cell is the basis, with the three-axis coordinate system as represented in Figure 3.4. In all but the hexagonal crystal system, crystallographic planes
are specified by three Miller indices as (hkl). Any two planes parallel to each other
1496T_c03_38-79 11/11/05 19:05 Page 56
REVISED PAGES
56 • Chapter 3 / The Structure of Crystalline Solids
are equivalent and have identical indices. The procedure employed in determination of the h, k, and l index numbers is as follows:
1. If the plane passes through the selected origin, either another parallel plane
must be constructed within the unit cell by an appropriate translation, or a
new origin must be established at the corner of another unit cell.
2. At this point the crystallographic plane either intersects or parallels each of
the three axes; the length of the planar intercept for each axis is determined
in terms of the lattice parameters a, b, and c.
3. The reciprocals of these numbers are taken. A plane that parallels an axis
may be considered to have an infinite intercept, and, therefore, a zero index.
4. If necessary, these three numbers are changed to the set of smallest integers
by multiplication or division by a common factor.3
5. Finally, the integer indices, not separated by commas, are enclosed within
parentheses, thus: (hkl).
Crystallographic
Planes
An intercept on the negative side of the origin is indicated by a bar or minus sign positioned over the appropriate index. Furthermore, reversing the directions of all indices specifies another plane parallel to, on the opposite side of
and equidistant from, the origin. Several low-index planes are represented in
Figure 3.9.
One interesting and unique characteristic of cubic crystals is that planes and directions having the same indices are perpendicular to one another; however, for
other crystal systems there are no simple geometrical relationships between planes
and directions having the same indices.
EXAMPLE PROBLEM 3.9
Determination of Planar (Miller) Indices
Determine the Miller indices for the plane shown in the accompanying
sketch (a).
z
z
z⬘
c/2
c
y
O
O
O⬘
y
a
(012) Plane
b
x
x
(a)
3
–b
x⬘
(b)
On occasion, index reduction is not carried out (e.g., for x-ray diffraction studies that are
described in Section 3.16); for example, (002) is not reduced to (001). In addition, for ceramic materials, the ionic arrangement for a reduced-index plane may be different from
that for a nonreduced one.
1496T_c03_38-79 11/11/05 19:05 Page 57
REVISED PAGES
3.10 Crystallographic Planes • 57
Solution
Since the plane passes through the selected origin O, a new origin must be
chosen at the corner of an adjacent unit cell, taken as O¿ and shown in sketch
(b). This plane is parallel to the x axis, and the intercept may be taken as q a.
The y and z axes intersections, referenced to the new origin O¿ , are b and
c 2, respectively. Thus, in terms of the lattice parameters a, b, and c, these intersections are q, 1, and 12. The reciprocals of these numbers are 0, 1, and
2; and since all are integers, no further reduction is necessary. Finally, enclosure in parentheses yields 10122.
These steps are briefly summarized below:
Intercepts
Intercepts (in terms of lattice parameters)
Reciprocals
Reductions (unnecessary)
Enclosure
z
x
y
z
qa
q
0
b
1
1
c2
1
2
2
10122
(001) Plane referenced to
the origin at point O
z
(110) Plane referenced to the
origin at point O
y
O
y
O
Other equivalent
(001) planes
x
Other equivalent
(110) planes
x
(a)
z
(b)
(111) Plane referenced to
the origin at point O
y
O
Other equivalent
(111) planes
x
(c)
Figure 3.9 Representations of a series each of (a) (001), (b) (110), and (c) (111)
crystallographic planes.
1496T_c03_38-79 12/20/05 7:38 Page 58
2nd REVISE PAGES
58 • Chapter 3 / The Structure of Crystalline Solids
EXAMPLE PROBLEM 3.10
Construction of Specified Cr ystallographic Plane
Construct a (011) plane within a cubic unit cell.
z
z
e
(011) Plane
c
f
b
O
–y
g
y
y
a
Point of intersection
along y axis
h
b
x
x
(a)
(b)
Solution
To solve this problem, carry out the procedure used in the preceding example in
reverse order. To begin, the indices are removed from the parentheses, and reciprocals are taken, which yields q , 1, and 1. This means that the particular plane
parallels the x axis while intersecting the y and z axes at b and c, respectively,
as indicated in the accompanying sketch (a). This plane has been drawn in sketch
(b). A plane is indicated by lines representing its intersections with the planes
that constitute the faces of the unit cell or their extensions. For example, in this
figure, line ef is the intersection between the (011) plane and the top face of the
unit cell; also, line gh represents the intersection between this same (011) plane
and the plane of the bottom unit cell face extended. Similarly, lines eg and fh are
the intersections between (011) and back and front cell faces, respectively.
Atomic Arrangements
Planar Atomic
Arrangements
The atomic arrangement for a crystallographic plane, which is often of interest, depends on the crystal structure. The (110) atomic planes for FCC and BCC crystal
structures are represented in Figures 3.10 and 3.11; reduced-sphere unit cells are
also included. Note that the atomic packing is different for each case. The circles
represent atoms lying in the crystallographic planes as would be obtained from a
slice taken through the centers of the full-sized hard spheres.
A “family” of planes contains all those planes that are crystallographically
equivalent—that is, having the same atomic packing; and a family is designated by
C
A
B
A
B
C
D
E
F
E
F
D
(a)
(b)
Figure 3.10 (a) Reducedsphere FCC unit cell with
(110) plane. (b) Atomic
packing of an FCC (110)
plane. Corresponding
atom positions from
(a) are indicated.
1496T_c03_38-79 12/20/05 7:38 Page 59
2nd REVISE PAGES
3.10 Crystallographic Planes • 59
Figure 3.11
(a) Reduced-sphere
BCC unit cell with
(110) plane.
(b) Atomic packing
of a BCC (110)
plane. Corresponding
atom positions from
(a) are indicated.
B⬘
A⬘
A⬘
B⬘
C⬘
C⬘
D⬘
E⬘
E⬘
D⬘
(a)
(b)
indices that are enclosed in braces—such as {100}. For example, in cubic crystals the
(111), (111), (111), (111), (111), (111), (111), and (111) planes all belong to the {111}
family. On the other hand, for tetragonal crystal structures, the {100} family would
contain only the (100), (100), (010), and (010) since the (001) and (001) planes are
not crystallographically equivalent. Also, in the cubic system only, planes having
the same indices, irrespective of order and sign, are equivalent. For example, both
(123) and (312) belong to the {123} family.
Hexagonal Cr ystals
For crystals having hexagonal symmetry, it is desirable that equivalent planes have
the same indices; as with directions, this is accomplished by the Miller–Bravais
system shown in Figure 3.7. This convention leads to the four-index (hkil) scheme,
which is favored in most instances, since it more clearly identifies the orientation
of a plane in a hexagonal crystal. There is some redundancy in that i is determined
by the sum of h and k through
i 1h k2
(3.7)
Otherwise the three h, k, and l indices are identical for both indexing systems.
Figure 3.8b presents several of the common planes that are found for crystals having hexagonal symmetry.
EXAMPLE PROBLEM 3.11
Determination of Miller–Bravais Indices for a Plane
Within a Hexagonal Unit Cell
Determine the Miller–Bravais indices for the plane shown in the hexagonal
unit cell.
z
H
F
G
E
a2
c
D
a3
a
B
C
a
A
a1
1496T_c03_38-79 11/11/05 19:05 Page 60
REVISED PAGES
60 • Chapter 3 / The Structure of Crystalline Solids
Solution
To determine these Miller–Bravais indices, consider the plane in the figure
referenced to the parallelepiped labeled with the letters A through H at its
corners. This plane intersects the a1 axis at a distance a from the origin of the
a1-a2-a3-z coordinate axes system (point C). Furthermore, its intersections with
the a2 and z axes are a and c, respectively. Therefore, in terms of the lattice
parameters, these intersections are 1, 1, and 1. Furthermore, the reciprocals
of these numbers are also 1, 1, and 1. Hence
h1
k 1
l1
and, from Equation 3.7
i 1h k2
11 12 0
Therefore the (hkil) indices are (1101).
Notice that the third index is zero (i.e., its reciprocal q ), which means
that this plane parallels the a3 axis. Upon inspection of the above figure, it
may be noted that this is indeed the case.
3.11 LINEAR AND PLANAR DENSITIES
The two previous sections discussed the equivalency of nonparallel crystallographic
directions and planes. Directional equivalency is related to linear density in the sense
that, for a particular material, equivalent directions have identical linear densities.
The corresponding parameter for crystallographic planes is planar density, and
planes having the same planar density values are also equivalent.
Linear density (LD) is defined as the number of atoms per unit length whose
centers lie on the direction vector for a specific crystallographic direction; that is,
LD
number of atoms centered on direction vector
length of direction vector
(3.8)
Of course, the units of linear density are reciprocal length (e.g., nm1, m1).
For example, let us determine the linear density of the [110] direction for the
FCC crystal structure. An FCC unit cell (reduced sphere) and the [110] direction
therein are shown in Figure 3.12a. Represented in Figure 3.12b are those five atoms
that lie on the bottom face of this unit cell; here the [110] direction vector passes
from the center of atom X, through atom Y, and finally to the center of atom Z.
With regard to the numbers of atoms, it is necessary to take into account the sharing
of atoms with adjacent unit cells (as discussed in Section 3.4 relative to atomic packing factor computations). Each of the X and Z corner atoms are also shared with
one other adjacent unit cell along this [110] direction (i.e., one-half of each of these
atoms belongs to the unit cell being considered), while atom Y lies entirely within
the unit cell. Thus, there is an equivalence of two atoms along the [110] direction
vector in the unit cell. Now, the direction vector length is equal to 4R (Figure 3.12b);
thus, from Equation 3.8, the [110] linear density for FCC is
LD110
2 atoms
1
4R
2R
(3.9)
1496T_c03_38-79 12/20/05 7:38 Page 61
2nd REVISE PAGES
3.12 Close-Packed Crystal Structures • 61
R
X
Y
X
Z
Y
[110]
Z
(a)
Figure 3.12 (a) Reducedsphere FCC unit cell with
the [110] direction
indicated. (b) The bottom
face-plane of the FCC unit
cell in (a) on which is
shown the atomic spacing
in the [110] direction,
through atoms labeled X,
Y, and Z.
(b)
In an analogous manner, planar density (PD) is taken as the number of atoms
per unit area that are centered on a particular crystallographic plane, or
PD
number of atoms centered on a plane
area of plane
(3.10)
The units for planar density are reciprocal area (e.g., nm2, m2).
For example, consider the section of a (110) plane within an FCC unit cell as
represented in Figures 3.10a and 3.10b. Although six atoms have centers that lie on
this plane (Figure 3.10b), only one-quarter of each of atoms A, C, D, and F, and onehalf of atoms B and E, for a total equivalence of just 2 atoms are on that plane.
Furthermore, the area of this rectangular section is equal to the product of its length
and width. From Figure 3.10b, the length (horizontal dimension) is equal to 4R,
whereas the width (vertical dimension) is equal to 2R12, since it corresponds to
the FCC unit cell edge length (Equation 3.1). Thus, the area of this planar region
is 14R212R 122 8R2 12, and the planar density is determined as follows:
PD110
2 atoms
1
2
2
8R 12
4R 12
(3.11)
Linear and planar densities are important considerations relative to the process of
slip—that is, the mechanism by which metals plastically deform (Section 7.4). Slip
occurs on the most densely packed crystallographic planes and, in those planes,
along directions having the greatest atomic packing.
3.12 CLOSE-PACKED CRYSTAL STRUCTURES
Close-Packed
Structures (Metals)
You may remember from the discussion on metallic crystal structures that both
face-centered cubic and hexagonal close-packed crystal structures have atomic
packing factors of 0.74, which is the most efficient packing of equal-sized spheres
or atoms. In addition to unit cell representations, these two crystal structures may
be described in terms of close-packed planes of atoms (i.e., planes having a maximum atom or sphere-packing density); a portion of one such plane is illustrated in
Figure 3.13a. Both crystal structures may be generated by the stacking of these
close-packed planes on top of one another; the difference between the two structures lies in the stacking sequence.
Let the centers of all the atoms in one close-packed plane be labeled A. Associated with this plane are two sets of equivalent triangular depressions formed by
three adjacent atoms, into which the next close-packed plane of atoms may rest.
Those having the triangle vertex pointing up are arbitrarily designated as B positions, while the remaining depressions are those with the down vertices, which are
marked C in Figure 3.13a.
1496T_c03_38-79 12/20/05 7:38 Page 62
2nd REVISE PAGES
62 • Chapter 3 / The Structure of Crystalline Solids
B
C
B
C
B
C
B
C
B
C
C
B
B
C
C
C
(a)
C
C
C
C
C
C
C
C
Figure 3.13 (a) A portion
of a close-packed plane of
atoms; A, B, and C positions
are indicated. (b) The AB
stacking sequence for closepacked atomic planes.
(Adapted from W. G.
Moffatt, G. W. Pearsall, and
J. Wulff, The Structure and
Properties of Materials, Vol.
I, Structure, p. 50. Copyright
© 1964 by John Wiley &
Sons, New York. Reprinted
by permission of John Wiley
& Sons, Inc.)
C
(b)
A second close-packed plane may be positioned with the centers of its
atoms over either B or C sites; at this point both are equivalent. Suppose that
the B positions are arbitrarily chosen; the stacking sequence is termed AB, which
is illustrated in Figure 3.13b. The real distinction between FCC and HCP lies in
where the third close-packed layer is positioned. For HCP, the centers of this
layer are aligned directly above the original A positions. This stacking sequence,
ABABAB . . . , is repeated over and over. Of course, the ACACAC . . . arrangement would be equivalent. These close-packed planes for HCP are (0001)-type
planes, and the correspondence between this and the unit cell representation is
shown in Figure 3.14.
For the face-centered crystal structure, the centers of the third plane are situated
over the C sites of the first plane (Figure 3.15a). This yields an ABCABCABC . . .
A
B
A
B
A
Figure 3.14 Close-packed plane stacking
sequence for hexagonal close-packed.
(Adapted from W. G. Moffatt, G. W. Pearsall,
and J. Wulff, The Structure and Properties of
Materials, Vol. I, Structure, p. 51. Copyright ©
1964 by John Wiley & Sons, New York.
Reprinted by permission of John Wiley & Sons,
Inc.)
1496T_c03_38-79 12/20/05 7:38 Page 63
2nd REVISE PAGES
3.13 Single Crystals • 63
B
A
C
B
A
C
B
A
(a)
(b)
Figure 3.15 (a) Close-packed stacking sequence for face-centered cubic.
(b) A corner has been removed to show the relation between the stacking of
close-packed planes of atoms and the FCC crystal structure; the heavy triangle
outlines a (111) plane. [Figure (b) from W. G. Moffatt, G. W. Pearsall, and J. Wulff,
The Structure and Properties of Materials, Vol. I, Structure, p. 51. Copyright ©
1964 by John Wiley & Sons, New York. Reprinted by permission of John Wiley &
Sons, Inc.]
stacking sequence; that is, the atomic alignment repeats every third plane. It is more
difficult to correlate the stacking of close-packed planes to the FCC unit cell.
However, this relationship is demonstrated in Figure 3.15b. These planes are of
the (111) type; an FCC unit cell is outlined on the upper left-hand front face of
Figure 3.15b, in order to provide a perspective. The significance of these FCC and
HCP close-packed planes will become apparent in Chapter 7.
The concepts detailed in the previous four sections also relate to crystalline ceramic and polymeric materials, which are discussed in Chapters 12 and 14. We may
specify crystallographic planes and directions in terms of directional and Miller indices; furthermore, on occasion it is important to ascertain the atomic and ionic
arrangements of particular crystallographic planes. Also, the crystal structures of a
number of ceramic materials may be generated by the stacking of close-packed
planes of ions (Section 12.2).
C r ys t a l l i n e a n d
N o n c r ys t a l l i n e M a t e r i a l s
3.13 SINGLE CRYSTALS
single crystal
For a crystalline solid, when the periodic and repeated arrangement of atoms is perfect or extends throughout the entirety of the specimen without interruption, the
result is a single crystal. All unit cells interlock in the same way and have the same
1496T_c03_38-79 11/11/05 19:05 Page 64
REVISED PAGES
64 • Chapter 3 / The Structure of Crystalline Solids
Figure 3.16
Photograph of a
garnet single crystal
that was found in
Tongbei, Fujian
Province, China.
(Photograph courtesy
of Irocks.com, Megan
Foreman photo.)
orientation. Single crystals exist in nature, but they may also be produced artificially. They are ordinarily difficult to grow, because the environment must be carefully controlled.
If the extremities of a single crystal are permitted to grow without any external constraint, the crystal will assume a regular geometric shape having flat faces,
as with some of the gem stones; the shape is indicative of the crystal structure. A
photograph of a garnet single crystal is shown in Figure 3.16. Within the past few
years, single crystals have become extremely important in many of our modern technologies, in particular electronic microcircuits, which employ single crystals of silicon and other semiconductors.
3.14 POLYCRYSTALLINE MATERIALS
grain
polycrystalline
grain boundary
Most crystalline solids are composed of a collection of many small crystals or grains;
such materials are termed polycrystalline. Various stages in the solidification of a
polycrystalline specimen are represented schematically in Figure 3.17. Initially, small
crystals or nuclei form at various positions. These have random crystallographic orientations, as indicated by the square grids. The small grains grow by the successive
addition from the surrounding liquid of atoms to the structure of each. The extremities of adjacent grains impinge on one another as the solidification process approaches completion. As indicated in Figure 3.17, the crystallographic orientation
varies from grain to grain. Also, there exists some atomic mismatch within the region where two grains meet; this area, called a grain boundary, is discussed in more
detail in Section 4.6.
3.15 ANISOTROPY
anisotropy
isotropic
The physical properties of single crystals of some substances depend on the crystallographic direction in which measurements are taken. For example, the elastic
modulus, the electrical conductivity, and the index of refraction may have different
values in the [100] and [111] directions. This directionality of properties is termed
anisotropy, and it is associated with the variance of atomic or ionic spacing with
crystallographic direction. Substances in which measured properties are independent of the direction of measurement are isotropic. The extent and magnitude of
1496T_c03_38-79 11/11/05 19:05 Page 65
REVISED PAGES
3.15 Anisotropy • 65
(a)
(b)
(c)
(d)
Figure 3.17 Schematic diagrams of the various stages in the solidification of a
polycrystalline material; the square grids depict unit cells. (a) Small crystallite nuclei.
(b) Growth of the crystallites; the obstruction of some grains that are adjacent to one
another is also shown. (c) Upon completion of solidification, grains having irregular
shapes have formed. (d) The grain structure as it would appear under the microscope;
dark lines are the grain boundaries. (Adapted from W. Rosenhain, An Introduction to
the Study of Physical Metallurgy, 2nd edition, Constable & Company Ltd., London,
1915.)
anisotropic effects in crystalline materials are functions of the symmetry of the
crystal structure; the degree of anisotropy increases with decreasing structural
symmetry—triclinic structures normally are highly anisotropic. The modulus of elasticity values at [100], [110], and [111] orientations for several materials are presented
in Table 3.3.
For many polycrystalline materials, the crystallographic orientations of the individual grains are totally random. Under these circumstances, even though each grain
may be anisotropic, a specimen composed of the grain aggregate behaves isotropically.
Also, the magnitude of a measured property represents some average of the directional
values. Sometimes the grains in polycrystalline materials have a preferential crystallographic orientation, in which case the material is said to have a “texture.”
The magnetic properties of some iron alloys used in transformer cores are
anisotropic—that is, grains (or single crystals) magnetize in a 81009-type direction
easier than any other crystallographic direction. Energy losses in transformer cores
are minimized by utilizing polycrystalline sheets of these alloys into which have
been introduced a “magnetic texture”: most of the grains in each sheet have a
1496T_c03_38-79 11/11/05 19:05 Page 66
REVISED PAGES
66 • Chapter 3 / The Structure of Crystalline Solids
Table 3.3 Modulus of Elasticity Values for
Several Metals at Various Crystallographic Orientations
Modulus of Elasticity (GPa)
Metal
[100]
[110]
[111]
Aluminum
Copper
Iron
Tungsten
63.7
66.7
125.0
384.6
72.6
130.3
210.5
384.6
76.1
191.1
272.7
384.6
Source: R. W. Hertzberg, Deformation and Fracture
Mechanics of Engineering Materials, 3rd edition.
Copyright © 1989 by John Wiley & Sons, New York.
Reprinted by permission of John Wiley & Sons, Inc.
81009-type crystallographic direction that is aligned (or almost aligned) in the same
direction, which direction is oriented parallel to the direction of the applied magnetic field. Magnetic textures for iron alloys are discussed in detail in the Materials of Importance piece in Chapter 20 following Section 20.9.
3.16 X-RAY DIFFRACTION: DETERMINATION OF
CRYSTAL STRUCTURES
Historically, much of our understanding regarding the atomic and molecular
arrangements in solids has resulted from x-ray diffraction investigations; furthermore, x-rays are still very important in developing new materials. We will now give
a brief overview of the diffraction phenomenon and how, using x-rays, atomic
interplanar distances and crystal structures are deduced.
The Diffraction Phenomenon
diffraction
Diffraction occurs when a wave encounters a series of regularly spaced obstacles
that (1) are capable of scattering the wave, and (2) have spacings that are comparable in magnitude to the wavelength. Furthermore, diffraction is a consequence of
specific phase relationships established between two or more waves that have been
scattered by the obstacles.
Consider waves 1 and 2 in Figure 3.18a which have the same wavelength 1l2
and are in phase at point O–O¿ . Now let us suppose that both waves are scattered
in such a way that they traverse different paths. The phase relationship between
the scattered waves, which will depend upon the difference in path length, is important. One possibility results when this path length difference is an integral number of wavelengths. As noted in Figure 3.18a, these scattered waves (now labeled
1¿ and 2¿ ) are still in phase. They are said to mutually reinforce (or constructively
interfere with) one another; and, when amplitudes are added, the wave shown on
the right side of the figure results. This is a manifestation of diffraction, and we refer to a diffracted beam as one composed of a large number of scattered waves that
mutually reinforce one another.
Other phase relationships are possible between scattered waves that will not
lead to this mutual reinforcement. The other extreme is that demonstrated in
Figure 3.18b, wherein the path length difference after scattering is some integral
number of half wavelengths. The scattered waves are out of phase—that is, corresponding amplitudes cancel or annul one another, or destructively interfere (i.e., the
1496T_c03_38-79 11/11/05 19:05 Page 67
REVISED PAGES
3.16 X-Ray Diffraction: Determination of Crystal Structures • 67
O
Wave 1
Scattering
event
Wave 1'
Amplitude
A
A
2A
A
+
A
Wave 2
O'
Wave 2'
Position
(a)
P
Wave 3
Scattering
event
A
Amplitude
Figure 3.18
(a) Demonstration
of how two waves
(labeled 1 and 2)
that have the same
wavelength l and
remain in phase
after a scattering
event (waves 1¿ and
2¿ ) constructively
interfere with one
another. The
amplitudes of the
scattered waves add
together in the
resultant wave.
(b) Demonstration
of how two waves
(labeled 3 and 4)
that have the same
wavelength and
become out of phase
after a scattering
event (waves 3¿ and
4¿ ) destructively
interfere with one
another. The
amplitudes of the two
scattered waves
cancel one another.
Wave 3'
A
+
A
A
Wave 4
Wave 4'
P'
Position
(b)
resultant wave has zero amplitude), as indicated on the extreme right side of the
figure. Of course, phase relationships intermediate between these two extremes
exist, resulting in only partial reinforcement.
X-Ray Diffraction and Bragg’s Law
X-rays are a form of electromagnetic radiation that have high energies and short
wavelengths—wavelengths on the order of the atomic spacings for solids. When a
beam of x-rays impinges on a solid material, a portion of this beam will be scattered
in all directions by the electrons associated with each atom or ion that lies within
the beam’s path. Let us now examine the necessary conditions for diffraction of
x-rays by a periodic arrangement of atoms.
Consider the two parallel planes of atoms A–A¿ and B–B¿ in Figure 3.19, which
have the same h, k, and l Miller indices and are separated by the interplanar spacing dhkl. Now assume that a parallel, monochromatic, and coherent (in-phase) beam
of x-rays of wavelength l is incident on these two planes at an angle u. Two rays in
this beam, labeled 1 and 2, are scattered by atoms P and Q. Constructive interference of the scattered rays 1¿ and 2¿ occurs also at an angle u to the planes, if the
path length difference between 1–P–1¿ and 2–Q–2¿ (i.e., SQ QT ) is equal to a
whole number, n, of wavelengths. That is, the condition for diffraction is
nl SQ QT
(3.12)
1496T_c03_38-79 11/11/05 19:05 Page 68
REVISED PAGES
68 • Chapter 3 / The Structure of Crystalline Solids
Figure 3.19
Diffraction of x-rays
by planes of atoms
(A–A¿ and B–B¿ ).
1
Incident
beam
2
A
P
B
S
1'
Diffracted
beam
2'
A'
dhkl
T
B'
Q
Bragg’s law—
relationship among
x-ray wavelength,
interatomic spacing,
and angle of
diffraction for
constructive
interference
Bragg’s law
Interplanar
separation for a
plane having indices
h, k, and l
or
nl dhkl sin u dhkl sin u
2dhkl sin u
(3.13)
Equation 3.13 is known as Bragg’s law; also, n is the order of reflection, which
may be any integer (1, 2, 3, . . . ) consistent with sin u not exceeding unity. Thus, we
have a simple expression relating the x-ray wavelength and interatomic spacing to
the angle of the diffracted beam. If Bragg’s law is not satisfied, then the interference will be nonconstructive in nature so as to yield a very low-intensity diffracted
beam.
The magnitude of the distance between two adjacent and parallel planes of
atoms (i.e., the interplanar spacing dhkl) is a function of the Miller indices (h, k,
and l) as well as the lattice parameter(s). For example, for crystal structures that have
cubic symmetry,
a
(3.14)
dhkl
2h2 k2 l 2
in which a is the lattice parameter (unit cell edge length). Relationships similar to
Equation 3.14, but more complex, exist for the other six crystal systems noted in
Table 3.2.
Bragg’s law, Equation 3.13, is a necessary but not sufficient condition for diffraction by real crystals. It specifies when diffraction will occur for unit cells having atoms positioned only at cell corners. However, atoms situated at other sites
(e.g., face and interior unit cell positions as with FCC and BCC) act as extra scattering centers, which can produce out-of-phase scattering at certain Bragg angles.
The net result is the absence of some diffracted beams that, according to Equation
3.13, should be present. For example, for the BCC crystal structure, h k l must
be even if diffraction is to occur, whereas for FCC, h, k, and l must all be either odd
or even.
1496T_c03_38-79 11/11/05 19:05 Page 69
REVISED PAGES
3.16 X-Ray Diffraction: Determination of Crystal Structures • 69
Figure 3.20 Schematic diagram of an
x-ray diffractometer; T x-ray source,
S specimen, C detector, and O
the axis around which the specimen
and detector rotate.
O
S
0°
T
20°
160
°
2
°
40
14
0°
C
0°
60
°
80°
100°
12
Concept Check 3.2
For cubic crystals, as values of the planar indices h, k, and l increase, does the distance between adjacent and parallel planes (i.e., the interplanar spacing) increase
or decrease? Why?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Diffraction Techniques
One common diffraction technique employs a powdered or polycrystalline specimen consisting of many fine and randomly oriented particles that are exposed to
monochromatic x-radiation. Each powder particle (or grain) is a crystal, and having
a large number of them with random orientations ensures that some particles are
properly oriented such that every possible set of crystallographic planes will be
available for diffraction.
The diffractometer is an apparatus used to determine the angles at which diffraction occurs for powdered specimens; its features are represented schematically
in Figure 3.20. A specimen S in the form of a flat plate is supported so that rotations about the axis labeled O are possible; this axis is perpendicular to the plane
of the page. The monochromatic x-ray beam is generated at point T, and the intensities of diffracted beams are detected with a counter labeled C in the figure. The
specimen, x-ray source, and counter are all coplanar.
The counter is mounted on a movable carriage that may also be rotated about
the O axis; its angular position in terms of 2u is marked on a graduated scale.4 Carriage and specimen are mechanically coupled such that a rotation of the specimen
through u is accompanied by a 2u rotation of the counter; this assures that the incident and reflection angles are maintained equal to one another (Figure 3.20).
4
Note that the symbol u has been used in two different contexts for this discussion. Here,
u represents the angular locations of both x-ray source and counter relative to the specimen surface. Previously (e.g., Equation 3.13), it denoted the angle at which the Bragg
criterion for diffraction is satisfied.
1496T_c03_38-79 11/11/05 19:05 Page 70
REVISED PAGES
70 • Chapter 3 / The Structure of Crystalline Solids
(111)
Intensity
Figure 3.21
Diffraction pattern
for powdered lead.
(Courtesy of Wesley
L. Holman.)
0.0
(311)
(200)
(220)
(222)
(400) (331) (420)
10.0
20.0
30.0
40.0
50.0
60.0
70.0
80.0
90.0
(422)
100.0
Diffraction angle 2
Collimators are incorporated within the beam path to produce a well-defined and
focused beam. Utilization of a filter provides a near-monochromatic beam.
As the counter moves at constant angular velocity, a recorder automatically
plots the diffracted beam intensity (monitored by the counter) as a function of 2u;
2u is termed the diffraction angle, which is measured experimentally. Figure 3.21
shows a diffraction pattern for a powdered specimen of lead. The high-intensity
peaks result when the Bragg diffraction condition is satisfied by some set of crystallographic planes. These peaks are plane-indexed in the figure.
Other powder techniques have been devised wherein diffracted beam intensity and
position are recorded on a photographic film instead of being measured by a counter.
One of the primary uses of x-ray diffractometry is for the determination of crystal structure. The unit cell size and geometry may be resolved from the angular positions of the diffraction peaks, whereas arrangement of atoms within the unit cell
is associated with the relative intensities of these peaks.
X-rays, as well as electron and neutron beams, are also used in other types of
material investigations. For example, crystallographic orientations of single crystals
are possible using x-ray diffraction (or Laue) photographs. In the (a) chapter-opening
photograph for this chapter is shown a photograph that was generated using an incident x-ray beam that was directed on a magnesium crystal; each spot (with the
exception of the darkest one near the center) resulted from an x-ray beam that was
diffracted by a specific set of crystallographic planes. Other uses of x-rays include
qualitative and quantitative chemical identifications and the determination of residual
stresses and crystal size.
EXAMPLE PROBLEM 3.12
Interplanar Spacing and Diffraction Angle Computations
For BCC iron, compute (a) the interplanar spacing, and (b) the diffraction angle for the (220) set of planes. The lattice parameter for Fe is 0.2866 nm. Also,
assume that monochromatic radiation having a wavelength of 0.1790 nm is
used, and the order of reflection is 1.
Solution
(a) The value of the interplanar spacing dhkl is determined using Equation 3.14,
with a 0.2866 nm, and h 2, k 2, and l 0, since we are considering the
(220) planes. Therefore,
dhkl
a
2h k2 l 2
2
0.2866 nm
2122 2 122 2 102 2
0.1013 nm
1496T_c03_38-79 11/11/05 19:05 Page 71
REVISED PAGES
3.17 Noncrystalline Solids • 71
(b) The value of u may now be computed using Equation 3.13, with n 1,
since this is a first-order reflection:
sin u
11210.1790 nm2
nl
0.884
2dhkl
12210.1013 nm2
u sin1 10.8842 62.13
The diffraction angle is 2u, or
2u 122162.132 124.26
3.17 NONCRYSTALLINE SOLIDS
noncrystalline
amorphous
It has been mentioned that noncrystalline solids lack a systematic and regular
arrangement of atoms over relatively large atomic distances. Sometimes such materials are also called amorphous (meaning literally without form), or supercooled
liquids, inasmuch as their atomic structure resembles that of a liquid.
An amorphous condition may be illustrated by comparison of the crystalline
and noncrystalline structures of the ceramic compound silicon dioxide (SiO2), which
may exist in both states. Figures 3.22a and 3.22b present two-dimensional schematic
diagrams for both structures of SiO2. Even though each silicon ion bonds to three
oxygen ions for both states, beyond this, the structure is much more disordered and
irregular for the noncrystalline structure.
Whether a crystalline or amorphous solid forms depends on the ease with which
a random atomic structure in the liquid can transform to an ordered state during
solidification. Amorphous materials, therefore, are characterized by atomic or molecular structures that are relatively complex and become ordered only with some
difficulty. Furthermore, rapidly cooling through the freezing temperature favors the
formation of a noncrystalline solid, since little time is allowed for the ordering
process.
Metals normally form crystalline solids, but some ceramic materials are crystalline,
whereas others, the inorganic glasses, are amorphous. Polymers may be completely
Silicon atom
Oxygen atom
(a)
(b)
Figure 3.22 Two-dimensional schemes of the structure of (a) crystalline silicon dioxide
and (b) noncrystalline silicon dioxide.
1496T_c03_38-79 11/11/05 19:05 Page 72
REVISED PAGES
72 • Chapter 3 / The Structure of Crystalline Solids
noncrystalline and semicrystalline consisting of varying degrees of crystallinity.
More about the structure and properties of amorphous ceramics and polymers is
contained in Chapters 12 and 14.
Concept Check 3.3
Do noncrystalline materials display the phenomenon of allotropy (or polymorphism)? Why or why not?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
SUMMARY
Fundamental Concepts
Unit Cells
Atoms in crystalline solids are positioned in orderly and repeated patterns that are
in contrast to the random and disordered atomic distribution found in noncrystalline or amorphous materials. Atoms may be represented as solid spheres, and,
for crystalline solids, crystal structure is just the spatial arrangement of these spheres.
The various crystal structures are specified in terms of parallelepiped unit cells,
which are characterized by geometry and atom positions within.
Metallic Crystal Structures
Most common metals exist in at least one of three relatively simple crystal structures: face-centered cubic (FCC), body-centered cubic (BCC), and hexagonal closepacked (HCP). Two features of a crystal structure are coordination number (or
number of nearest-neighbor atoms) and atomic packing factor (the fraction of solid
sphere volume in the unit cell). Coordination number and atomic packing factor
are the same for both FCC and HCP crystal structures, each of which may be generated by the stacking of close-packed planes of atoms.
Point Coordinates
Crystallographic Directions
Crystallographic Planes
Crystallographic points, directions, and planes are specified in terms of indexing
schemes. The basis for the determination of each index is a coordinate axis system
defined by the unit cell for the particular crystal structure. The location of a point
within a unit cell is specified using coordinates that are fractional multiples of the
cell edge lengths. Directional indices are computed in terms of the vector projection
on each of the coordinate axes, whereas planar indices are determined from the reciprocals of axial intercepts. For hexagonal unit cells, a four-index scheme for both
directions and planes is found to be more convenient.
Linear and Planar Densities
Crystallographic directional and planar equivalencies are related to atomic linear
and planar densities, respectively.The atomic packing (i.e., planar density) of spheres
1496T_c03_38-79 11/11/05 19:05 Page 73
REVISED PAGES
References • 73
in a crystallographic plane depends on the indices of the plane as well as the crystal structure. For a given crystal structure, planes having identical atomic packing
yet different Miller indices belong to the same family.
Single Crystals
Polycrystalline Materials
Single crystals are materials in which the atomic order extends uninterrupted over
the entirety of the specimen; under some circumstances, they may have flat faces
and regular geometric shapes. The vast majority of crystalline solids, however, are
polycrystalline, being composed of many small crystals or grains having different
crystallographic orientations.
Crystal Systems
Polymorphism and Allotropy
Anisotropy
Other concepts introduced in this chapter were: crystal system (a classification
scheme for crystal structures on the basis of unit cell geometry), polymorphism (or
allotropy) (when a specific material can have more than one crystal structure), and
anisotropy (the directionality dependence of properties).
X-Ray Diffraction: Determination of Crystal Structures
X-ray diffractometry is used for crystal structure and interplanar spacing determinations. A beam of x-rays directed on a crystalline material may experience diffraction (constructive interference) as a result of its interaction with a series of
parallel atomic planes according to Bragg’s law. Interplanar spacing is a function of
the Miller indices and lattice parameter(s) as well as the crystal structure.
I M P O R TA N T T E R M S A N D C O N C E P T S
Allotropy
Amorphous
Anisotropy
Atomic packing factor (APF)
Body-centered cubic (BCC)
Bragg’s law
Coordination number
Crystal structure
Crystal system
Crystalline
Diffraction
Face-centered cubic (FCC)
Grain
Grain boundary
Hexagonal close-packed (HCP)
Isotropic
Lattice
Lattice parameters
Miller indices
Noncrystalline
Polycrystalline
Polymorphism
Single crystal
Unit cell
REFERENCES
Azaroff, L. F., Elements of X-Ray Crystallography,
McGraw-Hill, New York, 1968. Reprinted by
TechBooks, Marietta, OH, 1990.
Buerger, M. J., Elementary Crystallography, Wiley,
New York, 1956.
Cullity, B. D., and S. R. Stock, Elements of X-Ray
Diffraction, 3rd edition, Prentice Hall, Upper
Saddle River, NJ, 2001.
1496T_c03_38-79 11/11/05 19:05 Page 74
REVISED PAGES
74 • Chapter 3 / The Structure of Crystalline Solids
QUESTIONS AND PROBLEMS
Fundamental Concepts
3.1 What is the difference between atomic structure and crystal structure?
Unit Cells
Metallic Crystal Structures
3.2 If the atomic radius of lead is 0.175 nm, calculate the volume of its unit cell in cubic meters.
3.3 Show for the body-centered cubic crystal
structure that the unit cell edge length a and
the atomic radius R are related through
a 4R 13.
3.4 For the HCP crystal structure, show that the
ideal ca ratio is 1.633.
3.5 Show that the atomic packing factor for BCC
is 0.68.
3.6 Show that the atomic packing factor for HCP
is 0.74.
Density Computations
3.7 Molybdenum has a BCC crystal structure, an
atomic radius of 0.1363 nm, and an atomic
weight of 95.94 g/mol. Compute and compare
its theoretical density with the experimental
value found inside the front cover.
3.8 Calculate the radius of a palladium atom,
given that Pd has an FCC crystal structure, a
density of 12.0 g/cm3, and an atomic weight of
106.4 g/mol.
3.9 Calculate the radius of a tantalum atom,
given that Ta has a BCC crystal structure, a
density of 16.6 g/cm3, and an atomic weight of
180.9 g/mol.
3.10 Some hypothetical metal has the simple cubic
crystal structure shown in Figure 3.23. If its
atomic weight is 74.5 g/mol and the atomic
radius is 0.145 nm, compute its density.
3.11 Titanium has an HCP crystal structure and a
density of 4.51 g/cm3.
(a) What is the volume of its unit cell in cubic
meters?
(b) If the c a ratio is 1.58, compute the values
of c and a.
3.12 Using atomic weight, crystal structure, and
atomic radius data tabulated inside the front
cover, compute the theoretical densities of
aluminum, nickel, magnesium, and tungsten,
and then compare these values with the measured densities listed in this same table. The
ca ratio for magnesium is 1.624.
3.13 Niobium has an atomic radius of 0.1430 nm and
a density of 8.57 g/cm3. Determine whether it
has an FCC or BCC crystal structure.
3.14 Below are listed the atomic weight, density,
and atomic radius for three hypothetical alloys. For each determine whether its crystal
structure is FCC, BCC, or simple cubic and
then justify your determination. A simple
cubic unit cell is shown in Figure 3.23.
Alloy
A
B
C
Atomic
Weight
(g/mol)
Density
(g/cm3)
Atomic
Radius
(nm)
43.1
184.4
91.6
6.40
12.30
9.60
0.122
0.146
0.137
3.15 The unit cell for uranium has orthorhombic
symmetry, with a, b, and c lattice parameters
of 0.286, 0.587, and 0.495 nm, respectively. If
its density, atomic weight, and atomic radius
are 19.05 g/cm3, 238.03 g/mol, and 0.1385 nm,
respectively, compute the atomic packing
factor.
Figure 3.23 Hard-sphere unit cell representation of the
simple cubic crystal structure.
1496T_c03_38-79 12/20/05 7:38 Page 75
2nd REVISE PAGES
Questions and Problems • 75
3.16 Indium has a tetragonal unit cell for which
the a and c lattice parameters are 0.459 and
0.495 nm, respectively.
(a) If the atomic packing factor and atomic radius are 0.693 and 0.1625 nm, respectively, determine the number of atoms in each unit cell.
(b) The atomic weight of indium is 114.82
g/mol; compute its theoretical density.
3.17 Beryllium has an HCP unit cell for which the
ratio of the lattice parameters c a is 1.568. If
the radius of the Be atom is 0.1143 nm, (a)
determine the unit cell volume, and (b) calculate the theoretical density of Be and compare it with the literature value.
3.18 Magnesium has an HCP crystal structure, a
ca ratio of 1.624, and a density of 1.74 g/cm3.
Compute the atomic radius for Mg.
3.19 Cobalt has an HCP crystal structure, an atomic
radius of 0.1253 nm, and a ca ratio of 1.623.
Compute the volume of the unit cell for Co.
Crystal Systems
3.20 Below is a unit cell for a hypothetical metal.
(a) To which crystal system does this unit cell
belong?
(b) What would this crystal structure be
called?
(c) Calculate the density of the material,
given that its atomic weight is 141 g/mol.
3.23 List the point coordinates of both the sodium
and chlorine ions for a unit cell of the sodium
chloride crystal structure (Figure 12.2).
3.24 List the point coordinates of both the zinc and
sulfur atoms for a unit cell of the zinc blende
crystal structure (Figure 12.4).
3.25 Sketch a tetragonal unit cell, and within that
cell indicate locations of the 1 1 21 and 21 41 12
point coordinates.
3.26 Using the Molecule Definition Utility found
in both “Metallic Crystal Structures and Crystallography” and “Ceramic Crystal Structures” modules of VMSE, located on the
book’s web site [www.wiley.com/college/callister (Student Companion Site)], generate
(and print out) a three-dimensional unit cell
for b tin given the following: (1) the unit cell
is tetragonal with a 0.583 nm and c 0.318
nm, and (2) Sn atoms are located at the following point coordinates:
000
011
1
3
100
2 0 4
1
3
110
2 1 4
1 1
124
010
0 21 14
001
1 1 1
2 2 2
101
111
Crystallographic Directions
3.27 Draw an orthorhombic unit cell, and within
that cell a [211] direction.
3.28 Sketch a monoclinic unit cell, and within that
cell a [101] direction.
3.29 What are the indices for the directions indicated by the two vectors in the sketch below?
+z
90°
+z
0.45 nm
O
90°
90°
+y
0.35 nm
+x
0.4 nm
Direction 2
0.35 nm
3.21 Sketch a unit cell for the face-centered orthorhombic crystal structure.
+y
0.3 nm
Point Coordinates
3.22 List the point coordinates for all atoms that are
associated with the FCC unit cell (Figure 3.1).
+x
Direction 1
0.5 nm
1496T_c03_38-79 12/20/05 7:38 Page 76
2nd REVISE PAGES
76 • Chapter 3 / The Structure of Crystalline Solids
3.30 Within a cubic unit cell, sketch the following
directions:
(a) [101],
(e) [111],
(b) [211],
(f) [212],
(c) [102],
(g) [312],
(d) [313],
(h) [301].
3.31 Determine the indices for the directions shown
in the following cubic unit cell:
3.35 Determine the indices for the directions shown
in the following hexagonal unit cell:
z
a2
A
+z
B
a3
D
C
A
a1
1
2
C
1, 1
2 2
1
2
B
D
+y
3.36 Using Equations 3.6a, 3.6b, 3.6c, and 3.6d, derive expressions for each of the three primed
indices set (u¿, v¿, and w¿ ) in terms of the four
unprimed indices (u, v, t, and w).
Crystallographic Planes
+x
3.32 Determine the indices for the directions shown
in the following cubic unit cell:
+z
3.37 (a) Draw an orthorhombic unit cell, and within
that cell a (021) plane.
(b) Draw a monoclinic unit cell, and within
that cell a (200) plane.
3.38 What are the indices for the two planes drawn
in the sketch below?
2
3
Plane 2
1
3
C
1, 1
2 2
1
3
D
Plane 1
1
2
+y
1, 1
2 2
+y
0.2 nm
A
1
2
+z
B
2
3
0.4 nm
+x
0.4 nm
+x
3.33 For tetragonal crystals, cite the indices of
directions that are equivalent to each of the
following directions:
(a) [011]
(b) [100]
3.34 Convert the [110] and [001] directions into the
four-index Miller–Bravais scheme for hexagonal unit cells.
3.39 Sketch within a cubic unit cell the following
planes:
(a) (101),
(e) (111),
(b) (211),
(f) (212),
(c) (012),
(g) (312),
(d) (313),
(h) (301).
1496T_c03_38-79 11/11/05 19:05 Page 77
REVISED PAGES
Questions and Problems • 77
1
3
1
2
B
A
+y
+x
3.41 Determine the Miller indices for the planes
shown in the following unit cell:
+z
0.4 nm
1
2
1
2
0.4 nm
A
0.4 nm
(100)
(001)
B
+y
1
2
+x
3.42 Determine the Miller indices for the planes
shown in the following unit cell:
0.55 nm
+z
and (111) planes, (b) (110) and (110) planes,
and (c) (111) and (001) planes.
3.44 Sketch the atomic packing of (a) the (100)
plane for the FCC crystal structure, and (b)
the (111) plane for the BCC crystal structure
(similar to Figures 3.10b and 3.11b).
3.45 Consider the reduced-sphere unit cell shown
in Problem 3.20, having an origin of the coordinate system positioned at the atom labeled with an O. For the following sets of
planes, determine which are equivalent:
(a) (100), (010), and (001)
(b) (110), (101), (011), and (101)
(c) (111), (111), (111), and (111)
3.46 Here are three different crystallographic
planes for a unit cell of a hypothetical metal.
The circles represent atoms:
0.55 nm
3.40 Determine the Miller indices for the planes
shown in the following unit cell:
(110)
(a) To what crystal system does the unit cell
belong?
(b) What would this crystal structure be called?
3.47 Below are shown three different crystallographic planes for a unit cell of some hypothetical metal. The circles represent atoms:
0.32 nm
0.39 nm
1
2
1
2
0.36 nm
+z
A
0.30 nm
0.20 nm
(110)
B
+y
+x
3.43 Cite the indices of the direction that results
from the intersection of each of the following
pair of planes within a cubic crystal: (a) (110)
0.25 nm
(101)
(011)
(a) To what crystal system does the unit cell
belong?
(b) What would this crystal structure be called?
(c) If the density of this metal is 18.91 g/cm3,
determine its atomic weight.
3.48 Convert the (111) and (012) planes into the
four-index Miller–Bravais scheme for hexagonal unit cells.
1496T_c03_38-79 12/20/05 7:38 Page 78
2nd REVISE PAGES
78 • Chapter 3 / The Structure of Crystalline Solids
3.49 Determine the indices for the planes shown
in the hexagonal unit cells below:
z
3.50 Sketch the (0111) and (2110) planes in a
hexagonal unit cell.
Linear and Planar Densities
a2
a3
a1
(a)
z
a2
a3
a1
(b)
z
3.51 (a) Derive linear density expressions for FCC
[100] and [111] directions in terms of the
atomic radius R.
(b) Compute and compare linear density values for these same two planes for copper.
3.52 (a) Derive linear density expressions for
BCC [110] and [111] directions in terms of the
atomic radius R.
(b) Compute and compare linear density values for these same two planes for iron.
3.53 (a) Derive planar density expressions for
FCC (100) and (111) planes in terms of the
atomic radius R.
(b) Compute and compare planar density values for these same two planes for aluminum.
3.54 (a) Derive planar density expressions for
BCC (100) and (110) planes in terms of the
atomic radius R.
(b) Compute and compare planar density values for these same two planes for molybdenum.
3.55 (a) Derive the planar density expression for
the HCP (0001) plane in terms of the atomic
radius R.
(b) Compute the planar density value for this
same plane for titanium.
Polycrystalline Materials
a2
3.56 Explain why the properties of polycrystalline
materials are most often isotropic.
a3
a1
(c)
z
a2
a3
(d)
a1
X-Ray Diffraction: Determination of
Crystal Structures
3.57 Using the data for aluminum in Table 3.1,
compute the interplanar spacing for the (110)
set of planes.
3.58 Determine the expected diffraction angle for
the first-order reflection from the (310) set of
planes for BCC chromium when monochromatic radiation of wavelength 0.0711 nm is
used.
3.59 Using the data for -iron in Table 3.1, compute the interplanar spacings for the (111) and
(211) sets of planes.
1496T_c03_38-79 11/11/05 19:05 Page 79
REVISED PAGES
Questions and Problems • 79
Intensity
Figure 3.24 Diffraction pattern
for powdered tungsten. (Courtesy
of Wesley L. Holman.)
0.0
20.0
40.0
60.0
80.0
100.0
Diffraction angle 2
3.60 The metal rhodium has an FCC crystal structure. If the angle of diffraction for the (311) set
of planes occurs at 36.12 (first-order reflection) when monochromatic x-radiation having
a wavelength of 0.0711 nm is used, compute
(a) the interplanar spacing for this set of planes,
and (b) the atomic radius for a rhodium atom.
3.61 The metal niobium has a BCC crystal structure.
If the angle of diffraction for the (211) set of
planes occurs at 75.99 (first-order reflection)
when monochromatic x-radiation having a
wavelength of 0.1659 nm is used, compute
(a) the interplanar spacing for this set of planes,
and (b) the atomic radius for the niobium atom.
3.62 For which set of crystallographic planes will a
first-order diffraction peak occur at a diffraction angle of 44.53 for FCC nickel when
monochromatic radiation having a wavelength of 0.1542 nm is used?
3.63 Figure 3.21 shows an x-ray diffraction pattern
for lead taken using a diffractometer and
monochromatic x-radiation having a wavelength of 0.1542 nm; each diffraction peak on
the pattern has been indexed. Compute the
interplanar spacing for each set of planes
indexed; also determine the lattice parameter
of Pb for each of the peaks.
3.64 The diffraction peaks shown in Figure 3.21 are
indexed according to the reflection rules for
FCC (i.e., h, k, and l must all be either odd or
even). Cite the h, k, and l indices of the first
four diffraction peaks for BCC crystals consistent with h k l being even.
3.65 Figure 3.24 shows the first five peaks of the
x-ray diffraction pattern for tungsten, which
has a BCC crystal structure; monochromatic
x-radiation having a wavelength of 0.1542 nm
was used.
(a) Index (i.e., give h, k, and l indices) for
each of these peaks.
(b) Determine the interplanar spacing for
each of the peaks.
(c) For each peak, determine the atomic radius for W and compare these with the value
presented in Table 3.1.
Noncrystalline Solids
3.66 Would you expect a material in which the
atomic bonding is predominantly ionic in
nature to be more or less likely to form
a noncrystalline solid upon solidification
than a covalent material? Why? (See Section 2.6.)
1496T_c04_80-108 12/20/05 7:46 Page 80
Chapter
4
2nd REVISE PAGES
Imperfections in Solids
A
scanning probe micrograph
(generated using a scanning-tunneling
microscope) that shows a (111)-type
surface plane* for silicon. The arrow
points to the location of a silicon
atom that was removed using a tungsten nanotip probe. This site from
which an atom is missing is the surface analogue of a vacancy defect—
that is, a vacant lattice site within
the bulk material. Approximately
20,000,000. (Micrograph courtesy
of D. Huang, Stanford University.)
WHY STUDY Imperfections in Solids?
The properties of some materials are profoundly influenced by the presence of imperfections. Consequently,
it is important to have a knowledge about the types of
imperfections that exist and the roles they play in affecting the behavior of materials. For example, the mechanical properties of pure metals experience significant
alterations when alloyed (i.e., when impurity atoms are
added)—for example, brass (70% copper–30% zinc) is
much harder and stronger than pure copper (Section
7.9).
Also, integrated circuit microelectronic devices
found in our computers, calculators, and home appliances function because of highly controlled concentrations of specific impurities that are incorporated
into small, localized regions of semiconducting materials (Sections 18.11 and 18.15).
*The plane shown here is termed a “Si(111)-77 reconstructed surface.” The twodimensional arrangement of atoms in a surface plane is different than the atomic arrangement for the equivalent plane within the interior of the material (i.e., the surface plane
has been “reconstructed” by atomic displacements). The “77” notation pertains to the
displacement magnitude. Furthermore, the diamond shape that has been drawn indicates
a unit cell for this 77 structure.
80 •
1496T_c04_80-108 12/20/05 7:46 Page 81
2nd REVISE PAGES
Learning Objectives
After studying this chapter you should be able to do the following:
1. Describe both vacancy and self-interstitial crys5. For each of edge, screw, and mixed dislocations:
talline defects.
(a) describe and make a drawing of the
2. Calculate the equilibrium number of vacancies
dislocation.
in a material at some specified temperature,
(b) note the location of the dislocation line, and
given the relevant constants.
(c) indicate the direction along which the dislo3. Name the two types of solid solutions, and procation line extends.
vide a brief written definition and/or schematic
6. Describe the atomic structure within the vicinity
sketch of each.
of (a) a grain boundary, and (b) a twin boundary.
4. Given the masses and atomic weights of two or
more elements in a metal alloy, calculate the
weight percent and atom percent for each
element.
4.1 INTRODUCTION
imperfection
point defect
Thus far it has been tacitly assumed that perfect order exists throughout crystalline
materials on an atomic scale. However, such an idealized solid does not exist; all
contain large numbers of various defects or imperfections. As a matter of fact, many
of the properties of materials are profoundly sensitive to deviations from crystalline
perfection; the influence is not always adverse, and often specific characteristics are
deliberately fashioned by the introduction of controlled amounts or numbers of
particular defects, as detailed in succeeding chapters.
By “crystalline defect” is meant a lattice irregularity having one or more of its
dimensions on the order of an atomic diameter. Classification of crystalline imperfections is frequently made according to geometry or dimensionality of the defect.
Several different imperfections are discussed in this chapter, including point defects
(those associated with one or two atomic positions), linear (or one-dimensional)
defects, as well as interfacial defects, or boundaries, which are two-dimensional.
Impurities in solids are also discussed, since impurity atoms may exist as point
defects. Finally, techniques for the microscopic examination of defects and the
structure of materials are briefly described.
Po i n t D e f e c t s
4.2 VACANCIES AND SELF-INTERSTITIALS
vacancy
Temperaturedependence of the
equilibrium number
of vacancies
The simplest of the point defects is a vacancy, or vacant lattice site, one normally
occupied from which an atom is missing (Figure 4.1). All crystalline solids contain
vacancies and, in fact, it is not possible to create such a material that is free of these
defects. The necessity of the existence of vacancies is explained using principles of
thermodynamics; in essence, the presence of vacancies increases the entropy (i.e.,
the randomness) of the crystal.
The equilibrium number of vacancies Nv for a given quantity of material
depends on and increases with temperature according to
Nv N exp a
Qv
b
kT
(4.1)
1496T_c04_80-108 11/14/05 9:53 Page 82
REVISED PAGES
82 • Chapter 4 / Imperfections in Solids
Self-interstitial
Boltzmann’s constant
self-interstitial
Vacancy
Figure 4.1 Two-dimensional
representations of a vacancy and a
self-interstitial. (Adapted from W. G.
Moffatt, G. W. Pearsall, and J. Wulff,
The Structure and Properties of
Materials, Vol. I, Structure, p. 77.
Copyright © 1964 by John Wiley
& Sons, New York. Reprinted by
permission of John Wiley & Sons,
Inc.)
In this expression, N is the total number of atomic sites, Qv is the energy required
for the formation of a vacancy, T is the absolute temperature1 in kelvins, and k is
the gas or Boltzmann’s constant. The value of k is 1.38 10 23 J/atom-K, or
8.62 105 eV/atom-K, depending on the units of Qv.2 Thus, the number of vacancies increases exponentially with temperature; that is, as T in Equation 4.1 increases, so also does the expression exp (QvkT ). For most metals, the fraction of
vacancies NvN just below the melting temperature is on the order of 104; that is,
one lattice site out of 10,000 will be empty. As ensuing discussions indicate, a number of other material parameters have an exponential dependence on temperature
similar to that of Equation 4.1.
A self-interstitial is an atom from the crystal that is crowded into an interstitial site, a small void space that under ordinary circumstances is not occupied. This
kind of defect is also represented in Figure 4.1. In metals, a self-interstitial introduces relatively large distortions in the surrounding lattice because the atom is substantially larger than the interstitial position in which it is situated. Consequently,
the formation of this defect is not highly probable, and it exists in very small concentrations, which are significantly lower than for vacancies.
EXAMPLE PROBLEM 4.1
Number of Vacancies Computation at a Specified
Temperature
Calculate the equilibrium number of vacancies per cubic meter for copper at
1000C. The energy for vacancy formation is 0.9 eV/atom; the atomic weight
and density (at 1000C) for copper are 63.5 g/mol and 8.4 g/cm3, respectively.
Absolute temperature in kelvins (K) is equal to C 273.
Boltzmann’s constant per mole of atoms becomes the gas constant R; in such a case
R 8.31 J/mol-K.
1
2
1496T_c04_80-108 11/14/05 9:53 Page 83
REVISED PAGES
4.3 Impurities in Solids • 83
Solution
This problem may be solved by using Equation 4.1; it is first necessary, however, to determine the value of N, the number of atomic sites per cubic meter
for copper, from its atomic weight ACu, its density r, and Avogadro’s number
NA, according to
Number of atoms
per unit volume for
a metal
N
NA r
ACu
(4.2)
16.023 1023 atoms/mol218.4 g/cm3 21106 cm3/m3 2
63.5 g/mol
28
8.0 10 atoms/m3
Thus, the number of vacancies at 1000C (1273 K) is equal to
Nv N exp a
Qv
b
kT
18.0 1028 atoms/m3 2 exp c
2.2 1025 vacancies/m3
10.9 eV2
18.62 105 eV/K211273 K2
d
4.3 IMPURITIES IN SOLIDS
alloy
solid solution
solute, solvent
A pure metal consisting of only one type of atom just isn’t possible; impurity or
foreign atoms will always be present, and some will exist as crystalline point defects. In fact, even with relatively sophisticated techniques, it is difficult to refine
metals to a purity in excess of 99.9999%. At this level, on the order of 1022 to 1023
impurity atoms will be present in one cubic meter of material. Most familiar metals are not highly pure; rather, they are alloys, in which impurity atoms have been
added intentionally to impart specific characteristics to the material. Ordinarily, alloying is used in metals to improve mechanical strength and corrosion resistance.
For example, sterling silver is a 92.5% silver–7.5% copper alloy. In normal ambient
environments, pure silver is highly corrosion resistant, but also very soft. Alloying
with copper significantly enhances the mechanical strength without depreciating the
corrosion resistance appreciably.
The addition of impurity atoms to a metal will result in the formation of a solid
solution and/or a new second phase, depending on the kinds of impurity, their concentrations, and the temperature of the alloy. The present discussion is concerned
with the notion of a solid solution; treatment of the formation of a new phase is
deferred to Chapter 9.
Several terms relating to impurities and solid solutions deserve mention. With
regard to alloys, solute and solvent are terms that are commonly employed.“Solvent”
represents the element or compound that is present in the greatest amount; on
occasion, solvent atoms are also called host atoms. “Solute” is used to denote an
element or compound present in a minor concentration.
Solid Solutions
A solid solution forms when, as the solute atoms are added to the host material,
the crystal structure is maintained, and no new structures are formed. Perhaps it is
1496T_c04_80-108 11/14/05 9:53 Page 84
REVISED PAGES
84 • Chapter 4 / Imperfections in Solids
Substitutional
impurity atom
Figure 4.2 Two-dimensional
schematic representations of
substitutional and interstitial
impurity atoms. (Adapted from
W. G. Moffatt, G. W. Pearsall,
and J. Wulff, The Structure and
Properties of Materials, Vol. I,
Structure, p. 77. Copyright © 1964
by John Wiley & Sons, New York.
Reprinted by permission of John
Wiley & Sons, Inc.)
Interstitial
impurity atom
substitutional solid
solution
interstitial solid
solution
useful to draw an analogy with a liquid solution. If two liquids, soluble in each other
(such as water and alcohol) are combined, a liquid solution is produced as the
molecules intermix, and its composition is homogeneous throughout. A solid solution is also compositionally homogeneous; the impurity atoms are randomly and
uniformly dispersed within the solid.
Impurity point defects are found in solid solutions, of which there are two types:
substitutional and interstitial. For the substitutional type, solute or impurity atoms
replace or substitute for the host atoms (Figure 4.2). There are several features of
the solute and solvent atoms that determine the degree to which the former dissolves in the latter, as follows:
1. Atomic size factor. Appreciable quantities of a solute may be accommodated in this type of solid solution only when the difference in atomic
radii between the two atom types is less than about 15% . Otherwise
the solute atoms will create substantial lattice distortions and a new phase
will form.
2. Crystal structure. For appreciable solid solubility the crystal structures for
metals of both atom types must be the same.
3. Electronegativity. The more electropositive one element and the more electronegative the other, the greater is the likelihood that they will form an
intermetallic compound instead of a substitutional solid solution.
4. Valences. Other factors being equal, a metal will have more of a tendency to
dissolve another metal of higher valency than one of a lower valency.
An example of a substitutional solid solution is found for copper and nickel.
These two elements are completely soluble in one another at all proportions. With
regard to the aforementioned rules that govern degree of solubility, the atomic radii
for copper and nickel are 0.128 and 0.125 nm, respectively, both have the FCC crystal structure, and their electronegativities are 1.9 and 1.8 (Figure 2.7); finally, the
most common valences are 1 for copper (although it sometimes can be 2) and
2 for nickel.
For interstitial solid solutions, impurity atoms fill the voids or interstices among
the host atoms (see Figure 4.2). For metallic materials that have relatively high
atomic packing factors, these interstitial positions are relatively small. Consequently,
1496T_c04_80-108 11/14/05 9:53 Page 85
REVISED PAGES
4.4 Specification of Composition • 85
the atomic diameter of an interstitial impurity must be substantially smaller than
that of the host atoms. Normally, the maximum allowable concentration of interstitial impurity atoms is low (less than 10%). Even very small impurity atoms are
ordinarily larger than the interstitial sites, and as a consequence they introduce some
lattice strains on the adjacent host atoms. Problem 4.5 calls for determination of
the radii of impurity atoms (in terms of R, the host atom radius) that will just fit
into interstitial positions without introducing any lattice strains for both FCC and
BCC crystal structures.
Carbon forms an interstitial solid solution when added to iron; the maximum
concentration of carbon is about 2%. The atomic radius of the carbon atom is much
less than that for iron: 0.071 nm versus 0.124 nm. Solid solutions are also possible
for ceramic materials, as discussed in Section 12.5.
4.4 SPECIFICATION OF COMPOSITION
composition
weight percent
Computation of
weight percent (for a
two-element alloy)
atom percent
It is often necessary to express the composition (or concentration)3 of an alloy in
terms of its constituent elements. The two most common ways to specify composition are weight (or mass) percent and atom percent. The basis for weight percent
(wt%) is the weight of a particular element relative to the total alloy weight. For
an alloy that contains two hypothetical atoms denoted by 1 and 2, the concentration of 1 in wt%, C1, is defined as
C1
m1
100
m1 m2
(4.3)
where m1 and m2 represent the weight (or mass) of elements 1 and 2, respectively.
The concentration of 2 would be computed in an analogous manner.
The basis for atom percent (at%) calculations is the number of moles of an element in relation to the total moles of the elements in the alloy. The number of
moles in some specified mass of a hypothetical element 1, nm1, may be computed
as follows:
m1¿
nm1
(4.4)
A1
Here, m¿1 and A1 denote the mass (in grams) and atomic weight, respectively, for
element 1.
Concentration in terms of atom percent of element 1 in an alloy containing
1 and 2 atoms, C¿1, is defined by4
Computation of
atom percent (for a
two-element alloy)
C¿1
nm1
100
nm1 nm2
(4.5)
In like manner, the atom percent of 2 may be determined.
3
The terms composition and concentration will be assumed to have the same meaning in
this book (i.e., the relative content of a specific element or constituent in an alloy) and will
be used interchangeably.
4
In order to avoid confusion in notations and symbols that are being used in this section,
we should point out that the prime (as in C¿1 and m¿1) is used to designate both composition, in atom percent, as well as mass of material in units of grams.
1496T_c04_80-108 11/14/05 9:53 Page 86
REVISED PAGES
86 • Chapter 4 / Imperfections in Solids
Atom percent computations also can be carried out on the basis of the number of atoms instead of moles, since one mole of all substances contains the same
number of atoms.
Composition Conversions
Sometimes it is necessary to convert from one composition scheme to another—
for example, from weight percent to atom percent. We will now present equations
for making these conversions in terms of the two hypothetical elements 1 and 2.
Using the convention of the previous section (i.e., weight percents denoted by C1
and C2, atom percents by C¿1 and C¿2, and atomic weights as A1 and A2), these conversion expressions are as follows:
Conversion of
weight percent to
atom percent (for a
two-element alloy)
Conversion of atom
percent to weight
percent (for a twoelement alloy)
C¿1
C1A2
100
C1A2 C2A1
(4.6a)
C¿2
C2A1
100
C1A2 C2A1
(4.6b)
C1
C¿1A1
100
C¿1A1 C¿2A2
(4.7a)
C2
C¿2A2
100
C¿1A1 C¿2A2
(4.7b)
Since we are considering only two elements, computations involving the preceding equations are simplified when it is realized that
C1 C2 100
C¿1 C¿2 100
(4.8a)
(4.8b)
In addition, it sometimes becomes necessary to convert concentration from
weight percent to mass of one component per unit volume of material (i.e., from
units of wt% to kg/m3); this latter composition scheme is often used in diffusion computations (Section 5.3). Concentrations in terms of this basis will be denoted using a double prime (i.e., C–1 and C–2), and the relevant equations are as
follows:
Conversion of
weight percent to
mass per unit
volume (for a twoelement alloy)
C–1
C1
103
C2 ¢
° C1
r1
r2
(4.9a)
C–2
C2
103
C2 ¢
° C1
r1
r2
(4.9b)
For density r in units of g/cm3, these expressions yield C–1 and C–2 in kg/m3.
Furthermore, on occasion we desire to determine the density and atomic weight
of a binary alloy given the composition in terms of either weight percent or atom
1496T_c04_80-108 11/14/05 9:53 Page 87
REVISED PAGES
4.4 Specification of Composition • 87
percent. If we represent alloy density and atomic weight by rave and Aave, respectively, then
Computation of
density (for a twoelement metal alloy)
rave
rave
Computation of
atomic weight (for a
two-element metal
alloy)
Aave
Aave
100
(4.10a)
C1
C2
r1
r2
C¿1A1 C¿2A2
C¿1A1
C¿2A2
r1
r2
100
C1
C2
A1
A2
(4.10b)
(4.11a)
C¿1A1 C¿2A2
100
(4.11b)
It should be noted that Equations 4.9 and 4.11 are not always exact. In their derivations, it is assumed that total alloy volume is exactly equal to the sum of the
volumes of the individual elements. This normally is not the case for most alloys;
however, it is a reasonably valid assumption and does not lead to significant errors
for dilute solutions and over composition ranges where solid solutions exist.
EXAMPLE PROBLEM 4.2
Derivation of Composition-Conversion Equation
Derive Equation 4.6a.
Solution
To simplify this derivation, we will assume that masses are expressed in units
of grams, and denoted with a prime (e.g., m¿1). Furthermore, the total alloy
mass (in grams) M¿ is
M¿ m¿1 m¿2
(4.12)
Using the definition of C¿1 (Equation 4.5) and incorporating the expression for nm1, Equation 4.4, and the analogous expression for nm2 yields
C¿1
nm1
100
nm1 nm2
m¿1
A1
m¿1
m¿2
A1
A2
100
(4.13)
Rearrangement of the mass-in-grams equivalent of Equation 4.3 leads to
m¿1
C1M¿
100
(4.14)
1496T_c04_80-108 11/14/05 9:53 Page 88
REVISED PAGES
88 • Chapter 4 / Imperfections in Solids
Substitution of this expression and its m¿2 equivalent into Equation 4.13 gives
C1M¿
100A1
C¿1
100
C1M¿
C2M¿
100A1
100A2
(4.15)
Upon simplification we have
C¿1
C1A2
100
C1A2 C2A1
which is identical to Equation 4.6a.
EXAMPLE PROBLEM 4.3
Composition Conversion—From Weight Percent to
Atom Percent
Determine the composition, in atom percent, of an alloy that consists of
97 wt% aluminum and 3 wt% copper.
Solution
If we denote the respective weight percent compositions as CA1 97 and
CCu 3, substitution into Equations 4.6a and 4.6b yields
C¿A1
CA1ACu
100
CA1ACu CCu AA1
1972163.55 g/mol2
100
CCu AA1
100
CCu AA1 CA1ACu
132126.98 g/mol2
100
1972163.55 g/mol2 132126.98 g/mol2
98.7 at%
and
C¿Cu
132126.98 g/mol2 1972163.55 g/mol2
1.30 at%
Miscellaneous Imperfections
4.5 DISLOCATIONS—LINEAR DEFECTS
edge dislocation
A dislocation is a linear or one-dimensional defect around which some of the atoms
are misaligned. One type of dislocation is represented in Figure 4.3: an extra portion of a plane of atoms, or half-plane, the edge of which terminates within the
crystal. This is termed an edge dislocation; it is a linear defect that centers around
the line that is defined along the end of the extra half-plane of atoms. This is
1496T_c04_80-108 12/20/05 7:46 Page 89
2nd REVISE PAGES
4.5 Dislocations—Linear Defects • 89
Burgers vector
b
Figure 4.3 The atom positions around an
edge dislocation; extra half-plane of atoms
shown in perspective. (Adapted from
A. G. Guy, Essentials of Materials Science,
McGraw-Hill Book Company, New York,
1976, p. 153.)
Edge
dislocation
line
dislocation line
Edge
screw dislocation
Screw
Mixed
mixed dislocation
Burgers vector
sometimes termed the dislocation line, which, for the edge dislocation in Figure 4.3,
is perpendicular to the plane of the page. Within the region around the dislocation
line there is some localized lattice distortion. The atoms above the dislocation line
in Figure 4.3 are squeezed together, and those below are pulled apart; this is reflected in the slight curvature for the vertical planes of atoms as they bend around
this extra half-plane. The magnitude of this distortion decreases with distance away
from the dislocation line; at positions far removed, the crystal lattice is virtually perfect. Sometimes the edge dislocation in Figure 4.3 is represented by the symbol ,
which also indicates the position of the dislocation line. An edge dislocation may
also be formed by an extra half-plane of atoms that is included in the bottom portion of the crystal; its designation is a .
Another type of dislocation, called a screw dislocation, exists, which may be
thought of as being formed by a shear stress that is applied to produce the distortion shown in Figure 4.4a: the upper front region of the crystal is shifted one atomic
distance to the right relative to the bottom portion. The atomic distortion associated with a screw dislocation is also linear and along a dislocation line, line AB in
Figure 4.4b. The screw dislocation derives its name from the spiral or helical path
or ramp that is traced around the dislocation line by the atomic planes of atoms.
Sometimes the symbol
is used to designate a screw dislocation.
Most dislocations found in crystalline materials are probably neither pure edge
nor pure screw, but exhibit components of both types; these are termed mixed dislocations. All three dislocation types are represented schematically in Figure 4.5;
the lattice distortion that is produced away from the two faces is mixed, having varying degrees of screw and edge character.
The magnitude and direction of the lattice distortion associated with a dislocation is expressed in terms of a Burgers vector, denoted by a b. Burgers vectors
are indicated in Figures 4.3 and 4.4 for edge and screw dislocations, respectively.
Furthermore, the nature of a dislocation (i.e., edge, screw, or mixed) is defined by
the relative orientations of dislocation line and Burgers vector. For an edge, they
are perpendicular (Figure 4.3), whereas for a screw, they are parallel (Figure 4.4);
they are neither perpendicular nor parallel for a mixed dislocation. Also, even
though a dislocation changes direction and nature within a crystal (e.g., from edge
to mixed to screw), the Burgers vector will be the same at all points along its line.
For example, all positions of the curved dislocation in Figure 4.5 will have the Burgers
vector shown. For metallic materials, the Burgers vector for a dislocation will point
in a close-packed crystallographic direction and will be of magnitude equal to the
interatomic spacing.
1496T_c04_80-108 12/20/05 7:46 Page 90
2nd REVISE PAGES
90 • Chapter 4 / Imperfections in Solids
Figure 4.4 (a) A
screw dislocation
within a crystal.
(b) The screw
dislocation in (a) as
viewed from above.
The dislocation line
extends along line
AB. Atom positions
above the slip plane
are designated by
open circles, those
below by solid
circles. [Figure
(b) from W. T. Read,
Jr., Dislocations in
Crystals, McGrawHill Book Company,
New York, 1953.]
C
A
D
Dislocation
line
Burgers vector b
(a)
A
B
b
D
C
(b)
1496T_c04_80-108 12/20/05 7:46 Page 91
2nd REVISE PAGES
4.5 Dislocations—Linear Defects • 91
Figure 4.5
(a) Schematic
representation of a
dislocation that has
edge, screw, and
mixed character.
(b) Top view, where
open circles denote
atom positions
above the slip plane.
Solid circles, atom
positions below.
At point A, the
dislocation is pure
screw, while at point
B, it is pure edge.
For regions in
between where there
is curvature in the
dislocation line, the
character is mixed
edge and screw.
[Figure (b) from
W. T. Read, Jr.,
Dislocations in
Crystals, McGrawHill Book Company,
New York, 1953.]
b
B
A
b
C
(a)
B
b
b
C
A
b
(b)
1496T_c04_80-108 11/14/05 9:53 Page 92
REVISED PAGES
92 • Chapter 4 / Imperfections in Solids
Figure 4.6 A transmission electron
micrograph of a titanium alloy in which the
dark lines are dislocations. 51,450.
(Courtesy of M. R. Plichta, Michigan
Technological University.)
As we note in Section 7.4, the permanent deformation of most crystalline materials is by the motion of dislocations. In addition, the Burgers vector is an element
of the theory that has been developed to explain this type of deformation.
Dislocations can be observed in crystalline materials using electron-microscopic
techniques. In Figure 4.6, a high-magnification transmission electron micrograph,
the dark lines are the dislocations.
Virtually all crystalline materials contain some dislocations that were introduced during solidification, during plastic deformation, and as a consequence of
thermal stresses that result from rapid cooling. Dislocations are involved in the plastic deformation of crystalline materials, both metals and ceramics, as discussed in
Chapters 7 and 12. They have also been observed in polymeric materials, and are
discussed in Section 14.13.
4.6 INTERFACIAL DEFECTS
Interfacial defects are boundaries that have two dimensions and normally separate
regions of the materials that have different crystal structures and/or crystallographic
orientations. These imperfections include external surfaces, grain boundaries, twin
boundaries, stacking faults, and phase boundaries.
External Surfaces
One of the most obvious boundaries is the external surface, along which the crystal structure terminates. Surface atoms are not bonded to the maximum number of
nearest neighbors, and are therefore in a higher energy state than the atoms at interior positions. The bonds of these surface atoms that are not satisfied give rise to a
surface energy, expressed in units of energy per unit area (J/m2 or erg/cm2). To reduce this energy, materials tend to minimize, if at all possible, the total surface area.
For example, liquids assume a shape having a minimum area—the droplets become
spherical. Of course, this is not possible with solids, which are mechanically rigid.
Grain Boundaries
Another interfacial defect, the grain boundary, was introduced in Section 3.14 as the
boundary separating two small grains or crystals having different crystallographic
1496T_c04_80-108 11/14/05 9:53 Page 93
REVISED PAGES
4.6 Interfacial Defects • 93
Figure 4.7 Schematic
diagram showing smalland high-angle grain
boundaries and the
adjacent atom positions.
Angle of misalignment
High-angle
grain boundary
Small-angle
grain boundary
Angle of misalignment
orientations in polycrystalline materials. A grain boundary is represented schematically from an atomic perspective in Figure 4.7. Within the boundary region,
which is probably just several atom distances wide, there is some atomic mismatch in a transition from the crystalline orientation of one grain to that of an
adjacent one.
Various degrees of crystallographic misalignment between adjacent grains are
possible (Figure 4.7). When this orientation mismatch is slight, on the order of a
few degrees, then the term small- (or low- ) angle grain boundary is used. These
boundaries can be described in terms of dislocation arrays. One simple smallangle grain boundary is formed when edge dislocations are aligned in the manner
of Figure 4.8. This type is called a tilt boundary; the angle of misorientation, u, is
also indicated in the figure. When the angle of misorientation is parallel to the
boundary, a twist boundary results, which can be described by an array of screw
dislocations.
The atoms are bonded less regularly along a grain boundary (e.g., bond angles
are longer), and consequently, there is an interfacial or grain boundary energy similar to the surface energy described above. The magnitude of this energy is a function of the degree of misorientation, being larger for high-angle boundaries. Grain
boundaries are more chemically reactive than the grains themselves as a consequence of this boundary energy. Furthermore, impurity atoms often preferentially
segregate along these boundaries because of their higher energy state. The total interfacial energy is lower in large or coarse-grained materials than in fine-grained
ones, since there is less total boundary area in the former. Grains grow at elevated
temperatures to reduce the total boundary energy, a phenomenon explained in
Section 7.13.
In spite of this disordered arrangement of atoms and lack of regular bonding
along grain boundaries, a polycrystalline material is still very strong; cohesive forces
within and across the boundary are present. Furthermore, the density of a polycrystalline specimen is virtually identical to that of a single crystal of the same
material.
1496T_c04_80-108 11/14/05 9:53 Page 94
REVISED PAGES
94 • Chapter 4 / Imperfections in Solids
Figure 4.8 Demonstration of how a tilt boundary
having an angle of misorientation u results from an
alignment of edge dislocations.
b
Twin Boundaries
A twin boundary is a special type of grain boundary across which there is a specific
mirror lattice symmetry; that is, atoms on one side of the boundary are located in
mirror-image positions of the atoms on the other side (Figure 4.9). The region of
material between these boundaries is appropriately termed a twin. Twins result from
atomic displacements that are produced from applied mechanical shear forces
(mechanical twins), and also during annealing heat treatments following deformation (annealing twins). Twinning occurs on a definite crystallographic plane and in
a specific direction, both of which depend on the crystal structure. Annealing twins
are typically found in metals that have the FCC crystal structure, while mechanical
twins are observed in BCC and HCP metals. The role of mechanical twins in the
deformation process is discussed in Section 7.7. Annealing twins may be observed
Twin plane (boundary)
Figure 4.9 Schematic diagram
showing a twin plane or boundary
and the adjacent atom positions
(colored circles).
1496T_c04_80-108 12/20/05 7:46 Page 95
2nd REVISE PAGES
4.6 Interfacial Defects • 95
MATERIALS OF IMPORTANCE
Catalysts (and Surface Defects)
A
catalyst is a substance that speeds up the rate
of a chemical reaction without participating
in the reaction itself (i.e., it is not consumed). One
type of catalyst exists as a solid; reactant molecules
in a gas or liquid phase are adsorbed5 onto the catalytic surface, at which point some type of interaction occurs that promotes an increase in their
chemical reactivity rate.
Adsorption sites on a catalyst are normally surface defects associated with planes of atoms; an interatomic/intermolecular bond is formed between
a defect site and an adsorbed molecular species.
Several types of surface defects, represented
schematically in Figure 4.10, include ledges, kinks,
terraces, vacancies, and individual adatoms (i.e.,
atoms adsorbed on the surface).
One important use of catalysts is in catalytic
converters on automobiles, which reduce the emission of exhaust gas pollutants such as carbon
monoxide (CO), nitrogen oxides (NOx, where x is
variable), and unburned hydrocarbons. Air is introduced into the exhaust emissions from the automobile engine; this mixture of gases then passes
over the catalyst, which adsorbs on its surface molecules of CO, NOx, and O2. The NOx dissociates
into N and O atoms, whereas the O2 dissociates
into its atomic species. Pairs of nitrogen atoms
combine to form N2 molecules, and carbon
monoxide is oxidized to form carbon dioxide (CO2).
Furthermore, any unburned hydrocarbons are also
oxidized to CO2 and H2O.
One of the materials used as a catalyst in this
application is (Ce0.5Zr0.5)O2. Figure 4.11 is a highresolution transmission electron micrograph, which
shows several single crystals of this material. Individual atoms are resolved in this micrograph as well
as some of the defects presented in Figure 4.10.
These surface defects act as adsorption sites for the
atomic and molecular species noted in the previous
paragraph. Consequently, dissociation, combination, and oxidation reactions involving these species
are facilitated, such that the content of pollutant
species (CO, NOx, and unburned hydrocarbons) in
the exhaust gas stream is reduced significantly.
Terrace
Kink
Ledge
Adatom
Step
Vacancy
Figure 4.10 Schematic representations of surface
defects that are potential adsorption sites for catalysis.
Individual atom sites are represented as cubes.
(From BOUDART, MICHEL, KINETICS OF
HETEROGENEOUS CATALYTIC REACTIONS.
© 1984 Princeton University Press. Reprinted by
permission of Princeton University Press.)
5
Figure 4.11 High-resolution transmission electron
micrograph that shows single crystals of (Ce0.5Zr0.5)O2;
this material is used in catalytic converters for
automobiles. Surface defects represented schematically
in Figure 4.10 are noted on the crystals. [From W. J.
Stark, L. Mädler, M. Maciejewski, S. E. Pratsinis, A.
Baiker, “Flame-Synthesis of Nanocrystalline
Ceria/Zirconia: Effect of Carrier Liquid,” Chem.
Comm., 588–589 (2003). Reproduced by permission of
The Royal Society of Chemistry.]
Adsorption is the adhesion of molecules of a gas or liquid to a solid surface. It should not be confused with
absorption which is the assimilation of molecules into a solid or liquid.
1496T_c04_80-108 11/14/05 9:53 Page 96
REVISED PAGES
96 • Chapter 4 / Imperfections in Solids
in the photomicrograph of the polycrystalline brass specimen shown in Figure 4.13c.
The twins correspond to those regions having relatively straight and parallel sides
and a different visual contrast than the untwinned regions of the grains within which
they reside. An explanation for the variety of textural contrasts in this photomicrograph is provided in Section 4.10.
Miscellaneous Interfacial Defects
Other possible interfacial defects include stacking faults, phase boundaries, and ferromagnetic domain walls. Stacking faults are found in FCC metals when there is an
interruption in the ABCABCABC . . . stacking sequence of close-packed planes
(Section 3.12). Phase boundaries exist in multiphase materials (Section 9.3) across
which there is a sudden change in physical and/or chemical characteristics. For ferromagnetic and ferrimagnetic materials, the boundary that separates regions having different directions of magnetization is termed a domain wall, which is discussed
in Section 20.7.
Associated with each of the defects discussed in this section is an interfacial energy, the magnitude of which depends on boundary type, and which will vary from
material to material. Normally, the interfacial energy will be greatest for external
surfaces and least for domain walls.
Concept Check 4.1
The surface energy of a single crystal depends on crystallographic orientation. Does
this surface energy increase or decrease with an increase in planar density. Why?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
4.7 BULK OR VOLUME DEFECTS
Other defects exist in all solid materials that are much larger than those heretofore
discussed. These include pores, cracks, foreign inclusions, and other phases. They
are normally introduced during processing and fabrication steps. Some of these defects and their effects on the properties of materials are discussed in subsequent
chapters.
4.8 ATOMIC VIBRATIONS
atomic vibration
Every atom in a solid material is vibrating very rapidly about its lattice position
within the crystal. In a sense, these atomic vibrations may be thought of as imperfections or defects. At any instant of time not all atoms vibrate at the same frequency and amplitude, nor with the same energy. At a given temperature there will
exist a distribution of energies for the constituent atoms about an average energy.
Over time the vibrational energy of any specific atom will also vary in a random
manner. With rising temperature, this average energy increases, and, in fact, the temperature of a solid is really just a measure of the average vibrational activity of
atoms and molecules. At room temperature, a typical vibrational frequency is on
the order of 1013 vibrations per second, whereas the amplitude is a few thousandths
of a nanometer.
Many properties and processes in solids are manifestations of this vibrational
atomic motion. For example, melting occurs when the vibrations are vigorous enough
to rupture large numbers of atomic bonds. A more detailed discussion of atomic vibrations and their influence on the properties of materials is presented in Chapter 19.
1496T_c04_80-108 11/14/05 9:53 Page 97
REVISED PAGES
4.9 General • 97
M i c ro s c o p i c E x a m i n a t i o n
4.9 GENERAL
microstructure
microscopy
photomicrograph
On occasion it is necessary or desirable to examine the structural elements and defects that influence the properties of materials. Some structural elements are of
macroscopic dimensions; that is, they are large enough to be observed with the unaided eye. For example, the shape and average size or diameter of the grains for a
polycrystalline specimen are important structural characteristics. Macroscopic
grains are often evident on aluminum streetlight posts and also on highway guard
rails. Relatively large grains having different textures are clearly visible on the surface of the sectioned lead ingot shown in Figure 4.12. However, in most materials
the constituent grains are of microscopic dimensions, having diameters that may
be on the order of microns,6 and their details must be investigated using some type
of microscope. Grain size and shape are only two features of what is termed the
microstructure; these and other microstructural characteristics are discussed in subsequent chapters.
Optical, electron, and scanning probe microscopes are commonly used in
microscopy. These instruments aid in investigations of the microstructural features
of all material types. Some of these techniques employ photographic equipment in
conjunction with the microscope; the photograph on which the image is recorded
is called a photomicrograph. In addition, many microstructural images are computer generated and/or enhanced.
Microscopic examination is an extremely useful tool in the study and characterization of materials. Several important applications of microstructural examinations are as follows: to ensure that the associations between the properties and
structure (and defects) are properly understood, to predict the properties of materials once these relationships have been established, to design alloys with new
property combinations, to determine whether or not a material has been correctly
heat treated, and to ascertain the mode of mechanical fracture. Several techniques
that are commonly used in such investigations are discussed next.
Figure 4.12 High-purity
polycrystalline lead ingot in which
the individual grains may be
discerned. 0.7. (Reproduced with
permission from Metals Handbook,
Vol. 9, 9th edition, Metallography
and Microstructures, American
Society for Metals, Metals Park,
OH, 1985.)
6
A micron ( m), sometimes called a micrometer, is 106 m.
1496T_c04_80-108 11/14/05 9:53 Page 98
REVISED PAGES
98 • Chapter 4 / Imperfections in Solids
4.10 MICROSCOPIC TECHNIQUES
Optical Microscopy
With optical microscopy, the light microscope is used to study the microstructure; optical and illumination systems are its basic elements. For materials that
are opaque to visible light (all metals and many ceramics and polymers), only
the surface is subject to observation, and the light microscope must be used in a
reflecting mode. Contrasts in the image produced result from differences in
reflectivity of the various regions of the microstructure. Investigations of this
type are often termed metallographic, since metals were first examined using this
technique.
Normally, careful and meticulous surface preparations are necessary to reveal
the important details of the microstructure. The specimen surface must first be
ground and polished to a smooth and mirrorlike finish. This is accomplished by using successively finer abrasive papers and powders. The microstructure is revealed
by a surface treatment using an appropriate chemical reagent in a procedure termed
etching. The chemical reactivity of the grains of some single-phase materials depends on crystallographic orientation. Consequently, in a polycrystalline specimen,
etching characteristics vary from grain to grain. Figure 4.13b shows how normally
incident light is reflected by three etched surface grains, each having a different orientation. Figure 4.13a depicts the surface structure as it might appear when viewed
with the microscope; the luster or texture of each grain depends on its reflectance
properties. A photomicrograph of a polycrystalline specimen exhibiting these characteristics is shown in Figure 4.13c.
Also, small grooves form along grain boundaries as a consequence of etching.
Since atoms along grain boundary regions are more chemically active, they dissolve
at a greater rate than those within the grains. These grooves become discernible
when viewed under a microscope because they reflect light at an angle different
from that of the grains themselves; this effect is displayed in Figure 4.14a. Figure
4.14b is a photomicrograph of a polycrystalline specimen in which the grain boundary grooves are clearly visible as dark lines.
When the microstructure of a two-phase alloy is to be examined, an etchant is
often chosen that produces a different texture for each phase so that the different
phases may be distinguished from each other.
Electron Microscopy
The upper limit to the magnification possible with an optical microscope is approximately 2000 times. Consequently, some structural elements are too fine or
small to permit observation using optical microscopy. Under such circumstances
the electron microscope, which is capable of much higher magnifications, may be
employed.
An image of the structure under investigation is formed using beams of electrons instead of light radiation. According to quantum mechanics, a high-velocity
electron will become wave-like, having a wavelength that is inversely proportional
to its velocity. When accelerated across large voltages, electrons can be made to
have wavelengths on the order of 0.003 nm (3 pm). High magnifications and resolving powers of these microscopes are consequences of the short wavelengths of
electron beams. The electron beam is focused and the image formed with magnetic
lenses; otherwise the geometry of the microscope components is essentially the same
as with optical systems. Both transmission and reflection beam modes of operation
are possible for electron microscopes.
1496T_c04_80-108 11/14/05 9:53 Page 99
REVISED PAGES
4.10 Microscopic Techniques • 99
(a)
Microscope
Polished and
etched surface
Figure 4.13
(a) Polished and etched
grains as they might
appear when viewed
with an optical
microscope. (b) Section
taken through these
grains showing how the
etching characteristics
and resulting surface
texture vary from
grain to grain because
of differences in
crystallographic
orientation.
(c) Photomicrograph
of a polycrystalline
brass specimen. 60.
(Photomicrograph
courtesy of J. E. Burke,
General Electric Co.)
(b)
(c)
Transmission Electron Microscopy
transmission electron
microscope (TEM)
The image seen with a transmission electron microscope (TEM) is formed by an
electron beam that passes through the specimen. Details of internal microstructural
features are accessible to observation; contrasts in the image are produced by differences in beam scattering or diffraction produced between various elements of
the microstructure or defect. Since solid materials are highly absorptive to electron
beams, a specimen to be examined must be prepared in the form of a very thin foil;
1496T_c04_80-108 11/14/05 9:53 Page 100
REVISED PAGES
100 • Chapter 4 / Imperfections in Solids
Microscope
Polished and
etched surface
Surface
groove
Figure 4.14 (a) Section of a grain
boundary and its surface groove
produced by etching; the light reflection
characteristics in the vicinity of the groove
are also shown. (b) Photomicrograph of
the surface of a polished and etched
polycrystalline specimen of an ironchromium alloy in which the grain
boundaries appear dark. 100.
[Photomicrograph courtesy of L. C. Smith
and C. Brady, the National Bureau of
Standards, Washington, DC (now the
National Institute of Standards and
Technology, Gaithersburg, MD.)]
Grain boundary
(a)
(b)
this ensures transmission through the specimen of an appreciable fraction of the
incident beam. The transmitted beam is projected onto a fluorescent screen or a
photographic film so that the image may be viewed. Magnifications approaching
1,000,000 are possible with transmission electron microscopy, which is frequently
utilized in the study of dislocations.
Scanning Electron Microscopy
scanning electron
microscope (SEM)
A more recent and extremely useful investigative tool is the scanning electron
microscope (SEM). The surface of a specimen to be examined is scanned with an
electron beam, and the reflected (or back-scattered) beam of electrons is collected,
then displayed at the same scanning rate on a cathode ray tube (similar to a CRT
television screen). The image on the screen, which may be photographed, represents
the surface features of the specimen. The surface may or may not be polished and
etched, but it must be electrically conductive; a very thin metallic surface coating
1496T_c04_80-108 11/14/05 12:38 Page 101
REVISED PAGES
4.10 Microscopic Techniques • 101
must be applied to nonconductive materials. Magnifications ranging from 10 to in
excess of 50,000 times are possible, as are also very great depths of field. Accessory
equipment permits qualitative and semiquantitative analysis of the elemental composition of very localized surface areas.
Scanning Probe Microscopy
scanning probe
microscope (SPM)
In the past decade and a half, the field of microscopy has experienced a revolution
with the development of a new family of scanning probe microscopes. This scanning
probe microscope (SPM), of which there are several varieties, differs from the optical and electron microscopes in that neither light nor electrons is used to form an
image. Rather, the microscope generates a topographical map, on an atomic scale,
that is a representation of surface features and characteristics of the specimen being examined. Some of the features that differentiate the SPM from other microscopic techniques are as follows:
• Examination on the nanometer scale is possible inasmuch as magnifications
as high as 109 are possible; much better resolutions are attainable than with
other microscopic techniques.
• Three-dimensional magnified images are generated that provide topographical information about features of interest.
• Some SPMs may be operated in a variety of environments (e.g., vacuum, air,
liquid); thus, a particular specimen may be examined in its most suitable
environment.
Scanning probe microscopes employ a tiny probe with a very sharp tip that is
brought into very close proximity (i.e., to within on the order of a nanometer) of the
specimen surface. This probe is then raster-scanned across the plane of the surface.
During scanning, the probe experiences deflections perpendicular to this plane, in response to electronic or other interactions between the probe and specimen surface.
The in-surface-plane and out-of-plane motions of the probe are controlled by piezoelectric (Section 18.25) ceramic components that have nanometer resolutions. Furthermore, these probe movements are monitored electronically, and transferred to
and stored in a computer, which then generates the three-dimensional surface image.
Specific scanning probe microscopic techniques differ from one another with
regard to the type of interaction that is monitored. A scanning probe micrograph
in which may be observed the atomic structure and a missing atom on the surface
of silicon is shown in the chapter-opening photograph for this chapter.
These new SPMs, which allow examination of the surface of materials at the
atomic and molecular level, have provided a wealth of information about a host of
materials, from integrated circuit chips to biological molecules. Indeed, the advent
of the SPMs has helped to usher in the era of nanomaterials—materials whose properties are designed by engineering atomic and molecular structures.
Figure 4.15a is a bar-chart showing dimensional size ranges for several types of
structures found in materials (note that the axes are scaled logarithmically). Likewise, the useful dimensional resolution ranges for the several microscopic techniques
discussed in this chapter (plus the naked eye) are presented in the bar-chart of
Figure 4.15b. For three of these techniques (viz. SPM, TEM, and SEM), an upper
resolution value is not imposed by the characteristics of the microscope, and,
therefore, is somewhat arbitrary and not well defined. Furthermore, by comparing
Figures 4.15a and 4.15b, it is possible to decide which microscopic technique(s) is
(are) best suited for examination of each of the structure types.
1496T_c04_80-108 11/14/05 9:53 Page 102
REVISED PAGES
102 • Chapter 4 / Imperfections in Solids
Dimensions of structural feature (m)
1014
1012
1010
108
106
104
102
Subatomic particles
Atom/ion diameters
Unit cell edge lengths
Dislocations (width)
Second phase particles
Grains
Macrostructural features
(porosity, voids, cracks)
106
104
102
102
1
104
106
108
Dimensions of structural feature (nm)
Useful resolution ranges (m)
1012
1010
108
106
104
102
1
Scanning probe microscopes
Transmission electron microscopes
Scanning electron microscopes
Optical microscopes
Naked eye
102
1
102
104
106
108
Useful resolution ranges (nm)
Figure 4.15 (a) Bar-chart showing size ranges for several structural features found in
materials. (b) Bar-chart showing the useful resolution ranges for four microscopic
techniques discussed in this chapter, in addition to the naked eye. (Courtesy of Prof. Sidnei
Paciornik, DCMM PUC-Rio, Rio de Janeiro, Brazil, and Prof. Carlos Pérez Bergmann,
Federal University of Rio Grande do Sul, Porto Alegre, Brazil.)
4.11 GRAIN SIZE DETERMINATION
grain size
The grain size is often determined when the properties of a polycrystalline material are under consideration. In this regard, there exist a number of techniques by
which size is specified in terms of average grain volume, diameter, or area. Grain
size may be estimated by using an intercept method, described as follows. Straight
lines all the same length are drawn through several photomicrographs that show
the grain structure. The grains intersected by each line segment are counted; the
line length is then divided by an average of the number of grains intersected, taken
over all the line segments. The average grain diameter is found by dividing this result
by the linear magnification of the photomicrographs.
Probably the most common method utilized, however, is that devised by the
American Society for Testing and Materials (ASTM).7 The ASTM has prepared
several standard comparison charts, all having different average grain sizes. To each
7
ASTM Standard E 112, “Standard Methods for Estimating the Average Grain Size for
Metals.”
1496T_c04_80-108 11/14/05 9:53 Page 103
REVISED PAGES
4.11 Grain Size Determination • 103
Relationship
between ASTM
grain size number
and number of
grains per square
inch (at 100)
is assigned a number ranging from 1 to 10, which is termed the grain size number.
A specimen must be properly prepared to reveal the grain structure, which is photographed at a magnification of 100. Grain size is expressed as the grain size number of the chart that most nearly matches the grains in the micrograph. Thus, a relatively simple and convenient visual determination of grain size number is possible.
Grain size number is used extensively in the specification of steels.
The rationale behind the assignment of the grain size number to these various
charts is as follows. Let n represent the grain size number, and N the average number
of grains per square inch at a magnification of 100. These two parameters are
related to each other through the expression
N 2n1
(4.16)
Concept Check 4.2
Does the grain size number (n of Equation 4.16) increase or decrease with decreasing
grain size? Why?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
EXAMPLE PROBLEM 4.4
Computations of ASTM Grain Size Number and Number of
Grains Per Unit Area
(a) Determine the ASTM grain size number of a metal specimen if 45 grains
per square inch are measured at a magnification of 100.
(b) For this same specimen, how many grains per square inch will there be at
a magnification of 85?
Solution
(a) In order to determine the ASTM grain size number (n) it is necessary
to employ Equation 4.16. Taking logarithms of both sides of this expression
leads to
log N 1n 12 log 2
And, solving for n yields
n
log N
1
log 2
From the problem statement, N 45, and, therefore
n
log 45
1 6.5
log 2
(b) At magnifications other than 100, use of the following modified form
of Equation 4.16 is necessary:
NM a
M 2
b 2n1
100
(4.17)
1496T_c04_80-108 11/14/05 9:53 Page 104
REVISED PAGES
104 • Chapter 4 / Imperfections in Solids
In this expression NM the number of grains per square inch at magnification M. In addition, the inclusion of the 1M1002 2 term makes use of the fact
that, while magnification is a length parameter, area is expressed in terms of
units of length squared. As a consequence, the number of grains per unit area
increases with the square of the increase in magnification.
Solving Equation 4.17 for NM, realizing that M 85 and n 6.5 leads to
100 2
b
M
100 2
b 62.6 grains/in.2
216.512 a
85
NM 2n1 a
SUMMARY
Vacancies and Self-Interstitials
All solid materials contain large numbers of imperfections or deviations from crystalline perfection. The several types of imperfection are categorized on the basis of
their geometry and size. Point defects are those associated with one or two atomic
positions, including vacancies (or vacant lattice sites), self-interstitials (host atoms
that occupy interstitial sites), and impurity atoms.
Impurities in Solids
A solid solution may form when impurity atoms are added to a solid, in which case
the original crystal structure is retained and no new phases are formed. For substitutional solid solutions, impurity atoms substitute for host atoms, and appreciable
solubility is possible only when atomic diameters and electronegativities for both
atom types are similar, when both elements have the same crystal structure, and
when the impurity atoms have a valence that is the same as or less than the host
material. Interstitial solid solutions form for relatively small impurity atoms that
occupy interstitial sites among the host atoms.
Specification of Composition
Composition of an alloy may be specified in weight percent or atom percent. The
basis for weight percent computations is the weight (or mass) of each alloy constituent relative to the total alloy weight. Atom percents are calculated in terms of
the number of moles for each constituent relative to the total moles of all the elements in the alloy.
Dislocations—Linear Defects
Dislocations are one-dimensional crystalline defects of which there are two pure
types: edge and screw. An edge may be thought of in terms of the lattice distortion
along the end of an extra half-plane of atoms; a screw, as a helical planar ramp. For
mixed dislocations, components of both pure edge and screw are found. The magnitude and direction of lattice distortion associated with a dislocation are specified
by its Burgers vector. The relative orientations of Burgers vector and dislocation
line are (1) perpendicular for edge, (2) parallel for screw, and (3) neither perpendicular nor parallel for mixed.
1496T_c04_80-108 11/14/05 9:53 Page 105
REVISED PAGES
References • 105
Interfacial Defects
Bulk or Volume Defects
Atomic Vibrations
Other imperfections include interfacial defects [external surfaces, grain boundaries
(both small- and high-angle), twin boundaries, etc.], volume defects (cracks, pores,
etc.), and atomic vibrations. Each type of imperfection has some influence on the
properties of a material.
Microscopic Techniques
Many of the important defects and structural elements of materials are of microscopic dimensions, and observation is possible only with the aid of a microscope.
Both optical and electron microscopes are employed, usually in conjunction with
photographic equipment. Transmissive and reflective modes are possible for each
microscope type; preference is dictated by the nature of the specimen as well as the
structural element or defect to be examined.
More recent scanning probe microscopic techniques have been developed that
generate topographical maps representing the surface features and characteristics
of the specimen. Examinations on the atomic and molecular levels are possible using
these techniques.
Grain Size Determination
Grain size of polycrystalline materials is frequently determined using photomicrographic techniques. Two methods are commonly employed: intercept and standard
comparison charts.
I M P O R TA N T T E R M S A N D C O N C E P T S
Alloy
Atom percent
Atomic vibration
Boltzmann’s constant
Burgers vector
Composition
Dislocation line
Edge dislocation
Grain size
Imperfection
Interstitial solid solution
Microscopy
Microstructure
Mixed dislocation
Photomicrograph
Point defect
Scanning electron microscope (SEM)
Scanning probe microscope (SPM)
Screw dislocation
Self-interstitial
Solid solution
Solute
Solvent
Substitutional solid solution
Transmission electron microscope (TEM)
Vacancy
Weight percent
REFERENCES
ASM Handbook, Vol. 9, Metallography and Microstructures, ASM International, Materials Park,
OH, 2004.
Brandon, D., and W. D. Kaplan, Microstructural
Characterization of Materials, Wiley, New York,
1999.
DeHoff, R. T., and F. N. Rhines, Quantitative Microscopy, TechBooks, Marietta, OH, 1991.
Van Bueren, H. G., Imperfections in Crystals, NorthHolland, Amsterdam (Wiley-Interscience, New
York), 1960.
Vander Voort, G. F., Metallography, Principles and
Practice, ASM International, Materials Park,
OH, 1984.
1496T_c04_80-108 11/14/05 9:53 Page 106
REVISED PAGES
106 • Chapter 4 / Imperfections in Solids
QUESTIONS AND PROBLEMS
Vacancies and Self-Interstitials
4.1 Calculate the fraction of atom sites that are
vacant for copper at its melting temperature
of 1084C (1357 K). Assume an energy for
vacancy formation of 0.90 eV/atom.
4.2 Calculate the number of vacancies per cubic
meter in gold at 900C. The energy for vacancy
formation is 0.98 eV/atom. Furthermore, the
density and atomic weight forAu are 18.63 g/cm3
(at 900°C) and 196.9 g/mol, respectively.
4.3 Calculate the energy for vacancy formation in
silver, given that the equilibrium number of
vacancies at 800C (1073 K) is 3.6 1023 m3.
The atomic weight and density (at 800C)
for silver are, respectively, 107.9 g/mol and
9.5 g/cm3.
Impurities in Solids
4.4 Below, atomic radius, crystal structure, electronegativity, and the most common valence
are tabulated, for several elements; for those
that are nonmetals, only atomic radii are
indicated.
Element
Atomic
Radius
(nm)
Ni
C
H
O
Ag
Al
Co
Cr
Fe
Pt
Zn
0.1246
0.071
0.046
0.060
0.1445
0.1431
0.1253
0.1249
0.1241
0.1387
0.1332
Crystal
Structure
Electronegativity
Valence
FCC
1.8
2
FCC
FCC
HCP
BCC
BCC
FCC
HCP
1.9
1.5
1.8
1.6
1.8
2.2
1.6
1
3
2
3
2
2
2
Which of these elements would you expect to
form the following with nickel:
(a) A substitutional solid solution having
complete solubility
(b) A substitutional solid solution of incomplete solubility
(c) An interstitial solid solution
4.5 For both FCC and BCC crystal structures,
there are two different types of interstitial
sites. In each case, one site is larger than the
other, and is normally occupied by impurity
atoms. For FCC, this larger one is located at
the center of each edge of the unit cell; it is
termed an octahedral interstitial site. On the
other hand, with BCC the larger site type is
found at 0 12 14 positions—that is, lying on {100}
faces, and situated midway between two unit
cell edges on this face and one-quarter of the
distance between the other two unit cell edges;
it is termed a tetrahedral interstitial site. For
both FCC and BCC crystal structures, compute the radius r of an impurity atom that will
just fit into one of these sites in terms of the
atomic radius R of the host atom.
Specification of Composition
4.6 Derive the following equations:
(a) Equation 4.7a
(b) Equation 4.9a
(c) Equation 4.10a
(d) Equation 4.11b
4.7 What is the composition, in atom percent, of
an alloy that consists of 92.5 wt% Ag and
7.5 wt% Cu?
4.8 What is the composition, in weight percent,
of an alloy that consists of 5 at% Cu and
95 at% Pt?
4.9 Calculate the composition, in weight percent,
of an alloy that contains 105 kg of iron, 0.2 kg
of carbon, and 1.0 kg of chromium.
4.10 What is the composition, in atom percent, of an
alloy that contains 33 g copper and 47 g zinc?
4.11 What is the composition, in atom percent,
of an alloy that contains 44.5 lbm of silver,
83.7 lbm of gold, and 5.3 lbm of Cu?
4.12 What is the composition, in atom percent,
of an alloy that consists of 5.5 wt% Pb and
94.5 wt% Sn?
4.13 Convert the atom percent composition in
Problem 4.11 to weight percent.
4.14 Calculate the number of atoms per cubic
meter in lead.
4.15 The concentration of silicon in an iron-silicon
alloy is 0.25 wt%. What is the concentration in
kilograms of silicon per cubic meter of alloy?
1496T_c04_80-108 12/20/05 7:46 Page 107
2nd REVISE PAGES
Questions and Problems • 107
4.16 Determine the approximate density of a
Ti-6Al-4V titanium alloy that has a composition of 90 wt% Ti, 6 wt% Al, and 4 wt% V.
4.17 Calculate the unit cell edge length for an
80 wt% Ag-20 wt% Pd alloy. All of the palladium is in solid solution, the crystal structure for this alloy is FCC, and the roomtemperature density of Pd is 12.02 g/cm3.
4.18 Some hypothetical alloy is composed of
25 wt% of metal A and 75 wt% of metal B.
If the densities of metals A and B are 6.17 and
8.00 g/cm3, respectively, whereas their respective atomic weights are 171.3 and 162.0 g/mol,
determine whether the crystal structure for
this alloy is simple cubic, face-centered cubic,
or body-centered cubic. Assume a unit cell
edge length of 0.332 nm.
4.19 For a solid solution consisting of two elements
(designated as 1 and 2), sometimes it is desirable to determine the number of atoms per cubic centimeter of one element in a solid solution, N1, given the concentration of that element
specified in weight percent, C1. This computation is possible using the following expression:
N1
NAC1
C1A1
A1
1100 C1 2
r1
r2
(4.18)
where
NA Avogadro’s number
r1 and r2 densities of the two elements
A1 the atomic weight of element 1
Derive Equation 4.18 using Equation 4.2 and
expressions contained in Section 4.4.
4.20 Molybdenum forms a substitutional solid
solution with tungsten. Compute the number
of molybdenum atoms per cubic centimeter
for a molybdenum-tungsten alloy that contains 16.4 wt% Mo and 83.6 wt% W. The densities of pure molybdenum and tungsten are
10.22 and 19.30 g/cm3, respectively.
4.21 Niobium forms a substitutional solid solution with vanadium. Compute the number of
niobium atoms per cubic centimeter for a
niobium-vanadium alloy that contains 24
wt% Nb and 76 wt% V. The densities of pure
niobium and vanadium are 8.57 and 6.10
g/cm3, respectively.
4.22 Sometimes it is desirable to be able to determine the weight percent of one element, C1,
that will produce a specified concentration in
terms of the number of atoms per cubic centimeter, N1, for an alloy composed of two
types of atoms. This computation is possible
using the following expression:
C1
100
NA r2
r2
1
r1
N1A1
(4.19)
where
NA Avogadro’s number
r1 and r2 densities of the two elements
A1 and A2 the atomic weights of the two
elements
Derive Equation 4.19 using Equation 4.2 and
expressions contained in Section 4.4.
4.23 Gold forms a substitutional solid solution
with silver. Compute the weight percent of
gold that must be added to silver to yield an
alloy that contains 5.5 1021 Au atoms per
cubic centimeter. The densities of pure Au
and Ag are 19.32 and 10.49 g/cm3, respectively.
4.24 Germanium forms a substitutional solid solution with silicon. Compute the weight percent
of germanium that must be added to silicon to
yield an alloy that contains 2.43 1021 Ge
atoms per cubic centimeter. The densities of
pure Ge and Si are 5.32 and 2.33 g/cm3,
respectively.
4.25 Iron and vanadium both have the BCC crystal structure, and V forms a substitutional
solid solution for concentrations up to approximately 20 wt% V at room temperature.
Compute the unit cell edge length for a
90 wt% Fe–10 wt% V alloy.
Dislocations—Linear Defects
4.26 Cite the relative Burgers vector–dislocation
line orientations for edge, screw, and mixed
dislocations.
1496T_c04_80-108 11/14/05 9:53 Page 108
REVISED PAGES
108 • Chapter 4 / Imperfections in Solids
Interfacial Defects
Grain Size Determination
4.27 For an FCC single crystal, would you expect the
surface energy for a (100) plane to be greater
or less than that for a (111) plane? Why? (Note:
You may want to consult the solution to Problem 3.53 at the end of Chapter 3.)
4.28 For a BCC single crystal, would you expect the
surface energy for a (100) plane to be greater
or less than that for a (110) plane? Why?
(Note: You may want to consult the solution
to Problem 3.54 at the end of Chapter 3.)
4.29 (a) For a given material, would you expect the
surface energy to be greater than, the same as,
or less than the grain boundary energy? Why?
(b) The grain boundary energy of a smallangle grain boundary is less than for a highangle one. Why is this so?
4.30 (a) Briefly describe a twin and a twin
boundary.
(b) Cite the difference between mechanical
and annealing twins.
4.31 For each of the following stacking sequences
found in FCC metals, cite the type of planar
defect that exists:
(a) . . . A B C A B C B A C B A . . .
(b) . . . A B C A B C B C A B C . . .
Now, copy the stacking sequences and indicate the position(s) of planar defect(s) with a
vertical dashed line.
4.32 (a) Using the intercept method, determine
the average grain size, in millimeters, of the
specimen whose microstructure is shown in
Figure 4.14(b); use at least seven straight-line
segments.
(b) Estimate the ASTM grain size number
for this material.
4.33 (a) Employing the intercept technique, determine the average grain size for the steel
specimen whose microstructure is shown in
Figure 9.25(a); use at least seven straight-line
segments.
(b) Estimate the ASTM grain size number
for this material.
4.34 For an ASTM grain size of 6, approximately
how many grains would there be per square
inch at
(a) a magnification of 100, and
(b) without any magnification?
4.35 Determine the ASTM grain size number if
30 grains per square inch are measured at a
magnification of 250.
4.36 Determine the ASTM grain size number if
25 grains per square inch are measured at a
magnification of 75.
DESIGN PROBLEMS
Specification of Composition
4.D1 Aluminum–lithium alloys have been developed by the aircraft industry to reduce the
weight and improve the performance of its
aircraft. A commercial aircraft skin material
having a density of 2.47 g/cm3 is desired. Compute the concentration of Li (in wt%) that is
required.
4.D2 Copper and platinum both have the FCC
crystal structure, and Cu forms a substitutional solid solution for concentrations up
to approximately 6 wt% Cu at room temperature. Determine the concentration in
weight percent of Cu that must be added to
platinum to yield a unit cell edge length of
0.390 nm.
1496T_c05_109-130 12/20/05 7:52 Page 109
2nd REVISE PAGES
Chapter
P
5
Diffusion
hotograph of a steel gear that
has been “case hardened.” The outer
surface layer was selectively hardened
by a high-temperature heat treatment
during which carbon from the surrounding atmosphere diffused into
the surface. The “case” appears as the
dark outer rim of that segment of the
gear that has been sectioned. Actual
size. (Photograph courtesy of Surface
Division Midland-Ross.)
WHY STUDY Diffusion?
Materials of all types are often heat treated to improve their properties. The phenomena that occur
during a heat treatment almost always involve atomic
diffusion. Often an enhancement of diffusion rate is
desired; on occasion measures are taken to reduce it.
Heat-treating temperatures and times, and/or cooling
rates are often predictable using the mathematics of
diffusion and appropriate diffusion constants. The
steel gear shown on this page has been case hardened
(Section 8.10); that is, its hardness and resistance to
failure by fatigue have been enhanced by diffusing
excess carbon or nitrogen into the outer surface layer.
• 109
1496T_c05_109-130 12/20/05 7:52 Page 110
2nd REVISE PAGES
Learning Objectives
After studying this chapter you should be able to do the following:
1. Name and describe the two atomic mechanisms
4. Write the solution to Fick’s second law for
of diffusion.
diffusion into a semi-infinite solid when the
2. Distinguish between steady-state and
concentration of diffusing species at the surface
nonsteady-state diffusion.
is held constant. Define all parameters in this
3. (a) Write Fick’s first and second laws in equaequation.
tion form, and define all parameters.
5. Calculate the diffusion coefficient for some
(b) Note the kind of diffusion for which each
material at a specified temperature, given the
of these equations is normally applied.
appropriate diffusion constants.
5.1 INTRODUCTION
Cu
Ni
(a)
(b)
Concentration of Ni, Cu
diffusion
Many reactions and processes that are important in the treatment of materials rely
on the transfer of mass either within a specific solid (ordinarily on a microscopic level)
or from a liquid, a gas, or another solid phase. This is necessarily accomplished by
diffusion, the phenomenon of material transport by atomic motion. This chapter discusses the atomic mechanisms by which diffusion occurs, the mathematics of diffusion, and the influence of temperature and diffusing species on the rate of diffusion.
The phenomenon of diffusion may be demonstrated with the use of a diffusion
couple, which is formed by joining bars of two different metals together so that
there is intimate contact between the two faces; this is illustrated for copper and
nickel in Figure 5.1, which includes schematic representations of atom positions and
composition across the interface. This couple is heated for an extended period at
an elevated temperature (but below the melting temperature of both metals), and
100
Cu
Ni
0
Position
(c)
Figure 5.1 (a) A copper–nickel diffusion couple
before a high-temperature heat treatment.
(b) Schematic representations of Cu (red
circles) and Ni (blue circles) atom locations within
the diffusion couple. (c) Concentrations of copper
and nickel as a function of position across the
couple.
1496T_c05_109-130 11/14/05 13:07 Page 111
REVISED PAGES
5.2 Diffusion Mechanisms • 111
Diffusion of Cu atoms
Cu
Cu-Ni alloy
Diffusion of Ni atoms
(a)
Ni
Figure 5.2 (a) A copper–nickel diffusion couple
after a high-temperature heat treatment, showing the
alloyed diffusion zone. (b) Schematic representations
of Cu (red circles) and Ni (blue circles) atom
locations within the couple. (c) Concentrations of
copper and nickel as a function of position across
the couple.
Concentration of Ni, Cu
(b)
100
Cu
Ni
0
Position
(c)
interdiffusion
impurity diffusion
self-diffusion
cooled to room temperature. Chemical analysis will reveal a condition similar to
that represented in Figure 5.2—namely, pure copper and nickel at the two extremities of the couple, separated by an alloyed region. Concentrations of both metals
vary with position as shown in Figure 5.2c. This result indicates that copper atoms
have migrated or diffused into the nickel, and that nickel has diffused into copper.
This process, whereby atoms of one metal diffuse into another, is termed interdiffusion, or impurity diffusion.
Interdiffusion may be discerned from a macroscopic perspective by changes in
concentration which occur over time, as in the example for the Cu–Ni diffusion couple. There is a net drift or transport of atoms from high- to low-concentration regions. Diffusion also occurs for pure metals, but all atoms exchanging positions are
of the same type; this is termed self-diffusion. Of course, self-diffusion is not normally subject to observation by noting compositional changes.
5.2 DIFFUSION MECHANISMS
From an atomic perspective, diffusion is just the stepwise migration of atoms from
lattice site to lattice site. In fact, the atoms in solid materials are in constant motion, rapidly changing positions. For an atom to make such a move, two conditions
must be met: (1) there must be an empty adjacent site, and (2) the atom must have
sufficient energy to break bonds with its neighbor atoms and then cause some lattice distortion during the displacement. This energy is vibrational in nature (Section 4.8). At a specific temperature some small fraction of the total number of atoms
is capable of diffusive motion, by virtue of the magnitudes of their vibrational
energies. This fraction increases with rising temperature.
Several different models for this atomic motion have been proposed; of these
possibilities, two dominate for metallic diffusion.
1496T_c05_109-130 11/14/05 10:12 Page 112
REVISED PAGES
112 • Chapter 5 / Diffusion
Figure 5.3
Schematic
representations of
(a) vacancy diffusion
and (b) interstitial
diffusion.
Motion of a host or
substitutional atom
Vacancy
Vacancy
(a )
Position of interstitial
atom before diffusion
Position of interstitial
atom after diffusion
(b)
Vacancy Diffusion
vacancy diffusion
One mechanism involves the interchange of an atom from a normal lattice position
to an adjacent vacant lattice site or vacancy, as represented schematically in Figure 5.3a. This mechanism is aptly termed vacancy diffusion. Of course, this process
necessitates the presence of vacancies, and the extent to which vacancy diffusion
can occur is a function of the number of these defects that are present; significant
concentrations of vacancies may exist in metals at elevated temperatures (Section 4.2).
Since diffusing atoms and vacancies exchange positions, the diffusion of atoms in
one direction corresponds to the motion of vacancies in the opposite direction. Both
self-diffusion and interdiffusion occur by this mechanism; for the latter, the impurity atoms must substitute for host atoms.
Interstitial Diffusion
interstitial diffusion
The second type of diffusion involves atoms that migrate from an interstitial position to a neighboring one that is empty. This mechanism is found for interdiffusion
of impurities such as hydrogen, carbon, nitrogen, and oxygen, which have atoms
that are small enough to fit into the interstitial positions. Host or substitutional impurity atoms rarely form interstitials and do not normally diffuse via this mechanism. This phenomenon is appropriately termed interstitial diffusion (Figure 5.3b).
In most metal alloys, interstitial diffusion occurs much more rapidly than diffusion by the vacancy mode, since the interstitial atoms are smaller and thus more mobile. Furthermore, there are more empty interstitial positions than vacancies; hence,
the probability of interstitial atomic movement is greater than for vacancy diffusion.
5.3 STEADY-STATE DIFFUSION
Diffusion is a time-dependent process—that is, in a macroscopic sense, the quantity of an element that is transported within another is a function of time. Often it
1496T_c05_109-130 11/14/05 10:12 Page 113
REVISED PAGES
5.3 Steady-State Diffusion • 113
diffusion flux
is necessary to know how fast diffusion occurs, or the rate of mass transfer. This
rate is frequently expressed as a diffusion flux (J), defined as the mass (or, equivalently, the number of atoms) M diffusing through and perpendicular to a unit
cross-sectional area of solid per unit of time. In mathematical form, this may be
represented as
Definition of
diffusion flux
J
M
At
(5.1a)
where A denotes the area across which diffusion is occurring and t is the elapsed
diffusion time. In differential form, this expression becomes
J
steady-state diffusion
concentration profile
concentration
gradient
1 dM
A dt
(5.1b)
The units for J are kilograms or atoms per meter squared per second (kg/m2-s or
atoms/m2-s).
If the diffusion flux does not change with time, a steady-state condition exists.
One common example of steady-state diffusion is the diffusion of atoms of a gas
through a plate of metal for which the concentrations (or pressures) of the diffusing species on both surfaces of the plate are held constant. This is represented
schematically in Figure 5.4a.
When concentration C is plotted versus position (or distance) within the solid
x, the resulting curve is termed the concentration profile; the slope at a particular
point on this curve is the concentration gradient:
concentration gradient
dC
dx
(5.2a)
In the present treatment, the concentration profile is assumed to be linear, as
depicted in Figure 5.4b, and
CA CB
¢C
concentration gradient
(5.2b)
xA xB
¢x
PA > PB
and constant
Thin metal plate
Gas at
pressure PB
Gas at
pressure PA
Direction of
diffusion of
gaseous species
Concentration of diffusing species, C
Figure 5.4
(a) Steady-state
diffusion across a thin
plate. (b) A linear
concentration profile
for the diffusion
situation in (a).
CA
CB
xA
xB
Position, x
Area, A
(a)
(b)
1496T_c05_109-130 11/14/05 10:12 Page 114
REVISED PAGES
114 • Chapter 5 / Diffusion
For diffusion problems, it is sometimes convenient to express concentration in terms
of mass of diffusing species per unit volume of solid (kg/m3 or g/cm3).1
The mathematics of steady-state diffusion in a single (x) direction is relatively
simple, in that the flux is proportional to the concentration gradient through the
expression
Fick’s first law—
diffusion flux for
steady-state diffusion
(in one direction)
diffusion coefficient
Fick’s first law
driving force
J D
dC
dx
(5.3)
The constant of proportionality D is called the diffusion coefficient, which is expressed in square meters per second. The negative sign in this expression indicates
that the direction of diffusion is down the concentration gradient, from a high to a
low concentration. Equation 5.3 is sometimes called Fick’s first law.
Sometimes the term driving force is used in the context of what compels a reaction to occur. For diffusion reactions, several such forces are possible; but when
diffusion is according to Equation 5.3, the concentration gradient is the driving force.
One practical example of steady-state diffusion is found in the purification of
hydrogen gas. One side of a thin sheet of palladium metal is exposed to the impure
gas composed of hydrogen and other gaseous species such as nitrogen, oxygen, and
water vapor. The hydrogen selectively diffuses through the sheet to the opposite
side, which is maintained at a constant and lower hydrogen pressure.
EXAMPLE PROBLEM 5.1
Diffusion Flux Computation
A plate of iron is exposed to a carburizing (carbon-rich) atmosphere on one
side and a decarburizing (carbon-deficient) atmosphere on the other side at
700C (1300F). If a condition of steady state is achieved, calculate the diffusion flux of carbon through the plate if the concentrations of carbon at positions
of 5 and 10 mm (5 103 and 102 m) beneath the carburizing surface are 1.2
and 0.8 kg/m3, respectively. Assume a diffusion coefficient of 3 1011 m2/s
at this temperature.
Solution
Fick’s first law, Equation 5.3, is utilized to determine the diffusion flux. Substitution of the values above into this expression yields
J D
11.2 0.82 kg/m3
CA CB
13 1011 m2/s2
xA xB
15 103 102 2 m
2.4 109 kg/m2-s
5.4 NONSTEADY-STATE DIFFUSION
Most practical diffusion situations are nonsteady-state ones. That is, the diffusion
flux and the concentration gradient at some particular point in a solid vary with
time, with a net accumulation or depletion of the diffusing species resulting. This is
illustrated in Figure 5.5, which shows concentration profiles at three different
1
Conversion of concentration from weight percent to mass per unit volume (in kg/m3) is
possible using Equation 4.9.
1496T_c05_109-130 11/14/05 10:12 Page 115
REVISED PAGES
Concentration of diffusing species
5.4 Nonsteady-State Diffusion • 115
t3 > t2 > t1
Figure 5.5 Concentration profiles for
nonsteady-state diffusion taken at three
different times, t1, t2, and t3.
t3
t2
t1
Distance
diffusion times. Under conditions of nonsteady state, use of Equation 5.3 is no longer
convenient; instead, the partial differential equation
0C
0
0C
aD b
0t
0x
0x
Fick’s second law
Fick’s second law—
diffusion equation
for nonsteady-state
diffusion (in one
direction)
(5.4a)
known as Fick’s second law, is used. If the diffusion coefficient is independent
of composition (which should be verified for each particular diffusion situation),
Equation 5.4a simplifies to
0C
0 2C
D 2
0t
0x
(5.4b)
Solutions to this expression (concentration in terms of both position and time) are
possible when physically meaningful boundary conditions are specified. Comprehensive collections of these are given by Crank, and Carslaw and Jaeger (see References).
One practically important solution is for a semi-infinite solid2 in which the surface concentration is held constant. Frequently, the source of the diffusing species
is a gas phase, the partial pressure of which is maintained at a constant value. Furthermore, the following assumptions are made:
1. Before diffusion, any of the diffusing solute atoms in the solid are uniformly
distributed with concentration of C0.
2. The value of x at the surface is zero and increases with distance into the solid.
3. The time is taken to be zero the instant before the diffusion process begins.
These boundary conditions are simply stated as
For t 0, C C0 at 0 x q
For t 7 0, C Cs 1the constant surface concentration2 at x 0
C C0 at x q
Application of these boundary conditions to Equation 5.4b yields the solution
Solution to Fick’s
second law for the
condition of constant
surface concentration
(for a semi-infinite
solid)
Cx C0
x
1 erf a
b
Cs C0
21Dt
2
(5.5)
A bar of solid is considered to be semi-infinite if none of the diffusing atoms reaches the
bar end during the time over which diffusion takes place. A bar of length l is considered to
be semi-infinite when l 7 101Dt.
1496T_c05_109-130 11/14/05 10:12 Page 116
REVISED PAGES
116 • Chapter 5 / Diffusion
Table 5.1 Tabulation of Error Function Values
z
erf(z)
z
erf(z)
z
erf(z)
0
0.025
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.50
0
0.0282
0.0564
0.1125
0.1680
0.2227
0.2763
0.3286
0.3794
0.4284
0.4755
0.5205
0.55
0.60
0.65
0.70
0.75
0.80
0.85
0.90
0.95
1.0
1.1
1.2
0.5633
0.6039
0.6420
0.6778
0.7112
0.7421
0.7707
0.7970
0.8209
0.8427
0.8802
0.9103
1.3
1.4
1.5
1.6
1.7
1.8
1.9
2.0
2.2
2.4
2.6
2.8
0.9340
0.9523
0.9661
0.9763
0.9838
0.9891
0.9928
0.9953
0.9981
0.9993
0.9998
0.9999
where Cx represents the concentration at depth x after time t. The expression
erf(x21Dt) is the Gaussian error function,3 values of which are given in mathematical tables for various x 21Dt values; a partial listing is given in Table 5.1. The
concentration parameters that appear in Equation 5.5 are noted in Figure 5.6, a
concentration profile taken at a specific time. Equation 5.5 thus demonstrates the
relationship between concentration, position, and time—namely, that Cx, being a
function of the dimensionless parameter x 1Dt, may be determined at any time
and position if the parameters C0, Cs, and D are known.
Suppose that it is desired to achieve some specific concentration of solute, C1,
in an alloy; the left-hand side of Equation 5.5 now becomes
C1 C0
constant
Cs C0
This being the case, the right-hand side of this same expression is also a constant,
and subsequently
x
constant
(5.6a)
21Dt
Figure 5.6 Concentration profile for
nonsteady-state diffusion; concentration
parameters relate to Equation 5.5.
Concentration, C
Cs
Cs – C0
Cx
Cx – C0
C0
Distance from interface, x
3
This Gaussian error function is defined by
erf1z2
2 z y2
e dy
1p 0
where x2 1Dt has been replaced by the variable z.
1496T_c05_109-130 11/14/05 10:12 Page 117
REVISED PAGES
5.4 Nonsteady-State Diffusion • 117
or
x2
constant
Dt
(5.6b)
Some diffusion computations are thus facilitated on the basis of this relationship, as demonstrated in Example Problem 5.3.
EXAMPLE PROBLEM 5.2
Nonsteady-State Diffusion Time Computation I
carburizing
For some applications, it is necessary to harden the surface of a steel (or ironcarbon alloy) above that of its interior. One way this may be accomplished is
by increasing the surface concentration of carbon in a process termed carburizing; the steel piece is exposed, at an elevated temperature, to an atmosphere rich in a hydrocarbon gas, such as methane 1CH4 2.
Consider one such alloy that initially has a uniform carbon concentration
of 0.25 wt% and is to be treated at 950 C (1750 F). If the concentration of
carbon at the surface is suddenly brought to and maintained at 1.20 wt%, how
long will it take to achieve a carbon content of 0.80 wt% at a position 0.5 mm
below the surface? The diffusion coefficient for carbon in iron at this temperature is 1.6 1011 m2/s; assume that the steel piece is semi-infinite.
Solution
Since this is a nonsteady-state diffusion problem in which the surface composition is held constant, Equation 5.5 is used. Values for all the parameters
in this expression except time t are specified in the problem as follows:
C0 0.25 wt% C
Cs 1.20 wt% C
Cx 0.80 wt% C
x 0.50 mm 5 104 m
D 1.6 1011 m2/s
Thus,
15 104 m2
Cx C0
0.80 0.25
1 erf c
d
Cs C0
1.20 0.25
2211.6 1011 m2/s21t2
62.5 s1 2
b
0.4210 erf a
1t
We must now determine from Table 5.1 the value of z for which the error
function is 0.4210. An interpolation is necessary, as
z
erf(z)
0.35
z
0.40
0.3794
0.4210
0.4284
z 0.35
0.4210 0.3794
0.40 0.35
0.4284 0.3794
1496T_c05_109-130 11/14/05 10:12 Page 118
REVISED PAGES
118 • Chapter 5 / Diffusion
or
z 0.392
Therefore,
62.5 s1 2
0.392
1t
and solving for t,
ta
62.5 s1 2 2
b 25,400 s 7.1 h
0.392
EXAMPLE PROBLEM 5.3
Nonsteady-State Diffusion Time Computation II
The diffusion coefficients for copper in aluminum at 500 and 600 C are
4.8 1014 and 5.3 1013 m2/s, respectively. Determine the approximate time
at 500 C that will produce the same diffusion result (in terms of concentration of Cu at some specific point in Al) as a 10-h heat treatment at 600°C.
Solution
This is a diffusion problem in which Equation 5.6b may be employed. The
composition in both diffusion situations will be equal at the same position (i.e.,
x is also a constant), thus
Dt constant
(5.7)
at both temperatures. That is,
D500 t 500 D600 t 600
or
t 500
15.3 1013 m2/s2110 h2
D600 t 600
110.4 h
D500
4.8 1014 m2/s
5.5 FACTORS THAT INFLUENCE DIFFUSION
Diffusing Species
The magnitude of the diffusion coefficient D is indicative of the rate at which atoms
diffuse. Coefficients, both self- and interdiffusion, for several metallic systems are listed
in Table 5.2. The diffusing species as well as the host material influence the diffusion coefficient. For example, there is a significant difference in magnitude between
self-diffusion and carbon interdiffusion in a iron at 500 C, the D value being greater
for the carbon interdiffusion (3.0 1021 vs. 2.4 1012 m2/s). This comparison also
provides a contrast between rates of diffusion via vacancy and interstitial modes
as discussed above. Self-diffusion occurs by a vacancy mechanism, whereas carbon
diffusion in iron is interstitial.
Temperature
Temperature has a most profound influence on the coefficients and diffusion rates.
For example, for the self-diffusion of Fe in a-Fe, the diffusion coefficient increases
approximately six orders of magnitude (from 3.0 1021 to 1.8 1015 m2/s2 in
rising temperature from 500 to 900 C (Table 5.2). The temperature dependence of
1496T_c05_109-130 12/20/05 7:52 Page 119
2nd REVISE PAGES
5.5 Factors That Influence Diffusion • 119
Table 5.2 A Tabulation of Diffusion Data
Diffusing
Species
Host
Metal
Activation Energy Qd
2
Calculated Values
kJ/mol
eV/atom
T(C )
D(m2/s)
4
D0(m /s)
Fe
-Fe
(BCC)
2.8 10
251
2.60
500
900
3.0 1021
1.8 1015
Fe
-Fe
(FCC)
5.0 105
284
2.94
900
1100
1.1 1017
7.8 1016
C
-Fe
6.2 107
80
0.83
500
900
2.4 1012
1.7 1010
C
-Fe
2.3 105
148
1.53
900
1100
5.9 1012
5.3 1011
Cu
Cu
7.8 105
211
2.19
500
4.2 1019
Zn
Cu
2.4 105
189
1.96
500
4.0 1018
Al
Al
4
2.3 10
144
1.49
500
4.2 1014
Cu
Al
6.5 105
136
1.41
500
4.1 1014
Mg
Al
4
1.2 10
131
1.35
500
1.9 1013
Cu
Ni
2.7 105
256
2.65
500
1.3 1022
Source: E. A. Brandes and G. B. Brook (Editors), Smithells Metals Reference Book, 7th edition, ButterworthHeinemann, Oxford, 1992.
the diffusion coefficients is
Dependence of the
diffusion coefficient
on temperature
D D0 exp a
Qd
b
RT
(5.8)
where
activation energy
D0 a temperature-independent preexponential 1m2/s2
Qd the activation energy for diffusion (J/mol or eV/atom)
R the gas constant, 8.31 J/mol-K or 8.62 105 eV/atom-K
T absolute temperature (K)
The activation energy may be thought of as that energy required to produce
the diffusive motion of one mole of atoms. A large activation energy results in a
relatively small diffusion coefficient. Table 5.2 also contains a listing of D0 and Qd
values for several diffusion systems.
Taking natural logarithms of Equation 5.8 yields
ln D ln D0
Qd 1
a b
R T
(5.9a)
Qd 1
a b
2.3 R T
(5.9b)
or in terms of logarithms to the base 10
log D log D0
Since D0, Qd, and R are all constants, Equation 5.9b takes on the form of an equation of a straight line:
y b mx
where y and x are analogous, respectively, to the variables log D and 1 T. Thus, if
log D is plotted versus the reciprocal of the absolute temperature, a straight line
1496T_c05_109-130 11/14/05 10:12 Page 120
REVISED PAGES
120 • Chapter 5 / Diffusion
Figure 5.7 Plot of
the logarithm of the
diffusion coefficient
versus the reciprocal of
absolute temperature
for several metals.
[Data taken from E. A.
Brandes and G. B.
Brook (Editors),
Smithells Metals
Reference Book, 7th
edition, ButterworthHeinemann, Oxford,
1992.]
Temperature (°C)
1500 1200 1000 800
600
500
400
300
10–8
10–10
C in ␣ –Fe
Diffusion coefficient (m2/s)
C in ␥ –Fe
10–12
Zn in Cu
10–14
Fe in ␥ –Fe
10–16
Al in Al
Fe in ␣ –Fe
Cu in Cu
10–18
10–20
0.5
1.0
1.5
2.0
Reciprocal temperature (1000/K)
should result, having slope and intercept of Qd 2.3R and log D0, respectively. This
is, in fact, the manner in which the values of Qd and D0 are determined experimentally. From such a plot for several alloy systems (Figure 5.7), it may be noted
that linear relationships exist for all cases shown.
Concept Check 5.1
Rank the magnitudes of the diffusion coefficients from greatest to least for the following systems:
N in Fe at 700°C
Cr in Fe at 700°C
N in Fe at 900°C
Cr in Fe at 900°C
Now justify this ranking. (Note: Both Fe and Cr have the BCC crystal structure, and
the atomic radii for Fe, Cr, and N are 0.124, 0.125, and 0.065 nm, respectively. You
may also want to refer to Section 4.3.)
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Concept Check 5.2
Consider the self-diffusion of two hypothetical metals A and B. On a schematic
graph of ln D versus 1 T , plot (and label) lines for both metals given that
D0 1A2 7 D0 1B2 and also that Qd 1A2 7 Qd 1B2.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
1496T_c05_109-130 11/14/05 13:07 Page 121
REVISED PAGES
5.5 Factors That Influence Diffusion • 121
EXAMPLE PROBLEM 5.4
Diffusion Coefficient Determination
Using the data in Table 5.2, compute the diffusion coefficient for magnesium
in aluminum at 550 C.
Solution
This diffusion coefficient may be determined by applying Equation 5.8; the
values of D0 and Qd from Table 5.2 are 1.2 104 m2/s and 131 kJ/mol,
respectively. Thus,
1131,000 J/mol2
D 11.2 104 m2/s2 exp c
d
18.31 J/mol-K21550 273 K2
5.8 1013 m2/s
EXAMPLE PROBLEM 5.5
Diffusion Coefficient Activation Energy and
Preexponential Calculations
In Figure 5.8 is shown a plot of the logarithm (to the base 10) of the diffusion
coefficient versus reciprocal of absolute temperature, for the diffusion of copper in gold. Determine values for the activation energy and the preexponential.
Solution
From Equation 5.9b the slope of the line segment in Figure 5.8 is equal to
Qd 2.3R, and the intercept at 1T 0 gives the value of log D0. Thus, the
activation energy may be determined as
D0 and Qd from
Experimental Data
Qd 2.3R 1slope2 2.3R
2.3R
£
¢ 1log D2
£
§
1
¢a b
T
log D1 log D2
§
1
1
T1
T2
where D1 and D2 are the diffusion coefficient values at 1T1 and 1T2, respectively. Let us arbitrarily take 1T1 0.8 103 1K2 1 and 1T2 1.1
103 1K2 1. We may now read the corresponding log D1 and log D2 values from
the line segment in Figure 5.8.
[Before this is done, however, a parenthetic note of caution is offered. The
vertical axis in Figure 5.8 is scaled logarithmically (to the base 10); however,
the actual diffusion coefficient values are noted on this axis. For example, for
D 1014 m2/s, the logarithm of D is 14.0 not 1014. Furthermore, this logarithmic scaling affects the readings between decade values; for example, at a
location midway between 1014 and 1015, the value is not 5 1015 but, rather,
1014.5 3.2 1015].
Thus, from Figure 5.8, at 1T1 0.8 103 1K2 1, log D1 12.40, while
for 1T2 1.1 103 1K2 1, log D2 15.45, and the activation energy, as
1496T_c05_109-130 11/14/05 10:12 Page 122
REVISED PAGES
122 • Chapter 5 / Diffusion
Figure 5.8 Plot of the
logarithm of the diffusion
coefficient versus the
reciprocal of absolute
temperature for the
diffusion of copper in gold.
Diffusion coefficient (m2/s)
10–12
10–13
10–14
10–15
10–16
10–17
0.7
0.8
0.9
1.0
1.1
1.2
Reciprocal temperature (1000/K)
determined from the slope of the line segment in Figure 5.8 is
Qd 2.3R
£
log D1 log D2
§
1
1
T1
T2
2.3 18.31 J/mol-K2 c
12.40 115.452
0.8 103 1K2 1 1.1 103 1K2 1
194,000 J/mol 194 kJ/mol
d
Now, rather than trying to make a graphical extrapolation to determine
D0, a more accurate value is obtained analytically using Equation 5.9b, and a
specific value of D (or log D) and its corresponding T (or 1T ) from Figure 5.8. Since we know that log D 15.45 at 1 T 1.1 103 (K)1, then
Qd 1
log D0 log D
a b
2.3R T
1194,000 J/mol211.1 103 3K 4 1 2
15.45
12.3218.31 J/mol-K2
4.28
Thus, D0 104.28 m2/s 5.2 105 m2/s.
DESIGN EXAMPLE 5.1
Diffusion Temperature–Time Heat Treatment Specification
The wear resistance of a steel gear is to be improved by hardening its surface.
This is to be accomplished by increasing the carbon content within an outer
surface layer as a result of carbon diffusion into the steel; the carbon is to be
supplied from an external carbon-rich gaseous atmosphere at an elevated and
constant temperature. The initial carbon content of the steel is 0.20 wt%, whereas
the surface concentration is to be maintained at 1.00 wt%. For this treatment to
be effective, a carbon content of 0.60 wt% must be established at a position
0.75 mm below the surface. Specify an appropriate heat treatment in terms of
temperature and time for temperatures between 900 C and 1050 C. Use data in
Table 5.2 for the diffusion of carbon in g-iron.
1496T_c05_109-130 12/20/05 7:52 Page 123
2nd REVISE PAGES
5.5 Factors That Influence Diffusion • 123
Solution
Since this is a nonsteady-state diffusion situation, let us first of all employ Equation 5.5, utilizing the following values for the concentration parameters:
C0 0.20 wt% C
Cs 1.00 wt% C
Cx 0.60 wt% C
Therefore
Cx C0
0.60 0.20
x
1 erf a
b
Cs C0
1.00 0.20
21Dt
and thus
0.5 erf a
x
b
21Dt
Using an interpolation technique as demonstrated in Example Problem 5.2 and
the data presented in Table 5.1,
x
0.4747
21Dt
(5.10)
The problem stipulates that x 0.75 mm 7.5 104 m. Therefore
7.5 104 m
0.4747
21Dt
This leads to
Dt 6.24 107 m2
Furthermore, the diffusion coefficient depends on temperature according to
Equation 5.8; and, from Table 5.2 for the diffusion of carbon in g-iron,
D0 2.3 105 m2/s and Qd 148,000 J/mol. Hence
Dt D0 exp a
12.3 105 m2/s2 exp c
Qd
b 1t2 6.24 107 m2
RT
148,000 J/mol
d 1t2 6.24 107 m2
18.31 J/mol-K21T 2
and solving for the time t
t 1in s2
0.0271
17,810
exp a
b
T
Thus, the required diffusion time may be computed for some specified temperature (in K). Below are tabulated t values for four different temperatures that
lie within the range stipulated in the problem.
Time
Temperature
(C )
s
h
900
950
1000
1050
106,400
57,200
32,300
19,000
29.6
15.9
9.0
5.3
1496T_c05_109-130 11/14/05 10:13 Page 124
REVISED PAGES
124 • Chapter 5 / Diffusion
MATERIAL OF IMPORTANCE
Aluminum for Integrated Circuit Interconnects
T
he heart of all computers and other electronic
devices is the integrated circuit (or IC).4 Each
integrated circuit chip is a thin square wafer having dimensions on the order of 6 mm by 6 mm by
0.4 mm; furthermore, literally millions of interconnected electronic components and circuits are
embedded in one of the chip faces. The base material for ICs is silicon, to which has been added
very specific and extremely minute and controlled
concentrations of impurities that are confined to
very small and localized regions. For some ICs, the
impurities are added using high-temperature diffusion heat treatments.
One important step in the IC fabrication
process is the deposition of very thin and narrow
conducting circuit paths to facilitate the passage
of current from one device to another; these paths
are called “interconnects,” and several are shown
in Figure 5.9, a scanning electron micrograph of
an IC chip. Of course the material to be used for
interconnects must have a high electrical conductivity—a metal, since, of all materials, metals have
the highest conductivities. Table 5.3 cites values
for silver, copper, gold, and aluminum, the most
conductive metals.
Table 5.3 Room-Temperature Electrical Conductivity Values for Silver, Copper,
Gold, and Aluminum (the Four
Most Conductive Metals)
Electrical Conductivity
[(ohm-meters)1]
Metal
6.8 107
6.0 107
4.3 107
3.8 107
Silver
Copper
Gold
Aluminum
On the basis of these conductivities, and discounting material cost, Ag is the metal of choice,
followed by Cu, Au, and Al.
Once these interconnects have been deposited, it is still necessary to subject the IC chip
to other heat treatments, which may run as high as
500 C. If, during these treatments, there is significant diffusion of the interconnect metal into the
silicon, the electrical functionality of the IC will be
destroyed. Thus, since the extent of diffusion is dependent on the magnitude of the diffusion coefficient, it is necessary to select an interconnect metal
that has a small value of D in silicon. Figure 5.10
Interconnects
Temperature (°C)
1200 1000 900 800 700 600 500 400
10–12
Cu in Si
Diffusion coefficient (m2/s)
Au in Si
1014
4 × 1013
2.5 × 1015
Ag in Si
4.2 × 1017
1016
Al in Si
1018
2.5 × 1021
1020
1022
0.6
Figure 5.9 Scanning electron micrograph of an integrated circuit chip, on which is noted aluminum interconnect regions. Approximately 2000. (Photograph
courtesy of National Semiconductor Corporation.)
4
0.8
1.0
1.2
Reciprocal temperature (1000/K)
1.4
Figure 5.10 Logarithm of D-versus-1T (K) curves
(lines) for the diffusion of copper, gold, silver, and
aluminum in silicon. Also noted are D values at 500 C.
Integrated circuits, their components and materials, are discussed in Section 18.15 and Sections 22.15 through 22.20.
1496T_c05_109-130 11/14/05 10:13 Page 125
REVISED PAGES
Summary • 125
plots the logarithm of D versus 1 T for the diffusion, into silicon, of copper, gold, silver, and aluminum. Also, a dashed vertical line has been
constructed at 500 C, from which values of D, for
the four metals are noted at this temperature. Here
it may be seen that the diffusion coefficient for aluminum in silicon (2.5 1021 m2/s) is at least four
orders of magnitude (i.e., a factor of 104) lower
than the values for the other three metals.
Aluminum is indeed used for interconnects in
some integrated circuits; even though its electrical
conductivity is slightly lower than the values for
silver, copper, and gold, its extremely low diffusion
coefficient makes it the material of choice for this
application. An aluminum-copper-silicon alloy
(Al-4 wt% Cu-1.5 wt% Si) is sometimes also used
for interconnects; it not only bonds easily to the
surface of the chip, but is also more corrosion resistant than pure aluminum.
More recently, copper interconnects have also
been used. However, it is first necessary to deposit
a very thin layer of tantalum or tantalum nitride
beneath the copper, which acts as a barrier to deter
diffusion of Cu into the silicon.
5.6 OTHER DIFFUSION PATHS
Atomic migration may also occur along dislocations, grain boundaries, and external surfaces. These are sometimes called “short-circuit” diffusion paths inasmuch as
rates are much faster than for bulk diffusion. However, in most situations shortcircuit contributions to the overall diffusion flux are insignificant because the crosssectional areas of these paths are extremely small.
SUMMARY
Diffusion Mechanisms
Solid-state diffusion is a means of mass transport within solid materials by stepwise
atomic motion. The term “self-diffusion” refers to the migration of host atoms; for
impurity atoms, the term “interdiffusion” is used. Two mechanisms are possible: vacancy and interstitial. For a given host metal, interstitial atomic species generally
diffuse more rapidly.
Steady-State Diffusion
Nonsteady-State Diffusion
For steady-state diffusion, the concentration profile of the diffusing species is time
independent, and the flux or rate is proportional to the negative of the concentration gradient according to Fick’s first law. The mathematics for nonsteady state
are described by Fick’s second law, a partial differential equation. The solution for
a constant surface composition boundary condition involves the Gaussian error
function.
Factors That Influence Diffusion
The magnitude of the diffusion coefficient is indicative of the rate of atomic motion, being strongly dependent on and increasing exponentially with increasing
temperature.
1496T_c05_109-130 11/14/05 10:13 Page 126
REVISED PAGES
126 • Chapter 5 / Diffusion
I M P O R TA N T T E R M S A N D C O N C E P T S
Activation energy
Carburizing
Concentration gradient
Concentration profile
Diffusion
Diffusion coefficient
Diffusion flux
Driving force
Fick’s first and second laws
Interdiffusion (impurity diffusion)
Interstitial diffusion
Nonsteady-state diffusion
Self-diffusion
Steady-state diffusion
Vacancy diffusion
REFERENCES
Gale, W. F. and T. C. Totemeier, (Editors), Smithells
Metals Reference Book, 8th edition,ButterworthHeinemann Ltd, Woburn, UK, 2004.
Carslaw, H. S. and J. C. Jaeger, Conduction of Heat
in Solids, 2nd edition, Oxford University Press,
Oxford, 1986.
Crank, J., The Mathematics of Diffusion, 2nd edition, Oxford University Press, Oxford, 1980.
Glicksman, M., Diffusion in Solids, WileyInterscience, New York, 2000.
Shewmon, P. G., Diffusion in Solids, 2nd edition,
The Minerals, Metals and Materials Society,
Warrendale, PA, 1989.
QUESTIONS AND PROBLEMS
Introduction
5.1 Briefly explain the difference between selfdiffusion and interdiffusion.
5.2 Self-diffusion involves the motion of atoms
that are all of the same type; therefore it is
not subject to observation by compositional
changes, as with interdiffusion. Suggest one
way in which self-diffusion may be monitored.
Diffusion Mechanisms
5.3 (a) Compare interstitial and vacancy atomic
mechanisms for diffusion.
(b) Cite two reasons why interstitial diffusion
is normally more rapid than vacancy diffusion.
Steady-State Diffusion
5.4 Briefly explain the concept of steady state as
it applies to diffusion.
5.5 (a) Briefly explain the concept of a driving
force.
(b) What is the driving force for steady-state
diffusion?
5.6 The purification of hydrogen gas by diffusion
through a palladium sheet was discussed in
Section 5.3. Compute the number of kilograms of hydrogen that pass per hour through
a 6-mm-thick sheet of palladium having an
area of 0.25 m2 at 600C. Assume a diffusion
coefficient of 1.7 108 m2/s, that the concentrations at the high- and low-pressure
sides of the plate are 2.0 and 0.4 kg of hydrogen per cubic meter of palladium, and that
steady-state conditions have been attained.
5.7 A sheet of steel 2.5 mm thick has nitrogen
atmospheres on both sides at 900C and
is permitted to achieve a steady-state diffusion condition. The diffusion coefficient for
nitrogen in steel at this temperature is
1.2 1010 m2/s, and the diffusion flux is
found to be 1.0 107 kg/m2-s. Also, it is
known that the concentration of nitrogen
in the steel at the high-pressure surface is
2 kg/m3. How far into the sheet from this highpressure side will the concentration be 0.5
kg/m3? Assume a linear concentration profile.
5.8 A sheet of BCC iron 2 mm thick was exposed
to a carburizing gas atmosphere on one side
and a decarburizing atmosphere on the other
side at 675C. After having reached steady
state, the iron was quickly cooled to room temperature.The carbon concentrations at the two
surfaces of the sheet were determined to be
0.015 and 0.0068 wt%. Compute the diffusion
1496T_c05_109-130 12/20/05 7:52 Page 127
2nd REVISE PAGES
Questions and Problems • 127
coefficient if the diffusion flux is 7.36 109
kg/m2-s. Hint: Use Equation 4.9 to convert the
concentrations from weight percent to kilograms of carbon per cubic meter of iron.
5.9 When a-iron is subjected to an atmosphere of
nitrogen gas, the concentration of nitrogen in
the iron, CN (in weight percent), is a function
of hydrogen pressure, pN2(in MPa), and absolute temperature (T) according to
CN 4.90 103 1pN2 exp a
37.6 kJ/mol
b
RT
(5.11)
Furthermore, the values of D0 and Qd for
this diffusion system are 3.0 107 m2/s
and 76,150 J/mol, respectively. Consider a thin
iron membrane 1.5 mm thick that is at 300C.
Compute the diffusion flux through this membrane if the nitrogen pressure on one side of
the membrane is 0.10 MPa (0.99 atm), and on
the other side 5.0 MPa (49.3 atm).
Nonsteady-State Diffusion
5.10 Show that
Cx
B
x2
exp a
b
4Dt
1Dt
is also a solution to Equation 5.4b. The parameter B is a constant, being independent of
both x and t.
5.11 Determine the carburizing time necessary to
achieve a carbon concentration of 0.30 wt%
at a position 4 mm into an iron–carbon alloy
that initially contains 0.10 wt% C. The surface
concentration is to be maintained at 0.90 wt%
C, and the treatment is to be conducted at
1100C. Use the diffusion data for g-Fe in
Table 5.2.
5.12 An FCC iron–carbon alloy initially containing 0.55 wt% C is exposed to an oxygen-rich
and virtually carbon-free atmosphere at 1325 K
(1052C). Under these circumstances the carbon diffuses from the alloy and reacts at the
surface with the oxygen in the atmosphere;
that is, the carbon concentration at the surface
position is maintained essentially at 0 wt% C.
(This process of carbon depletion is termed
decarburization.) At what position will the
carbon concentration be 0.25 wt% after a
10-h treatment? The value of D at 1325 K is
4.3 1011 m2/s.
5.13 Nitrogen from a gaseous phase is to be diffused into pure iron at 675C. If the surface
concentration is maintained at 0.2 wt% N,
what will be the concentration 2 mm from the
surface after 25 h? The diffusion coefficient for
nitrogen in iron at 675C is 1.9 1011 m2/s.
5.14 Consider a diffusion couple composed of two
semi-infinite solids of the same metal, and
that each side of the diffusion couple has a
different concentration of the same elemental
impurity; furthermore, assume each impurity
level is constant throughout its side of the diffusion couple. For this situation, the solution
to Fick’s second law (assuming that the diffusion coefficient for the impurity is independent of concentration), is as follows:
Cx a
C1 C2
C1 C2
x
ba
b erf a
b
2
2
21Dt
(5.12)
In this expression, when the x 0 position is
taken as the initial diffusion couple interface,
then C1 is the impurity concentration for x 6 0;
likewise, C2 is the impurity content for x 7 0.
A diffusion couple composed of two
platinum-gold alloys is formed; these alloys
have compositions of 99.0 wt% Pt-1.0 wt%
Au and 96.0 wt% Pt-4.0 wt% Au. Determine
the time this diffusion couple must be heated
at 1000C (1273 K) in order for the composition to be 2.8 wt% Au at the 10 mm position
into the 4.0 wt% Au side of the diffusion couple. Preexponential and activation energy
values for Au diffusion in Pt are 1.3 105
m2/s and 252,000 J/mol, respectively.
5.15 For a steel alloy it has been determined that a
carburizing heat treatment of 15 h duration will
raise the carbon concentration to 0.35 wt% at
a point 2.0 mm from the surface. Estimate the
time necessary to achieve the same concentration at a 6.0-mm position for an identical steel
and at the same carburizing temperature.
Factors That Influence Diffusion
5.16 Cite the values of the diffusion coefficients
for the interdiffusion of carbon in both a-iron
(BCC) and g-iron (FCC) at 900C. Which is
larger? Explain why this is the case.
5.17 Using the data in Table 5.2, compute the value
of D for the diffusion of magnesium in aluminum at 400C.
1496T_c05_109-130 12/20/05 7:52 Page 128
2nd REVISE PAGES
128 • Chapter 5 / Diffusion
5.18 At what temperature will the diffusion coefficient for the diffusion of zinc in copper have
a value of 2.6 1016 m2/s? Use the diffusion
data in Table 5.2.
5.19 The preexponential and activation energy
for the diffusion of chromium in nickel are
1.1 104 m2/s and 272,000 J/mol, respectively. At what temperature will the diffusion
coefficient have a value of 1.2 1014 m2/s?
5.20 The activation energy for the diffusion of copper in silver is 193,000 J/mol. Calculate the diffusion coefficient at 1200 K (927C), given
that D at 1000 K (727C) is 1.0 1014 m2/s.
5.21 The diffusion coefficients for nickel in iron
are given at two temperatures:
T(K)
D(m2/s)
1473
1673
2.2 1015
4.8 1014
(a) Determine the values of D0 and the activation energy Qd.
(b) What is the magnitude of D at 1300C
(1573 K)?
5.22 The diffusion coefficients for carbon in nickel
are given at two temperatures:
T(C)
D(m2/s)
600
700
5.5 1014
3.9 1013
Diffusion coefficient (m2/s)
(a) Determine the values of D0 and Qd.
(b) What is the magnitude of D at 850C?
5.23 Below is shown a plot of the logarithm (to
the base 10) of the diffusion coefficient
10 –13
10 –14
10 –15
0.8
0.9
1.0
Reciprocal temperature (1000/K)
versus reciprocal of the absolute temperature, for the diffusion of gold in silver. Determine values for the activation energy and
preexponential.
5.24 Carbon is allowed to diffuse through a steel
plate 10 mm thick. The concentrations of carbon at the two faces are 0.85 and 0.40 kg
C/cm3 Fe, which are maintained constant. If
the preexponential and activation energy are
6.2 107 m2/s and 80,000 J/mol, respectively,
compute the temperature at which the diffusion flux is 6.3 1010 kg/m2-s.
5.25 The steady-state diffusion flux through a
metal plate is 7.8 108 kg/m2-s at a temperature of 1200C (1473 K) and when the
concentration gradient is 500 kg/m4. Calculate the diffusion flux at 1000C (1273 K) for
the same concentration gradient and assuming an activation energy for diffusion of
145,000 J/mol.
5.26 At approximately what temperature would a
specimen of g-iron have to be carburized for
4 h to produce the same diffusion result as at
1000C for 12 h?
5.27 (a) Calculate the diffusion coefficient for
magnesium in aluminum at 450C.
(b) What time will be required at 550C to
produce the same diffusion result (in terms of
concentration at a specific point) as for 15 h
at 450C?
5.28 A copper–nickel diffusion couple similar
to that shown in Figure 5.1a is fashioned.After
a 500-h heat treatment at 1000C (1273 K) the
concentration of Ni is 3.0 wt% at the 1.0-mm
position within the copper. At what temperature should the diffusion couple be heated to
produce this same concentration (i.e., 3.0 wt%
Ni) at a 2.0-mm position after 500 h? The preexponential and activation energy for the diffusion of Ni in Cu are 2.7 104 m2/s and
236,000 J/mol, respectively.
5.29 A diffusion couple similar to that shown
in Figure 5.1a is prepared using two hypothetical metals A and B. After a 20-h heat
treatment at 800C (and subsequently cooling
to room temperature) the concentration of B
in A is 2.5 wt% at the 5.0-mm position within
metal A. If another heat treatment is conducted on an identical diffusion couple, only
1496T_c05_109-130 11/14/05 10:13 Page 129
REVISED PAGES
Design Problems • 129
at 1000C for 20 h, at what position will the
composition be 2.5 wt% B? Assume that the
preexponential and activation energy for the
diffusion coefficient are 1.5 104 m2/s and
125,000 J/mol, respectively.
5.30 The outer surface of a steel gear is to
be hardened by increasing its carbon content; the carbon is to be supplied from an
external carbon-rich atmosphere that is
maintained at an elevated temperature. A
diffusion heat treatment at 600C (873 K)
for 100 min increases the carbon concentration to 0.75 wt% at a position 0.5 mm below
the surface. Estimate the diffusion time
required at 900C (1173 K) to achieve this
same concentration also at a 0.5-mm position. Assume that the surface carbon
content is the same for both heat treatments,
which is maintained constant. Use the
diffusion data in Table 5.2 for C diffusion
in a-Fe.
5.31 An FCC iron–carbon alloy initially containing 0.10 wt% C is carburized at an elevated
temperature and in an atmosphere wherein
the surface carbon concentration is maintained at 1.10 wt%. If after 48 h the concentration of carbon is 0.30 wt% at a position
3.5 mm below the surface, determine the
temperature at which the treatment was carried out.
DESIGN PROBLEMS
Steady-State Diffusion
(Factors That Influence Diffusion)
5.D1 It is desired to enrich the partial pressure of
hydrogen in a hydrogen–nitrogen gas mixture for which the partial pressures of both
gases are 0.1013 MPa (1 atm). It has been
proposed to accomplish this by passing both
gases through a thin sheet of some metal at
an elevated temperature; inasmuch as hydrogen diffuses through the plate at a higher
rate than does nitrogen, the partial pressure
of hydrogen will be higher on the exit side of
the sheet. The design calls for partial pressures of 0.051 MPa (0.5 atm) and 0.01013
MPa (0.1 atm), respectively, for hydrogen
and nitrogen. The concentrations of hydrogen and nitrogen (CH and CN, in mol/m3) in
this metal are functions of gas partial pressures ( pH2 and pN2, in MPa) and absolute temperature and are given by the following expressions:
CH 2.5 103 1pH2 exp a
27.8 kJ/mol
b
RT
(5.13a)
37.6 kJ/mol
CN 2.75 103 1pN2 exp a
b
RT
(5.13b)
Furthermore, the diffusion coefficients for
the diffusion of these gases in this metal are
functions of the absolute temperature as follows:
13.4 kJ/mol
b
DH 1m2/s2 1.4 107 exp a
RT
(5.14a)
76.15 kJ/mol
DN 1m2/s2 3.0 107 exp a
b
RT
(5.14b)
Is it possible to purify hydrogen gas in this
manner? If so, specify a temperature at which
the process may be carried out, and also the
thickness of metal sheet that would be required. If this procedure is not possible, then
state the reason(s) why.
5.D2 A gas mixture is found to contain two diatomic A and B species (A2 and B2) for which
the partial pressures of both are 0.1013 MPa
(1 atm). This mixture is to be enriched in the
partial pressure of the A species by passing
both gases through a thin sheet of some
metal at an elevated temperature. The resulting enriched mixture is to have a partial
pressure of 0.051 MPa (0.5 atm) for gas A
and 0.0203 MPa (0.2 atm) for gas B. The concentrations of A and B (CA and CB, in
1496T_c05_109-130 12/20/05 7:52 Page 130
2nd REVISE PAGES
130 • Chapter 5 / Diffusion
mol/m3) are functions of gas partial pressures
( pA2 and pB2, in MPa) and absolute temperature according to the following expressions:
CA 1.5 103 1pA2 exp a
20.0 kJ/mol
b
RT
(5.15a)
27.0 kJ/mol
CB 2.0 103 1pB2 exp a
b
RT
(5.15b)
Furthermore, the diffusion coefficients for the
diffusion of these gases in the metal are functions of the absolute temperature as follows:
DA 1m2/s2 5.0 107 exp a
13.0 kJ/mol
b
RT
(5.16a)
21.0
kJ/mol
DB 1m2/s2 3.0 106 exp a
b
RT
(5.16b)
Is it possible to purify the A gas in this manner? If so, specify a temperature at which the
process may be carried out, and also the
thickness of metal sheet that would be required. If this procedure is not possible, then
state the reason(s) why.
Nonsteady-State Diffusion
(Factors That Influence Diffusion)
5.D3 The wear resistance of a steel shaft is to be
improved by hardening its surface. This is to
be accomplished by increasing the nitrogen
content within an outer surface layer as a result of nitrogen diffusion into the steel; the
nitrogen is to be supplied from an external
nitrogen-rich gas at an elevated and constant
temperature. The initial nitrogen content of
the steel is 0.0025 wt%, whereas the surface
concentration is to be maintained at 0.45 wt%.
For this treatment to be effective, a nitrogen
content of 0.12 wt% must be established at
a position 0.45 mm below the surface. Specify an appropriate heat treatment in terms of
temperature and time for a temperature
between 475C and 625C. The preexponential and activation energy for the diffusion of nitrogen in iron are 3 107 m2/s and
76,150 J/mol, respectively, over this temperature range.
5.D4 The wear resistance of a steel gear is to be
improved by hardening its surface, as described in Design Example 5.1. However, in
this case the initial carbon content of the
steel is 0.15 wt%, and a carbon content of
0.75 wt% is to be established at a position
0.65 mm below the surface. Furthermore, the
surface concentration is to be maintained
constant, but may be varied between 1.2 and
1.4 wt% C. Specify an appropriate heat treatment in terms of surface carbon concentration and time, and for a temperature between
1000C and 1200C.
1496T_c06_131-173 12/21/05 8:26 Page 131
Chapter
A
2nd REVISE PAGES
6
Mechanical Properties
of Metals
modern Rockwell hardness tester. (Photograph courtesy of
Wilson Instruments Division, Instron Corporation, originator of the
Rockwell® Hardness Tester.)
WHY STUDY The Mechanical Properties of Metals?
It is incumbent on engineers to understand how the
various mechanical properties are measured and what
these properties represent; they may be called upon
to design structures/components using predetermined
materials such that unacceptable levels of deformation
and/or failure will not occur. We demonstrate this
procedure with respect to the design of a tensiletesting apparatus in Design Example 6.1.
• 131
1496T_c06_131-173 12/21/05 7:43 Page 132
2nd REVISE PAGES
Learning Objectives
After studying this chapter you should be able to do the following:
1. Define engineering stress and engineering strain.
7. Give brief definitions of and the units for
2. State Hooke’s law, and note the conditions
modulus of resilience and toughness (static).
under which it is valid.
8. For a specimen being loaded in tension, given
3. Define Poisson’s ratio.
the applied load, the instantaneous cross4. Given an engineering stress–strain diagram,
sectional dimensions, as well as original and
determine (a) the modulus of elasticity,
instantaneous lengths, be able to compute
(b) the yield strength (0.002 strain offset),
true stress and true strain values.
and (c) the tensile strength, and (d) estimate
9. Name the two most common hardness-testing
the percent elongation.
techniques; note two differences between
5. For the tensile deformation of a ductile cylinthem.
drical specimen, describe changes in specimen
10. (a) Name and briefly describe the two differprofile to the point of fracture.
ent microindentation hardness testing tech6. Compute ductility in terms of both percent
niques, and (b) cite situations for which these
elongation and percent reduction of area
techniques are generally used.
for a material that is loaded in tension to
11. Compute the working stress for a ductile
fracture.
material.
6.1 INTRODUCTION
Many materials, when in service, are subjected to forces or loads; examples include
the aluminum alloy from which an airplane wing is constructed and the steel in an
automobile axle. In such situations it is necessary to know the characteristics of the
material and to design the member from which it is made such that any resulting
deformation will not be excessive and fracture will not occur. The mechanical behavior of a material reflects the relationship between its response or deformation
to an applied load or force. Important mechanical properties are strength, hardness,
ductility, and stiffness.
The mechanical properties of materials are ascertained by performing carefully
designed laboratory experiments that replicate as nearly as possible the service
conditions. Factors to be considered include the nature of the applied load and its
duration, as well as the environmental conditions. It is possible for the load to be
tensile, compressive, or shear, and its magnitude may be constant with time, or it may
fluctuate continuously. Application time may be only a fraction of a second, or it
may extend over a period of many years. Service temperature may be an important
factor.
Mechanical properties are of concern to a variety of parties (e.g., producers and
consumers of materials, research organizations, government agencies) that have differing interests. Consequently, it is imperative that there be some consistency in the
manner in which tests are conducted, and in the interpretation of their results. This
consistency is accomplished by using standardized testing techniques. Establishment
and publication of these standards are often coordinated by professional societies. In
the United States the most active organization is the American Society for Testing
and Materials (ASTM). Its Annual Book of ASTM Standards (http://www.astm.org)
comprises numerous volumes, which are issued and updated yearly; a large number of
these standards relate to mechanical testing techniques. Several of these are referenced
by footnote in this and subsequent chapters.
The role of structural engineers is to determine stresses and stress distributions
within members that are subjected to well-defined loads. This may be accomplished
1496T_c06_131-173 11/16/05 17:06 Page 133
REVISED PAGES
6.2 Concepts of Stress and Strain • 133
by experimental testing techniques and/or by theoretical and mathematical stress
analyses. These topics are treated in traditional stress analysis and strength of
materials texts.
Materials and metallurgical engineers, on the other hand, are concerned with
producing and fabricating materials to meet service requirements as predicted by
these stress analyses. This necessarily involves an understanding of the relationships
between the microstructure (i.e., internal features) of materials and their mechanical
properties.
Materials are frequently chosen for structural applications because they have
desirable combinations of mechanical characteristics. The present discussion is confined primarily to the mechanical behavior of metals; polymers and ceramics are
treated separately because they are, to a large degree, mechanically dissimilar to
metals. This chapter discusses the stress–strain behavior of metals and the related
mechanical properties, and also examines other important mechanical characteristics. Discussions of the microscopic aspects of deformation mechanisms and methods to strengthen and regulate the mechanical behavior of metals are deferred to
later chapters.
6.2 CONCEPTS OF STRESS AND STRAIN
If a load is static or changes relatively slowly with time and is applied uniformly
over a cross section or surface of a member, the mechanical behavior may be ascertained by a simple stress–strain test; these are most commonly conducted for
metals at room temperature. There are three principal ways in which a load may
be applied: namely, tension, compression, and shear (Figures 6.1a, b, c). In engineering practice many loads are torsional rather than pure shear; this type of loading is illustrated in Figure 6.1d.
Tension Tests 1
One of the most common mechanical stress–strain tests is performed in tension. As
will be seen, the tension test can be used to ascertain several mechanical properties of materials that are important in design. A specimen is deformed, usually to
fracture, with a gradually increasing tensile load that is applied uniaxially along the
long axis of a specimen. A standard tensile specimen is shown in Figure 6.2. Normally, the cross section is circular, but rectangular specimens are also used. This
“dogbone” specimen configuration was chosen so that, during testing, deformation
is confined to the narrow center region (which has a uniform cross section along
its length), and, also, to reduce the likelihood of fracture at the ends of the specimen. The standard diameter is approximately 12.8 mm (0.5 in.), whereas the reduced section length should be at least four times this diameter; 60 mm (214 in.) is
common. Gauge length is used in ductility computations, as discussed in Section 6.6;
the standard value is 50 mm (2.0 in.). The specimen is mounted by its ends into the
holding grips of the testing apparatus (Figure 6.3). The tensile testing machine is
designed to elongate the specimen at a constant rate, and to continuously and
simultaneously measure the instantaneous applied load (with a load cell) and the
resulting elongations (using an extensometer).A stress–strain test typically takes several minutes to perform and is destructive; that is, the test specimen is permanently
deformed and usually fractured.
1
ASTM Standards E 8 and E 8M, “Standard Test Methods for Tension Testing of Metallic
Materials.”
1496T_c06_131-173 11/16/05 17:21 Page 134
REVISED PAGES
134 • Chapter 6 / Mechanical Properties of Metals
Figure 6.1
(a) Schematic
illustration of how a
tensile load produces
an elongation and
positive linear
strain. Dashed lines
represent the shape
before deformation;
solid lines, after
deformation.
(b) Schematic
illustration of how a
compressive load
produces contraction
and a negative linear
strain. (c) Schematic
representation of
shear strain g,
where g tan u.
(d) Schematic
representation
of torsional
deformation (i.e.,
angle of twist f)
produced by an
applied torque T.
F
F
A0
l
l0
l0
l
A0
F
F
(a)
( b)
T
A0
T
F
F
F
(c )
engineering stress
engineering strain
(d)
The output of such a tensile test is recorded (usually on a computer) as load
or force versus elongation. These load–deformation characteristics are dependent
on the specimen size. For example, it will require twice the load to produce the same
elongation if the cross-sectional area of the specimen is doubled. To minimize these
geometrical factors, load and elongation are normalized to the respective parameters
of engineering stress and engineering strain. Engineering stress s is defined by the
relationship
Definition of
engineering stress
(for tension and
compression)
Figure 6.2 A
standard tensile
specimen with
circular cross
section.
s
F
A0
Reduced section
1
2 "
4
3"
Diameter
4
0.505" Diameter
2"
Gauge length
3"
8
Radius
(6.1)
1496T_c06_131-173 11/16/05 17:06 Page 135
REVISED PAGES
6.2 Concepts of Stress and Strain • 135
Load cell
Figure 6.3 Schematic representation of
the apparatus used to conduct tensile
stress–strain tests. The specimen is elongated
by the moving crosshead; load cell and
extensometer measure, respectively, the
magnitude of the applied load and the
elongation. (Adapted from H. W. Hayden,
W. G. Moffatt, and J. Wulff, The Structure and
Properties of Materials, Vol. III, Mechanical
Behavior, p. 2. Copyright © 1965 by John
Wiley & Sons, New York. Reprinted by
permission of John Wiley & Sons, Inc.)
Extensometer
Specimen
Moving
crosshead
in which F is the instantaneous load applied perpendicular to the specimen cross
section, in units of newtons (N) or pounds force 1lbf 2 , and A0 is the original crosssectional area before any load is applied (m2 or in.2). The units of engineering
stress (referred to subsequently as just stress) are megapascals, MPa (SI) (where
1 MPa 106 N/m2), and pounds force per square inch, psi (Customary U.S.).2
Engineering strain is defined according to
Definition of
engineering strain
(for tension and
compression)
li l0
¢l
l0
l0
(6.2)
in which l0 is the original length before any load is applied, and li is the instantaneous length. Sometimes the quantity li l0 is denoted as ¢l, and is the deformation elongation or change in length at some instant, as referenced to the original
length. Engineering strain (subsequently called just strain) is unitless, but meters
per meter or inches per inch are often used; the value of strain is obviously independent of the unit system. Sometimes strain is also expressed as a percentage, in
which the strain value is multiplied by 100.
Compression Tests 3
Compression stress–strain tests may be conducted if in-service forces are of this
type. A compression test is conducted in a manner similar to the tensile test, except
that the force is compressive and the specimen contracts along the direction of the
stress. Equations 6.1 and 6.2 are utilized to compute compressive stress and strain,
respectively. By convention, a compressive force is taken to be negative, which yields
a negative stress. Furthermore, since l0 is greater than li, compressive strains computed from Equation 6.2 are necessarily also negative. Tensile tests are more common because they are easier to perform; also, for most materials used in structural
applications, very little additional information is obtained from compressive tests.
2
Conversion from one system of stress units to the other is accomplished by the relationship
145 psi 1 MPa.
3
ASTM Standard E 9, “Standard Test Methods of Compression Testing of Metallic Materials
at Room Temperature.”
1496T_c06_131-173 11/16/05 17:06 Page 136
REVISED PAGES
136 • Chapter 6 / Mechanical Properties of Metals
Compressive tests are used when a material’s behavior under large and permanent
(i.e., plastic) strains is desired, as in manufacturing applications, or when the material is brittle in tension.
Shear and Torsional Tests 4
For tests performed using a pure shear force as shown in Figure 6.1c, the shear stress
t is computed according to
Definition of shear
stress
t
F
A0
(6.3)
where F is the load or force imposed parallel to the upper and lower faces, each of
which has an area of A0. The shear strain g is defined as the tangent of the strain
angle u, as indicated in the figure. The units for shear stress and strain are the same
as for their tensile counterparts.
Torsion is a variation of pure shear, wherein a structural member is twisted in
the manner of Figure 6.1d; torsional forces produce a rotational motion about the
longitudinal axis of one end of the member relative to the other end. Examples of
torsion are found for machine axles and drive shafts, and also for twist drills. Torsional tests are normally performed on cylindrical solid shafts or tubes. A shear
stress t is a function of the applied torque T, whereas shear strain g is related to
the angle of twist, f in Figure 6.1d.
Geometric Considerations of the Stress State
Stresses that are computed from the tensile, compressive, shear, and torsional force
states represented in Figure 6.1 act either parallel or perpendicular to planar faces
of the bodies represented in these illustrations. Note that the stress state is a function of the orientations of the planes upon which the stresses are taken to act. For
example, consider the cylindrical tensile specimen of Figure 6.4 that is subjected to
a tensile stress s applied parallel to its axis. Furthermore, consider also the plane
p-p¿ that is oriented at some arbitrary angle u relative to the plane of the specimen
end-face. Upon this plane p-p¿ , the applied stress is no longer a pure tensile one.
Rather, a more complex stress state is present that consists of a tensile (or normal)
stress s¿ that acts normal to the p-p¿ plane and, in addition, a shear stress t¿ that
acts parallel to this plane; both of these stresses are represented in the figure. Using
mechanics of materials principles,5 it is possible to develop equations for s¿ and t¿
in terms of s and u, as follows:
1 cos 2u
s¿ s cos2 u s a
b
(6.4a)
2
sin 2u
b
t¿ s sin u cos u s a
(6.4b)
2
These same mechanics principles allow the transformation of stress components
from one coordinate system to another coordinate system that has a different orientation. Such treatments are beyond the scope of the present discussion.
4
ASTM Standard E 143, “Standard Test for Shear Modulus.”
See, for example, W. F. Riley, L. D. Sturges, and D. H. Morris, Mechanics of Materials, 5th
edition, John Wiley & Sons, New York, 1999.
5
1496T_c06_131-173 11/16/05 17:06 Page 137
REVISED PAGES
6.3 Stress–Strain Behavior • 137
Figure 6.4 Schematic representation showing normal (s¿ ) and shear
(t¿ ) stresses that act on a plane oriented at an angle u relative to the
plane taken perpendicular to the direction along which a pure tensile
stress (s) is applied.
p⬘
⬘
⬘
p
Elastic Deformation
6.3 STRESS–STRAIN BEHAVIOR
The degree to which a structure deforms or strains depends on the magnitude of
an imposed stress. For most metals that are stressed in tension and at relatively low
levels, stress and strain are proportional to each other through the relationship
Hooke’s law—
relationship between
engineering stress
and engineering
strain for elastic
deformation (tension
and compression)
modulus of elasticity
s E
(6.5)
This is known as Hooke’s law, and the constant of proportionality E (GPa or psi)6
is the modulus of elasticity, or Young’s modulus. For most typical metals the magnitude of this modulus ranges between 45 GPa (6.5 106 psi), for magnesium, and
407 GPa (59 106 psi), for tungsten. Modulus of elasticity values for several metals at room temperature are presented in Table 6.1.
Table 6.1 Room-Temperature Elastic and Shear Moduli, and Poisson’s Ratio
for Various Metal Alloys
Modulus of
Elasticity
Shear Modulus
Metal Alloy
GPa
10 psi
GPa
106 psi
Poisson’s
Ratio
Aluminum
Brass
Copper
Magnesium
Nickel
Steel
Titanium
Tungsten
69
97
110
45
207
207
107
407
10
14
16
6.5
30
30
15.5
59
25
37
46
17
76
83
45
160
3.6
5.4
6.7
2.5
11.0
12.0
6.5
23.2
0.33
0.34
0.34
0.29
0.31
0.30
0.34
0.28
6
The SI unit for the modulus of elasticity is gigapascal, GPa, where 1 GPa 109 N/m2
103 MPa.
6
1496T_c06_131-173 12/21/05 7:43 Page 138
2nd REVISE PAGES
138 • Chapter 6 / Mechanical Properties of Metals
Figure 6.5 Schematic stress–strain diagram showing linear
elastic deformation for loading and unloading cycles.
Stress
Unload
Slope = modulus
of elasticity
Load
0
0
Strain
Metal Alloys
Deformation in which stress and strain are proportional is called elastic deformation; a plot of stress (ordinate) versus strain (abscissa) results in a linear relationship, as shown in Figure 6.5. The slope of this linear segment corresponds to the
modulus of elasticity E. This modulus may be thought of as stiffness, or a material’s
resistance to elastic deformation. The greater the modulus, the stiffer the material, or
the smaller the elastic strain that results from the application of a given stress. The
modulus is an important design parameter used for computing elastic deflections.
Elastic deformation is nonpermanent, which means that when the applied load
is released, the piece returns to its original shape. As shown in the stress–strain plot
(Figure 6.5), application of the load corresponds to moving from the origin up and
along the straight line. Upon release of the load, the line is traversed in the opposite direction, back to the origin.
There are some materials (e.g., gray cast iron, concrete, and many polymers) for
which this elastic portion of the stress–strain curve is not linear (Figure 6.6); hence, it
is not possible to determine a modulus of elasticity as described above. For this nonlinear behavior, either tangent or secant modulus is normally used. Tangent modulus
is taken as the slope of the stress–strain curve at some specified level of stress, while
secant modulus represents the slope of a secant drawn from the origin to some given
point of the s– curve. The determination of these moduli is illustrated in Figure 6.6.
Figure 6.6 Schematic
stress–strain diagram showing
non-linear elastic behavior, and
how secant and tangent moduli
are determined.
2
⌬
⌬⑀
Stress
elastic deformation
= Tangent modulus (at 2)
1
⌬
⌬⑀
= Secant modulus
(between origin and 1)
Strain ⑀
1496T_c06_131-173 11/16/05 17:06 Page 139
REVISED PAGES
6.3 Stress–Strain Behavior • 139
Strongly
bonded
Force F
Figure 6.7 Force
versus interatomic
separation for
weakly and strongly
bonded atoms. The
magnitude of the
modulus of elasticity
is proportional to
the slope of each
curve at the
equilibrium
interatomic
separation r0.
dF
dr r
0
Separation r
0
Weakly
bonded
On an atomic scale, macroscopic elastic strain is manifested as small changes
in the interatomic spacing and the stretching of interatomic bonds. As a consequence, the magnitude of the modulus of elasticity is a measure of the resistance
to separation of adjacent atoms, that is, the interatomic bonding forces. Furthermore, this modulus is proportional to the slope of the interatomic force–separation
curve (Figure 2.8a) at the equilibrium spacing:
E r a
dF
b
dr r0
(6.6)
Figure 6.7 shows the force–separation curves for materials having both strong and
weak interatomic bonds; the slope at r0 is indicated for each.
Values of the modulus of elasticity for ceramic materials are about the same as
for metals; for polymers they are lower (Figure 1.4). These differences are a direct
consequence of the different types of atomic bonding in the three materials types.
Furthermore, with increasing temperature, the modulus of elasticity diminishes, as
is shown for several metals in Figure 6.8.
Temperature (°F)
–400
0
400
800
1200
1600
70
Tungsten
50
300
40
Steel
200
30
20
100
Aluminum
10
0
–200
0
200
400
Temperature (°C)
600
800
0
Modulus of elasticity (106 psi)
60
400
Modulus of elasticity (GPa)
Figure 6.8 Plot of
modulus of elasticity
versus temperature
for tungsten, steel,
and aluminum.
(Adapted from
K. M. Ralls, T. H.
Courtney, and
J. Wulff, Introduction
to Materials Science
and Engineering.
Copyright © 1976
by John Wiley &
Sons, New York.
Reprinted by
permission of John
Wiley & Sons, Inc.)
1496T_c06_131-173 11/16/05 17:06 Page 140
REVISED PAGES
140 • Chapter 6 / Mechanical Properties of Metals
As would be expected, the imposition of compressive, shear, or torsional stresses
also evokes elastic behavior. The stress–strain characteristics at low stress levels are
virtually the same for both tensile and compressive situations, to include the magnitude of the modulus of elasticity. Shear stress and strain are proportional to each
other through the expression
Relationship
between shear stress
and shear strain for
elastic deformation
6.4
t Gg
(6.7)
where G is the shear modulus, the slope of the linear elastic region of the shear
stress–strain curve. Table 6.1 also gives the shear moduli for a number of the common metals.
ANELASTICITY
anelasticity
Up to this point, it has been assumed that elastic deformation is time independent—
that is, that an applied stress produces an instantaneous elastic strain that remains constant over the period of time the stress is maintained. It has also been assumed that
upon release of the load the strain is totally recovered—that is, that the strain immediately returns to zero. In most engineering materials, however, there will also exist a
time-dependent elastic strain component. That is, elastic deformation will continue after the stress application, and upon load release some finite time is required for complete recovery. This time-dependent elastic behavior is known as anelasticity, and it is
due to time-dependent microscopic and atomistic processes that are attendant to the
deformation. For metals the anelastic component is normally small and is often neglected. However, for some polymeric materials its magnitude is significant; in this case
it is termed viscoelastic behavior, which is the discussion topic of Section 15.4.
EXAMPLE PROBLEM 6.1
Elongation (Elastic) Computation
A piece of copper originally 305 mm (12 in.) long is pulled in tension with a
stress of 276 MPa (40,000 psi). If the deformation is entirely elastic, what will
be the resultant elongation?
Solution
Since the deformation is elastic, strain is dependent on stress according to Equation 6.5. Furthermore, the elongation ¢l is related to the original length l0 through
Equation 6.2. Combining these two expressions and solving for ¢l yields
s E a
¢l
¢l
bE
l0
sl0
E
The values of s and l0 are given as 276 MPa and 305 mm, respectively, and
the magnitude of E for copper from Table 6.1 is 110 GPa (16 106 psi). Elongation is obtained by substitution into the expression above as
¢l
1276 MPa21305 mm2
110 103 MPa
0.77 mm 10.03 in.2
1496T_c06_131-173 11/16/05 17:06 Page 141
REVISED PAGES
6.5 Elastic Properties of Materials • 141
6.5 ELASTIC PROPERTIES OF MATERIALS
Poisson’s ratio
When a tensile stress is imposed on a metal specimen, an elastic elongation and accompanying strain z result in the direction of the applied stress (arbitrarily taken
to be the z direction), as indicated in Figure 6.9. As a result of this elongation, there
will be constrictions in the lateral (x and y) directions perpendicular to the applied
stress; from these contractions, the compressive strains x and y may be determined.
If the applied stress is uniaxial (only in the z direction), and the material is isotropic,
then x y. A parameter termed Poisson’s ratio n is defined as the ratio of the
lateral and axial strains, or
Definition of
Poisson’s ratio in
terms of lateral and
axial strains
n
y
x
z
z
(6.8)
The negative sign is included in the expression so that n will always be positive,
since x and z will always be of opposite sign. Theoretically, Poisson’s ratio for
isotropic materials should be 14 ; furthermore, the maximum value for n (or that value
for which there is no net volume change) is 0.50. For many metals and other alloys,
values of Poisson’s ratio range between 0.25 and 0.35. Table 6.1 shows n values for
several common metallic materials.
For isotropic materials, shear and elastic moduli are related to each other and
to Poisson’s ratio according to
Relationship among
elastic parameters—
modulus of elasticity,
shear modulus, and
Poisson’s ratio
E 2G 11 n2
(6.9)
In most metals G is about 0.4E; thus, if the value of one modulus is known, the
other may be approximated.
Many materials are elastically anisotropic; that is, the elastic behavior (e.g., the
magnitude of E) varies with crystallographic direction (see Table 3.3). For these materials the elastic properties are completely characterized only by the specification
⌬lz
2
z
l0x
⌬lx
2
l0z
z
⑀z
⌬lz/2
=
2
l0z
–
⑀x
⌬lx/2
=
2
l0x
z
y
x
Figure 6.9 Axial (z) elongation
(positive strain) and lateral (x and y)
contractions (negative strains) in
response to an imposed tensile stress.
Solid lines represent dimensions after
stress application; dashed lines, before.
1496T_c06_131-173 11/16/05 17:06 Page 142
REVISED PAGES
142 • Chapter 6 / Mechanical Properties of Metals
of several elastic constants, their number depending on characteristics of the crystal
structure. Even for isotropic materials, for complete characterization of the elastic
properties, at least two constants must be given. Since the grain orientation is random in most polycrystalline materials, these may be considered to be isotropic; inorganic ceramic glasses are also isotropic. The remaining discussion of mechanical
behavior assumes isotropy and polycrystallinity because such is the character of
most engineering materials.
EXAMPLE PROBLEM 6.2
Computation of Load to Produce Specified
Diameter Change
A tensile stress is to be applied along the long axis of a cylindrical brass rod
that has a diameter of 10 mm (0.4 in.). Determine the magnitude of the load
required to produce a 2.5 103 mm (104 in.) change in diameter if the
deformation is entirely elastic.
Solution
This deformation situation is represented in the accompanying drawing.
F
d0
di
z
li
x
l0
⑀z =
⌬l
l0
=
⑀ x = ⌬d =
d0
li – l0
l0
d i – d0
d0
F
When the force F is applied, the specimen will elongate in the z direction and
at the same time experience a reduction in diameter, ¢d, of 2.5 103 mm in
the x direction. For the strain in the x direction,
x
¢d
2.5 103 mm
2.5 104
d0
10 mm
which is negative, since the diameter is reduced.
1496T_c06_131-173 11/16/05 17:06 Page 143
REVISED PAGES
6.5 Elastic Properties of Materials • 143
It next becomes necessary to calculate the strain in the z direction using
Equation 6.8. The value for Poisson’s ratio for brass is 0.34 (Table 6.1), and thus
z
12.5 104 2
x
7.35 104
n
0.34
The applied stress may now be computed using Equation 6.5 and the modulus of elasticity, given in Table 6.1 as 97 GPa (14 106 psi), as
s zE 17.35 104 2197 103 MPa2 71.3 MPa
Finally, from Equation 6.1, the applied force may be determined as
F sA0 s a
d0 2
bp
2
171.3 106 N/m2 2a
10 103 m 2
b p 5600 N 11293 lbf 2
2
Plastic Deformation
Elastic Plastic
Upper yield
point
y
Stress
Figure 6.10
(a) Typical stress–
strain behavior for
a metal showing
elastic and plastic
deformations, the
proportional limit P,
and the yield strength
sy, as determined
using the 0.002
strain offset method.
(b) Representative
stress–strain behavior
found for some
steels demonstrating
the yield point
phenomenon.
Stress
plastic deformation
For most metallic materials, elastic deformation persists only to strains of about
0.005. As the material is deformed beyond this point, the stress is no longer proportional to strain (Hooke’s law, Equation 6.5, ceases to be valid), and permanent,
nonrecoverable, or plastic deformation occurs. Figure 6.10a plots schematically the
tensile stress–strain behavior into the plastic region for a typical metal. The transition from elastic to plastic is a gradual one for most metals; some curvature results
at the onset of plastic deformation, which increases more rapidly with rising stress.
P
y
Lower yield
point
Strain
Strain
0.002
(a)
(b)
1496T_c06_131-173 12/21/05 7:43 Page 144
2nd REVISE PAGES
144 • Chapter 6 / Mechanical Properties of Metals
From an atomic perspective, plastic deformation corresponds to the breaking
of bonds with original atom neighbors and then reforming bonds with new neighbors as large numbers of atoms or molecules move relative to one another; upon
removal of the stress they do not return to their original positions. The mechanism
of this deformation is different for crystalline and amorphous materials. For crystalline solids, deformation is accomplished by means of a process called slip, which
involves the motion of dislocations as discussed in Section 7.2. Plastic deformation
in noncrystalline solids (as well as liquids) occurs by a viscous flow mechanism,
which is outlined in Section 12.10.
6.6 TENSILE PROPERTIES
Yielding and Yield Strength
yielding
proportional limit
yield strength
Metal Alloys
Most structures are designed to ensure that only elastic deformation will result when
a stress is applied. A structure or component that has plastically deformed, or experienced a permanent change in shape, may not be capable of functioning as intended.
It is therefore desirable to know the stress level at which plastic deformation begins, or where the phenomenon of yielding occurs. For metals that experience this
gradual elastic–plastic transition, the point of yielding may be determined as the
initial departure from linearity of the stress–strain curve; this is sometimes called
the proportional limit, as indicated by point P in Figure 6.10a. In such cases the
position of this point may not be determined precisely. As a consequence, a convention has been established wherein a straight line is constructed parallel to the
elastic portion of the stress–strain curve at some specified strain offset, usually 0.002.
The stress corresponding to the intersection of this line and the stress–strain curve
as it bends over in the plastic region is defined as the yield strength y.7 This is
demonstrated in Figure 6.10a. Of course, the units of yield strength are MPa or psi.8
For those materials having a nonlinear elastic region (Figure 6.6), use of the
strain offset method is not possible, and the usual practice is to define the yield
strength as the stress required to produce some amount of strain (e.g., 0.005).
Some steels and other materials exhibit the tensile stress–strain behavior as
shown in Figure 6.10b. The elastic–plastic transition is very well defined and occurs
abruptly in what is termed a yield point phenomenon.At the upper yield point, plastic
deformation is initiated with an actual decrease in stress. Continued deformation
fluctuates slightly about some constant stress value, termed the lower yield point;
stress subsequently rises with increasing strain. For metals that display this effect,
the yield strength is taken as the average stress that is associated with the lower yield
point, since it is well defined and relatively insensitive to the testing procedure.9 Thus,
it is not necessary to employ the strain offset method for these materials.
The magnitude of the yield strength for a metal is a measure of its resistance
to plastic deformation. Yield strengths may range from 35 MPa (5000 psi) for a lowstrength aluminum to over 1400 MPa (200,000 psi) for high-strength steels.
7
“Strength” is used in lieu of “stress” because strength is a property of the metal, whereas
stress is related to the magnitude of the applied load.
8
For Customary U.S. units, the unit of kilopounds per square inch (ksi) is sometimes used
for the sake of convenience, where
1 ksi 1000 psi
9
Note that to observe the yield point phenomenon, a “stiff” tensile-testing apparatus must
be used; by stiff is meant that there is very little elastic deformation of the machine during
loading.
1496T_c06_131-173 11/16/05 17:06 Page 145
REVISED PAGES
6.6 Tensile Properties • 145
Concept Check 6.1
Cite the primary differences between elastic, anelastic, and plastic deformation
behaviors.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Tensile Strength
Figure 6.11 Typical
engineering stress–
strain behavior to
fracture, point F.
The tensile strength
TS is indicated at
point M. The circular
insets represent the
geometry of the
deformed specimen
at various points
along the curve.
TS
M
F
Stress
tensile strength
After yielding, the stress necessary to continue plastic deformation in metals increases to a maximum, point M in Figure 6.11, and then decreases to the eventual
fracture, point F. The tensile strength TS (MPa or psi) is the stress at the maximum on the engineering stress–strain curve (Figure 6.11). This corresponds to the
maximum stress that can be sustained by a structure in tension; if this stress is
applied and maintained, fracture will result. All deformation up to this point is
uniform throughout the narrow region of the tensile specimen. However, at this
maximum stress, a small constriction or neck begins to form at some point, and all
subsequent deformation is confined at this neck, as indicated by the schematic
specimen insets in Figure 6.11. This phenomenon is termed “necking,” and fracture ultimately occurs at the neck. The fracture strength corresponds to the stress
at fracture.
Tensile strengths may vary anywhere from 50 MPa (7000 psi) for an aluminum
to as high as 3000 MPa (450,000 psi) for the high-strength steels. Ordinarily, when
the strength of a metal is cited for design purposes, the yield strength is used. This
is because by the time a stress corresponding to the tensile strength has been applied, often a structure has experienced so much plastic deformation that it is useless. Furthermore, fracture strengths are not normally specified for engineering
design purposes.
Strain
1496T_c06_131-173 11/16/05 17:06 Page 146
REVISED PAGES
146 • Chapter 6 / Mechanical Properties of Metals
EXAMPLE PROBLEM 6.3
Mechanical Property Determinations from
Stress–Strain Plot
From the tensile stress–strain behavior for the brass specimen shown in
Figure 6.12, determine the following:
(a) The modulus of elasticity
(b) The yield strength at a strain offset of 0.002
(c) The maximum load that can be sustained by a cylindrical specimen having an original diameter of 12.8 mm (0.505 in.)
(d) The change in length of a specimen originally 250 mm (10 in.) long that
is subjected to a tensile stress of 345 MPa (50,000 psi)
Solution
(a) The modulus of elasticity is the slope of the elastic or initial linear portion of the stress–strain curve. The strain axis has been expanded in the inset,
Figure 6.12, to facilitate this computation. The slope of this linear region is the
rise over the run, or the change in stress divided by the corresponding change
in strain; in mathematical terms,
E slope
s2 s1
¢s
2 1
¢
(6.10)
500
70
Tensile strength
450 MPa (65,000 psi)
60
400
300
50
103 psi
MPa 40
40
200
30
Yield strength
250 MPa (36,000 psi)
200
20
100
30
20
10
100
10
0
0
0
0.10
0
0
0.005
0.20
0.30
0
0.40
Strain
Figure 6.12 The stress–strain behavior for the brass specimen
discussed in Example Problem 6.3.
Stress (103 psi)
Stress (MPa)
A
1496T_c06_131-173 11/16/05 17:06 Page 147
REVISED PAGES
6.6 Tensile Properties • 147
Inasmuch as the line segment passes through the origin, it is convenient to
take both s1 and 1 as zero. If s2 is arbitrarily taken as 150 MPa, then 2 will
have a value of 0.0016. Therefore,
E
1150 02 MPa
93.8 GPa 113.6 106 psi2
0.0016 0
which is very close to the value of 97 GPa (14 106 psi) given for brass in
Table 6.1.
(b) The 0.002 strain offset line is constructed as shown in the inset; its intersection with the stress–strain curve is at approximately 250 MPa (36,000 psi),
which is the yield strength of the brass.
(c) The maximum load that can be sustained by the specimen is calculated
by using Equation 6.1, in which s is taken to be the tensile strength, from
Figure 6.12, 450 MPa (65,000 psi). Solving for F, the maximum load, yields
F sA0 s a
d0 2
bp
2
1450 106 N/m2 2 a
12.8 103 m 2
b p 57,900 N 113,000 lbf 2
2
(d) To compute the change in length, ¢l, in Equation 6.2, it is first necessary to determine the strain that is produced by a stress of 345 MPa. This is
accomplished by locating the stress point on the stress–strain curve, point A,
and reading the corresponding strain from the strain axis, which is approximately 0.06. Inasmuch as l0 250 mm, we have
¢l l0 10.0621250 mm2 15 mm 10.6 in.2
Ductility
Ductility is another important mechanical property. It is a measure of the degree
of plastic deformation that has been sustained at fracture. A material that experiences very little or no plastic deformation upon fracture is termed brittle. The tensile stress–strain behaviors for both ductile and brittle materials are schematically
illustrated in Figure 6.13.
Brittle
B
Figure 6.13 Schematic representations of
tensile stress–strain behavior for brittle and
ductile materials loaded to fracture.
Ductile
Stress
ductility
B⬘
A
C
C⬘
Strain
1496T_c06_131-173 11/16/05 17:06 Page 148
REVISED PAGES
148 • Chapter 6 / Mechanical Properties of Metals
Ductility may be expressed quantitatively as either percent elongation or percent
reduction in area. The percent elongation %EL is the percentage of plastic strain
at fracture, or
Ductility, as percent
elongation
%EL a
lf l0
l0
b 100
(6.11)
where lf is the fracture length10 and l0 is the original gauge length as above. Inasmuch as a significant proportion of the plastic deformation at fracture is confined
to the neck region, the magnitude of %EL will depend on specimen gauge length.
The shorter l0, the greater is the fraction of total elongation from the neck and, consequently, the higher the value of %EL. Therefore, l0 should be specified when percent elongation values are cited; it is commonly 50 mm (2 in.).
Percent reduction in area %RA is defined as
Ductility, as percent
reduction in area
%RA a
A0 Af
A0
b 100
(6.12)
where A0 is the original cross-sectional area and Af is the cross-sectional area at the
point of fracture.10 Percent reduction in area values are independent of both l0 and A0.
Furthermore, for a given material the magnitudes of %EL and %RA will, in general,
be different. Most metals possess at least a moderate degree of ductility at room temperature; however, some become brittle as the temperature is lowered (Section 8.6).
A knowledge of the ductility of materials is important for at least two reasons.
First, it indicates to a designer the degree to which a structure will deform plastically before fracture. Second, it specifies the degree of allowable deformation during fabrication operations. We sometimes refer to relatively ductile materials as
being “forgiving,” in the sense that they may experience local deformation without
fracture should there be an error in the magnitude of the design stress calculation.
Brittle materials are approximately considered to be those having a fracture
strain of less than about 5%.
Thus, several important mechanical properties of metals may be determined from
tensile stress–strain tests. Table 6.2 presents some typical room-temperature values of
Table 6.2 Typical Mechanical Properties of Several Metals and Alloys in an
Annealed State
Metal Alloy
Aluminum
Copper
Brass (70Cu–30Zn)
Iron
Nickel
Steel (1020)
Titanium
Molybdenum
10
Yield Strength
MPa (ksi)
Tensile Strength
MPa (ksi)
Ductility, %EL
[in 50 mm (2 in.)]
35 (5)
69 (10)
75 (11)
130 (19)
138 (20)
180 (26)
450 (65)
565 (82)
90 (13)
200 (29)
300 (44)
262 (38)
480 (70)
380 (55)
520 (75)
655 (95)
40
45
68
45
40
25
25
35
Both lf and Af are measured subsequent to fracture and after the two broken ends have
been repositioned back together.
1496T_c06_131-173 11/16/05 17:06 Page 149
REVISED PAGES
6.6 Tensile Properties • 149
120
800
–200°C
100
80
60
–100°C
400
40
Stress (103 psi)
600
Stress (MPa)
Figure 6.14
Engineering stress–
strain behavior
for iron at three
temperatures.
25°C
200
20
0
0
0.1
0.2
0.3
0.4
0
0.5
Strain
yield strength, tensile strength, and ductility for several of the common metals. These
properties are sensitive to any prior deformation, the presence of impurities, and/or
any heat treatment to which the metal has been subjected. The modulus of elasticity
is one mechanical parameter that is insensitive to these treatments. As with modulus
of elasticity, the magnitudes of both yield and tensile strengths decline with increasing
temperature; just the reverse holds for ductility—it usually increases with temperature.
Figure 6.14 shows how the stress–strain behavior of iron varies with temperature.
Resilience
resilience
Resilience is the capacity of a material to absorb energy when it is deformed elastically and then, upon unloading, to have this energy recovered.The associated property is the modulus of resilience, Ur, which is the strain energy per unit volume
required to stress a material from an unloaded state up to the point of yielding.
Computationally, the modulus of resilience for a specimen subjected to a uniaxial tension test is just the area under the engineering stress–strain curve taken to
yielding (Figure 6.15), or
Definition of
modulus of
resilience
Ur
y
sd
(6.13a)
0
Figure 6.15 Schematic representation showing how modulus
of resilience (corresponding to the shaded area) is determined
from the tensile stress–strain behavior of a material.
Stress
σy
0.002
y
Strain
1496T_c06_131-173 11/16/05 17:06 Page 150
REVISED PAGES
150 • Chapter 6 / Mechanical Properties of Metals
Assuming a linear elastic region,
Modulus of
resilience for linear
elastic behavior
Ur 12 syy
(6.13b)
in which y is the strain at yielding.
The units of resilience are the product of the units from each of the two axes
of the stress–strain plot. For SI units, this is joules per cubic meter (J/m3, equivalent to Pa), whereas with Customary U.S. units it is inch-pounds force per cubic inch
(in.-lbf/in.3, equivalent to psi). Both joules and inch-pounds force are units of energy, and thus this area under the stress–strain curve represents energy absorption
per unit volume (in cubic meters or cubic inches) of material.
Incorporation of Equation 6.5 into Equation 6.13b yields
Modulus of
resilience for linear
elastic behavior,
and incorporating
Hooke’s law
Ur 12 syy 12 sy a
sy
E
b
s2y
2E
(6.14)
Thus, resilient materials are those having high yield strengths and low moduli of
elasticity; such alloys would be used in spring applications.
Toughness
toughness
Toughness is a mechanical term that is used in several contexts; loosely speaking,
it is a measure of the ability of a material to absorb energy up to fracture. Specimen geometry as well as the manner of load application are important in toughness determinations. For dynamic (high strain rate) loading conditions and when a
notch (or point of stress concentration) is present, notch toughness is assessed by
using an impact test, as discussed in Section 8.6. Furthermore, fracture toughness is
a property indicative of a material’s resistance to fracture when a crack is present
(Section 8.5).
For the static (low strain rate) situation, toughness may be ascertained from
the results of a tensile stress–strain test. It is the area under the s– curve up to
the point of fracture. The units for toughness are the same as for resilience (i.e.,
energy per unit volume of material). For a material to be tough, it must display
both strength and ductility; often, ductile materials are tougher than brittle ones.
This is demonstrated in Figure 6.13, in which the stress–strain curves are plotted
for both material types. Hence, even though the brittle material has higher yield
and tensile strengths, it has a lower toughness than the ductile one, by virtue of
lack of ductility; this is deduced by comparing the areas ABC and AB¿C¿ in
Figure 6.13.
Concept Check 6.2
Of those metals listed in Table 6.3,
(a) Which will experience the greatest percent reduction in area? Why?
(b) Which is the strongest? Why?
(c) Which is the stiffest? Why?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
1496T_c06_131-173 11/16/05 17:06 Page 151
REVISED PAGES
6.7 True Stress and Strain • 151
Table 6.3 Tensile Stress–Strain Data for Several Hypothetical Metals to be
Used with Concept Checks 6.2 and 6.4
Material
A
B
C
D
E
Yield
Strength
(MPa)
310
100
415
700
Tensile
Strength
(MPa)
Strain at
Fracture
340
120
550
850
Fractures before yielding
0.23
0.40
0.15
0.14
Fracture
Strength
(MPa)
Elastic
Modulus
(GPa)
265
105
500
720
650
210
150
310
210
350
6.7 TRUE STRESS AND STRAIN
true stress
Definition of true
stress
true strain
Definition of true
strain
From Figure 6.11, the decline in the stress necessary to continue deformation past
the maximum, point M, seems to indicate that the metal is becoming weaker. This
is not at all the case; as a matter of fact, it is increasing in strength. However, the
cross-sectional area is decreasing rapidly within the neck region, where deformation is
occurring. This results in a reduction in the load-bearing capacity of the specimen.
The stress, as computed from Equation 6.1, is on the basis of the original crosssectional area before any deformation, and does not take into account this reduction
in area at the neck.
Sometimes it is more meaningful to use a true stress–true strain scheme. True
stress sT is defined as the load F divided by the instantaneous cross-sectional area
Ai over which deformation is occurring (i.e., the neck, past the tensile point), or
sT
F
Ai
(6.15)
Furthermore, it is occasionally more convenient to represent strain as true strain
T, defined by
T ln
li
l0
(6.16)
If no volume change occurs during deformation—that is, if
Aili A0l0
(6.17)
true and engineering stress and strain are related according to
Conversion of
engineering stress
to true stress
Conversion of
engineering strain
to true strain
sT s11 2
(6.18a)
T ln11 2
(6.18b)
Equations 6.18a and 6.18b are valid only to the onset of necking; beyond this point
true stress and strain should be computed from actual load, cross-sectional area,
and gauge length measurements.
A schematic comparison of engineering and true stress–strain behaviors is made
in Figure 6.16. It is worth noting that the true stress necessary to sustain increasing
strain continues to rise past the tensile point M¿.
1496T_c06_131-173 11/16/05 17:06 Page 152
REVISED PAGES
152 • Chapter 6 / Mechanical Properties of Metals
Figure 6.16 A comparison of typical
tensile engineering stress–strain and
true stress–strain behaviors. Necking
begins at point M on the engineering
curve, which corresponds to M¿ on the
true curve. The “corrected” true stress–
strain curve takes into account the
complex stress state within the neck
region.
True
Stress
M⬘
Corrected
M
Engineering
Strain
Coincident with the formation of a neck is the introduction of a complex stress
state within the neck region (i.e., the existence of other stress components in addition to the axial stress). As a consequence, the correct stress (axial) within the neck
is slightly lower than the stress computed from the applied load and neck crosssectional area. This leads to the “corrected” curve in Figure 6.16.
For some metals and alloys the region of the true stress–strain curve from the
onset of plastic deformation to the point at which necking begins may be approximated by
True stress-true
strain relationship
in plastic region of
deformation (to
point of necking)
sT KTn
(6.19)
In this expression, K and n are constants; these values will vary from alloy to alloy,
and will also depend on the condition of the material (i.e., whether it has been plastically deformed, heat treated, etc.). The parameter n is often termed the strainhardening exponent and has a value less than unity. Values of n and K for several
alloys are contained in Table 6.4.
Table 6.4 Tabulation of n and K Values (Equation 6.19) for
Several Alloys
K
Material
n
MPa
psi
Low-carbon steel
(annealed)
0.21
600
87,000
4340 steel alloy
(tempered @ 315C)
0.12
2650
385,000
304 stainless steel
(annealed)
0.44
1400
205,000
Copper
(annealed)
0.44
530
76,500
Naval brass
(annealed)
0.21
585
85,000
2024 aluminum alloy
(heat treated—T3)
0.17
780
113,000
AZ-31B magnesium alloy
(annealed)
0.16
450
66,000
1496T_c06_131-173 11/16/05 17:06 Page 153
REVISED PAGES
6.7 True Stress and Strain • 153
EXAMPLE PROBLEM 6.4
Ductility and True-Stress-At-Fracture Computations
A cylindrical specimen of steel having an original diameter of 12.8 mm
(0.505 in.) is tensile tested to fracture and found to have an engineering fracture strength sf of 460 MPa (67,000 psi). If its cross-sectional diameter at
fracture is 10.7 mm (0.422 in.), determine:
(a) The ductility in terms of percent reduction in area
(b) The true stress at fracture
Solution
(a) Ductility is computed using Equation 6.12, as
10.7 mm 2
12.8 mm 2
bpa
bp
2
2
%RA
100
12.8 mm 2
a
bp
2
2
128.7 mm 89.9 mm2
100 30%
128.7 mm2
a
(b) True stress is defined by Equation 6.15, where in this case the area is
taken as the fracture area Af. However, the load at fracture must first be
computed from the fracture strength as
1 m2
F sf A0 1460 106 N/m2 21128.7 mm2 2 a 6
b 59,200 N
10 mm2
Thus, the true stress is calculated as
sT
F
Af
59,200 N
1 m2
189.9 mm2 2 a 6
b
10 mm2
6.6 108 N/m2 660 MPa 195,700 psi2
EXAMPLE PROBLEM 6.5
Calculation of Strain-Hardening Exponent
Compute the strain-hardening exponent n in Equation 6.19 for an alloy in
which a true stress of 415 MPa (60,000 psi) produces a true strain of 0.10;
assume a value of 1035 MPa (150,000 psi) for K.
Solution
This requires some algebraic manipulation of Equation 6.19 so that n becomes
the dependent parameter. This is accomplished by taking logarithms and
rearranging. Solving for n yields
log sT log K
log T
log1415 MPa2 log11035 MPa2
0.40
log10.12
n
1496T_c06_131-173 11/16/05 17:06 Page 154
REVISED PAGES
154 • Chapter 6 / Mechanical Properties of Metals
6.8 ELASTIC RECOVERY AFTER PLASTIC
DEFORMATION
Upon release of the load during the course of a stress–strain test, some fraction of
the total deformation is recovered as elastic strain. This behavior is demonstrated
in Figure 6.17, a schematic engineering stress–strain plot. During the unloading cycle, the curve traces a near straight-line path from the point of unloading (point D),
and its slope is virtually identical to the modulus of elasticity, or parallel to the initial elastic portion of the curve. The magnitude of this elastic strain, which is regained during unloading, corresponds to the strain recovery, as shown in Figure 6.17.
If the load is reapplied, the curve will traverse essentially the same linear portion
in the direction opposite to unloading; yielding will again occur at the unloading
stress level where the unloading began. There will also be an elastic strain recovery associated with fracture.
6.9 COMPRESSIVE, SHEAR, AND TORSIONAL
DEFORMATION
Of course, metals may experience plastic deformation under the influence of applied compressive, shear, and torsional loads. The resulting stress–strain behavior
into the plastic region will be similar to the tensile counterpart (Figure 6.10a: yielding and the associated curvature). However, for compression, there will be no
maximum, since necking does not occur; furthermore, the mode of fracture will be
different from that for tension.
Concept Check 6.3
Make a schematic plot showing the tensile engineering stress–strain behavior for a
typical metal alloy to the point of fracture. Now superimpose on this plot a schematic
compressive engineering stress–strain curve for the same alloy. Explain any differences between the two curves.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Figure 6.17 Schematic tensile
stress–strain diagram showing the
phenomena of elastic strain recovery
and strain hardening. The initial yield
strength is designated as sy0; syi is the
yield strength after releasing the load
at point D, and then upon reloading.
D
yi
y0
Stress
Unload
Reapply
load
Strain
Elastic strain
recovery
1496T_c06_131-173 11/16/05 17:06 Page 155
REVISED PAGES
6.10 Hardness • 155
6.10 HARDNESS
hardness
Another mechanical property that may be important to consider is hardness, which
is a measure of a material’s resistance to localized plastic deformation (e.g., a small
dent or a scratch). Early hardness tests were based on natural minerals with a scale
constructed solely on the ability of one material to scratch another that was softer.
A qualitative and somewhat arbitrary hardness indexing scheme was devised,
termed the Mohs scale, which ranged from 1 on the soft end for talc to 10 for diamond. Quantitative hardness techniques have been developed over the years in
which a small indenter is forced into the surface of a material to be tested, under
controlled conditions of load and rate of application. The depth or size of the resulting indentation is measured, which in turn is related to a hardness number;
the softer the material, the larger and deeper is the indentation, and the lower
the hardness index number. Measured hardnesses are only relative (rather than
absolute), and care should be exercised when comparing values determined by
different techniques.
Hardness tests are performed more frequently than any other mechanical test
for several reasons:
1. They are simple and inexpensive—ordinarily no special specimen need be
prepared, and the testing apparatus is relatively inexpensive.
2. The test is nondestructive—the specimen is neither fractured nor excessively
deformed; a small indentation is the only deformation.
3. Other mechanical properties often may be estimated from hardness data,
such as tensile strength (see Figure 6.19).
Rockwell Hardness Tests 11
The Rockwell tests constitute the most common method used to measure hardness
because they are so simple to perform and require no special skills. Several different scales may be utilized from possible combinations of various indenters and different loads, which permit the testing of virtually all metal alloys (as well as some
polymers). Indenters include spherical and hardened steel balls having diameters
of 161 , 18, 14, and 12 in. (1.588, 3.175, 6.350, and 12.70 mm), and a conical diamond (Brale)
indenter, which is used for the hardest materials.
With this system, a hardness number is determined by the difference in depth
of penetration resulting from the application of an initial minor load followed by a
larger major load; utilization of a minor load enhances test accuracy. On the basis
of the magnitude of both major and minor loads, there are two types of tests: Rockwell and superficial Rockwell. For Rockwell, the minor load is 10 kg, whereas major
loads are 60, 100, and 150 kg. Each scale is represented by a letter of the alphabet;
several are listed with the corresponding indenter and load in Tables 6.5 and 6.6a.
For superficial tests, 3 kg is the minor load; 15, 30, and 45 kg are the possible major load values. These scales are identified by a 15, 30, or 45 (according to load),
followed by N, T, W, X, or Y, depending on indenter. Superficial tests are frequently
performed on thin specimens. Table 6.6b presents several superficial scales.
When specifying Rockwell and superficial hardnesses, both hardness number
and scale symbol must be indicated. The scale is designated by the symbol HR
11
ASTM Standard E 18, “Standard Test Methods for Rockwell Hardness and Rockwell
Superficial Hardness of Metallic Materials.”
Diamond
pyramid
Knoop
microhardness
120°
l/b = 7.11
b/t = 4.00
136°
d
D
Side View
t
Shape of Indentation
d1
l
d
b
d1
Top View
15 kg
30 kg ¶ Superficial Rockwell
45 kg
60 kg
100 kg ¶ Rockwell
150 kg
P
P
P
Load
pD 3 D 2D2 d 2 4
2P
HK 14.2P l 2
HV 1.854Pd 21
HB
Formula for
Hardness Numbera
For the hardness formulas given, P (the applied load) is in kg, while D, d, d1, and l are all in mm.
Source: Adapted from H. W. Hayden, W. G. Moffatt, and J. Wulff, The Structure and Properties of Materials, Vol. III, Mechanical Behavior. Copyright ©
1965 by John Wiley & Sons, New York. Reprinted by permission of John Wiley & Sons, Inc.
a
Diamond
cone;
1 1 1 1
16 , 8 , 4 , 2 in.
diameter
steel spheres
Diamond
pyramid
Vickers
microhardness
Rockwell and
Superficial
Rockwell
10-mm sphere
of steel or
tungsten carbide
Indenter
Brinell
Test
Table 6.5 Hardness-Testing Techniques
1496T_c06_131-173 11/16/05 17:06 Page 156
REVISED PAGES
156 • Chapter 6 / Mechanical Properties of Metals
1496T_c06_131-173 11/16/05 17:06 Page 157
REVISED PAGES
6.10 Hardness • 157
Table 6.6a Rockwell Hardness Scales
Scale Symbol
Indenter
Major Load (kg)
A
B
C
D
E
F
G
H
K
Diamond
1
16 -in. ball
Diamond
Diamond
1
8 -in. ball
1
16 -in. ball
1
16 -in. ball
1
8 -in. ball
1
8 -in. ball
60
100
150
100
100
60
150
60
150
Table 6.6b Superficial Rockwell Hardness Scales
Scale Symbol
Indenter
Major Load (kg)
15N
30N
45N
15T
30T
45T
15W
30W
45W
Diamond
Diamond
Diamond
1
16 -in. ball
1
16 -in. ball
1
16 -in. ball
1
8 -in. ball
1
8 -in. ball
1
8 -in. ball
15
30
45
15
30
45
15
30
45
followed by the appropriate scale identification.12 For example, 80 HRB represents a Rockwell hardness of 80 on the B scale, and 60 HR30W indicates a superficial hardness of 60 on the 30W scale.
For each scale, hardnesses may range up to 130; however, as hardness values
rise above 100 or drop below 20 on any scale, they become inaccurate; and because
the scales have some overlap, in such a situation it is best to utilize the next harder
or softer scale.
Inaccuracies also result if the test specimen is too thin, if an indentation is made
too near a specimen edge, or if two indentations are made too close to one another.
Specimen thickness should be at least ten times the indentation depth, whereas
allowance should be made for at least three indentation diameters between the
center of one indentation and the specimen edge, or to the center of a second indentation. Furthermore, testing of specimens stacked one on top of another is not
recommended. Also, accuracy is dependent on the indentation being made into a
smooth flat surface.
The modern apparatus for making Rockwell hardness measurements (see the
chapter-opening photograph for this chapter) is automated and very simple to use;
hardness is read directly, and each measurement requires only a few seconds.
The modern testing apparatus also permits a variation in the time of load application. This variable must also be considered in interpreting hardness data.
12
Rockwell scales are also frequently designated by an R with the appropriate scale letter
as a subscript, for example, RC denotes the Rockwell C scale.
1496T_c06_131-173 11/16/05 17:06 Page 158
REVISED PAGES
158 • Chapter 6 / Mechanical Properties of Metals
Brinell Hardness Tests 13
In Brinell tests, as in Rockwell measurements, a hard, spherical indenter is forced
into the surface of the metal to be tested. The diameter of the hardened steel (or
tungsten carbide) indenter is 10.00 mm (0.394 in.). Standard loads range between
500 and 3000 kg in 500-kg increments; during a test, the load is maintained constant
for a specified time (between 10 and 30 s). Harder materials require greater applied loads. The Brinell hardness number, HB, is a function of both the magnitude
of the load and the diameter of the resulting indentation (see Table 6.5).14 This diameter is measured with a special low-power microscope, utilizing a scale that is
etched on the eyepiece. The measured diameter is then converted to the appropriate HB number using a chart; only one scale is employed with this technique.
Semiautomatic techniques for measuring Brinell hardness are available. These
employ optical scanning systems consisting of a digital camera mounted on a flexible probe, which allows positioning of the camera over the indentation. Data
from the camera are transferred to a computer that analyzes the indentation, determines its size, and then calculates the Brinell hardness number. For this technique, surface finish requirements are normally more stringent that for manual
measurements.
Maximum specimen thickness as well as indentation position (relative to specimen edges) and minimum indentation spacing requirements are the same as for
Rockwell tests. In addition, a well-defined indentation is required; this necessitates
a smooth flat surface in which the indentation is made.
Knoop and Vickers Microindentation Hardness Tests 15
Two other hardness-testing techniques are Knoop (pronounced nup) and Vickers
(sometimes also called diamond pyramid). For each test a very small diamond indenter having pyramidal geometry is forced into the surface of the specimen. Applied loads are much smaller than for Rockwell and Brinell, ranging between 1 and
1000 g. The resulting impression is observed under a microscope and measured; this
measurement is then converted into a hardness number (Table 6.5). Careful specimen surface preparation (grinding and polishing) may be necessary to ensure a
well-defined indentation that may be accurately measured. The Knoop and Vickers
hardness numbers are designated by HK and HV, respectively,16 and hardness scales
for both techniques are approximately equivalent. Knoop and Vickers are referred
to as microindentation-testing methods on the basis of indenter size. Both are well
suited for measuring the hardness of small, selected specimen regions; furthermore,
Knoop is used for testing brittle materials such as ceramics.
The modern microindentation hardness-testing equipment has been automated
by coupling the indenter apparatus to an image analyzer that incorporates a computer and software package. The software controls important system functions to
include indent location, indent spacing, computation of hardness values, and plotting of data.
13
ASTM Standard E 10, “Standard Test Method for Brinell Hardness of Metallic Materials.”
The Brinell hardness number is also represented by BHN.
15
ASTM Standard E 92, “Standard Test Method for Vickers Hardness of Metallic Materials,” and ASTM Standard E 384, “Standard Test for Microhardness of Materials.”
16
Sometimes KHN and VHN are used to denote Knoop and Vickers hardness numbers,
respectively.
14
1496T_c06_131-173 12/21/05 7:43 Page 159
2nd REVISE PAGES
6.10 Hardness • 159
Other hardness-testing techniques are frequently employed but will not be discussed here; these include ultrasonic microhardness, dynamic (Scleroscope), durometer (for plastic and elastomeric materials), and scratch hardness tests. These are
described in references provided at the end of the chapter.
Hardness Conversion
The facility to convert the hardness measured on one scale to that of another is
most desirable. However, since hardness is not a well-defined material property, and
because of the experimental dissimilarities among the various techniques, a comprehensive conversion scheme has not been devised. Hardness conversion data have
been determined experimentally and found to be dependent on material type and
characteristics. The most reliable conversion data exist for steels, some of which are
presented in Figure 6.18 for Knoop, Brinell, and two Rockwell scales; the Mohs
scale is also included. Detailed conversion tables for various other metals and
Figure 6.18
Comparison of
several hardness
scales. (Adapted
from G. F. Kinney,
Engineering
Properties and
Applications of
Plastics, p. 202.
Copyright © 1957 by
John Wiley & Sons,
New York.
Reprinted by
permission of John
Wiley & Sons, Inc.)
10,000
10
Diamond
5,000
2,000
Nitrided steels
1,000
80
1000
800
600
Cutting tools
60
110
40
400
200
100
100
20
200
100
80
0
60
Rockwell
C
40
20
0
Knoop
hardness
50
20
8
Topaz
7
Quartz
6
Orthoclase
5
Apatite
4
3
Fluorite
Calcite
2
Gypsum
1
Talc
File hard
500
300
9
Corundum
or
sapphire
Rockwell
B
Easily
machined
steels
Brasses
and
aluminum
alloys
Most
plastics
10
5
Brinell
hardness
Mohs
hardness
1496T_c06_131-173 11/16/05 17:06 Page 160
REVISED PAGES
160 • Chapter 6 / Mechanical Properties of Metals
alloys are contained in ASTM Standard E 140, “Standard Hardness Conversion
Tables for Metals.” In light of the preceding discussion, care should be exercised in
extrapolation of conversion data from one alloy system to another.
Correlation Between Hardness and Tensile Strength
Both tensile strength and hardness are indicators of a metal’s resistance to plastic
deformation. Consequently, they are roughly proportional, as shown in Figure 6.19,
for tensile strength as a function of the HB for cast iron, steel, and brass. The same
proportionality relationship does not hold for all metals, as Figure 6.19 indicates.
As a rule of thumb for most steels, the HB and the tensile strength are related
according to
For steel alloys,
conversion of Brinell
hardness to tensile
strength
TS1MPa2 3.45 HB
(6.20a)
TS1psi2 500 HB
(6.20b)
Concept Check 6.4
Of those metals listed in Table 6.3, which is the hardest? Why?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Rockwell hardness
60 70 80 90
100 HRB
20
30
40
50 HRC
250
1500
150
1000
100
500
Brass
Cast iron (nodular)
50
0
0
100
200
300
Brinell hardness number
400
0
500
Tensile strength (103 psi)
Tensile strength (MPa)
200
Steels
Figure 6.19 Relationships
between hardness and tensile
strength for steel, brass, and
cast iron. [Data taken from
Metals Handbook: Properties
and Selection: Irons and
Steels, Vol. 1, 9th edition,
B. Bardes (Editor), American
Society for Metals, 1978,
pp. 36 and 461; and Metals
Handbook: Properties and
Selection: Nonferrous Alloys
and Pure Metals, Vol. 2, 9th
edition, H. Baker (Managing
Editor), American Society for
Metals, 1979, p. 327.]
1496T_c06_131-173 11/16/05 17:06 Page 161
REVISED PAGES
6.11 Variability of Material Properties • 161
P ro p e r t y Va r i a b i l i t y a n d
D e s i g n / S a f e t y Fa c t o r s
6.11 VARIABILITY OF MATERIAL PROPERTIES
At this point it is worthwhile to discuss an issue that sometimes proves troublesome
to many engineering students—namely, that measured material properties are not exact quantities. That is, even if we have a most precise measuring apparatus and a highly
controlled test procedure, there will always be some scatter or variability in the data
that are collected from specimens of the same material. For example, consider a number of identical tensile samples that are prepared from a single bar of some metal alloy, which samples are subsequently stress–strain tested in the same apparatus. We
would most likely observe that each resulting stress–strain plot is slightly different from
the others. This would lead to a variety of modulus of elasticity, yield strength, and tensile strength values. A number of factors lead to uncertainties in measured data. These
include the test method, variations in specimen fabrication procedures, operator bias,
and apparatus calibration. Furthermore, inhomogeneities may exist within the same
lot of material, and/or slight compositional and other differences from lot to lot. Of
course, appropriate measures should be taken to minimize the possibility of measurement error, and also to mitigate those factors that lead to data variability.
It should also be mentioned that scatter exists for other measured material properties such as density, electrical conductivity, and coefficient of thermal expansion.
It is important for the design engineer to realize that scatter and variability of
materials properties are inevitable and must be dealt with appropriately. On occasion, data must be subjected to statistical treatments and probabilities determined.
For example, instead of asking the question, “What is the fracture strength of this
alloy?” the engineer should become accustomed to asking the question, “What is
the probability of failure of this alloy under these given circumstances?”
It is often desirable to specify a typical value and degree of dispersion (or scatter) for some measured property; such is commonly accomplished by taking the
average and the standard deviation, respectively.
Computation of Average and Standard Deviation Values
An average value is obtained by dividing the sum of all measured values by the
number of measurements taken. In mathematical terms, the average x of some
parameter x is
n
a xi
Computation of
average value
x
i1
(6.21)
n
where n is the number of observations or measurements and xi is the value of a
discrete measurement.
Furthermore, the standard deviation s is determined using the following
expression:
n
Computation of
standard deviation
s £
2
a 1xi x2
i1
n1
§
1 2
(6.22)
1496T_c06_131-173 11/16/05 17:06 Page 162
REVISED PAGES
162 • Chapter 6 / Mechanical Properties of Metals
where xi, x, and n are defined above. A large value of the standard deviation corresponds to a high degree of scatter.
EXAMPLE PROBLEM 6.6
Average and Standard Deviation Computations
The following tensile strengths were measured for four specimens of the same
steel alloy:
Sample Number
Tensile Strength
(MPa)
1
2
3
4
520
512
515
522
(a) Compute the average tensile strength.
(b) Determine the standard deviation.
Solution
(a) The average tensile strength 1TS2 is computed using Equation 6.21 with
n 4:
4
TS
a 1TS2 i
i1
4
520 512 515 522
4
517 MPa
525
525
Tensile strength (MPa)
Tensile strength (MPa)
僓僒
TS + s
520
515
520
僓僒
TS
515
僓僒
TS – s
510
1
2
3
4
510
Sample number
(a)
(b)
Figure 6.20 (a) Tensile strength data associated with Example Problem 6.6. (b) The
manner in which these data could be plotted. The data point corresponds to the
average value of the tensile strength 1TS2; error bars that indicate the degree of scatter
correspond to the average value plus and minus the standard deviation 1TS s2.
1496T_c06_131-173 11/16/05 17:06 Page 163
REVISED PAGES
6.12 Design/Safety Factors • 163
(b) For the standard deviation, using Equation 6.22,
4
s
£
2
a 51TS2 i TS6
i1
41
§
1 2
1520 5172 2 1512 5172 2 1515 5172 2 1522 5172 2 1 2
d
41
4.6 MPa
c
Figure 6.20 presents the tensile strength by specimen number for this
example problem and also how the data may be represented in graphical form.
The tensile strength data point (Figure 6.20b) corresponds to the average value
TS, whereas scatter is depicted by error bars (short horizontal lines) situated
above and below the data point symbol and connected to this symbol by vertical lines. The upper error bar is positioned at a value of the average value
plus the standard deviation 1TS s2, whereas the lower error bar corresponds
to the average minus the standard deviation 1TS s2.
6.12 DESIGN/SAFETY FACTORS
design stress
There will always be uncertainties in characterizing the magnitude of applied
loads and their associated stress levels for in-service applications; ordinarily load
calculations are only approximate. Furthermore, as noted in the previous section,
virtually all engineering materials exhibit a variability in their measured mechanical properties. Consequently, design allowances must be made to protect
against unanticipated failure. One way this may be accomplished is by establishing, for the particular application, a design stress, denoted as sd. For static
situations and when ductile materials are used, sd is taken as the calculated stress
level sc (on the basis of the estimated maximum load) multiplied by a design
factor, N¿ ; that is,
sd N¿sc
safe stress
Computation of safe
(or working) stress
(6.23)
where N¿ is greater than unity. Thus, the material to be used for the particular
application is chosen so as to have a yield strength at least as high as this value
of sd.
Alternatively, a safe stress or working stress, sw, is used instead of design stress.
This safe stress is based on the yield strength of the material and is defined as the
yield strength divided by a factor of safety, N, or
sw
sy
N
(6.24)
Utilization of design stress (Equation 6.23) is usually preferred since it is based
on the anticipated maximum applied stress instead of the yield strength of the material; normally there is a greater uncertainty in estimating this stress level than in
the specification of the yield strength. However, in the discussion of this text, we
are concerned with factors that influence the yield strengths of metal alloys and not
in the determination of applied stresses; therefore, the succeeding discussion will
deal with working stresses and factors of safety.
1496T_c06_131-173 12/21/05 7:43 Page 164
2nd REVISE PAGES
164 • Chapter 6 / Mechanical Properties of Metals
Case Study:
“Materials Selection
for a Torsionally
Stressed Cylindrical
Shaft,” Chapter 22,
which may be found
at www.wiley.com/
college/callister
(Student Companion
Site).
The choice of an appropriate value of N is necessary. If N is too large, then
component overdesign will result; that is, either too much material or an alloy having a higher-than-necessary strength will be used. Values normally range between
1.2 and 4.0. Selection of N will depend on a number of factors, including economics, previous experience, the accuracy with which mechanical forces and material
properties may be determined, and, most important, the consequences of failure in
terms of loss of life and/or property damage.
DESIGN EXAMPLE 6.1
Specification of Support Post Diameter
A tensile-testing apparatus is to be constructed that must withstand a maximum load of 220,000 N (50,000 lbf). The design calls for two cylindrical support posts, each of which is to support half of the maximum load. Furthermore,
plain-carbon (1045) steel ground and polished shafting rounds are to be used;
the minimum yield and tensile strengths of this alloy are 310 MPa (45,000 psi)
and 565 MPa (82,000 psi), respectively. Specify a suitable diameter for these
support posts.
Solution
The first step in this design process is to decide on a factor of safety, N, which
then allows determination of a working stress according to Equation 6.24. In
addition, to ensure that the apparatus will be safe to operate, we also want to
minimize any elastic deflection of the rods during testing; therefore, a relatively
conservative factor of safety is to be used, say N 5. Thus, the working stress
sw is just
sy
sw
N
310 MPa
62 MPa 19000 psi2
5
From the definition of stress, Equation 6.1,
d 2
F
A0 a b p
sw
2
where d is the rod diameter and F is the applied force; furthermore, each of the
two rods must support half of the total force or 110,000 N (25,000 psi). Solving
for d leads to
F
d2
B psw
2
110,000 N
B p162 106 N/m2 2
4.75 102 m 47.5 mm 11.87 in.2
Therefore, the diameter of each of the two rods should be 47.5 mm or 1.87 in.
1496T_c06_131-173 11/16/05 17:06 Page 165
REVISED PAGES
Summary • 165
SUMMARY
Concepts of Stress and Strain
Stress–Strain Behavior
Elastic Properties of Materials
True Stress and Strain
A number of the important mechanical properties of materials, predominantly
metals, have been discussed in this chapter. Concepts of stress and strain were first
introduced. Stress is a measure of an applied mechanical load or force, normalized
to take into account cross-sectional area. Two different stress parameters were
defined—engineering stress and true stress. Strain represents the amount of deformation induced by a stress; both engineering and true strains are used.
Some of the mechanical characteristics of metals can be ascertained by simple
stress–strain tests. There are four test types: tension, compression, torsion, and shear.
Tensile are the most common. A material that is stressed first undergoes elastic, or
nonpermanent, deformation, wherein stress and strain are proportional. The constant of proportionality is the modulus of elasticity for tension and compression,
and is the shear modulus when the stress is shear. Poisson’s ratio represents the
negative ratio of transverse and longitudinal strains.
Tensile Properties
The phenomenon of yielding occurs at the onset of plastic or permanent deformation; yield strength is determined by a strain offset method from the stress–strain
behavior, which is indicative of the stress at which plastic deformation begins. Tensile strength corresponds to the maximum tensile stress that may be sustained by
a specimen, whereas percents elongation and reduction in area are measures of
ductility—the amount of plastic deformation that has occurred at fracture. Resilience is the capacity of a material to absorb energy during elastic deformation;
modulus of resilience is the area beneath the engineering stress–strain curve up
to the yield point. Also, static toughness represents the energy absorbed during
the fracture of a material, and is taken as the area under the entire engineering
stress–strain curve. Ductile materials are normally tougher than brittle ones.
Hardness
Hardness is a measure of the resistance to localized plastic deformation. In several
popular hardness-testing techniques (Rockwell, Brinell, Knoop, and Vickers) a small
indenter is forced into the surface of the material, and an index number is determined on the basis of the size or depth of the resulting indentation. For many metals, hardness and tensile strength are approximately proportional to each other.
Variability of Material Properties
Measured mechanical properties (as well as other material properties) are not exact and precise quantities, in that there will always be some scatter for the measured data. Typical material property values are commonly specified in terms of
averages, whereas magnitudes of scatter may be expressed as standard deviations.
Design/Safety Factors
As a result of uncertainties in both measured mechanical properties and inservice
applied stresses, design or safe stresses are normally utilized for design purposes. For
ductile materials, safe stress is the ratio of the yield strength and the factor of safety.
1496T_c06_131-173 11/16/05 17:06 Page 166
REVISED PAGES
166 • Chapter 6 / Mechanical Properties of Metals
I M P O R TA N T T E R M S A N D C O N C E P T S
Anelasticity
Design stress
Ductility
Elastic deformation
Elastic recovery
Engineering strain
Engineering stress
Hardness
Modulus of elasticity
Plastic deformation
Poisson’s ratio
Proportional limit
Resilience
Safe stress
Shear
Tensile strength
Toughness
True strain
True stress
Yielding
Yield strength
REFERENCES
ASM Handbook, Vol. 8, Mechanical Testing and
Evaluation, ASM International, Materials Park,
OH, 2000.
Boyer, H. E. (Editor), Atlas of Stress–Strain Curves,
2nd edition,ASM International, Materials Park,
OH, 2002.
Chandler, H. (Editor), Hardness Testing, 2nd edition,
ASM International, Materials Park, OH, 2000.
Courtney, T. H., Mechanical Behavior of Materials,
2nd edition, McGraw-Hill Higher Education,
Burr Ridge, IL, 2000.
Davis, J. R. (Editor), Tensile Testing, 2nd edition,
ASM International, Materials Park, OH, 2004.
Dieter, G. E., Mechanical Metallurgy, 3rd edition,
McGraw-Hill Book Company, New York, 1986.
Dowling, N. E., Mechanical Behavior of Materials, 2nd edition, Prentice Hall PTR, Paramus,
NJ, 1998.
McClintock, F. A. and A. S. Argon, Mechanical
Behavior of Materials, Addison-Wesley Publishing Co., Reading, MA, 1966. Reprinted by
CBLS Publishers, Marietta, OH, 1993.
Meyers, M. A. and K. K. Chawla, Mechanical
Behavior of Materials, Prentice Hall PTR,
Paramus, NJ, 1999.
QUESTIONS AND PROBLEMS
Concepts of Stress and Strain
6.1 Using mechanics of materials principles (i.e.,
equations of mechanical equilibrium applied
to a free-body diagram), derive Equations
6.4a and 6.4b.
6.2 (a) Equations 6.4a and 6.4b are expressions
for normal 1s¿2 and shear 1t¿2 stresses, respectively, as a function of the applied tensile stress 1s2 and the inclination angle of
the plane on which these stresses are taken
(u of Figure 6.4). Make a plot on which
is presented the orientation parameters of
these expressions (i.e., cos2 u and sin u cos u)
versus u.
(b) From this plot, at what angle of inclination is the normal stress a maximum?
(c) Also, at what inclination angle is the shear
stress a maximum?
Stress–Strain Behavior
6.3 A specimen of copper having a rectangular
cross section 15.2 mm 19.1 mm (0.60 in.
0.75 in.) is pulled in tension with 44,500 N
(10,000 lbf) force, producing only elastic deformation. Calculate the resulting strain.
6.4 A cylindrical specimen of a nickel alloy having an elastic modulus of 207 GPa (30 106
psi) and an original diameter of 10.2 mm
(0.40 in.) will experience only elastic deformation when a tensile load of 8900 N (2000 lbf)
is applied. Compute the maximum length
of the specimen before deformation if the
maximum allowable elongation is 0.25 mm
(0.010 in.).
6.5 An aluminum bar 125 mm (5.0 in.) long and
having a square cross section 16.5 mm (0.65 in.)
on an edge is pulled in tension with a load
1496T_c06_131-173 12/21/05 7:43 Page 167
2nd REVISE PAGES
Questions and Problems • 167
6.8
6.9
Figure 6.21 Tensile
stress–strain
behavior for an
alloy steel.
tension. Determine its elongation when a load
of 65,250 N (14,500 lbf ) is applied.
6.10 Figure 6.22 shows, for a gray cast iron, the
tensile engineering stress–strain curve in the
elastic region. Determine (a) the tangent
modulus at 25 MPa (3625 psi), and (b) the
secant modulus taken to 35 MPa (5000 psi).
6.11 As noted in Section 3.15, for single crystals of
some substances, the physical properties are
anisotropic; that is, they are dependent on crystallographic direction. One such property is the
modulus of elasticity. For cubic single crystals,
the modulus of elasticity in a general [uvw]
direction, Euvw, is described by the relationship
1
1
1
1
3a
b
Euvw
EH100I
EH100I
EH111I
1a2b2 b2g2 g2a2 2
where EH100I and EH111I are the moduli of elasticity in [100] and [111] directions, respectively;
a, b, and g are the cosines of the angles between [uvw] and the respective [100], [010],
and [001] directions. Verify that the EH110I values for aluminum, copper, and iron in Table 3.3
are correct.
6.12 In Section 2.6 it was noted that the net bonding energy EN between two isolated positive
and negative ions is a function of interionic distance r as follows:
B
A
EN n
(6.25)
r
r
300
2000
Stress (MPa)
103 psi
300
MPa
2000
200
200
1000
1000
100
0
0
0.000
0.020
100
0
0.000
0.040
Strain
0.005
0.010
Strain
0.060
0.015
0.080
Stress (103 psi)
6.7
Stress
6.6
of 66,700 N (15,000 lbf), and experiences an
elongation of 0.43 mm (1.7 102 in.). Assuming that the deformation is entirely elastic, calculate the modulus of elasticity of the
aluminum.
Consider a cylindrical nickel wire 2.0 mm
(0.08 in.) in diameter and 3 104 mm (1200 in.)
long. Calculate its elongation when a load of
300 N (67 lbf ) is applied. Assume that the deformation is totally elastic.
For a brass alloy, the stress at which plastic
deformation begins is 345 MPa (50,000 psi),
and the modulus of elasticity is 103 GPa
(15.0 106 psi).
(a) What is the maximum load that may be
applied to a specimen with a cross-sectional
area of 130 mm2 (0.2 in.2) without plastic
deformation?
(b) If the original specimen length is 76 mm
(3.0 in.), what is the maximum length to which
it may be stretched without causing plastic
deformation?
A cylindrical rod of steel (E 207 GPa, 30
106 psi) having a yield strength of 310 MPa
(45,000 psi) is to be subjected to a load of
11,100 N (2500 lbf ). If the length of the rod is
500 mm (20.0 in.), what must be the diameter
to allow an elongation of 0.38 mm (0.015 in.)?
Consider a cylindrical specimen of a steel alloy (Figure 6.21) 8.5 mm (0.33 in.) in diameter and 80 mm (3.15 in.) long that is pulled in
1496T_c06_131-173 11/16/05 17:06 Page 168
REVISED PAGES
168 • Chapter 6 / Mechanical Properties of Metals
Figure 6.22 Tensile
stress–strain behavior
for a gray cast iron.
60
8
6
40
30
4
Stress (103 psi)
Stress (MPa)
50
20
2
10
0
0
0.0002
0.0004
0.0006
0
0.0008
Strain
where A, B, and n are constants for the particular ion pair. Equation 6.25 is also valid for
the bonding energy between adjacent ions in
solid materials. The modulus of elasticity E is
proportional to the slope of the interionic
force–separation curve at the equilibrium
interionic separation; that is,
E r a
dF
b
dr r0
Derive an expression for the dependence of
the modulus of elasticity on these A, B, and n
parameters (for the two-ion system) using the
following procedure:
1. Establish a relationship for the force F as
a function of r, realizing that
F
dEN
dr
2. Now take the derivative dFdr.
3. Develop an expression for r0, the equilibrium separation. Since r0 corresponds to the
value of r at the minimum of the EN-versus-r
curve (Figure 2.8b), take the derivative
dENdr, set it equal to zero, and solve for r,
which corresponds to r0.
4. Finally, substitute this expression for r0 into
the relationship obtained by taking dF dr.
6.13 Using the solution to Problem 6.12, rank the
magnitudes of the moduli of elasticity for the
following hypothetical X, Y, and Z materials
from the greatest to the least. The appropriate A, B, and n parameters (Equation 6.25)
for these three materials are tabulated below;
they yield EN in units of electron volts and r
in nanometers:
Material
X
Y
Z
A
1.5
2.0
3.5
B
n
6
7.0 10
1.0 105
4.0 106
8
9
7
Elastic Properties of Materials
6.14 A cylindrical specimen of steel having a diameter of 15.2 mm (0.60 in.) and length of
250 mm (10.0 in.) is deformed elastically in
tension with a force of 48,900 N (11,000 lbf ).
Using the data contained in Table 6.1, determine the following:
(a) The amount by which this specimen will
elongate in the direction of the applied stress.
(b) The change in diameter of the specimen.
Will the diameter increase or decrease?
6.15 A cylindrical bar of aluminum 19 mm (0.75 in.)
in diameter is to be deformed elastically by
application of a force along the bar axis. Using the data in Table 6.1, determine the force
that will produce an elastic reduction of 2.5
103 mm (1.0 104 in.) in the diameter.
6.16 A cylindrical specimen of some metal alloy
10 mm (0.4 in.) in diameter is stressed elastically in tension. A force of 15,000 N (3370 lbf )
produces a reduction in specimen diameter
of 7 103 mm (2.8 104 in.). Compute
Poisson’s ratio for this material if its elastic
modulus is 100 GPa (14.5 106 psi).
1496T_c06_131-173 11/16/05 17:06 Page 169
REVISED PAGES
Questions and Problems • 169
6.17 A cylindrical specimen of a hypothetical metal
alloy is stressed in compression. If its original
and final diameters are 30.00 and 30.04 mm,
respectively, and its final length is 105.20 mm,
compute its original length if the deformation is totally elastic. The elastic and shear
moduli for this alloy are 65.5 and 25.4 GPa,
respectively.
6.18 Consider a cylindrical specimen of some hypothetical metal alloy that has a diameter of
10.0 mm (0.39 in.). A tensile force of 1500 N
(340 lbf ) produces an elastic reduction in diameter of 6.7 104 mm (2.64 105 in.).
Compute the elastic modulus of this alloy,
given that Poisson’s ratio is 0.35.
6.19 A brass alloy is known to have a yield strength
of 240 MPa (35,000 psi), a tensile strength of
310 MPa (45,000 psi), and an elastic modulus
of 110 GPa (16.0 106 psi). A cylindrical
specimen of this alloy 15.2 mm (0.60 in.) in diameter and 380 mm (15.0 in.) long is stressed
in tension and found to elongate 1.9 mm
(0.075 in.). On the basis of the information
given, is it possible to compute the magnitude
of the load that is necessary to produce this
change in length? If so, calculate the load. If
not, explain why.
6.20 A cylindrical metal specimen 15.0 mm (0.59 in.)
in diameter and 150 mm (5.9 in.) long is to
be subjected to a tensile stress of 50 MPa
(7250 psi); at this stress level the resulting
deformation will be totally elastic.
(a) If the elongation must be less than
0.072 mm (2.83 103 in.), which of the metals in Table 6.1 are suitable candidates? Why?
(b) If, in addition, the maximum permissible
diameter decrease is 2.3 103 mm (9.1
105 in.) when the tensile stress of 50 MPa is
applied, which of the metals that satisfy the criterion in part (a) are suitable candidates? Why?
6.21 Consider the brass alloy for which the
stress–strain behavior is shown in Figure 6.12.
A cylindrical specimen of this material
10.0 mm (0.39 in.) in diameter and 101.6 mm
(4.0 in.) long is pulled in tension with a force
of 10,000 N (2250 lbf ). If it is known that this
alloy has a value for Poisson’s ratio of 0.35,
compute (a) the specimen elongation, and
(b) the reduction in specimen diameter.
6.22 A cylindrical rod 120 mm long and having a
diameter of 15.0 mm is to be deformed using
a tensile load of 35,000 N. It must not experience either plastic deformation or a diameter
reduction of more than 1.2 102 mm. Of
the materials listed below, which are possible
candidates? Justify your choice(s).
Material
Aluminum alloy
Titanium alloy
Steel alloy
Magnesium alloy
Modulus
of Elasticity
(GPa)
Yield
Strength
(MPa)
Poisson’s
Ratio
70
105
205
45
250
850
550
170
0.33
0.36
0.27
0.35
6.23 A cylindrical rod 500 mm (20.0 in.) long, having a diameter of 12.7 mm (0.50 in.), is to be
subjected to a tensile load. If the rod is to experience neither plastic deformation nor an
elongation of more than 1.3 mm (0.05 in.)
when the applied load is 29,000 N (6500 lbf ),
which of the four metals or alloys listed below are possible candidates? Justify your
choice(s).
Material
Aluminum alloy
Brass alloy
Copper
Steel alloy
Modulus
of Elasticity
(GPa)
Yield
Strength
(MPa)
Tensile
Strength
(MPa)
70
100
110
207
255
345
210
450
420
420
275
550
Tensile Properties
6.24 Figure 6.21 shows the tensile engineering
stress–strain behavior for a steel alloy.
(a) What is the modulus of elasticity?
(b) What is the proportional limit?
(c) What is the yield strength at a strain offset of 0.002?
(d) What is the tensile strength?
6.25 A cylindrical specimen of a brass alloy having a length of 100 mm (4 in.) must elongate
only 5 mm (0.2 in.) when a tensile load of
100,000 N (22,500 lbf) is applied. Under
these circumstances what must be the radius
of the specimen? Consider this brass alloy
to have the stress–strain behavior shown in
Figure 6.12.
1496T_c06_131-173 11/16/05 17:06 Page 170
REVISED PAGES
170 • Chapter 6 / Mechanical Properties of Metals
6.26 A load of 140,000 N (31,500 lbf) is applied to
a cylindrical specimen of a steel alloy (displaying the stress–strain behavior shown in
Figure 6.21) that has a cross-sectional diameter of 10 mm (0.40 in.).
(a) Will the specimen experience elastic and/
or plastic deformation? Why?
(b) If the original specimen length is 500 mm
(20 in.), how much will it increase in length
when this load is applied?
6.27 A bar of a steel alloy that exhibits the
stress–strain behavior shown in Figure 6.21 is
subjected to a tensile load; the specimen is
375 mm (14.8 in.) long and of square cross
section 5.5 mm (0.22 in.) on a side.
(a) Compute the magnitude of the load necessary to produce an elongation of 2.25 mm
(0.088 in.).
(b) What will be the deformation after the
load has been released?
6.28 A cylindrical specimen of stainless steel having a diameter of 12.8 mm (0.505 in.) and a
gauge length of 50.800 mm (2.000 in.) is pulled
in tension. Use the load–elongation characteristics tabulated below to complete parts (a)
through (f).
Load
N
0
12,700
25,400
38,100
50,800
76,200
89,100
92,700
102,500
107,800
119,400
128,300
149,700
159,000
160,400
159,500
151,500
124,700
Length
lbf
0
2,850
5,710
8,560
11,400
17,100
20,000
20,800
23,000
24,200
26,800
28,800
33,650
35,750
36,000
35,850
34,050
28,000
Fracture
mm
in.
50.800
50.825
50.851
50.876
50.902
50.952
51.003
51.054
51.181
51.308
51.562
51.816
52.832
53.848
54.356
54.864
55.880
56.642
2.000
2.001
2.002
2.003
2.004
2.006
2.008
2.010
2.015
2.020
2.030
2.040
2.080
2.120
2.140
2.160
2.200
2.230
(a) Plot the data as engineering stress versus
engineering strain.
(b) Compute the modulus of elasticity.
(c) Determine the yield strength at a strain
offset of 0.002.
(d) Determine the tensile strength of this alloy.
(e) What is the approximate ductility, in percent elongation?
(f) Compute the modulus of resilience.
6.29 A specimen of magnesium having a rectangular cross section of dimensions 3.2 mm
19.1 mm (18 in. 34 in.) is deformed in tension. Using the load–elongation data tabulated
as follows, complete parts (a) through (f).
Load
Length
lbf
N
0
310
625
1265
1670
1830
2220
2890
3170
3225
3110
2810
0
1380
2780
5630
7430
8140
9870
12,850
14,100
14,340
13,830
12,500
in.
2.500
2.501
2.502
2.505
2.508
2.510
2.525
2.575
2.625
2.675
2.725
2.775
Fracture
mm
63.50
63.53
63.56
63.62
63.70
63.75
64.14
65.41
66.68
67.95
69.22
70.49
(a) Plot the data as engineering stress versus
engineering strain.
(b) Compute the modulus of elasticity.
(c) Determine the yield strength at a strain
offset of 0.002.
(d) Determine the tensile strength of this alloy.
(e) Compute the modulus of resilience.
(f) What is the ductility, in percent elongation?
6.30 A cylindrical metal specimen having an
original diameter of 12.8 mm (0.505 in.) and
gauge length of 50.80 mm (2.000 in.) is pulled
in tension until fracture occurs. The diameter
at the point of fracture is 8.13 mm (0.320 in.),
and the fractured gauge length is 74.17 mm
(2.920 in.). Calculate the ductility in terms
of percent reduction in area and percent
elongation.
6.31 Calculate the moduli of resilience for the materials having the stress–strain behaviors
shown in Figures 6.12 and 6.21.
1496T_c06_131-173 11/16/05 17:06 Page 171
REVISED PAGES
Questions and Problems • 171
6.32 Determine the modulus of resilience for each
of the following alloys:
Yield Strength
Material
MPa
psi
830
380
275
690
120,000
55,000
40,000
100,000
Steel alloy
Brass alloy
Aluminum alloy
Titanium alloy
Use modulus of elasticity values in Table 6.1.
6.33 A steel alloy to be used for a spring application must have a modulus of resilience of at
least 2.07 MPa (300 psi). What must be its
minimum yield strength?
True Stress and Strain
6.34 Show that Equations 6.18a and 6.18b are valid
when there is no volume change during
deformation.
6.35 Demonstrate that Equation 6.16, the expression defining true strain, may also be represented by
A0
T ln a b
Ai
when specimen volume remains constant during deformation. Which of these two expressions is more valid during necking? Why?
6.36 Using the data in Problem 6.28 and Equations 6.15, 6.16, and 6.18a, generate a true
stress–true strain plot for stainless steel.
Equation 6.18a becomes invalid past the point
at which necking begins; therefore, measured
diameters are given below for the last three
data points, which should be used in true
stress computations.
Load
Length
Diameter
N
lbf
mm
in.
mm
in.
159,500
151,500
124,700
35,850
34,050
28,000
54.864
55.880
56.642
2.160
2.200
2.230
12.22
11.80
10.65
0.481
0.464
0.419
6.37 A tensile test is performed on a metal specimen, and it is found that a true plastic strain
of 0.16 is produced when a true stress of
500 MPa (72,500 psi) is applied; for the same
metal, the value of K in Equation 6.19 is
825 MPa (120,000 psi). Calculate the true
strain that results from the application of a
true stress of 600 MPa (87,000 psi).
6.38 For some metal alloy, a true stress of
345 MPa (50,000 psi) produces a plastic true
strain of 0.02. How much will a specimen of
this material elongate when a true stress
of 415 MPa (60,000 psi) is applied if the
original length is 500 mm (20 in.)? Assume
a value of 0.22 for the strain-hardening
exponent, n.
6.39 The following true stresses produce the corresponding true plastic strains for a brass alloy:
True Stress
( psi)
True Strain
60,000
70,000
0.15
0.25
What true stress is necessary to produce a true
plastic strain of 0.21?
6.40 For a brass alloy, the following engineering
stresses produce the corresponding plastic
engineering strains, prior to necking:
Engineering Stress
(MPa)
Engineering Strain
315
340
0.105
0.220
On the basis of this information, compute the
engineering stress necessary to produce an
engineering strain of 0.28.
6.41 Find the toughness (or energy to cause fracture) for a metal that experiences both elastic
and plastic deformation. Assume Equation 6.5
for elastic deformation, that the modulus of
elasticity is 103 GPa (15 106 psi), and that
elastic deformation terminates at a strain of
0.007. For plastic deformation, assume that the
relationship between stress and strain is described by Equation 6.19, in which the values
for K and n are 1520 MPa (221,000 psi) and
0.15, respectively. Furthermore, plastic deformation occurs between strain values of 0.007
and 0.60, at which point fracture occurs.
6.42 For a tensile test, it can be demonstrated that
necking begins when
dsT
sT
dT
(6.26)
1496T_c06_131-173 11/16/05 17:06 Page 172
REVISED PAGES
172 • Chapter 6 / Mechanical Properties of Metals
Using Equation 6.19, determine the value of
the true strain at this onset of necking.
6.43 Taking the logarithm of both sides of Equation 6.19 yields
log sT log K n log T
(6.27)
Thus, a plot of log sT versus log T in the plastic region to the point of necking should yield
a straight line having a slope of n and an
intercept (at log sT 0) of log K.
Using the appropriate data tabulated in
Problem 6.28, make a plot of log sT versus log
T and determine the values of n and K. It will
be necessary to convert engineering stresses
and strains to true stresses and strains using
Equations 6.18a and 6.18b.
Elastic Recovery After Plastic Deformation
6.44 A cylindrical specimen of a brass alloy 10.0 mm
(0.39 in.) in diameter and 120.0 mm (4.72 in.)
long is pulled in tension with a force of 11,750 N
(2640 lbf); the force is subsequently released.
(a) Compute the final length of the specimen
at this time. The tensile stress–strain behavior
for this alloy is shown in Figure 6.12.
(b) Compute the final specimen length when
the load is increased to 23,500 N (5280 lbf)
and then released.
6.45 A steel alloy specimen having a rectangular
cross section of dimensions 19 mm 3.2 mm
(34 in. 18 in.) has the stress–strain behavior
shown in Figure 6.21. If this specimen is
subjected to a tensile force of 110,000 N
(25,000 lbf) then
(a) Determine the elastic and plastic strain
values.
(b) If its original length is 610 mm (24.0 in.),
what will be its final length after the load in
part (a) is applied and then released?
Hardness
6.46 (a) A 10-mm-diameter Brinell hardness indenter produced an indentation 2.50 mm in diameter in a steel alloy when a load of 1000 kg
was used. Compute the HB of this material.
(b) What will be the diameter of an indentation to yield a hardness of 300 HB when a
500-kg load is used?
6.47 Estimate the Brinell and Rockwell hardnesses for the following:
(a) The naval brass for which the stress–strain
behavior is shown in Figure 6.12.
(b) The steel alloy for which the stress–strain
behavior is shown in Figure 6.21.
6.48 Using the data represented in Figure 6.19,
specify equations relating tensile strength and
Brinell hardness for brass and nodular cast iron,
similar to Equations 6.20a and 6.20b for steels.
Variability of Material Properties
6.49 Cite five factors that lead to scatter in measured material properties.
6.50 Below are tabulated a number of Rockwell G
hardness values that were measured on a
single steel specimen. Compute average and
standard deviation hardness values.
47.3
52.1
45.6
49.9
47.6
50.4
48.7
50.0
46.2
48.3
51.1
46.7
47.1
50.4
45.9
46.4
48.5
49.7
Design/Safety Factors
6.51 Upon what three criteria are factors of safety
based?
6.52 Determine working stresses for the two alloys
that have the stress–strain behaviors shown in
Figures 6.12 and 6.21.
DESIGN PROBLEMS
6.D1 A large tower is to be supported by a series
of steel wires; it is estimated that the load on
each wire will be 13,300 N (3000 lbf). Determine the minimum required wire diameter, assuming a factor of safety of 2 and a yield
strength of 860 MPa (125,000 psi) for the steel.
6.D2 (a) Gaseous hydrogen at a constant pressure
of 0.658 MPa (5 atm) is to flow within the
inside of a thin-walled cylindrical tube of
nickel that has a radius of 0.125 m. The temperature of the tube is to be 350C and the
pressure of hydrogen outside of the tube will
1496T_c06_131-173 11/16/05 17:06 Page 173
REVISED PAGES
Design Problems • 173
be maintained at 0.0127 MPa (0.125 atm).
Calculate the minimum wall thickness if the
diffusion flux is to be no greater than 1.25
107 mol/m2-s. The concentration of hydrogen
in the nickel, CH (in moles hydrogen per m3
of Ni) is a function of hydrogen pressure,
PH2 (in MPa) and absolute temperature (T)
according to
12.3 kJ/mol
CH 30.81pH2 exp a
b (6.28)
RT
Furthermore, the diffusion coefficient for
the diffusion of H in Ni depends on temperature as
39.56 kJ/mol
DH 1m2/s2 4.76 107 exp a
b
RT
(6.29)
(b) For thin-walled cylindrical tubes that are
pressurized, the circumferential stress is a
function of the pressure difference across the
wall ( ¢p), cylinder radius (r), and tube thickness ( ¢x) as
r¢p
s
(6.30)
4¢x
Compute the circumferential stress to which
the walls of this pressurized cylinder are
exposed.
(c) The room-temperature yield strength of
Ni is 100 MPa (15,000 psi) and, furthermore,
sy diminishes about 5 MPa for every 50C rise
in temperature. Would you expect the wall
thickness computed in part (b) to be suitable
for this Ni cylinder at 350C? Why or why not?
(d) If this thickness is found to be suitable,
compute the minimum thickness that could
be used without any deformation of the tube
walls. How much would the diffusion flux increase with this reduction in thickness? On
the other hand, if the thickness determined
in part (c) is found to be unsuitable, then
specify a minimum thickness that you would
use. In this case, how much of a diminishment
in diffusion flux would result?
6.D3 Consider the steady-state diffusion of hydrogen through the walls of a cylindrical
nickel tube as described in Problem 6.D2.
One design calls for a diffusion flux of
2.5 108 mol/m2-s, a tube radius of 0.100 m,
and inside and outside pressures of 1.015 MPa
(10 atm) and 0.01015 MPa (0.1 atm), respectively; the maximum allowable temperature
is 300C. Specify a suitable temperature and
wall thickness to give this diffusion flux and
yet ensure that the tube walls will not experience any permanent deformation.
1496T_c07_174-206 12/21/05 8:27 Page 174
Chapter
7
2nd REVISE PAGES
Dislocations and
Strengthening Mechanisms
I
n this photomicrograph of a lithium fluoride (LiF) single crystal, the small pyramidal pits repre-
sent those positions at which dislocations intersect the surface. The surface was polished and then
chemically treated; these “etch pits” result from localized chemical attack around the dislocations
and indicate the distribution of the dislocations. 750. (Photomicrograph courtesy of W. G.
Johnston, General Electric Co.)
WHY STUDY Dislocations and Strengthening Mechanisms?
With a knowledge of the nature of dislocations and
the role they play in the plastic deformation process,
we are able to understand the underlying mechanisms
of the techniques that are used to strengthen and
174 •
harden metals and their alloys. Thus, it becomes possible to design and tailor the mechanical properties of
materials—for example, the strength or toughness of a
metal–matrix composite.
1496T_c07_174-206 12/21/05 7:48 Page 175
2nd REVISE PAGES
Learning Objectives
After studying this chapter you should be able to do the following:
1. Describe edge and screw dislocation motion
6. Describe and explain solid-solution strengthenfrom an atomic perspective.
ing for substitutional impurity atoms in terms of
2. Describe how plastic deformation occurs by the
lattice strain interactions with dislocations.
motion of edge and screw dislocations in re7. Describe and explain the phenomenon of strain
sponse to applied shear stresses.
hardening (or cold working) in terms of disloca3. Define slip system and cite one example.
tions and strain field interactions.
4. Describe how the grain structure of a polycrys8. Describe recrystallization in terms of both the
talline metal is altered when it is plastically
alteration of microstructure and mechanical
deformed.
characteristics of the material.
5. Explain how grain boundaries impede dislocation
9. Describe the phenomenon of grain growth from
motion and why a metal having small grains is
both macroscopic and atomic perspectives.
stronger than one having large grains.
7.1 INTRODUCTION
Chapter 6 explained that materials may experience two kinds of deformation: elastic and plastic. Plastic deformation is permanent, and strength and hardness are
measures of a material’s resistance to this deformation. On a microscopic scale, plastic deformation corresponds to the net movement of large numbers of atoms in
response to an applied stress. During this process, interatomic bonds must be ruptured and then reformed. In crystalline solids, plastic deformation most often
involves the motion of dislocations, linear crystalline defects that were introduced
in Section 4.5. This chapter discusses the characteristics of dislocations and their involvement in plastic deformation. Twinning, another process by which some metals
plastically deform, is also treated. In addition, and probably most importantly, several techniques are presented for strengthening single-phase metals, the mechanisms
of which are described in terms of dislocations. Finally, the latter sections of this
chapter are concerned with recovery and recrystallization—processes that occur in
plastically deformed metals, normally at elevated temperatures—and, in addition,
grain growth.
Dislocations and Plastic Deformation
Early materials studies led to the computation of the theoretical strengths of perfect crystals, which were many times greater than those actually measured. During the 1930s it was theorized that this discrepancy in mechanical strengths could
be explained by a type of linear crystalline defect that has since come to be known
as a dislocation. It was not until the 1950s, however, that the existence of such
dislocation defects was established by direct observation with the electron microscope. Since then, a theory of dislocations has evolved that explains many of
the physical and mechanical phenomena in metals [as well as crystalline ceramics
(Section 12.10)].
7.2 BASIC CONCEPTS
Edge and screw are the two fundamental dislocation types. In an edge dislocation,
localized lattice distortion exists along the end of an extra half-plane of atoms, which
also defines the dislocation line (Figure 4.3). A screw dislocation may be thought of
1496T_c07_174-206 12/21/05 7:48 Page 176
2nd REVISE PAGES
176 • Chapter 7 / Dislocations and Strengthening Mechanisms
Edge
slip
as resulting from shear distortion; its dislocation line passes through the center of a
spiral, atomic plane ramp (Figure 4.4). Many dislocations in crystalline materials have
both edge and screw components; these are mixed dislocations (Figure 4.5).
Plastic deformation corresponds to the motion of large numbers of dislocations.
An edge dislocation moves in response to a shear stress applied in a direction perpendicular to its line; the mechanics of dislocation motion are represented in Figure 7.1. Let the initial extra half-plane of atoms be plane A. When the shear stress
is applied as indicated (Figure 7.1a), plane A is forced to the right; this in turn pushes
the top halves of planes B, C, D, and so on, in the same direction. If the applied
shear stress is of sufficient magnitude, the interatomic bonds of plane B are severed along the shear plane, and the upper half of plane B becomes the extra halfplane as plane A links up with the bottom half of plane B (Figure 7.1b). This process
is subsequently repeated for the other planes, such that the extra half-plane, by discrete steps, moves from left to right by successive and repeated breaking of bonds
and shifting by interatomic distances of upper half-planes. Before and after the
movement of a dislocation through some particular region of the crystal, the atomic
arrangement is ordered and perfect; it is only during the passage of the extra halfplane that the lattice structure is disrupted. Ultimately this extra half-plane may
emerge from the right surface of the crystal, forming an edge that is one atomic
distance wide; this is shown in Figure 7.1c.
The process by which plastic deformation is produced by dislocation motion is
termed slip; the crystallographic plane along which the dislocation line traverses
is the slip plane, as indicated in Figure 7.1. Macroscopic plastic deformation simply corresponds to permanent deformation that results from the movement of
dislocations, or slip, in response to an applied shear stress, as represented in
Figure 7.2a.
Dislocation motion is analogous to the mode of locomotion employed by a
caterpillar (Figure 7.3). The caterpillar forms a hump near its posterior end by
pulling in its last pair of legs a unit leg distance. The hump is propelled forward by
repeated lifting and shifting of leg pairs. When the hump reaches the anterior end,
Shear
stress
Shear
stress
A
B
C
D
Shear
stress
A
B
C
D
A
B
Slip plane
C
D
Unit step
of slip
Edge
dislocation
line
(a)
( b)
(c)
Figure 7.1 Atomic rearrangements that accompany the motion of an edge dislocation as it
moves in response to an applied shear stress. (a) The extra half-plane of atoms is labeled A.
(b) The dislocation moves one atomic distance to the right as A links up to the lower portion
of plane B; in the process, the upper portion of B becomes the extra half-plane. (c) A step
forms on the surface of the crystal as the extra half-plane exits. (Adapted from A. G. Guy,
Essentials of Materials Science, McGraw-Hill Book Company, New York, 1976, p. 153.)
1496T_c07_174-206 12/21/05 7:48 Page 177
2nd REVISE PAGES
7.2 Basic Concepts • 177
Figure 7.2 The formation of a step on
the surface of a crystal by the motion of
(a) an edge dislocation and (b) a screw
dislocation. Note that for an edge, the
dislocation line moves in the direction of
the applied shear stress t; for a screw, the
dislocation line motion is perpendicular
to the stress direction. (Adapted from
H. W. Hayden, W. G. Moffatt, and J. Wulff,
The Structure and Properties of
Materials,Vol. III, Mechanical Behavior,
p. 70. Copyright © 1965 by John Wiley &
Sons, New York. Reprinted by
permission of John Wiley & Sons, Inc.)
Direction
of motion
(a)
Direction
of motion
(b)
Screw
Mixed
dislocation density
the entire caterpillar has moved forward by the leg separation distance. The caterpillar hump and its motion correspond to the extra half-plane of atoms in the dislocation model of plastic deformation.
The motion of a screw dislocation in response to the applied shear stress is
shown in Figure 7.2b; the direction of movement is perpendicular to the stress direction. For an edge, motion is parallel to the shear stress. However, the net plastic
deformation for the motion of both dislocation types is the same (see Figure 7.2).
The direction of motion of the mixed dislocation line is neither perpendicular nor
parallel to the applied stress, but lies somewhere in between.
All metals and alloys contain some dislocations that were introduced during
solidification, during plastic deformation, and as a consequence of thermal stresses
that result from rapid cooling. The number of dislocations, or dislocation density in
a material, is expressed as the total dislocation length per unit volume or, equivalently, the number of dislocations that intersect a unit area of a random section. The
units of dislocation density are millimeters of dislocation per cubic millimeter or
just per square millimeter. Dislocation densities as low as 103 mm2 are typically
found in carefully solidified metal crystals. For heavily deformed metals, the density may run as high as 109 to 1010 mm2. Heat treating a deformed metal specimen
Figure 7.3 Representation of the analogy between caterpillar and dislocation motion.
1496T_c07_174-206 11/17/05 13:32 Page 178
REVISED PAGES
178 • Chapter 7 / Dislocations and Strengthening Mechanisms
can diminish the density to on the order of 105 to 106 mm2. By way of contrast, a
typical dislocation density for ceramic materials is between 102 and 104 mm2; also,
for silicon single crystals used in integrated circuits the value normally lies between
0.1 and 1 mm2.
7.3 CHARACTERISTICS OF DISLOCATIONS
lattice strain
Several characteristics of dislocations are important with regard to the mechanical
properties of metals. These include strain fields that exist around dislocations, which
are influential in determining the mobility of the dislocations, as well as their ability to multiply.
When metals are plastically deformed, some fraction of the deformation energy
(approximately 5%) is retained internally; the remainder is dissipated as heat. The
major portion of this stored energy is as strain energy associated with dislocations.
Consider the edge dislocation represented in Figure 7.4.As already mentioned, some
atomic lattice distortion exists around the dislocation line because of the presence
of the extra half-plane of atoms. As a consequence, there are regions in which compressive, tensile, and shear lattice strains are imposed on the neighboring atoms. For
example, atoms immediately above and adjacent to the dislocation line are squeezed
together. As a result, these atoms may be thought of as experiencing a compressive
strain relative to atoms positioned in the perfect crystal and far removed from the
dislocation; this is illustrated in Figure 7.4. Directly below the half-plane, the effect
is just the opposite; lattice atoms sustain an imposed tensile strain, which is as shown.
Shear strains also exist in the vicinity of the edge dislocation. For a screw dislocation, lattice strains are pure shear only. These lattice distortions may be considered
to be strain fields that radiate from the dislocation line. The strains extend into
the surrounding atoms, and their magnitude decreases with radial distance from the
dislocation.
The strain fields surrounding dislocations in close proximity to one another
may interact such that forces are imposed on each dislocation by the combined interactions of all its neighboring dislocations. For example, consider two edge dislocations that have the same sign and the identical slip plane, as represented in
Figure 7.5a. The compressive and tensile strain fields for both lie on the same side
of the slip plane; the strain field interaction is such that there exists between these
two isolated dislocations a mutual repulsive force that tends to move them apart.
On the other hand, two dislocations of opposite sign and having the same slip
plane will be attracted to one another, as indicated in Figure 7.5b, and dislocation
Compression
Tension
Figure 7.4 Regions of compression
(green) and tension (yellow) located
around an edge dislocation. (Adapted
from W. G. Moffatt, G. W. Pearsall, and
J. Wulff, The Structure and Properties of
Materials, Vol. I, Structure, p. 85. Copyright
© 1964 by John Wiley & Sons, New York.
Reprinted by permission of John Wiley &
Sons, Inc.)
1496T_c07_174-206 11/17/05 13:32 Page 179
REVISED PAGES
7.4 Slip Systems • 179
Figure 7.5 (a) Two edge dislocations of
the same sign and lying on the same slip
plane exert a repulsive force on each
other; C and T denote compression and
tensile regions, respectively. (b) Edge
dislocations of opposite sign and lying
on the same slip plane exert an attractive
force on each other. Upon meeting, they
annihilate each other and leave a region
of perfect crystal. (Adapted from H. W.
Hayden, W. G. Moffatt, and J. Wulff, The
Structure and Properties of Materials,
Vol. III, Mechanical Behavior, p. 75.
Copyright © 1965 by John Wiley & Sons,
New York. Reprinted by permission of
John Wiley & Sons.)
C
C
Repulsion
T
C
T
(a)
T
Dislocation
annihilation
Attraction
;
+
=
(Perfect crystal)
T
C
(b)
annihilation will occur when they meet. That is, the two extra half-planes of atoms
will align and become a complete plane. Dislocation interactions are possible between edge, screw, and/or mixed dislocations, and for a variety of orientations.These
strain fields and associated forces are important in the strengthening mechanisms
for metals.
During plastic deformation, the number of dislocations increases dramatically.
We know that the dislocation density in a metal that has been highly deformed may
be as high as 1010 mm2. One important source of these new dislocations is existing dislocations, which multiply; furthermore, grain boundaries, as well as internal
defects and surface irregularities such as scratches and nicks, which act as stress concentrations, may serve as dislocation formation sites during deformation.
7.4 SLIP SYSTEMS
slip system
Dislocations do not move with the same degree of ease on all crystallographic planes
of atoms and in all crystallographic directions. Ordinarily there is a preferred plane,
and in that plane there are specific directions along which dislocation motion occurs. This plane is called the slip plane; it follows that the direction of movement is
called the slip direction. This combination of the slip plane and the slip direction is
termed the slip system. The slip system depends on the crystal structure of the metal
and is such that the atomic distortion that accompanies the motion of a dislocation
is a minimum. For a particular crystal structure, the slip plane is the plane that has
the most dense atomic packing—that is, has the greatest planar density. The slip direction corresponds to the direction, in this plane, that is most closely packed with
atoms—that is, has the highest linear density. Planar and linear atomic densities
were discussed in Section 3.11.
Consider, for example, the FCC crystal structure, a unit cell of which is shown
in Figure 7.6a.There is a set of planes, the {111} family, all of which are closely packed.
A (111)-type plane is indicated in the unit cell; in Figure 7.6b, this plane is positioned
within the plane of the page, in which atoms are now represented as touching nearest neighbors.
1496T_c07_174-206 12/21/05 7:48 Page 180
2nd REVISE PAGES
180 • Chapter 7 / Dislocations and Strengthening Mechanisms
A
Figure 7.6 (a) A {111}
H110I slip system shown
within an FCC unit cell.
(b) The (111) plane
from (a) and three H110I
slip directions (as
indicated by arrows)
within that plane
comprise possible slip
systems.
A
C
B
B
C
F
E
D
E
F
D
(a)
(b)
Slip occurs along H110I-type directions within the {111} planes, as indicated by
arrows in Figure 7.6. Hence, {111}H110I represents the slip plane and direction combination, or the slip system for FCC. Figure 7.6b demonstrates that a given slip plane
may contain more than a single slip direction. Thus, several slip systems may exist
for a particular crystal structure; the number of independent slip systems represents
the different possible combinations of slip planes and directions. For example, for
face-centered cubic, there are 12 slip systems: four unique {111} planes and, within
each plane, three independent H110I directions.
The possible slip systems for BCC and HCP crystal structures are listed in Table
7.1. For each of these structures, slip is possible on more than one family of planes
(e.g., {110}, {211}, and {321} for BCC). For metals having these two crystal structures, some slip systems are often operable only at elevated temperatures.
Metals with FCC or BCC crystal structures have a relatively large number of
slip systems (at least 12). These metals are quite ductile because extensive plastic
deformation is normally possible along the various systems. Conversely, HCP metals, having few active slip systems, are normally quite brittle.
The Burgers vector concept was introduced in Section 4.5, and denoted by a b
for edge, screw, and mixed dislocations in Figures 4.3, 4.4, and 4.5, respectively. With
regard to the process of slip, a Burgers vector’s direction corresponds to a dislocation’s slip direction, whereas its magnitude is equal to the unit slip distance (or
interatomic separation in this direction). Of course, both the direction and the magnitude of b will depend on crystal structure, and it is convenient to specify a Burgers vector in terms of unit cell edge length (a) and crystallographic direction indices.
Table 7.1 Slip Systems for Face-Centered Cubic, Body-Centered Cubic, and
Hexagonal Close-Packed Metals
Metals
Cu, Al, Ni, Ag, Au
a-Fe, W, Mo
a-Fe, W
a-Fe, K
Cd, Zn, Mg, Ti, Be
Ti, Mg, Zr
Ti, Mg
Slip Plane
Face-Centered Cubic
51116
Body-Centered Cubic
51106
52116
53216
Hexagonal Close-Packed
500016
510106
510116
Slip Direction
Number of
Slip Systems
H110I
12
H111I
H111I
H111I
12
12
24
H1120I
H1120I
H1120I
3
3
6
1496T_c07_174-206 11/17/05 13:32 Page 181
REVISED PAGES
7.5 Slip in Single Crystals • 181
Burgers vectors for face-centered cubic, body-centered cubic, and hexagonal closepacked crystal structures are given as follows:
a
H110I
2
a
b1BCC2 H111I
2
a
b1HCP2 H1120I
3
b1FCC2
(7.1a)
(7.1b)
(7.1c)
Concept Check 7.1
Which of the following is the slip system for the simple cubic crystal structure? Why?
51006H110I
51106H110I
51006H010I
51106H111I
(Note: a unit cell for the simple cubic crystal structure is shown in Figure 3.23.)
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
7.5 SLIP IN SINGLE CRYSTALS
resolved shear stress
Resolved shear
stress—dependence
on applied stress and
orientation of stress
direction relative to
slip plane normal
and slip direction
A further explanation of slip is simplified by treating the process in single crystals,
then making the appropriate extension to polycrystalline materials. As mentioned
previously, edge, screw, and mixed dislocations move in response to shear stresses applied along a slip plane and in a slip direction. As noted in Section 6.2, even though
an applied stress may be pure tensile (or compressive), shear components exist at all
but parallel or perpendicular alignments to the stress direction (Equation 6.4b).These
are termed resolved shear stresses, and their magnitudes depend not only on the applied stress, but also on the orientation of both the slip plane and direction within
that plane. Let f represent the angle between the normal to the slip plane and the
applied stress direction, and l the angle between the slip and stress directions, as indicated in Figure 7.7; it can then be shown that for the resolved shear stress tR
tR s cos f cos l
where s is the applied stress. In general, f l 90, since it need not be the case that
the tensile axis, the slip plane normal, and the slip direction all lie in the same plane.
A metal single crystal has a number of different slip systems that are capable
of operating. The resolved shear stress normally differs for each one because the
orientation of each relative to the stress axis (f and l angles) also differs. However, one slip system is generally oriented most favorably—that is, has the largest
resolved shear stress, tR(max):
tR 1max2 s1cos f cos l2 max
critical resolved
shear stress
(7.2)
(7.3)
In response to an applied tensile or compressive stress, slip in a single crystal commences on the most favorably oriented slip system when the resolved shear stress
reaches some critical value, termed the critical resolved shear stress tcrss; it represents the minimum shear stress required to initiate slip, and is a property of the
1496T_c07_174-206 11/17/05 13:32 Page 182
REVISED PAGES
182 • Chapter 7 / Dislocations and Strengthening Mechanisms
Figure 7.7 Geometrical relationships between the
tensile axis, slip plane, and slip direction used in
calculating the resolved shear stress for a single crystal.
F
A
Normal to
slip plane
Slip
direction
F
material that determines when yielding occurs. The single crystal plastically deforms
or yields when tR(max) tcrss, and the magnitude of the applied stress required to
initiate yielding (i.e., the yield strength sy) is
Yield strength of
a single crystal—
dependence on the
critical resolved
shear stress and the
orientation of most
favorably oriented
slip system
sy
tcrss
1cos f cos l2 max
(7.4)
The minimum stress necessary to introduce yielding occurs when a single crystal is
oriented such that f l 45; under these conditions,
sy 2tcrss
(7.5)
For a single-crystal specimen that is stressed in tension, deformation will be as
in Figure 7.8, where slip occurs along a number of equivalent and most favorably
oriented planes and directions at various positions along the specimen length. This
slip deformation forms as small steps on the surface of the single crystal that are
parallel to one another and loop around the circumference of the specimen as indicated in Figure 7.8. Each step results from the movement of a large number of
Figure 7.8 Macroscopic slip in a single crystal.
Direction
of force
Slip plane
1496T_c07_174-206 11/17/05 13:32 Page 183
REVISED PAGES
7.5 Slip in Single Crystals • 183
Figure 7.9 Slip in a zinc single crystal. (From
C. F. Elam, The Distortion of Metal Crystals,
Oxford University Press, London, 1935.)
dislocations along the same slip plane. On the surface of a polished single crystal
specimen, these steps appear as lines, which are called slip lines. A zinc single crystal that has been plastically deformed to the degree that these slip markings are
discernible is shown in Figure 7.9.
With continued extension of a single crystal, both the number of slip lines and
the slip step width will increase. For FCC and BCC metals, slip may eventually begin along a second slip system, the system that is next most favorably oriented with
the tensile axis. Furthermore, for HCP crystals having few slip systems, if the stress
axis for the most favorable slip system is either perpendicular to the slip direction
(l 90) or parallel to the slip plane (f 90), the critical resolved shear stress
will be zero. For these extreme orientations the crystal ordinarily fractures rather
than deforming plastically.
Concept Check 7.2
Explain the difference between resolved shear stress and critical resolved shear stress.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
EXAMPLE PROBLEM 7.1
Resolved Shear Stress and Stress-to-Initiate-Yielding
Computations
Consider a single crystal of BCC iron oriented such that a tensile stress is
applied along a [010] direction.
(a) Compute the resolved shear stress along a (110) plane and in a [111]
direction when a tensile stress of 52 MPa (7500 psi) is applied.
1496T_c07_174-206 11/17/05 13:32 Page 184
REVISED PAGES
184 • Chapter 7 / Dislocations and Strengthening Mechanisms
(b) If slip occurs on a (110) plane and in a [111] direction, and the critical
resolved shear stress is 30 MPa (4350 psi), calculate the magnitude of the
applied tensile stress necessary to initiate yielding.
Solution
(a) A BCC unit cell along with the slip direction and plane as well as the
direction of the applied stress are shown in the accompanying diagram. In
order to solve this problem we must use Equation 7.2. However, it is first necessary to determine values for f and l, where, from the diagram below, f
is the angle between the normal to the (110) slip plane (i.e., the [110] direction) and the [010] direction, and l represents the angle between [111] and
[010] directions. In general, for cubic unit cells, an angle u between directions
1 and 2, represented by [u1v1w1] and [u2v2w2], respectively, is equal to
u cos1 c
u1u2 v1v2 w1w2
21u21
v21 w21 21u22 v22 w22 2
d
(7.6)
For the determination of the value of f, let [u1v1w1] [110] and [u2v2w2]
[010] such that
112102 112112 102102
f cos1 e
f
23 112 2 112 2 102 2 4 3 102 2 112 2 102 2 4
cos1 a
1
b 45
12
z
Slip plane
(110)
Normal to
slip plane
Slip
direction
[111]
y
Direction of
applied stress
[010]
x
However, for l, we take [u1v1w1] [111] and [u2v2w2] [010], and
l cos1 c
cos1 a
112102 112112 112102
2 3 112 112 2 112 2 4 3 102 2 112 2 102 2 4
2
1
b 54.7
13
Thus, according to Equation 7.2,
tR s cos f cos l 152 MPa21cos 4521cos 54.72
1
1
152 MPa2a
ba
b
12 13
21.3 MPa 13060 psi2
d
1496T_c07_174-206 11/17/05 13:32 Page 185
REVISED PAGES
7.7 Deformation by Twinning • 185
(b) The yield strength sy may be computed from Equation 7.4; f and l will
be the same as for part (a), and
sy
30 MPa
73.4 MPa 110,600 psi2
1cos 4521cos 54.72
7.6 PLASTIC DEFORMATION OF
POLYCRYSTALLINE MATERIALS
Deformation and slip in polycrystalline materials is somewhat more complex.
Because of the random crystallographic orientations of the numerous grains, the
direction of slip varies from one grain to another. For each, dislocation motion occurs
along the slip system that has the most favorable orientation, as defined above. This
is exemplified by a photomicrograph of a polycrystalline copper specimen that has
been plastically deformed (Figure 7.10); before deformation the surface was polished. Slip lines1 are visible, and it appears that two slip systems operated for most
of the grains, as evidenced by two sets of parallel yet intersecting sets of lines. Furthermore, variation in grain orientation is indicated by the difference in alignment
of the slip lines for the several grains.
Gross plastic deformation of a polycrystalline specimen corresponds to the comparable distortion of the individual grains by means of slip. During deformation,
mechanical integrity and coherency are maintained along the grain boundaries; that
is, the grain boundaries usually do not come apart or open up. As a consequence,
each individual grain is constrained, to some degree, in the shape it may assume by
its neighboring grains. The manner in which grains distort as a result of gross plastic deformation is indicated in Figure 7.11. Before deformation the grains are
equiaxed, or have approximately the same dimension in all directions. For this particular deformation, the grains become elongated along the direction in which the
specimen was extended.
Polycrystalline metals are stronger than their single-crystal equivalents, which
means that greater stresses are required to initiate slip and the attendant yielding.
This is, to a large degree, also a result of geometrical constraints that are imposed
on the grains during deformation. Even though a single grain may be favorably oriented with the applied stress for slip, it cannot deform until the adjacent and less
favorably oriented grains are capable of slip also; this requires a higher applied
stress level.
7.7 DEFORMATION BY TWINNING
In addition to slip, plastic deformation in some metallic materials can occur by the
formation of mechanical twins, or twinning. The concept of a twin was introduced in
Section 4.6; that is, a shear force can produce atomic displacements such that on one
side of a plane (the twin boundary), atoms are located in mirror-image positions of
atoms on the other side. The manner in which this is accomplished is demonstrated
1
These slip lines are microscopic ledges produced by dislocations (Figure 7.1c) that have
exited from a grain and appear as lines when viewed with a microscope. They are analogous
to the macroscopic steps found on the surfaces of deformed single crystals (Figures 7.8
and 7.9).
1496T_c07_174-206 11/17/05 13:32 Page 186
REVISED PAGES
186 • Chapter 7 / Dislocations and Strengthening Mechanisms
Figure 7.10 Slip lines on the surface
of a polycrystalline specimen of
copper that was polished and
subsequently deformed. 173.
[Photomicrograph courtesy of
C. Brady, National Bureau of
Standards (now the National Institute
of Standards and Technology,
Gaithersburg, MD).]
Figure 7.11 Alteration of the grain
structure of a polycrystalline metal
as a result of plastic deformation.
(a) Before deformation the grains
are equiaxed. (b) The deformation
has produced elongated grains.
170. (From W. G. Moffatt, G. W.
Pearsall, and J. Wulff, The Structure
and Properties of Materials, Vol. I,
Structure, p. 140. Copyright © 1964
by John Wiley & Sons, New York.
Reprinted by permission of John
Wiley & Sons, Inc.)
1496T_c07_174-206 11/17/05 13:32 Page 187
REVISED PAGES
7.7 Deformation by Twinning • 187
Polished surface
Twin plane
Twin plane
(a)
(b)
Figure 7.12 Schematic diagram showing how twinning results from an applied shear
stress t. In (b), open circles represent atoms that did not change position; dashed and
solid circles represent original and final atom positions, respectively. (From G. E. Dieter,
Mechanical Metallurgy, 3rd edition. Copyright © 1986 by McGraw-Hill Book Company,
New York. Reproduced with permission of McGraw-Hill Book Company.)
in Figure 7.12. Here, open circles represent atoms that did not move, and dashed and
solid circles represent original and final positions, respectively, of atoms within the
twinned region. As may be noted in this figure, the displacement magnitude within
the twin region (indicated by arrows) is proportional to the distance from the twin
plane. Furthermore, twinning occurs on a definite crystallographic plane and in a
specific direction that depend on crystal structure. For example, for BCC metals, the
twin plane and direction are (112) and [111], respectively.
Slip and twinning deformations are compared in Figure 7.13 for a single crystal
that is subjected to a shear stress t. Slip ledges are shown in Figure 7.13a, the formation
of which was described in Section 7.5; for twinning, the shear deformation is homogeneous (Figure 7.13b). These two processes differ from each other in several respects.
First, for slip, the crystallographic orientation above and below the slip plane is the
same both before and after the deformation; for twinning, there will be a reorientation across the twin plane. In addition, slip occurs in distinct atomic spacing multiples,
whereas the atomic displacement for twinning is less than the interatomic separation.
Mechanical twinning occurs in metals that have BCC and HCP crystal structures, at low temperatures, and at high rates of loading (shock loading), conditions
under which the slip process is restricted; that is, there are few operable slip systems.
Twin
planes
Slip
planes
Twin
(a)
(b)
Figure 7.13 For a single
crystal subjected to a shear
stress t, (a) deformation by
slip; (b) deformation by
twinning.
1496T_c07_174-206 11/17/05 13:49 Page 188
REVISED PAGES
188 • Chapter 7 / Dislocations and Strengthening Mechanisms
The amount of bulk plastic deformation from twinning is normally small relative
to that resulting from slip. However, the real importance of twinning lies with the
accompanying crystallographic reorientations; twinning may place new slip systems
in orientations that are favorable relative to the stress axis such that the slip process
can now take place.
M e c h a n i s m s o f St r e n g t h e n i n g i n M e t a l s
Metallurgical and materials engineers are often called on to design alloys having
high strengths yet some ductility and toughness; ordinarily, ductility is sacrificed when
an alloy is strengthened. Several hardening techniques are at the disposal of an engineer, and frequently alloy selection depends on the capacity of a material to be
tailored with the mechanical characteristics required for a particular application.
Important to the understanding of strengthening mechanisms is the relation between dislocation motion and mechanical behavior of metals. Because macroscopic
plastic deformation corresponds to the motion of large numbers of dislocations, the
ability of a metal to plastically deform depends on the ability of dislocations to move.
Since hardness and strength (both yield and tensile) are related to the ease with
which plastic deformation can be made to occur, by reducing the mobility of dislocations, the mechanical strength may be enhanced; that is, greater mechanical forces
will be required to initiate plastic deformation. In contrast, the more unconstrained
the dislocation motion, the greater is the facility with which a metal may deform,
and the softer and weaker it becomes. Virtually all strengthening techniques rely
on this simple principle: restricting or hindering dislocation motion renders a material harder and stronger.
The present discussion is confined to strengthening mechanisms for single-phase
metals, by grain size reduction, solid-solution alloying, and strain hardening. Deformation and strengthening of multiphase alloys are more complicated, involving
concepts beyond the scope of the present discussion; Chapter 10 and Section 11.9
treat techniques that are used to strengthen multiphase alloys.
7.8 STRENGTHENING BY GRAIN SIZE REDUCTION
The size of the grains, or average grain diameter, in a polycrystalline metal influences the mechanical properties. Adjacent grains normally have different crystallographic orientations and, of course, a common grain boundary, as indicated in
Figure 7.14. During plastic deformation, slip or dislocation motion must take place
Figure 7.14 The motion of a dislocation as
it encounters a grain boundary, illustrating
how the boundary acts as a barrier to
continued slip. Slip planes are discontinuous
and change directions across the boundary.
(From Van Vlack, A Textbook of Materials
Technology, 1st edition, © 1973, p. 53.
Adapted by permission of Pearson
Education, Inc., Upper Saddle River, NJ.)
Grain boundary
Slip plane
Grain A
Grain B
1496T_c07_174-206 11/17/05 13:32 Page 189
REVISED PAGES
7.8 Strengthening by Grain Size Reduction • 189
across this common boundary—say, from grain A to grain B in Figure 7.14. The
grain boundary acts as a barrier to dislocation motion for two reasons:
1. Since the two grains are of different orientations, a dislocation passing into
grain B will have to change its direction of motion; this becomes more difficult as the crystallographic misorientation increases.
2. The atomic disorder within a grain boundary region will result in a discontinuity of slip planes from one grain into the other.
It should be mentioned that, for high-angle grain boundaries, it may not be the case
that dislocations traverse grain boundaries during deformation; rather, dislocations
tend to “pile up” (or back up) at grain boundaries. These pile-ups introduce stress
concentrations ahead of their slip planes, which generate new dislocations in adjacent grains.
A fine-grained material (one that has small grains) is harder and stronger than
one that is coarse grained, since the former has a greater total grain boundary area
to impede dislocation motion. For many materials, the yield strength sy varies with
grain size according to
sy s0 kyd1 2
(7.7)
In this expression, termed the Hall-Petch equation, d is the average grain diameter,
and s0 and ky are constants for a particular material. Note that Equation 7.7 is not
valid for both very large (i.e., coarse) grain and extremely fine grain polycrystalline
materials. Figure 7.15 demonstrates the yield strength dependence on grain size for
a brass alloy. Grain size may be regulated by the rate of solidification from the liquid phase, and also by plastic deformation followed by an appropriate heat treatment, as discussed in Section 7.13.
It should also be mentioned that grain size reduction improves not only strength,
but also the toughness of many alloys.
Small-angle grain boundaries (Section 4.6) are not effective in interfering with
the slip process because of the slight crystallographic misalignment across the
boundary. On the other hand, twin boundaries (Section 4.6) will effectively block
Grain size, d (mm)
–1
10
10
–2
5 × 10
–3
30
200
150
20
100
10
50
0
0
4
8
d–1/2 (mm–1/ 2)
12
16
Yield strength (ksi)
Yield strength (MPa)
Hall-Petch
equation—
dependence of yield
strength on grain size
Figure 7.15 The influence of
grain size on the yield strength of
a 70 Cu–30 Zn brass alloy. Note
that the grain diameter increases
from right to left and is not
linear. (Adapted from H. Suzuki,
“The Relation Between the
Structure and Mechanical
Properties of Metals,” Vol. II,
National Physical Laboratory,
Symposium No. 15, 1963, p. 524.)
1496T_c07_174-206 1/9/06 12:56 Page 190
2nd REVISE PAGES
190 • Chapter 7 / Dislocations and Strengthening Mechanisms
slip and increase the strength of the material. Boundaries between two different
phases are also impediments to movements of dislocations; this is important in the
strengthening of more complex alloys.The sizes and shapes of the constituent phases
significantly affect the mechanical properties of multiphase alloys; these are the topics of discussion in Sections 10.7, 10.8, and 16.1.
7.9 SOLID-SOLUTION STRENGTHENING
Another technique to strengthen and harden metals is alloying with impurity atoms
that go into either substitutional or interstitial solid solution. Accordingly, this is
called solid-solution strengthening. High-purity metals are almost always softer and
weaker than alloys composed of the same base metal. Increasing the concentration
of the impurity results in an attendant increase in tensile and yield strengths, as indicated in Figures 7.16a and 7.16b for nickel in copper; the dependence of ductility on nickel concentration is presented in Figure 7.16c.
Alloys are stronger than pure metals because impurity atoms that go into solid
solution ordinarily impose lattice strains on the surrounding host atoms. Lattice strain
field interactions between dislocations and these impurity atoms result, and, consequently, dislocation movement is restricted. For example, an impurity atom that is
smaller than a host atom for which it substitutes exerts tensile strains on the surrounding crystal lattice, as illustrated in Figure 7.17a. Conversely, a larger substitutional
solid-solution
strengthening
180
25
60
40
Yield strength (MPa)
300
Tensile strength (ksi)
Tensile strength (MPa)
50
140
20
120
15
100
80
200
0
10
20
30
40
30
50
10
60
0
10
20
30
Nickel content (wt%)
Nickel content (wt%)
(a)
(b)
40
Elongation (% in 2 in.)
60
50
40
30
20
0
10
20
30
Nickel content (wt%)
(c)
40
50
Figure 7.16 Variation with nickel content of (a) tensile
strength, (b) yield strength, and (c) ductility (%EL) for
copper–nickel alloys, showing strengthening.
50
Yield strength (ksi)
160
400
1496T_c07_174-206 1/9/06 12:56 Page 191
2nd REVISE PAGES
7.10 Strain Hardening • 191
Figure 7.17 (a) Representation
of tensile lattice strains imposed
on host atoms by a smaller
substitutional impurity atom.
(b) Possible locations of smaller
impurity atoms relative to an
edge dislocation such that there
is partial cancellation of
impurity–dislocation lattice
strains.
(a)
(b)
atom imposes compressive strains in its vicinity (Figure 7.18a).These solute atoms tend
to diffuse to and segregate around dislocations in a way so as to reduce the overall
strain energy—that is, to cancel some of the strain in the lattice surrounding a dislocation. To accomplish this, a smaller impurity atom is located where its tensile strain
will partially nullify some of the dislocation’s compressive strain. For the edge dislocation in Figure 7.17b, this would be adjacent to the dislocation line and above the slip
plane. A larger impurity atom would be situated as in Figure 7.18b.
The resistance to slip is greater when impurity atoms are present because the
overall lattice strain must increase if a dislocation is torn away from them. Furthermore, the same lattice strain interactions (Figures 7.17b and 7.18b) will exist
between impurity atoms and dislocations that are in motion during plastic deformation. Thus, a greater applied stress is necessary to first initiate and then continue
plastic deformation for solid-solution alloys, as opposed to pure metals; this is evidenced by the enhancement of strength and hardness.
7.10 STRAIN HARDENING
strain hardening
cold working
Percent cold work—
dependence on
original and
deformed crosssectional areas
Strain hardening is the phenomenon whereby a ductile metal becomes harder and
stronger as it is plastically deformed. Sometimes it is also called work hardening,
or, because the temperature at which deformation takes place is “cold” relative to
the absolute melting temperature of the metal, cold working. Most metals strain
harden at room temperature.
It is sometimes convenient to express the degree of plastic deformation as percent cold work rather than as strain. Percent cold work (%CW) is defined as
%CW a
A0 Ad
b 100
A0
(7.8)
where A0 is the original area of the cross section that experiences deformation, and
Ad is the area after deformation.
Figure 7.18 (a) Representation
of compressive strains imposed
on host atoms by a larger
substitutional impurity atom.
(b) Possible locations of larger
impurity atoms relative to an
edge dislocation such that there
is partial cancellation of
impurity–dislocation lattice
strains.
(a)
(b)
1496T_c07_174-206 11/17/05 13:32 Page 192
REVISED PAGES
192 • Chapter 7 / Dislocations and Strengthening Mechanisms
140
120
900
800
1040 Steel
120
1040 Steel
800
80
500
Brass
400
60
700
100
600
Brass
80
500
Tensile strength (ksi)
600
Tensile strength (MPa)
100
Yield strength (ksi)
Yield strength (MPa)
700
Copper
300
40
60
400
Copper
200
300
40
20
100
0
10
20
30
40
50
60
70
200
0
10
20
30
40
Percent cold work
Percent cold work
(a)
(b)
50
60
70
70
60
Ductility (%EL)
50
40
Brass
30
20
1040 Steel
10
Copper
0
0
10
20
30
40
Percent cold work
(c)
50
60
70
Figure 7.19 For 1040 steel, brass, and copper, (a) the increase
in yield strength, (b) the increase in tensile strength, and (c) the
decrease in ductility (%EL) with percent cold work. [Adapted
from Metals Handbook: Properties and Selection: Irons and
Steels, Vol. 1, 9th edition, B. Bardes (Editor), American Society
for Metals, 1978, p. 226; and Metals Handbook: Properties and
Selection: Nonferrous Alloys and Pure Metals, Vol. 2, 9th edition,
H. Baker (Managing Editor), American Society for Metals, 1979,
pp. 276 and 327.]
Figures 7.19a and 7.19b demonstrate how steel, brass, and copper increase in yield
and tensile strength with increasing cold work. The price for this enhancement of
hardness and strength is in the ductility of the metal. This is shown in Figure 7.19c,
in which the ductility, in percent elongation, experiences a reduction with increasing percent cold work for the same three alloys. The influence of cold work on the
1496T_c07_174-206 11/17/05 13:32 Page 193
REVISED PAGES
7.10 Strain Hardening • 193
Figure 7.20 The influence of
cold work on the stress–strain
behavior of a low-carbon
steel; curves are shown for
0%CW, 4%CW, and 24%CW.
24%CW
600
4%CW
500
Stress (MPa)
0%CW
400
300
200
100
0
0
0.05
0.1
0.15
0.2
0.25
Strain
stress–strain behavior of a low-carbon steel is shown in Figure 7.20; here stress–strain
curves are plotted at 0%CW, 4%CW, and 24%CW.
Strain hardening is demonstrated in a stress–strain diagram presented earlier
(Figure 6.17). Initially, the metal with yield strength sy0 is plastically deformed to point
D. The stress is released, then reapplied with a resultant new yield strength, syi. The
metal has thus become stronger during the process because syi is greater than sy0.
The strain-hardening phenomenon is explained on the basis of dislocation–
dislocation strain field interactions similar to those discussed in Section 7.3. The
dislocation density in a metal increases with deformation or cold work, due to dislocation multiplication or the formation of new dislocations, as noted previously. Consequently, the average distance of separation between dislocations decreases—the
dislocations are positioned closer together. On the average, dislocation–dislocation
strain interactions are repulsive. The net result is that the motion of a dislocation
is hindered by the presence of other dislocations. As the dislocation density increases, this resistance to dislocation motion by other dislocations becomes more
pronounced. Thus, the imposed stress necessary to deform a metal increases with
increasing cold work.
Strain hardening is often utilized commercially to enhance the mechanical properties of metals during fabrication procedures. The effects of strain hardening may
be removed by an annealing heat treatment, as discussed in Section 11.7.
In passing, for the mathematical expression relating true stress and strain,
Equation 6.19, the parameter n is called the strain-hardening exponent, which is a
measure of the ability of a metal to strain harden; the larger its magnitude, the
greater the strain hardening for a given amount of plastic strain.
Concept Check 7.3
When making hardness measurements, what will be the effect of making an indentation very close to a preexisting indentation? Why?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
1496T_c07_174-206 11/17/05 13:32 Page 194
REVISED PAGES
194 • Chapter 7 / Dislocations and Strengthening Mechanisms
Concept Check 7.4
Would you expect a crystalline ceramic material to strain harden at room temperature? Why or why not?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
EXAMPLE PROBLEM 7.2
Tensile Strength and Ductility Determinations for
Cold-Worked Copper
Compute the tensile strength and ductility (%EL) of a cylindrical copper rod
if it is cold worked such that the diameter is reduced from 15.2 mm to 12.2 mm
(0.60 in. to 0.48 in.).
Solution
It is first necessary to determine the percent cold work resulting from the
deformation. This is possible using Equation 7.8:
15.2 mm 2
12.2 mm 2
bpa
bp
2
2
%CW
100 35.6%
15.2 mm 2
a
bp
2
a
The tensile strength is read directly from the curve for copper (Figure 7.19b)
as 340 MPa (50,000 psi). From Figure 7.19c, the ductility at 35.6%CW is about
7%EL.
In summary, we have just discussed the three mechanisms that may be used to
strengthen and harden single-phase metal alloys: strengthening by grain size reduction, solid-solution strengthening, and strain hardening. Of course they may be
used in conjunction with one another; for example, a solid-solution strengthened
alloy may also be strain hardened.
It should also be noted that the strengthening effects due to grain size reduction and strain hardening can be eliminated or at least reduced by an elevatedtemperature heat treatment (Sections 7.12 and 7.13). Conversely, solid-solution
strengthening is unaffected by heat treatment.
Re c ove r y, Re c r ys t a l l i z a t i o n ,
a n d G r a i n G row t h
As outlined in the preceding paragraphs of this chapter, plastically deforming a polycrystalline metal specimen at temperatures that are low relative to its absolute melting temperature produces microstructural and property changes that include (1) a
change in grain shape (Section 7.6), (2) strain hardening (Section 7.10), and (3) an increase in dislocation density (Section 7.3). Some fraction of the energy expended in
deformation is stored in the metal as strain energy, which is associated with tensile,
compressive, and shear zones around the newly created dislocations (Section 7.3).
1496T_c07_174-206 11/17/05 13:32 Page 195
REVISED PAGES
7.12 Recrystallization • 195
Furthermore, other properties such as electrical conductivity (Section 18.8) and corrosion resistance may be modified as a consequence of plastic deformation.
These properties and structures may revert back to the precold-worked states
by appropriate heat treatment (sometimes termed an annealing treatment). Such
restoration results from two different processes that occur at elevated temperatures:
recovery and recrystallization, which may be followed by grain growth.
7.11 RECOVERY
recovery
During recovery, some of the stored internal strain energy is relieved by virtue of
dislocation motion (in the absence of an externally applied stress), as a result of enhanced atomic diffusion at the elevated temperature. There is some reduction in
the number of dislocations, and dislocation configurations (similar to that shown in
Figure 4.8) are produced having low strain energies. In addition, physical properties such as electrical and thermal conductivities and the like are recovered to their
precold-worked states.
7.12 RECRYSTALLIZATION
recrystallization
Even after recovery is complete, the grains are still in a relatively high strain energy state. Recrystallization is the formation of a new set of strain-free and equiaxed
grains (i.e., having approximately equal dimensions in all directions) that have low
dislocation densities and are characteristic of the precold-worked condition. The
driving force to produce this new grain structure is the difference in internal energy between the strained and unstrained material. The new grains form as very
small nuclei and grow until they completely consume the parent material, processes
that involve short-range diffusion. Several stages in the recrystallization process
are represented in Figures 7.21a to 7.21d; in these photomicrographs, the small
(a)
( b)
Figure 7.21 Photomicrographs showing several stages of the recrystallization and grain
growth of brass. (a) Cold-worked (33%CW) grain structure. (b) Initial stage of recrystallization after heating 3 s at 580C (1075F); the very small grains are those that have
recrystallized. (c) Partial replacement of cold-worked grains by recrystallized ones (4 s at
580C). (d) Complete recrystallization (8 s at 580C). (e) Grain growth after 15 min at
580C. ( f ) Grain growth after 10 min at 700C (1290F). All photomicrographs 75.
(Photomicrographs courtesy of J. E. Burke, General Electric Company.)
1496T_c07_174-206 11/17/05 13:32 Page 196
REVISED PAGES
196 • Chapter 7 / Dislocations and Strengthening Mechanisms
(c)
(d)
(e)
(f )
Figure 7.21 (continued)
speckled grains are those that have recrystallized. Thus, recrystallization of coldworked metals may be used to refine the grain structure.
Also, during recrystallization, the mechanical properties that were changed as
a result of cold working are restored to their precold-worked values; that is, the
metal becomes softer, weaker, yet more ductile. Some heat treatments are designed
to allow recrystallization to occur with these modifications in the mechanical characteristics (Section 11.7).
Recrystallization is a process the extent of which depends on both time and
temperature. The degree (or fraction) of recrystallization increases with time, as
may be noted in the photomicrographs shown in Figures 7.21a–d. The explicit
1496T_c07_174-206 11/17/05 13:32 Page 197
REVISED PAGES
7.12 Recrystallization • 197
Annealing temperature (°F)
400
600
800
1000
1200
600
60
Tensile strength
500
40
400
Ductility (%EL)
Tensile strength (MPa)
50
30
Ductility
20
300
Recovery
Recrystallization
Grain growth
Cold-worked
and recovered
grains
Grain size (mm)
Figure 7.22 The
influence of
annealing
temperature (for an
annealing time of
1 h) on the tensile
strength and ductility
of a brass alloy. Grain
size as a function
of annealing
temperature is
indicated. Grain
structures during
recovery,
recrystallization,
and grain growth
stages are shown
schematically.
(Adapted from
G. Sachs and K. R.
Van Horn, Practical
Metallurgy, Applied
Metallurgy and the
Industrial Processing
of Ferrous and
Nonferrous Metals
and Alloys, American
Society for Metals,
1940, p. 139.)
New
grains
0.040
0.030
0.020
0.010
100
200
300
400
500
600
700
Annealing temperature (°C)
recrystallization
temperature
time dependence of recrystallization is addressed in more detail near the end of
Section 10.3.
The influence of temperature is demonstrated in Figure 7.22, which plots tensile strength and ductility (at room temperature) of a brass alloy as a function of
the temperature and for a constant heat treatment time of 1 h. The grain structures
found at the various stages of the process are also presented schematically.
The recrystallization behavior of a particular metal alloy is sometimes specified
in terms of a recrystallization temperature, the temperature at which recrystallization just reaches completion in 1 h. Thus, the recrystallization temperature for the
brass alloy of Figure 7.22 is about 450C 1850F2. Typically, it is between one-third
and one-half of the absolute melting temperature of a metal or alloy and depends
on several factors, including the amount of prior cold work and the purity of the
alloy. Increasing the percentage of cold work enhances the rate of recrystallization,
with the result that the recrystallization temperature is lowered, and approaches a
constant or limiting value at high deformations; this effect is shown in Figure 7.23.
Furthermore, it is this limiting or minimum recrystallization temperature that is
normally specified in the literature. There exists some critical degree of cold work
below which recrystallization cannot be made to occur, as shown in the figure; normally, this is between 2% and 20% cold work.
Recrystallization proceeds more rapidly in pure metals than in alloys. During
recrystallization, grain-boundary motion occurs as the new grain nuclei form and
then grow. It is believed that impurity atoms preferentially segregate at and interact
1496T_c07_174-206 11/17/05 13:32 Page 198
REVISED PAGES
198 • Chapter 7 / Dislocations and Strengthening Mechanisms
Figure 7.23 The
variation of
recrystallization
temperature with
percent cold work for
iron. For deformations
less than the critical
(about 5%CW),
recrystallization will
not occur.
900
1600
1400
700
1200
600
1000
500
800
400
300
0
10
20
Critical
deformation
30
40
Percent cold work
50
60
Recrystallization temperature (F)
Recrystallization temperature (C)
800
600
70
with these recrystallized grain boundaries so as to diminish their (i.e., grain boundary) mobilities; this results in a decrease of the recrystallization rate and raises the
recrystallization temperature, sometimes quite substantially. For pure metals, the recrystallization temperature is normally 0.3Tm, where Tm is the absolute melting temperature; for some commercial alloys it may run as high as 0.7Tm. Recrystallization
and melting temperatures for a number of metals and alloys are listed in Table 7.2.
Plastic deformation operations are often carried out at temperatures above the
recrystallization temperature in a process termed hot working, described in Section
11.4. The material remains relatively soft and ductile during deformation because
it does not strain harden, and thus large deformations are possible.
Concept Check 7.5
Briefly explain why some metals (i.e., lead and tin) do not strain harden when
deformed at room temperature.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Table 7.2 Recrystallization and Melting Temperatures for
Various Metals and Alloys
Recrystallization
Temperature
Metal
Lead
Tin
Zinc
Aluminum (99.999 wt%)
Copper (99.999 wt%)
Brass (60 Cu–40 Zn)
Nickel (99.99 wt%)
Iron
Tungsten
Melting
Temperature
C
F
C
F
4
4
10
80
120
475
370
450
1200
25
25
50
176
250
887
700
840
2200
327
232
420
660
1085
900
1455
1538
3410
620
450
788
1220
1985
1652
2651
2800
6170
1496T_c07_174-206 11/17/05 13:32 Page 199
REVISED PAGES
7.12 Recrystallization • 199
Concept Check 7.6
Would you expect it to be possible for ceramic materials to experience recrystallization? Why or why not?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
DESIGN EXAMPLE 7.1
Description of Diameter Reduction Procedure
A cylindrical rod of noncold-worked brass having an initial diameter of 6.4 mm
(0.25 in.) is to be cold worked by drawing such that the cross-sectional area is
reduced. It is required to have a cold-worked yield strength of at least 345 MPa
(50,000 psi) and a ductility in excess of 20%EL; in addition, a final diameter of
5.1 mm (0.20 in.) is necessary. Describe the manner in which this procedure may
be carried out.
Solution
Let us first consider the consequences (in terms of yield strength and ductility) of
cold working in which the brass specimen diameter is reduced from 6.4 mm (designated by d0) to 5.1 mm (di ). The %CW may be computed from Equation 7.8 as
d0 2
di 2
bpa bp
2
2
%CW
100
2
d0
a bp
2
5.1 mm 2
6.4 mm 2
bpa
bp
a
2
2
100 36.5%CW
6.4 mm 2
a
bp
2
a
From Figures 7.19a and 7.19c, a yield strength of 410 MPa (60,000 psi) and a ductility of 8%EL are attained from this deformation. According to the stipulated
criteria, the yield strength is satisfactory; however, the ductility is too low.
Another processing alternative is a partial diameter reduction, followed by
a recrystallization heat treatment in which the effects of the cold work are nullified. The required yield strength, ductility, and diameter are achieved through
a second drawing step.
Again, reference to Figure 7.19a indicates that 20%CW is required to give
a yield strength of 345 MPa. On the other hand, from Figure 7.19c, ductilities
greater than 20%EL are possible only for deformations of 23%CW or less. Thus
during the final drawing operation, deformation must be between 20%CW and
23%CW. Let’s take the average of these extremes, 21.5%CW, and then calculate
the final diameter for the first drawing d¿0, which becomes the original diameter
for the second drawing. Again, using Equation 7.8,
d¿0 2
5.1 mm 2
b pa
b p
2
2
21.5%CW
100
d¿0 2
a b p
2
a
1496T_c07_174-206 12/21/05 7:48 Page 200
2nd REVISE PAGES
200 • Chapter 7 / Dislocations and Strengthening Mechanisms
Now, solving for d¿0 from the expression above gives
d0¿ 5.8 mm 10.226 in.2
7.13 GRAIN GROWTH
grain growth
For grain growth,
dependence of grain
size on time
After recrystallization is complete, the strain-free grains will continue to grow if the
metal specimen is left at the elevated temperature (Figures 7.21d–f); this phenomenon is called grain growth. Grain growth does not need to be preceded by recovery and recrystallization; it may occur in all polycrystalline materials, metals and
ceramics alike.
An energy is associated with grain boundaries, as explained in Section 4.6. As
grains increase in size, the total boundary area decreases, yielding an attendant
reduction in the total energy; this is the driving force for grain growth.
Grain growth occurs by the migration of grain boundaries. Obviously, not all
grains can enlarge, but large ones grow at the expense of small ones that shrink.
Thus, the average grain size increases with time, and at any particular instant there
will exist a range of grain sizes. Boundary motion is just the short-range diffusion
of atoms from one side of the boundary to the other. The directions of boundary
movement and atomic motion are opposite to each other, as shown in Figure 7.24.
For many polycrystalline materials, the grain diameter d varies with time t
according to the relationship
d n d n0 Kt
(7.9)
where d0 is the initial grain diameter at t 0, and K and n are time-independent
constants; the value of n is generally equal to or greater than 2.
The dependence of grain size on time and temperature is demonstrated in
Figure 7.25, a plot of the logarithm of grain size as a function of the logarithm of
time for a brass alloy at several temperatures. At lower temperatures the curves are
linear. Furthermore, grain growth proceeds more rapidly as temperature increases;
that is, the curves are displaced upward to larger grain sizes. This is explained by
the enhancement of diffusion rate with rising temperature.
Atomic diffusion
across boundary
Direction of grain
boundary motion
Figure 7.24 Schematic representation of grain
growth via atomic diffusion. (From Van Vlack,
Lawrence H., Elements of Materials Science and
Engineering, © 1989, p. 221. Adapted by
permission of Pearson Education, Inc., Upper
Saddle River, NJ.)
1496T_c07_174-206 11/17/05 13:32 Page 201
REVISED PAGES
Summary • 201
850C
1.0
Grain diameter (mm)
(Logarithmic scale)
Figure 7.25 The logarithm of grain diameter
versus the logarithm of time for grain growth in
brass at several temperatures. (From J. E. Burke,
“Some Factors Affecting the Rate of Grain
Growth in Metals.” Reprinted with permission
from Metallurgical Transactions, Vol. 180, 1949, a
publication of The Metallurgical Society of
AIME, Warrendale, Pennsylvania.)
800C
700C
600C
0.1
500C
0.01
1
10
102
Time (min)
(Logarithmic scale)
103
104
The mechanical properties at room temperature of a fine-grained metal are
usually superior (i.e., higher strength and toughness) to those of coarse-grained ones.
If the grain structure of a single-phase alloy is coarser than that desired, refinement
may be accomplished by plastically deforming the material, then subjecting it to a
recrystallization heat treatment, as described above.
SUMMARY
Basic Concepts
Slip Systems
On a microscopic level, plastic deformation corresponds to the motion of dislocations in response to an externally applied shear stress, a process termed “slip.” Slip
occurs on specific crystallographic planes and within these planes only in certain
directions. A slip system represents a slip plane–slip direction combination, and
operable slip systems depend on the crystal structure of the material.
Slip in Single Crystals
The critical resolved shear stress is the minimum shear stress required to initiate
dislocation motion; the yield strength of a single crystal depends on both the magnitude of the critical resolved shear stress and the orientation of slip components
relative to the direction of the applied stress.
Plastic Deformation of Polycrystalline Materials
For polycrystalline materials, slip occurs within each grain along the slip systems
that are most favorably oriented with the applied stress; furthermore, during deformation, grains change shape in such a manner that coherency at the grain boundaries is maintained.
Deformation by Twinning
Under some circumstances limited plastic deformation may occur in BCC and HCP
metals by mechanical twinning. Normally, twinning is important to the degree that
accompanying crystallographic reorientations make the slip process more favorable.
1496T_c07_174-206 11/17/05 13:32 Page 202
REVISED PAGES
202 • Chapter 7 / Dislocations and Strengthening Mechanisms
Characteristics of Dislocations
Strengthening by Grain Size Reduction
Solid-Solution Strengthening
Strain Hardening
Since the ease with which a material is capable of plastic deformation is a function of
dislocation mobility, restricting dislocation motion increases hardness and strength. On
the basis of this principle, three different strengthening mechanisms were discussed.
Grain boundaries serve as barriers to dislocation motion; thus refining the grain size
of a polycrystalline material renders it harder and stronger. Solid-solution strengthening results from lattice strain interactions between impurity atoms and dislocations.
Finally, as a material is plastically deformed, the dislocation density increases, as does
also the extent of repulsive dislocation–dislocation strain field interactions; strain hardening is just the enhancement of strength with increased plastic deformation.
Recovery
Recrystallization
Grain Growth
The microstructural and mechanical characteristics of a plastically deformed metal
specimen may be restored to their predeformed states by an appropriate heat treatment, during which recovery, recrystallization, and grain growth processes are allowed to occur. During recovery there is a reduction in dislocation density and
alterations in dislocation configurations. Recrystallization is the formation of a new
set of grains that are strain free; in addition, the material becomes softer and more
ductile. Grain growth is the increase in average grain size of polycrystalline materials, which proceeds by grain boundary motion.
I M P O R TA N T T E R M S A N D C O N C E P T S
Cold working
Critical resolved shear stress
Dislocation density
Grain growth
Lattice strain
Recovery
Recrystallization
Recrystallization temperature
Resolved shear stress
Slip
Slip system
Solid-solution strengthening
Strain hardening
REFERENCES
Hirth, J. P., and J. Lothe, Theory of Dislocations,
2nd edition, Wiley-Interscience, New York,
1982. Reprinted by Krieger, Melbourne, FL,
1992.
Hull, D., Introduction to Dislocations, 3rd edition,
Butterworth-Heinemann, Woburn, UK, 1984.
Read, W. T., Jr., Dislocations in Crystals, McGrawHill, New York, 1953.
Weertman, J., and J. R.Weertman, Elementary Dislocation Theory, Macmillan, New York, 1964.
Reprinted by Oxford University Press, New
York, 1992.
QUESTIONS AND PROBLEMS
Basic Concepts
Characteristics of Dislocations
7.1 To provide some perspective on the dimensions of atomic defects, consider a metal spec-
imen that has a dislocation density of
105 mm2. Suppose that all the dislocations in
1000 mm3 11 cm3 2 were somehow removed
and linked end to end. How far (in miles)
1496T_c07_174-206 11/17/05 13:32 Page 203
REVISED PAGES
Questions and Problems • 203
would this chain extend? Now suppose that
the density is increased to 109 mm2 by cold
working. What would be the chain length of
dislocations in 1000 mm3 of material?
7.2 Consider two edge dislocations of opposite
sign and having slip planes that are separated
by several atomic distances as indicated in the
diagram. Briefly describe the defect that results when these two dislocations become
aligned with each other.
7.3 Is it possible for two screw dislocations of opposite sign to annihilate each other? Explain
your answer.
7.4 For each of edge, screw, and mixed dislocations,
cite the relationship between the direction of
the applied shear stress and the direction of
dislocation line motion.
Slip Systems
7.5 (a) Define a slip system.
(b) Do all metals have the same slip system?
Why or why not?
7.6 (a) Compare planar densities (Section 3.11
and Problem 3.53) for the (100), (110), and
(111) planes for FCC.
(b) Compare planar densities (Problem 3.54)
for the (100), (110), and (111) planes for BCC.
7.7 One slip system for the BCC crystal structure
is 51106H111I. In a manner similar to Figure
7.6b, sketch a {110}-type plane for the BCC
structure, representing atom positions with
circles. Now, using arrows, indicate two different H111I slip directions within this plane.
7.8 One slip system for the HCP crystal structure
is 500016H1120I. In a manner similar to Figure 7.6b, sketch a {0001}-type plane for the
HCP structure and, using arrows, indicate
three different H1120I slip directions within
this plane. You might find Figure 3.8 helpful.
7.9 Equations 7.1a and 7.1b, expressions for Burgers
vectors for FCC and BCC crystal structures, are
of the form
b
a
HuvwI
2
where a is the unit cell edge length. Also,
since the magnitudes of these Burgers vectors may be determined from the following
equation:
b
a 2
1u v2 w2 2 1 2
2
(7.10)
determine values of b for copper and iron.
You may want to consult Table 3.1.
7.10 (a) In the manner of Equations 7.1a, 7.1b, and
7.1c, specify the Burgers vector for the simple
cubic crystal structure. Its unit cell is shown
in Figure 3.23. Also, simple cubic is the crystal structure for the edge dislocation of
Figure 4.3, and for its motion as presented in
Figure 7.1. You may also want to consult the
answer to Concept Check 7.1.
(b) On the basis of Equation 7.10, formulate
an expression for the magnitude of the Burgers vector, b , for simple cubic.
Slip in Single Crystals
7.11 Sometimes cos cos in Equation 7.2 is
termed the Schmid factor. Determine the
magnitude of the Schmid factor for an FCC
single crystal oriented with its [120] direction
parallel to the loading axis.
7.12 Consider a metal single crystal oriented such
that the normal to the slip plane and the slip
direction are at angles of 60 and 35, respectively, with the tensile axis. If the critical resolved shear stress is 6.2 MPa (900 psi), will
an applied stress of 12 MPa (1750 psi) cause
the single crystal to yield? If not, what stress
will be necessary?
7.13 A single crystal of zinc is oriented for a tensile test such that its slip plane normal makes
an angle of 65 with the tensile axis.Three possible slip directions make angles of 30, 48,
and 78 with the same tensile axis.
(a) Which of these three slip directions is
most favored?
(b) If plastic deformation begins at a tensile
stress of 2.5 MPa (355 psi), determine the critical resolved shear stress for zinc.
7.14 Consider a single crystal of nickel oriented
such that a tensile stress is applied along a
[001] direction. If slip occurs on a (111) plane
and in a [101] direction, and is initiated at an
1496T_c07_174-206 11/17/05 13:32 Page 204
REVISED PAGES
204 • Chapter 7 / Dislocations and Strengthening Mechanisms
7.15
7.16
7.17
7.18
applied tensile stress of 13.9 MPa (2020 psi),
compute the critical resolved shear stress.
A single crystal of a metal that has the FCC
crystal structure is oriented such that a tensile stress is applied parallel to the [100] direction. If the critical resolved shear stress for
this material is 0.5 MPa, calculate the magnitude(s) of applied stress(es) necessary to
cause slip to occur on the (111) plane in each
of the [110], [101], and [011] directions.
(a) A single crystal of a metal that has the
BCC crystal structure is oriented such that a
tensile stress is applied in the [100] direction.
If the magnitude of this stress is 4.0 MPa, compute the resolved shear stress in the [111] direction on each of the (110), (011), and (101)
planes.
(b) On the basis of these resolved shear stress
values, which slip system(s) is (are) most favorably oriented?
Consider a single crystal of some hypothetical metal that has the BCC crystal structure
and is oriented such that a tensile stress
is applied along a [121] direction. If slip occurs on a (101) plane and in a [111] direction, compute the stress at which the crystal
yields if its critical resolved shear stress is
2.4 MPa.
The critical resolved shear stress for copper is
0.48 MPa (70 psi). Determine the maximum
possible yield strength for a single crystal of
Cu pulled in tension.
Deformation by Twinning
7.19 List four major differences between deformation by twinning and deformation by slip
relative to mechanism, conditions of occurrence, and final result.
Strengthening by Grain Size Reduction
7.20 Briefly explain why small-angle grain boundaries are not as effective in interfering with
the slip process as are high-angle grain
boundaries.
7.21 Briefly explain why HCP metals are typically
more brittle than FCC and BCC metals.
7.22 Describe in your own words the three strengthening mechanisms discussed in this chapter
(i.e., grain size reduction, solid-solution
strengthening, and strain hardening). Be sure
to explain how dislocations are involved in
each of the strengthening techniques.
7.23 (a) From the plot of yield strength versus
(grain diameter)1 2 for a 70 Cu–30 Zn cartridge brass, Figure 7.15, determine values for
the constants s0 and ky in Equation 7.7.
(b) Now predict the yield strength of this
alloy when the average grain diameter is
2.0 103 mm.
7.24 The lower yield point for an iron that has an
average grain diameter of 1 102 mm is
230 MPa (33,000 psi). At a grain diameter of
6 103 mm, the yield point increases to
275 MPa (40,000 psi). At what grain diameter will the lower yield point be 310 MPa
(45,000 psi)?
7.25 If it is assumed that the plot in Figure 7.15 is
for noncold-worked brass, determine the grain
size of the alloy in Figure 7.19; assume its composition is the same as the alloy in Figure 7.15.
Solid-Solution Strengthening
7.26 In the manner of Figures 7.17b and 7.18b, indicate the location in the vicinity of an edge
dislocation at which an interstitial impurity
atom would be expected to be situated. Now
briefly explain in terms of lattice strains why
it would be situated at this position.
Strain Hardening
7.27 (a) Show, for a tensile test, that
%CW a
b 100
1
if there is no change in specimen volume during the deformation process (i.e., A0l0 Adld).
(b) Using the result of part (a), compute the
percent cold work experienced by naval brass
(for which the stress–strain behavior is shown
in Figure 6.12) when a stress of 415 MPa
(60,000 psi) is applied.
7.28 Two previously undeformed cylindrical specimens of an alloy are to be strain hardened by
reducing their cross-sectional areas (while
maintaining their circular cross sections). For
one specimen, the initial and deformed radii
are 15 mm and 12 mm, respectively. The second specimen, with an initial radius of 11 mm,
1496T_c07_174-206 11/17/05 13:32 Page 205
REVISED PAGES
Questions and Problems • 205
must have the same deformed hardness as the
first specimen; compute the second specimen’s radius after deformation.
7.29 Two previously undeformed specimens of the
same metal are to be plastically deformed by
reducing their cross-sectional areas. One has
a circular cross section, and the other is rectangular; during deformation the circular
cross section is to remain circular, and the rectangular is to remain as such. Their original
and deformed dimensions are as follows:
Original dimensions
Deformed dimensions
Circular
(diameter, mm)
Rectangular
(mm)
18.0
15.9
20 50
13.7 55.1
Which of these specimens will be the hardest
after plastic deformation, and why?
7.30 A cylindrical specimen of cold-worked copper has a ductility (%EL) of 15%. If its coldworked radius is 6.4 mm (0.25 in.), what was
its radius before deformation?
7.31 (a) What is the approximate ductility (%EL)
of a brass that has a yield strength of 345 MPa
(50,000 psi)?
(b) What is the approximate Brinell hardness
of a 1040 steel having a yield strength of
620 MPa (90,000 psi)?
7.32 Experimentally, it has been observed for single crystals of a number of metals that the critical resolved shear stress tcrss is a function of
the dislocation density rD as
tcrss t0 A2rD
where t0 and A are constants. For copper, the
critical resolved shear stress is 0.69 MPa
(100 psi) at a dislocation density of 104 mm2.
If it is known that the value of t0 for copper
is 0.069 MPa (10 psi), compute the tcrss at a
dislocation density of 106 mm2.
Recovery
Recrystallization
Grain Growth
7.33 Briefly cite the differences between recovery
and recrystallization processes.
7.34 Estimate the fraction of recrystallization from
the photomicrograph in Figure 7.21c.
7.35 Explain the differences in grain structure for a
metal that has been cold worked and one that
has been cold worked and then recrystallized.
7.36 (a) What is the driving force for recrystallization?
(b) For grain growth?
7.37 (a) From Figure 7.25, compute the length of
time required for the average grain diameter
to increase from 0.03 to 0.3 mm at 600C for
this brass material.
(b) Repeat the calculation at 700C.
7.38 The average grain diameter for a brass material was measured as a function of time at
650C, which is tabulated below at two different times:
Time
(min)
40
100
Grain Diameter
(mm)
5.6 102
8.0 102
(a) What was the original grain diameter?
(b) What grain diameter would you predict
after 200 min at 650C?
7.39 An undeformed specimen of some alloy has
an average grain diameter of 0.050 mm. You
are asked to reduce its average grain diameter to 0.020 mm. Is this possible? If so, explain
the procedures you would use and name the
processes involved. If it is not possible, explain
why.
7.40 Grain growth is strongly dependent on
temperature (i.e., rate of grain growth increases with increasing temperature), yet temperature is not explicitly given as a part of
Equation 7.9.
(a) Into which of the parameters in this
expression would you expect temperature to
be included?
(b) On the basis of your intuition, cite an
explicit expression for this temperature
dependence.
7.41 An uncold-worked brass specimen of average
grain size 0.01 mm has a yield strength of
150 MPa (21,750 psi). Estimate the yield
strength of this alloy after it has been heated
to 500C for 1000 s, if it is known that the value
of s0 is 25 MPa (3625 psi).
1496T_c07_174-206 11/17/05 13:32 Page 206
REVISED PAGES
206 • Chapter 7 / Dislocations and Strengthening Mechanisms
DESIGN PROBLEMS
Strain Hardening
Recrystallization
7.D1 Determine whether or not it is possible to
cold work steel so as to give a minimum
Brinell hardness of 240, and at the same time
have a ductility of at least 15%EL. Justify
your decision.
7.D2 Determine whether or not it is possible to
cold work brass so as to give a minimum
Brinell hardness of 150, and at the same time
have a ductility of at least 20%EL. Justify
your decision.
7.D3 A cylindrical specimen of cold-worked steel
has a Brinell hardness of 240.
(a) Estimate its ductility in percent elongation.
(b) If the specimen remained cylindrical
during deformation and its original radius
was 10 mm (0.40 in.), determine its radius
after deformation.
7.D4 It is necessary to select a metal alloy for an
application that requires a yield strength of
at least 310 MPa (45,000 psi) while maintaining a minimum ductility (%EL) of 27%.
If the metal may be cold worked, decide
which of the following are candidates: copper, brass, and a 1040 steel. Why?
7.D5 A cylindrical rod of 1040 steel originally
11.4 mm (0.45 in.) in diameter is to be cold
worked by drawing; the circular cross section
will be maintained during deformation. A
cold-worked tensile strength in excess of
825 MPa (120,000 psi) and a ductility of at
least 12%EL are desired. Furthermore, the
final diameter must be 8.9 mm (0.35 in.).
Explain how this may be accomplished.
7.D6 A cylindrical rod of brass originally 10.2 mm
(0.40 in.) in diameter is to be cold worked by
drawing; the circular cross section will be
maintained during deformation. A coldworked yield strength in excess of 380 MPa
(55, 000 psi) and a ductility of at least 15%EL
are desired. Furthermore, the final diameter
must be 7.6 mm (0.30 in.). Explain how this
may be accomplished.
7.D7 A cylindrical brass rod having a minimum
tensile strength of 450 MPa (65,000 psi), a
ductility of at least 13%EL, and a final
diameter of 12.7 mm (0.50 in.) is desired.
Some 19.0 mm (0.75 in.) diameter brass stock
that has been cold worked 35% is available.
Describe the procedure you would follow
to obtain this material. Assume that brass
experiences cracking at 65%CW.
1496T_c08_207-251 12/21/05 8:28 Page 207
2nd REVISE PAGES
Chapter
A
8
Failure
n oil tanker that fractured in a brittle manner by crack propagation around its girth.
(Photography by Neal Boenzi. Reprinted with permission from The New York Times.)
WHY STUDY Failure?
The design of a component or structure often calls
upon the engineer to minimize the possibility of failure. Thus, it is important to understand the mechanics
of the various failure modes—i.e., fracture, fatigue,
and creep—and, in addition, be familiar with appro-
priate design principles that may be employed to
prevent in-service failures. For example, we discuss
in Sections 22.4 through 22.6 material selection and
processing issues relating to the fatigue of an automobile valve spring.
• 207
1496T_c08_207-251 12/21/05 7:51 Page 208
2nd REVISE PAGES
Learning Objectives
After studying this chapter you should be able to do the following:
1. Describe the mechanism of crack propagation
6. Define fatigue and specify the conditions under
for both ductile and brittle modes of fracture.
which it occurs.
2. Explain why the strengths of brittle materials are
7. From a fatigue plot for some material, determuch lower than predicted by theoretical calcumine (a) the fatigue lifetime (at a specified
lations.
stress level), and (b) the fatigue strength (at a
3. Define fracture toughness in terms of (a) a brief
specified number of cycles).
statement, and (b) an equation; define all pa8. Define creep and specify the conditions under
rameters in this equation.
which it occurs.
4. Make a distinction between fracture toughness
9. Given a creep plot for some material, determine
and plane strain fracture toughness.
(a) the steady-state creep rate, and (b) the
5. Name and describe the two impact fracture
rupture lifetime.
testing techniques.
8.1 INTRODUCTION
The failure of engineering materials is almost always an undesirable event for several
reasons; these include human lives that are put in jeopardy, economic losses, and the interference with the availability of products and services. Even though the causes of failure and the behavior of materials may be known, prevention of failures is difficult to
guarantee. The usual causes are improper materials selection and processing and inadequate design of the component or its misuse. It is the responsibility of the engineer
to anticipate and plan for possible failure and, in the event that failure does occur, to
assess its cause and then take appropriate preventive measures against future incidents.
The following topics are addressed in this chapter: simple fracture (both ductile and brittle modes), fundamentals of fracture mechanics, impact fracture testing,
the ductile-to-brittle transition, fatigue, and creep. These discussions include failure
mechanisms, testing techniques, and methods by which failure may be prevented or
controlled.
Concept Check 8.1
Cite two situations in which the possibility of failure is part of the design of a component or product.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Fr a c t u r e
8.2 FUNDAMENTALS OF FRACTURE
ductile, brittle
fracture
Simple fracture is the separation of a body into two or more pieces in response to
an imposed stress that is static (i.e., constant or slowly changing with time) and at
temperatures that are low relative to the melting temperature of the material. The
applied stress may be tensile, compressive, shear, or torsional; the present discussion
will be confined to fractures that result from uniaxial tensile loads. For engineering
materials, two fracture modes are possible: ductile and brittle. Classification is based
on the ability of a material to experience plastic deformation. Ductile materials typically exhibit substantial plastic deformation with high energy absorption before
fracture. On the other hand, there is normally little or no plastic deformation with
1496T_c08_207-251 11/18/05 11:50 Page 209
REVISED PAGES
8.3 Ductile Fracture • 209
low energy absorption accompanying a brittle fracture. The tensile stress–strain
behaviors of both fracture types may be reviewed in Figure 6.13.
“Ductile”and “brittle” are relative terms; whether a particular fracture is one
mode or the other depends on the situation. Ductility may be quantified in terms
of percent elongation (Equation 6.11) and percent reduction in area (Equation
6.12). Furthermore, ductility is a function of temperature of the material, the strain
rate, and the stress state. The disposition of normally ductile materials to fail in a
brittle manner is discussed in Section 8.6.
Any fracture process involves two steps—crack formation and propagation—in
response to an imposed stress. The mode of fracture is highly dependent on the mechanism of crack propagation. Ductile fracture is characterized by extensive plastic deformation in the vicinity of an advancing crack. Furthermore, the process proceeds
relatively slowly as the crack length is extended. Such a crack is often said to be stable.
That is, it resists any further extension unless there is an increase in the applied stress.
In addition, there will ordinarily be evidence of appreciable gross deformation at the
fracture surfaces (e.g., twisting and tearing). On the other hand, for brittle fracture,
cracks may spread extremely rapidly, with very little accompanying plastic deformation. Such cracks may be said to be unstable, and crack propagation, once started, will
continue spontaneously without an increase in magnitude of the applied stress.
Ductile fracture is almost always preferred for two reasons. First, brittle fracture occurs suddenly and catastrophically without any warning; this is a consequence
of the spontaneous and rapid crack propagation. On the other hand, for ductile fracture, the presence of plastic deformation gives warning that fracture is imminent,
allowing preventive measures to be taken. Second, more strain energy is required
to induce ductile fracture inasmuch as ductile materials are generally tougher. Under
the action of an applied tensile stress, most metal alloys are ductile, whereas ceramics
are notably brittle, and polymers may exhibit both types of fracture.
8.3 DUCTILE FRACTURE
Ductile fracture surfaces will have their own distinctive features on both macroscopic
and microscopic levels. Figure 8.1 shows schematic representations for two characteristic macroscopic fracture profiles. The configuration shown in Figure 8.1a is found
for extremely soft metals, such as pure gold and lead at room temperature, and other
metals, polymers, and inorganic glasses at elevated temperatures. These highly ductile materials neck down to a point fracture, showing virtually 100% reduction in area.
The most common type of tensile fracture profile for ductile metals is that represented in Figure 8.1b, where fracture is preceded by only a moderate amount of
Figure 8.1 (a) Highly ductile fracture in
which the specimen necks down to a point.
(b) Moderately ductile fracture after some
necking. (c) Brittle fracture without any
plastic deformation.
(a)
(b)
(c)
1496T_c08_207-251 11/18/05 11:50 Page 210
REVISED PAGES
210 • Chapter 8 / Failure
(a)
(b)
(c)
Shear
Fibrous
(d)
Figure 8.2 Stages in the cup-and-cone
fracture. (a) Initial necking. (b) Small
cavity formation. (c) Coalescence of
cavities to form a crack. (d) Crack
propagation. (e) Final shear fracture at a
45 angle relative to the tensile direction.
(From K. M. Ralls, T. H. Courtney, and
J. Wulff, Introduction to Materials Science
and Engineering, p. 468. Copyright © 1976
by John Wiley & Sons, New York.
Reprinted by permission of John Wiley &
Sons, Inc.)
(e)
necking. The fracture process normally occurs in several stages (Figure 8.2). First,
after necking begins, small cavities, or microvoids, form in the interior of the cross
section, as indicated in Figure 8.2b. Next, as deformation continues, these microvoids
enlarge, come together, and coalesce to form an elliptical crack, which has its long
axis perpendicular to the stress direction. The crack continues to grow in a direction
parallel to its major axis by this microvoid coalescence process (Figure 8.2c). Finally,
fracture ensues by the rapid propagation of a crack around the outer perimeter of
the neck (Figure 8.2d), by shear deformation at an angle of about 45 with the tensile axis—this is the angle at which the shear stress is a maximum. Sometimes a fracture having this characteristic surface contour is termed a cup-and-cone fracture
because one of the mating surfaces is in the form of a cup, the other like a cone. In
this type of fractured specimen (Figure 8.3a), the central interior region of the surface
has an irregular and fibrous appearance, which is indicative of plastic deformation.
Fractographic Studies
Much more detailed information regarding the mechanism of fracture is available
from microscopic examination, normally using scanning electron microscopy. Studies
(a)
(b)
Figure 8.3 (a) Cup-and-cone fracture in aluminum. (b) Brittle fracture in a mild steel.
1496T_c08_207-251 11/18/05 11:50 Page 211
REVISED PAGES
8.4 Brittle Fracture • 211
Figure 8.4 (a) Scanning electron fractograph showing spherical dimples characteristic
of ductile fracture resulting from uniaxial tensile loads. 3300. (b) Scanning electron
fractograph showing parabolic-shaped dimples characteristic of ductile fracture
resulting from shear loading. 5000. (From R. W. Hertzberg, Deformation and
Fracture Mechanics of Engineering Materials, 3rd edition. Copyright © 1989 by John
Wiley & Sons, New York. Reprinted by permission of John Wiley & Sons, Inc.)
of this type are termed fractographic. The scanning electron microscope is preferred
for fractographic examinations since it has a much better resolution and depth of field
than does the optical microscope; these characteristics are necessary to reveal the topographical features of fracture surfaces.
When the fibrous central region of a cup-and-cone fracture surface is examined
with the electron microscope at a high magnification, it will be found to consist of
numerous spherical “dimples” (Figure 8.4a); this structure is characteristic of fracture resulting from uniaxial tensile failure. Each dimple is one half of a microvoid
that formed and then separated during the fracture process. Dimples also form on
the 45 shear lip of the cup-and-cone fracture. However, these will be elongated or
C-shaped, as shown in Figure 8.4b. This parabolic shape may be indicative of shear
failure. Furthermore, other microscopic fracture surface features are also possible.
Fractographs such as those shown in Figures 8.4a and 8.4b provide valuable information in the analyses of fracture, such as the fracture mode, the stress state, and
the site of crack initiation.
8.4 BRITTLE FRACTURE
Brittle fracture takes place without any appreciable deformation, and by rapid crack
propagation. The direction of crack motion is very nearly perpendicular to the direction of the applied tensile stress and yields a relatively flat fracture surface, as
indicated in Figure 8.1c.
Fracture surfaces of materials that failed in a brittle manner will have their own
distinctive patterns; any signs of gross plastic deformation will be absent. For
example, in some steel pieces, a series of V-shaped “chevron” markings may form
near the center of the fracture cross section that point back toward the crack initiation site (Figure 8.5a). Other brittle fracture surfaces contain lines or ridges that
radiate from the origin of the crack in a fanlike pattern (Figure 8.5b). Often, both
of these marking patterns will be sufficiently coarse to be discerned with the naked
eye. For very hard and fine-grained metals, there will be no discernible fracture pattern. Brittle fracture in amorphous materials, such as ceramic glasses, yields a relatively shiny and smooth surface.
1496T_c08_207-251 12/21/05 7:51 Page 212
2nd REVISE PAGES
212 • Chapter 8 / Failure
(a)
(b)
Figure 8.5 (a) Photograph showing V-shaped “chevron” markings characteristic of brittle
fracture. Arrows indicate origin of crack. Approximately actual size. (b) Photograph of a
brittle fracture surface showing radial fan-shaped ridges. Arrow indicates origin of crack.
Approximately 2. [(a) From R. W. Hertzberg, Deformation and Fracture Mechanics of
Engineering Materials, 3rd edition. Copyright © 1989 by John Wiley & Sons, New York.
Reprinted by permission of John Wiley & Sons, Inc. Photograph courtesy of Roger Slutter,
Lehigh University. (b) Reproduced with permission from D. J. Wulpi, Understanding How
Components Fail, American Society for Metals, Materials Park, OH, 1985.]
transgranular
fracture
intergranular
fracture
For most brittle crystalline materials, crack propagation corresponds to the successive and repeated breaking of atomic bonds along specific crystallographic planes
(Figure 8.6a); such a process is termed cleavage. This type of fracture is said to be
transgranular (or transcrystalline), because the fracture cracks pass through the
grains. Macroscopically, the fracture surface may have a grainy or faceted texture
(Figure 8.3b), as a result of changes in orientation of the cleavage planes from grain
to grain. This cleavage feature is shown at a higher magnification in the scanning
electron micrograph of Figure 8.6b.
In some alloys, crack propagation is along grain boundaries (Figure 8.7a); this
fracture is termed intergranular. Figure 8.7b is a scanning electron micrograph
1496T_c08_207-251 11/18/05 13:43 Page 213
REVISED PAGES
8.4 Brittle Fracture • 213
SEM Micrograph
Path of crack propagation
Grains
(a)
(b)
Figure 8.6 (a) Schematic cross-section profile showing crack propagation through the
interior of grains for transgranular fracture. (b) Scanning electron fractograph of ductile
cast iron showing a transgranular fracture surface. Magnification unknown. [Figure (b)
from V. J. Colangelo and F. A. Heiser, Analysis of Metallurgical Failures, 2nd edition.
Copyright © 1987 by John Wiley & Sons, New York. Reprinted by permission of John
Wiley & Sons, Inc.]
1496T_c08_207-251 11/18/05 13:43 Page 214
REVISED PAGES
214 • Chapter 8 / Failure
SEM Micrograph
Grain boundaries
Path of crack propagation
(a)
(b)
Figure 8.7 (a) Schematic cross-section profile showing crack propagation along grain
boundaries for intergranular fracture. (b) Scanning electron fractograph showing an
intergranular fracture surface. 50. [Figure (b) reproduced with permission from ASM
Handbook, Vol. 12, Fractography, ASM International, Materials Park, OH, 1987.]
1496T_c08_207-251 11/18/05 11:50 Page 215
REVISED PAGES
8.5 Principles of Fracture Mechanics • 215
showing a typical intergranular fracture, in which the three-dimensional nature of
the grains may be seen. This type of fracture normally results subsequent to the
occurrence of processes that weaken or embrittle grain boundary regions.
8.5 PRINCIPLES OF FRACTURE MECHANICS
Brittle fracture of normally ductile materials, such as that shown in the chapteropening photograph for this chapter, has demonstrated the need for a better understanding of the mechanisms of fracture. Extensive research endeavors over
the past several decades have led to the evolution of the field of fracture mechanics. This subject allows quantification of the relationships between material
properties, stress level, the presence of crack-producing flaws, and crack propagation mechanisms. Design engineers are now better equipped to anticipate,
and thus prevent, structural failures. The present discussion centers on some of
the fundamental principles of the mechanics of fracture.
fracture mechanics
Stress Concentration
The measured fracture strengths for most brittle materials are significantly lower
than those predicted by theoretical calculations based on atomic bonding energies. This discrepancy is explained by the presence of very small, microscopic flaws
or cracks that always exist under normal conditions at the surface and within the
interior of a body of material. These flaws are a detriment to the fracture strength
because an applied stress may be amplified or concentrated at the tip, the magnitude of this amplification depending on crack orientation and geometry. This
phenomenon is demonstrated in Figure 8.8, a stress profile across a cross section
containing an internal crack. As indicated by this profile, the magnitude of this
localized stress diminishes with distance away from the crack tip. At positions far
σ0
σm
Stress
ρt
a
X
X'
x
2a
x'
σ0
x
x'
Position along X–X'
(a)
σ0
(b)
Figure 8.8 (a) The geometry of surface and internal cracks. (b) Schematic stress profile
along the line X–X ¿ in (a), demonstrating stress amplification at crack tip positions.
1496T_c08_207-251 11/18/05 11:50 Page 216
REVISED PAGES
216 • Chapter 8 / Failure
stress raiser
removed, the stress is just the nominal stress s0, or the applied load divided by
the specimen cross-sectional area (perpendicular to this load). Due to their ability to amplify an applied stress in their locale, these flaws are sometimes called
stress raisers.
If it is assumed that a crack is similar to an elliptical hole through a plate, and
is oriented perpendicular to the applied stress, the maximum stress, sm, occurs at
the crack tip and may be approximated by
For tensile loading,
computation of
maximum stress
at a crack tip
a 1 2
sm 2s0 a b
rt
(8.1)
where s0 is the magnitude of the nominal applied tensile stress, rt is the radius of
curvature of the crack tip (Figure 8.8a), and a represents the length of a surface
crack, or half of the length of an internal crack. For a relatively long microcrack
that has a small tip radius of curvature, the factor 1art 2 1 2 may be very large. This
will yield a value of sm that is many times the value of s0.
Sometimes the ratio sms0 is denoted as the stress concentration factor Kt:
Kt
sm
a 1 2
2a b
s0
rt
(8.2)
which is simply a measure of the degree to which an external stress is amplified at
the tip of a crack.
By way of comment, it should be said that stress amplification is not restricted
to these microscopic defects; it may occur at macroscopic internal discontinuities
(e.g., voids), at sharp corners, and at notches in large structures.
Furthermore, the effect of a stress raiser is more significant in brittle than in
ductile materials. For a ductile material, plastic deformation ensues when the maximum stress exceeds the yield strength. This leads to a more uniform distribution
of stress in the vicinity of the stress raiser and to the development of a maximum
stress concentration factor less than the theoretical value. Such yielding and stress
redistribution do not occur to any appreciable extent around flaws and discontinuities in brittle materials; therefore, essentially the theoretical stress concentration
will result.
Using principles of fracture mechanics, it is possible to show that the critical
stress sc required for crack propagation in a brittle material is described by the
expression
Critical stress for
crack propagation in
a brittle material
sc a
2Egs 1 2
b
pa
(8.3)
where
E modulus of elasticity
gs specific surface energy
a one half the length of an internal crack
All brittle materials contain a population of small cracks and flaws that have a
variety of sizes, geometries, and orientations. When the magnitude of a tensile stress
at the tip of one of these flaws exceeds the value of this critical stress, a crack forms
and then propagates, which results in fracture. Very small and virtually defect-free
metallic and ceramic whiskers have been grown with fracture strengths that approach
their theoretical values.
1496T_c08_207-251 11/18/05 11:50 Page 217
REVISED PAGES
8.5 Principles of Fracture Mechanics • 217
EXAMPLE PROBLEM 8.1
Maximum Flaw Length Computation
A relatively large plate of a glass is subjected to a tensile stress of 40 MPa. If
the specific surface energy and modulus of elasticity for this glass are 0.3 J/m2
and 69 GPa, respectively, determine the maximum length of a surface flaw
that is possible without fracture.
Solution
To solve this problem it is necessary to employ Equation 8.3. Rearrangement
of this expression such that a is the dependent variable, and realizing that
s 40 MPa, gs 0.3 J/m2, and E 69 GPa leads to
a
2Egs
ps2
122169 109 N/m2 210.3 N/m2
p 140 106 N/m2 2 2
8.2 106 m 0.0082 mm 8.2 mm
Fracture Toughness
Furthermore, using fracture mechanical principles, an expression has been developed
that relates this critical stress for crack propagation (sc) and crack length (a) as
Fracture toughness—
dependence on
critical stress for
crack propagation
and crack length
fracture toughness
Kc Ysc 1pa
(8.4)
In this expression Kc is the fracture toughness, a property that is a measure of a
material’s resistance to brittle fracture when a crack is present. Worth noting is that
Kc has the unusual units of MPa 1m or psi 1in. (alternatively, ksi 1in.). Furthermore, Y is a dimensionless parameter or function that depends on both crack and
specimen sizes and geometries, as well as the manner of load application.
Relative to this Y parameter, for planar specimens containing cracks that are
much shorter than the specimen width, Y has a value of approximately unity. For
example, for a plate of infinite width having a through-thickness crack (Figure 8.9a),
Figure 8.9 Schematic
representations of (a) an
interior crack in a plate of
infinite width, and (b) an
edge crack in a plate of
semi-infinite width.
2a
(a)
a
(b)
1496T_c08_207-251 11/18/05 11:50 Page 218
REVISED PAGES
218 • Chapter 8 / Failure
Figure 8.10 The
three modes of
crack surface
displacement.
(a) Mode I, opening
or tensile mode;
(b) mode II, sliding
mode; and (c) mode
III, tearing mode.
(a)
plane strain
plane strain fracture
toughness
Plane strain fracture
toughness for mode I
crack surface
displacement
(b)
(c)
Y 1.0; whereas for a plate of semi-infinite width containing an edge crack of length
a (Figure 8.9b), Y 1.1. Mathematical expressions for Y have been determined for a
variety of crack-specimen geometries; these expressions are often relatively complex.
For relatively thin specimens, the value of Kc will depend on specimen thickness. However, when specimen thickness is much greater than the crack dimensions,
Kc becomes independent of thickness; under these conditions a condition of plane
strain exists. By plane strain we mean that when a load operates on a crack in the
manner represented in Figure 8.9a, there is no strain component perpendicular to
the front and back faces. The Kc value for this thick-specimen situation is known as
the plane strain fracture toughness KIc; furthermore, it is also defined by
KIc Ys2pa
(8.5)
KIc is the fracture toughness cited for most situations. The I (i.e., Roman numeral
“one”) subscript for KIc denotes that the plane strain fracture toughness is for mode
I crack displacement, as illustrated in Figure 8.10a.1
Brittle materials, for which appreciable plastic deformation is not possible in
front of an advancing crack, have low KIc values and are vulnerable to catastrophic
failure. On the other hand, KIc values are relatively large for ductile materials. Fracture mechanics is especially useful in predicting catastrophic failure in materials
having intermediate ductilities. Plane strain fracture toughness values for a number
of different materials are presented in Table 8.1 (and Figure 1.6); a more extensive
list of KIc values is contained in Table B.5, Appendix B.
The plane strain fracture toughness KIc is a fundamental material property that
depends on many factors, the most influential of which are temperature, strain rate,
and microstructure. The magnitude of KIc diminishes with increasing strain rate and
decreasing temperature. Furthermore, an enhancement in yield strength wrought
by solid solution or dispersion additions or by strain hardening generally produces
a corresponding decrease in KIc. Furthermore, KIc normally increases with reduction
in grain size as composition and other microstructural variables are maintained constant. Yield strengths are included for some of the materials listed in Table 8.1.
Several different testing techniques are used to measure KIc.2 Virtually any specimen size and shape consistent with mode I crack displacement may be utilized,
1
Two other crack displacement modes denoted by II and III and as illustrated in Figures
8.10b and 8.10c are also possible; however, mode I is most commonly encountered.
2
See for example ASTM Standard E 399, “Standard Test Method for Plane Strain Fracture Toughness of Metallic Materials.”
1496T_c08_207-251 11/18/05 11:50 Page 219
REVISED PAGES
8.5 Principles of Fracture Mechanics • 219
Table 8.1 Room-Temperature Yield Strength and Plane Strain Fracture
Toughness Data for Selected Engineering Materials
Yield Strength
Material
MPa
KIc
ksi
MPa 1m
ksi 1in .
Metals
Aluminum Alloya
(7075-T651)
Aluminum Alloya
(2024-T3)
Titanium Alloya
(Ti-6Al-4V)
Alloy Steela
(4340 tempered @ 260C)
Alloy Steela
(4340 tempered @ 425C)
495
72
24
22
345
50
44
40
910
132
55
50
1640
238
50.0
45.8
1420
206
87.4
80.0
—
—
—
0.2–1.4
0.7–0.8
2.7–5.0
0.18–1.27
0.64–0.73
2.5–4.6
Ceramics
Concrete
Soda-Lime Glass
Aluminum Oxide
—
—
—
Polystyrene
(PS)
Poly(methyl methacrylate)
(PMMA)
Polycarbonate
(PC)
—
—
0.7–1.1
0.64–1.0
53.8–73.1
7.8–10.6
0.7–1.6
0.64–1.5
62.1
9.0
2.2
2.0
Polymers
a
Source: Reprinted with permission, Advanced Materials and Processes, ASM International, © 1990.
and accurate values will be realized provided that the Y scale parameter in Equation 8.5 has been properly determined.
Design Using Fracture Mechanics
According to Equations 8.4 and 8.5, three variables must be considered relative to
the possibility for fracture of some structural component—namely, the fracture
toughness (Kc) or plane strain fracture toughness (KIc), the imposed stress (s), and
the flaw size (a)—assuming, of course, that Y has been determined. When designing a component, it is first important to decide which of these variables are constrained by the application and which are subject to design control. For example,
material selection (and hence Kc or KIc) is often dictated by factors such as density
(for lightweight applications) or the corrosion characteristics of the environment.
Or, the allowable flaw size is either measured or specified by the limitations of
available flaw detection techniques. It is important to realize, however, that once
any combination of two of the above parameters is prescribed, the third becomes
fixed (Equations 8.4 and 8.5). For example, assume that KIc and the magnitude of
a are specified by application constraints; therefore, the design (or critical) stress sc
must be
Computation of
design stress
sc
KIc
Y1pa
(8.6)
1496T_c08_207-251 11/18/05 11:50 Page 220
REVISED PAGES
220 • Chapter 8 / Failure
Table 8.2 A List of Several Common Nondestructive Testing (NDT) Techniques
Technique
Scanning electron
microscopy (SEM)
Dye penetrant
Ultrasonics
Optical microscopy
Visual inspection
Acoustic emission
Radiography
(X-ray/gamma ray)
Defect Location
Defect Size
Sensitivity (mm)
Testing Location
Surface
0.001
Laboratory
Surface
Subsurface
Surface
Surface
Surface/subsurface
Subsurface
0.025–0.25
0.050
0.1–0.5
0.1
0.1
2% of specimen
thickness
Laboratory/in-field
Laboratory/in-field
Laboratory
Laboratory/in-field
Laboratory/in-field
Laboratory/in-field
On the other hand, if stress level and plane strain fracture toughness are fixed by
the design situation, then the maximum allowable flaw size ac is
Computation of
maximum allowable
flaw length
ac
1 KIc 2
a
b
p sY
(8.7)
A number of nondestructive test (NDT) techniques have been developed that
permit detection and measurement of both internal and surface flaws.3 Such techniques are used to examine structural components that are in service for defects
and flaws that could lead to premature failure; in addition, NDTs are used as a
means of quality control for manufacturing processes. As the name implies, these
techniques must not destroy the material/structure being examined. Furthermore,
some testing methods must be conducted in a laboratory setting; others may be
adapted for use in the field. Several commonly employed NDT techniques and their
characteristics are listed in Table 8.2.
One important example of the use of NDT is for the detection of cracks and
leaks in the walls of oil pipelines in remote areas such as Alaska. Ultrasonic analysis is utilized in conjunction with a “robotic analyzer” that can travel relatively long
distances within a pipeline.
DESIGN EXAMPLE 8.1
Material Specification for a Pressurized Spherical Tank
Consider the thin-walled spherical tank of radius r and thickness t (Figure 8.11)
that may be used as a pressure vessel.
(a) One design of such a tank calls for yielding of the wall material prior to failure as a result of the formation of a crack of critical size and its subsequent rapid
propagation. Thus, plastic distortion of the wall may be observed and the pressure within the tank released before the occurrence of catastrophic failure. Consequently, materials having large critical crack lengths are desired. On the basis
3
Sometimes the terms nondestructive evaluation (NDE) and nondestructive inspection
(NDI) are also used for these techniques.
1496T_c08_207-251 11/18/05 11:50 Page 221
REVISED PAGES
8.5 Principles of Fracture Mechanics • 221
2a
p
p
p
t
r
p
p
p
Figure 8.11 Schematic
diagram showing the
cross section of a
spherical tank that is
subjected to an
internal pressure p,
and that has a radial
crack of length 2a in
its wall.
p
p
of this criterion, rank the metal alloys listed in Table B.5,Appendix B, as to critical
crack size, from longest to shortest.
(b) An alternative design that is also often utilized with pressure vessels is
termed leak-before-break. Using principles of fracture mechanics, allowance is
made for the growth of a crack through the thickness of the vessel wall prior to
the occurrence of rapid crack propagation (Figure 8.11). Thus, the crack will completely penetrate the wall without catastrophic failure, allowing for its detection
by the leaking of pressurized fluid. With this criterion the critical crack length ac
(i.e., one-half of the total internal crack length) is taken to be equal to the pressure vessel thickness t. Allowance for ac t instead of ac t2 assures that fluid
leakage will occur prior to the buildup of dangerously high pressures. Using this
criterion, rank the metal alloys in Table B.5, Appendix B as to the maximum
allowable pressure.
For this spherical pressure vessel, the circumferential wall stress s is a function of the pressure p in the vessel and the radius r and wall thickness t according to
pr
s
(8.8)
2t
For both parts (a) and (b) assume a condition of plane strain.
Solution
(a) For the first design criterion, it is desired that the circumferential wall stress
be less than the yield strength of the material. Substitution of sy for s in Equation 8.5, and incorporation of a factor of safety N leads to
KIc Y a
sy
N
b 1pac
(8.9)
where ac is the critical crack length. Solving for ac yields the following expression:
ac
N 2 KIc 2
a
b
Y 2p sy
(8.10)
Therefore, the critical crack length is proportional to the square of the KIc-sy
ratio, which is the basis for the ranking of the metal alloys in Table B.5. The
1496T_c08_207-251 11/18/05 11:50 Page 222
REVISED PAGES
222 • Chapter 8 / Failure
Table 8.3 Ranking of Several Metal Alloys
Relative to Critical Crack Length
(Yielding Criterion) for a ThinWalled Spherical Pressure Vessel
Material
Medium carbon (1040) steel
AZ31B magnesium
2024 aluminum (T3)
Ti-5Al-2.5Sn titanium
4140 steel
(tempered @ 482C)
4340 steel
(tempered @ 425C)
Ti-6Al-4V titanium
17-7PH steel
7075 aluminum (T651)
4140 steel
(tempered @ 370C)
4340 steel
(tempered @ 260C)
a
KIc
Sy
2
b (mm)
43.1
19.6
16.3
6.6
5.3
3.8
3.7
3.4
2.4
1.6
0.93
ranking is provided in Table 8.3, where it may be seen that the medium carbon
(1040) steel with the largest ratio has the longest critical crack length, and, therefore, is the most desirable material on the basis of this criterion.
(b) As stated previously, the leak-before-break criterion is just met when onehalf of the internal crack length is equal to the thickness of the pressure vessel—
that is, when a t. Substitution of a t into Equation 8.5 gives
KIc Ys1pt
(8.11)
and, from Equation 8.8
t
pr
2s
(8.12)
The stress is replaced by the yield strength, inasmuch as the tank should be
designed to contain the pressure without yielding; furthermore, substitution of
Equation 8.12 into Equation 8.11, after some rearrangement, yields the following expression:
K 2Ic
2
(8.13)
p 2 a
b
Y pr sy
Hence, for some given spherical vessel of radius r, the maximum allowable pressure consistent with this leak-before-break criterion is proportional to K 2Ic sy.
The same several materials are ranked according to this ratio in Table 8.4; as may
be noted, the medium carbon steel will contain the greatest pressures.
Of the 11 metal alloys that are listed in Table B.5, the medium carbon steel
ranks first according to both yielding and leak-before-break criteria. For these
reasons, many pressure vessels are constructed of medium carbon steels, when
temperature extremes and corrosion need not be considered.
1496T_c08_207-251 11/18/05 11:50 Page 223
REVISED PAGES
8.6 Impact Fracture Testing • 223
Table 8.4 Ranking of Several Metal Alloys
Relative to Maximum Allowable
Pressure (Leak-Before-Break
Criterion) for a Thin-Walled
Spherical Pressure Vessel
KIc2
Material
Medium carbon (1040) steel
4140 steel
(tempered @ 482C)
Ti-5Al-2.5Sn titanium
2024 aluminum (T3)
4340 steel
(tempered @ 425C)
17-7PH steel
AZ31B magnesium
Ti-6Al-4V titanium
4140 steel
(tempered @ 370C)
4340 steel
(tempered @ 260C)
7075 aluminum (T651)
Sy
(MPa-m)
11.2
6.1
5.8
5.6
5.4
4.4
3.9
3.3
2.4
1.5
1.2
8.6 IMPACT FRACTURE TESTING
Prior to the advent of fracture mechanics as a scientific discipline, impact testing
techniques were established so as to ascertain the fracture characteristics of materials. It was realized that the results of laboratory tensile tests could not be extrapolated to predict fracture behavior; for example, under some circumstances normally
ductile metals fracture abruptly and with very little plastic deformation. Impact test
conditions were chosen to represent those most severe relative to the potential for
fracture—namely, (1) deformation at a relatively low temperature, (2) a high strain
rate (i.e., rate of deformation), and (3) a triaxial stress state (which may be introduced by the presence of a notch).
Impact Testing Techniques
Charpy, Izod tests
impact energy
Two standardized tests,4 the Charpy and Izod, were designed and are still used to
measure the impact energy, sometimes also termed notch toughness. The Charpy
V-notch (CVN) technique is most commonly used in the United States. For both
Charpy and Izod, the specimen is in the shape of a bar of square cross section, into
which a V-notch is machined (Figure 8.12a). The apparatus for making V-notch
impact tests is illustrated schematically in Figure 8.12b. The load is applied as an
impact blow from a weighted pendulum hammer that is released from a cocked
position at a fixed height h. The specimen is positioned at the base as shown. Upon
release, a knife edge mounted on the pendulum strikes and fractures the specimen
at the notch, which acts as a point of stress concentration for this high-velocity
4
ASTM Standard E 23, “Standard Test Methods for Notched Bar Impact Testing of Metallic
Materials.”
1496T_c08_207-251 11/18/05 11:50 Page 224
REVISED PAGES
224 • Chapter 8 / Failure
Figure 8.12
(a) Specimen used
for Charpy and Izod
impact tests. (b) A
schematic drawing of
an impact testing
apparatus. The
hammer is released
from fixed height h
and strikes the
specimen; the energy
expended in fracture
is reflected in the
difference between h
and the swing height
h¿. Specimen
placements for both
Charpy and Izod
tests are also shown.
[Figure (b) adapted
from H. W. Hayden,
W. G. Moffatt, and
J. Wulff, The
Structure and
Properties of
Materials, Vol. III,
Mechanical
Behavior, p. 13.
Copyright © 1965 by
John Wiley & Sons,
New York.
Reprinted by
permission of John
Wiley & Sons, Inc.]
8 mm
(0.32 in.)
(a)
10 mm
(0.39 in.)
10 mm
(0.39 in.)
Notch
Scale
Charpy
Izod
Pointer
Starting position
Hammer
End of swing
h
Specimen
h'
Anvil
(b)
impact blow. The pendulum continues its swing, rising to a maximum height h¿,
which is lower than h. The energy absorption, computed from the difference between h and h¿, is a measure of the impact energy. The primary difference between
the Charpy and Izod techniques lies in the manner of specimen support, as illustrated in Figure 8.12b. Furthermore, these are termed impact tests in light of the
manner of load application. Variables including specimen size and shape as well as
notch configuration and depth influence the test results.
1496T_c08_207-251 11/18/05 11:50 Page 225
REVISED PAGES
8.6 Impact Fracture Testing • 225
Both plane strain fracture toughness and these impact tests determine the fracture properties of materials. The former are quantitative in nature, in that a specific
property of the material is determined (i.e., KIc). The results of the impact tests, on
the other hand, are more qualitative and are of little use for design purposes. Impact
energies are of interest mainly in a relative sense and for making comparisons—
absolute values are of little significance. Attempts have been made to correlate
plane strain fracture toughnesses and CVN energies, with only limited success. Plane
strain fracture toughness tests are not as simple to perform as impact tests; furthermore, equipment and specimens are more expensive.
Ductile-to-Brittle Transition
Temperature (°F)
–40
0
40
80
120
160
200
240
280
100
A
80
100
Impact
energy
60
80
Shear
fracture
60
40
B
40
20
20
0
0
–40
–20
0
20
40
60
Temperature (°C)
80
100
120
140
Shear fracture (%)
Figure 8.13
Temperature
dependence of the
Charpy V-notch
impact energy
(curve A) and
percent shear
fracture (curve B)
for an A283 steel.
(Reprinted from
Welding Journal.
Used by permission
of the American
Welding Society.)
One of the primary functions of Charpy and Izod tests is to determine whether or
not a material experiences a ductile-to-brittle transition with decreasing temperature and, if so, the range of temperatures over which it occurs. The ductile-to-brittle
transition is related to the temperature dependence of the measured impact energy
absorption. This transition is represented for a steel by curve A in Figure 8.13. At
higher temperatures the CVN energy is relatively large, in correlation with a ductile mode of fracture. As the temperature is lowered, the impact energy drops suddenly over a relatively narrow temperature range, below which the energy has a
constant but small value; that is, the mode of fracture is brittle.
Alternatively, appearance of the failure surface is indicative of the nature of
fracture and may be used in transition temperature determinations. For ductile fracture this surface appears fibrous or dull (or of shear character), as in the steel specimen of Figure 8.14 that was tested at 79C. Conversely, totally brittle surfaces have
a granular (shiny) texture (or cleavage character) (the 59C specimen, Figure 8.14).
Over the ductile-to-brittle transition, features of both types will exist (in Figure 8.14,
displayed by specimens tested at 12C, 4C, 16C, and 24C). Frequently, the percent shear fracture is plotted as a function of temperature—curve B in Figure 8.13.
Impact energy (J)
ductile-to-brittle
transition
1496T_c08_207-251 11/18/05 11:50 Page 226
REVISED PAGES
226 • Chapter 8 / Failure
59
12
4
16
24
79
Figure 8.14 Photograph of fracture surfaces of A36 steel Charpy V-notch
specimens tested at indicated temperatures (in C). (From R. W. Hertzberg,
Deformation and Fracture Mechanics of Engineering Materials, 3rd edition,
Fig. 9.6, p. 329. Copyright © 1989 by John Wiley & Sons, Inc., New York.
Reprinted by permission of John Wiley & Sons, Inc.)
For many alloys there is a range of temperatures over which the ductile-to-brittle
transition occurs (Figure 8.13); this presents some difficulty in specifying a single
ductile-to-brittle transition temperature. No explicit criterion has been established,
and so this temperature is often defined as that temperature at which the CVN
energy assumes some value (e.g., 20 J or 15 ft-lbf), or corresponding to some given
fracture appearance (e.g., 50% fibrous fracture). Matters are further complicated
inasmuch as a different transition temperature may be realized for each of these
criteria. Perhaps the most conservative transition temperature is that at which the
fracture surface becomes 100% fibrous; on this basis, the transition temperature is
approximately 110C (230F) for the steel alloy that is the subject of Figure 8.13.
Structures constructed from alloys that exhibit this ductile-to-brittle behavior
should be used only at temperatures above the transition temperature, to avoid brittle and catastrophic failure. Classic examples of this type of failure occurred, with disastrous consequences, during World War II when a number of welded transport ships,
away from combat, suddenly and precipitously split in half.The vessels were constructed
of a steel alloy that possessed adequate ductility according to room-temperature tensile tests.The brittle fractures occurred at relatively low ambient temperatures, at about
4C (40F), in the vicinity of the transition temperature of the alloy. Each fracture
crack originated at some point of stress concentration, probably a sharp corner or fabrication defect, and then propagated around the entire girth of the ship.
In addition to the ductile-to-brittle transition represented in Figure 8.13, two other
general types of impact energy-versus-temperature behavior have been observed; these
are represented schematically by the upper and lower curves of Figure 8.15. Here it
Impact energy
Low-strength (FCC and HCP) metals
Low-strength steels (BCC)
High-strength materials
Temperature
Figure 8.15 Schematic curves for
the three general types of impact
energy-versus-temperature behavior.
1496T_c08_207-251 11/18/05 13:43 Page 227
REVISED PAGES
8.6 Impact Fracture Testing • 227
Temperature (°F)
0
200
–200
400
240
300
0.01
0.11
160
200
0.22
120
0.31
0.43
80
100
Impact energy (ft-lbf)
200
Impact energy (J)
Figure 8.16
Influence of carbon
content on the
Charpy V-notch
energy-versustemperature
behavior for steel.
(Reprinted with
permission from
ASM International,
Metals Park, OH
44073-9989, USA;
J. A. Reinbolt and
W. J. Harris, Jr.,
“Effect of Alloying
Elements on Notch
Toughness of
Pearlitic Steels,”
Transactions of ASM,
Vol. 43, 1951.)
0.53
0.63
40
0.67
0
–200
–100
0
100
Temperature (°C)
200
0
may be noted that low-strength FCC metals (some aluminum and copper alloys)
and most HCP metals do not experience a ductile-to-brittle transition (corresponding to the upper curve of Figure 8.15), and retain high impact energies (i.e., remain
ductile) with decreasing temperature. For high-strength materials (e.g., high-strength
steels and titanium alloys), the impact energy is also relatively insensitive to temperature (the lower curve of Figure 8.15); however, these materials are also very
brittle, as reflected by their low impact energy values. And, of course, the characteristic ductile-to-brittle transition is represented by the middle curve of Figure 8.15.
As noted, this behavior is typically found in low-strength steels that have the BCC
crystal structure.
For these low-strength steels, the transition temperature is sensitive to both alloy
composition and microstructure. For example, decreasing the average grain size results in a lowering of the transition temperature. Hence, refining the grain size both
strengthens (Section 7.8) and toughens steels. In contrast, increasing the carbon content, while increasing the strength of steels, also raises the CVN transition of steels,
as indicated in Figure 8.16.
Most ceramics and polymers also experience a ductile-to-brittle transition. For
ceramic materials, the transition occurs only at elevated temperatures, ordinarily
in excess of 1000C (1850F). This behavior as related to polymers is discussed in
Section 15.6.
Fa t i g u e
fatigue
Fatigue is a form of failure that occurs in structures subjected to dynamic and fluctuating stresses (e.g., bridges, aircraft, and machine components). Under these circumstances it is possible for failure to occur at a stress level considerably lower than
the tensile or yield strength for a static load. The term “fatigue” is used because this
type of failure normally occurs after a lengthy period of repeated stress or strain
cycling. Fatigue is important inasmuch as it is the single largest cause of failure in metals, estimated to comprise approximately 90% of all metallic failures; polymers and
1496T_c08_207-251 11/18/05 11:50 Page 228
REVISED PAGES
228 • Chapter 8 / Failure
ceramics (except for glasses) are also susceptible to this type of failure. Furthermore,
fatigue is catastrophic and insidious, occurring very suddenly and without warning.
Fatigue failure is brittlelike in nature even in normally ductile metals, in that
there is very little, if any, gross plastic deformation associated with failure. The
process occurs by the initiation and propagation of cracks, and ordinarily the fracture surface is perpendicular to the direction of an applied tensile stress.
8.7 CYCLIC STRESSES
max
+
0
–
Stress
Compression
Tension
The applied stress may be axial (tension-compression), flexural (bending), or torsional
(twisting) in nature. In general, three different fluctuating stress–time modes are possible. One is represented schematically by a regular and sinusoidal time dependence
in Figure 8.17a, wherein the amplitude is symmetrical about a mean zero stress level,
for example, alternating from a maximum tensile stress (smax) to a minimum compressive stress (smin) of equal magnitude; this is referred to as a reversed stress cycle.
min
Time
(a)
a
r
+
m
0
–
min
Time
(b)
+
–
Stress
Compression
Tension
Stress
Compression
Tension
max
Time
(c)
Figure 8.17 Variation
of stress with time that
accounts for fatigue
failures. (a) Reversed
stress cycle, in which the
stress alternates from a
maximum tensile stress
(+) to a maximum
compressive stress ()
of equal magnitude. (b)
Repeated stress cycle, in
which maximum and
minimum stresses are
asymmetrical relative to
the zero-stress level;
mean stress sm, range
of stress sr, and stress
amplitude sa are
indicated. (c) Random
stress cycle.
1496T_c08_207-251 11/18/05 13:43 Page 229
REVISED PAGES
8.8 The S–N Curve • 229
Another type, termed repeated stress cycle, is illustrated in Figure 8.17b; the maxima
and minima are asymmetrical relative to the zero stress level. Finally, the stress level
may vary randomly in amplitude and frequency, as exemplified in Figure 8.17c.
Also indicated in Figure 8.17b are several parameters used to characterize the
fluctuating stress cycle. The stress amplitude alternates about a mean stress sm,
defined as the average of the maximum and minimum stresses in the cycle, or
Mean stress for cyclic
loading—dependence
on maximum and
minimum stress
levels
Computation of
range of stress for
cyclic loading
sm
smax smin
2
(8.14)
Furthermore, the range of stress sr is just the difference between smax and smin —
namely,
sr smax smin
(8.15)
Stress amplitude sa is just one half of this range of stress, or
Computation of
stress amplitude for
cyclic loading
sa
smax smin
sr
2
2
(8.16)
Finally, the stress ratio R is just the ratio of minimum and maximum stress amplitudes:
Computation of
stress ratio
R
smin
smax
(8.17)
By convention, tensile stresses are positive and compressive stresses are negative.
For example, for the reversed stress cycle, the value of R is 1.
Concept Check 8.2
Make a schematic sketch of a stress-versus-time plot for the situation when the
stress ratio R has a value of 1.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Concept Check 8.3
Using Equations 8.16 and 8.17, demonstrate that increasing the value of the stress
ratio R produces a decrease in stress amplitude sa.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
8.8 THE S–N CURVE
As with other mechanical characteristics, the fatigue properties of materials can be
determined from laboratory simulation tests.5 A test apparatus should be designed
to duplicate as nearly as possible the service stress conditions (stress level, time
frequency, stress pattern, etc.). A schematic diagram of a rotating-bending test
5
See ASTM Standard E 466, “Standard Practice for Conducting Constant Amplitude
Axial Fatigue Tests of Metallic Materials,” and ASTM Standard E 468, “Standard Practice
for Presentation of Constant Amplitude Fatigue Test Results for Metallic Materials.”
1496T_c08_207-251 11/18/05 13:43 Page 230
REVISED PAGES
230 • Chapter 8 / Failure
Flexible coupling
Counter
Specimen
–
Bearing housing
ing
g hous
Bearin
High-speed
motor
+
Load
Load
Figure 8.18 Schematic diagram of fatigue-testing apparatus for making rotatingbending tests. (From Keyser, Materials Science in Engineering, 4th Edition, © 1986,
p. 88. Adapted by permission of Pearson Education, Inc., Upper Saddle River, NJ.)
fatigue limit
fatigue strength
fatigue life
apparatus, commonly used for fatigue testing, is shown in Figure 8.18; the compression and tensile stresses are imposed on the specimen as it is simultaneously
bent and rotated. Tests are also frequently conducted using an alternating uniaxial
tension-compression stress cycle.
A series of tests are commenced by subjecting a specimen to the stress cycling
at a relatively large maximum stress amplitude (smax), usually on the order of twothirds of the static tensile strength; the number of cycles to failure is counted. This
procedure is repeated on other specimens at progressively decreasing maximum
stress amplitudes. Data are plotted as stress S versus the logarithm of the number
N of cycles to failure for each of the specimens. The values of S are normally
taken as stress amplitudes (sa, Equation 8.16); on occasion, smax or smin values may
be used.
Two distinct types of S–N behavior are observed, which are represented
schematically in Figure 8.19. As these plots indicate, the higher the magnitude of
the stress, the smaller the number of cycles the material is capable of sustaining
before failure. For some ferrous (iron base) and titanium alloys, the S–N curve
(Figure 8.19a) becomes horizontal at higher N values; or there is a limiting stress
level, called the fatigue limit (also sometimes the endurance limit), below which
fatigue failure will not occur. This fatigue limit represents the largest value of fluctuating stress that will not cause failure for essentially an infinite number of cycles.
For many steels, fatigue limits range between 35% and 60% of the tensile strength.
Most nonferrous alloys (e.g., aluminum, copper, magnesium) do not have a
fatigue limit, in that the S–N curve continues its downward trend at increasingly
greater N values (Figure 8.19b). Thus, fatigue will ultimately occur regardless of the
magnitude of the stress. For these materials, the fatigue response is specified as
fatigue strength, which is defined as the stress level at which failure will occur for
some specified number of cycles (e.g., 107 cycles). The determination of fatigue
strength is also demonstrated in Figure 8.19b.
Another important parameter that characterizes a material’s fatigue behavior
is fatigue life Nf. It is the number of cycles to cause failure at a specified stress level,
as taken from the S–N plot (Figure 8.19b).
Unfortunately, there always exists considerable scatter in fatigue data—that is,
a variation in the measured N value for a number of specimens tested at the same
stress level. This variation may lead to significant design uncertainties when fatigue
life and/or fatigue limit (or strength) are being considered. The scatter in results is
a consequence of the fatigue sensitivity to a number of test and material parameters that are impossible to control precisely. These parameters include specimen fabrication and surface preparation, metallurgical variables, specimen alignment in the
apparatus, mean stress, and test frequency.
1496T_c08_207-251 12/21/05 7:51 Page 231
2nd REVISE PAGES
Fatigue
limit
103
104
105
106
107
108
109
1010
109
1010
Cycles to failure, N
(logarithmic scale)
(a)
Stress amplitude, S
Figure 8.19 Stress
amplitude (S) versus
logarithm of the
number of cycles to
fatigue failure (N)
for (a) a material
that displays a
fatigue limit, and
(b) a material that
does not display a
fatigue limit.
Stress amplitude, S
8.8 The S–N Curve • 231
S1
Fatigue strength
at N1 cycles
103
104
107
Fatigue life
at stress S1
N1 108
Cycles to failure, N
(logarithmic scale)
(b)
Fatigue S–N curves similar to those shown in Figure 8.19 represent “best fit”
curves that have been drawn through average-value data points. It is a little unsettling to realize that approximately one-half of the specimens tested actually failed
at stress levels lying nearly 25% below the curve (as determined on the basis of
statistical treatments).
Several statistical techniques have been developed to specify fatigue life and
fatigue limit in terms of probabilities. One convenient way of representing data
treated in this manner is with a series of constant probability curves, several of which
are plotted in Figure 8.20. The P value associated with each curve represents the
probability of failure. For example, at a stress of 200 MPa (30,000 psi), we would
expect 1% of the specimens to fail at about 106 cycles and 50% to fail at about
2 107 cycles, and so on. Remember that S–N curves represented in the literature
are normally average values, unless noted otherwise.
The fatigue behaviors represented in Figures 8.19a and 8.19b may be classified
into two domains. One is associated with relatively high loads that produce not only
elastic strain but also some plastic strain during each cycle. Consequently, fatigue
lives are relatively short; this domain is termed low-cycle fatigue and occurs at less
1496T_c08_207-251 11/18/05 11:50 Page 232
REVISED PAGES
232 • Chapter 8 / Failure
70
60
P = 0.99
50
P = 0.90
300
P = 0.50
40
P = 0.01
200
30
P = 0.10
Stress (103 psi)
400
Stress, S (MPa)
Figure 8.20 Fatigue
S–N probability of
failure curves for a
7075-T6 aluminum
alloy; P denotes the
probability of
failure. (From G. M.
Sinclair and T. J.
Dolan, Trans. ASME,
75, 1953, p. 867.
Reprinted with
permission of the
American Society of
Mechanical
Engineers.)
20
100
10
104
105
106
107
108
109
Cycles to failure, N
(logarithmic scale)
than about 104 to 105 cycles. For lower stress levels wherein deformations are totally
elastic, longer lives result. This is called high-cycle fatigue inasmuch as relatively
large numbers of cycles are required to produce fatigue failure. High-cycle fatigue
is associated with fatigue lives greater than about 104 to 105 cycles.
8.9 CRACK INITIATION AND PROPAGATION
The process of fatigue failure is characterized by three distinct steps: (1) crack
initiation, wherein a small crack forms at some point of high stress concentration;
(2) crack propagation, during which this crack advances incrementally with each
stress cycle; and (3) final failure, which occurs very rapidly once the advancing crack
has reached a critical size. Cracks associated with fatigue failure almost always initiate (or nucleate) on the surface of a component at some point of stress concentration. Crack nucleation sites include surface scratches, sharp fillets, keyways,
threads, dents, and the like. In addition, cyclic loading can produce microscopic surface discontinuities resulting from dislocation slip steps that may also act as stress
raisers, and therefore as crack initiation sites.
The region of a fracture surface that formed during the crack propagation step
may be characterized by two types of markings termed beachmarks and striations.
Both of these features indicate the position of the crack tip at some point in time
and appear as concentric ridges that expand away from the crack initiation site(s),
frequently in a circular or semicircular pattern. Beachmarks (sometimes also called
“clamshell marks”) are of macroscopic dimensions (Figure 8.21), and may be
observed with the unaided eye. These markings are found for components that
experienced interruptions during the crack propagation stage—for example, a
machine that operated only during normal work-shift hours. Each beachmark band
represents a period of time over which crack growth occurred.
On the other hand, fatigue striations are microscopic in size and subject to
observation with the electron microscope (either TEM or SEM). Figure 8.22 is an
electron fractograph that shows this feature. Each striation is thought to represent
the advance distance of a crack front during a single load cycle. Striation width
depends on, and increases with, increasing stress range.
At this point it should be emphasized that although both beachmarks and
striations are fatigue fracture surface features having similar appearances, they are
1496T_c08_207-251 12/21/05 7:51 Page 233
2nd REVISE PAGES
8.9 Crack Initiation and Propagation • 233
Figure 8.21 Fracture surface
of a rotating steel shaft that
experienced fatigue failure.
Beachmark ridges are
visible in the photograph.
(Reproduced with
permission from D. J.
Wulpi, Understanding How
Components Fail, American
Society for Metals, Materials
Park, OH, 1985.)
nevertheless different, both in origin and size. There may be literally thousands of
striations within a single beachmark.
Often the cause of failure may be deduced after examination of the failure surfaces. The presence of beachmarks and/or striations on a fracture surface confirms
that the cause of failure was fatigue. Nevertheless, the absence of either or both
does not exclude fatigue as the cause of failure.
One final comment regarding fatigue failure surfaces: Beachmarks and striations will not appear on that region over which the rapid failure occurs. Rather, the
Figure 8.22 Transmission
electron fractograph showing
fatigue striations in
aluminum. Magnification
unknown. (From V. J.
Colangelo and F. A. Heiser,
Analysis of Metallurgical
Failures, 2nd edition.
Copyright © 1987 by John
Wiley & Sons, New York.
Reprinted by permission of
John Wiley & Sons, Inc.)
1496T_c08_207-251 11/18/05 11:50 Page 234
REVISED PAGES
234 • Chapter 8 / Failure
Region of slow
crack propagation
Figure 8.23 Fatigue
failure surface. A crack
formed at the top edge.
The smooth region also
near the top corresponds
to the area over which the
crack propagated slowly.
Rapid failure occurred
over the area having a
dull and fibrous texture
(the largest area).
Approximately 0.5.
[Reproduced by
permission from Metals
Handbook: Fractography
and Atlas of Fractographs,
Vol. 9, 8th edition, H. E.
Boyer (Editor), American
Society for Metals, 1974.]
Region of rapid failure
rapid failure may be either ductile or brittle; evidence of plastic deformation will
be present for ductile, and absent for brittle, failure. This region of failure may be
noted in Figure 8.23.
Concept Check 8.4
Surfaces for some steel specimens that have failed by fatigue have a bright crystalline or grainy appearance. Laymen may explain the failure by saying that the
metal crystallized while in service. Offer a criticism for this explanation.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
8.10 FACTORS THAT AFFECT FATIGUE LIFE
As mentioned in Section 8.8, the fatigue behavior of engineering materials is highly
sensitive to a number of variables. Some of these factors include mean stress level,
geometrical design, surface effects, and metallurgical variables, as well as the environment. This section is devoted to a discussion of these factors and, in addition,
to measures that may be taken to improve the fatigue resistance of structural
components.
Mean Stress
The dependence of fatigue life on stress amplitude is represented on the S–N plot.
Such data are taken for a constant mean stress sm, often for the reversed cycle
1496T_c08_207-251 11/18/05 11:50 Page 235
REVISED PAGES
8.10 Factors That Affect Fatigue Life • 235
Figure 8.24 Demonstration of the
influence of mean stress sm on S–N fatigue
behavior.
Stress amplitude, a
m3 > m2 > m1
m1
m2
m3
Cycles to failure, N
(logarithmic scale)
situation (sm 0). Mean stress, however, will also affect fatigue life; this influence
may be represented by a series of S–N curves, each measured at a different sm, as
depicted schematically in Figure 8.24. As may be noted, increasing the mean stress
level leads to a decrease in fatigue life.
Surface Effects
For many common loading situations, the maximum stress within a component or
structure occurs at its surface. Consequently, most cracks leading to fatigue failure
originate at surface positions, specifically at stress amplification sites. Therefore, it
has been observed that fatigue life is especially sensitive to the condition and configuration of the component surface. Numerous factors influence fatigue resistance,
the proper management of which will lead to an improvement in fatigue life. These
include design criteria as well as various surface treatments.
Design Factors
The design of a component can have a significant influence on its fatigue characteristics. Any notch or geometrical discontinuity can act as a stress raiser and
fatigue crack initiation site; these design features include grooves, holes, keyways,
threads, and so on. The sharper the discontinuity (i.e., the smaller the radius of
curvature), the more severe the stress concentration. The probability of fatigue
failure may be reduced by avoiding (when possible) these structural irregularities, or by making design modifications whereby sudden contour changes leading
to sharp corners are eliminated—for example, calling for rounded fillets with large
radii of curvature at the point where there is a change in diameter for a rotating
shaft (Figure 8.25).
Fillet
(a)
( b)
Figure 8.25 Demonstration of how design
can reduce stress amplification. (a) Poor
design: sharp corner. (b) Good design:
fatigue lifetime improved by incorporating
rounded fillet into a rotating shaft at the
point where there is a change in diameter.
1496T_c08_207-251 11/18/05 13:54 Page 236
REVISED PAGES
236 • Chapter 8 / Failure
Figure 8.26 Schematic S–N fatigue curves
for normal and shot-peened steel.
Stress amplitude
Shot peened
Normal
Cycles to failure
(logarithmic scale)
Surface Treatments
case hardening
Two Case Studies:
“Automobile Valve
Spring,” and “Failure
of an Automobile
Rear Axle,” Chapter
22, may be found at
www.wiley.com/
college/callister
(Student Companion
Site).
During machining operations, small scratches and grooves are invariably introduced
into the workpiece surface by cutting tool action. These surface markings can limit
the fatigue life. It has been observed that improving the surface finish by polishing
will enhance fatigue life significantly.
One of the most effective methods of increasing fatigue performance is by
imposing residual compressive stresses within a thin outer surface layer. Thus, a
surface tensile stress of external origin will be partially nullified and reduced in
magnitude by the residual compressive stress. The net effect is that the likelihood
of crack formation and therefore of fatigue failure is reduced.
Residual compressive stresses are commonly introduced into ductile metals mechanically by localized plastic deformation within the outer surface region. Commercially, this is often accomplished by a process termed shot peening. Small, hard
particles (shot) having diameters within the range of 0.1 to 1.0 mm are projected
at high velocities onto the surface to be treated. The resulting deformation induces
compressive stresses to a depth of between one-quarter and one-half of the shot
diameter. The influence of shot peening on the fatigue behavior of steel is demonstrated schematically in Figure 8.26.
Case hardening is a technique by which both surface hardness and fatigue
life are enhanced for steel alloys. This is accomplished by a carburizing or nitriding process whereby a component is exposed to a carbonaceous or nitrogenous
atmosphere at an elevated temperature. A carbon- or nitrogen-rich outer surface
layer (or “case”) is introduced by atomic diffusion from the gaseous phase. The
case is normally on the order of 1 mm deep and is harder than the inner core of
material. (The influence of carbon content on hardness for Fe–C alloys is demonstrated in Figure 10.29a.) The improvement of fatigue properties results from
increased hardness within the case, as well as the desired residual compressive
stresses the formation of which attends the carburizing or nitriding process. A
carbon-rich outer case may be observed for the gear shown in the chapter-opening
photograph for Chapter 5; it appears as a dark outer rim within the sectioned
segment. The increase in case hardness is demonstrated in the photomicrograph
appearing in Figure 8.27. The dark and elongated diamond shapes are Knoop
microhardness indentations. The upper indentation, lying within the carburized
layer, is smaller than the core indentation.
1496T_c08_207-251 11/18/05 11:50 Page 237
REVISED PAGES
8.11 Environmental Effects • 237
Case
Core
region
Figure 8.27 Photomicrograph showing
both core (bottom) and carburized
outer case (top) regions of a casehardened steel. The case is harder as
attested by the smaller microhardness
indentation. 100. (From R. W.
Hertzberg, Deformation and Fracture
Mechanics of Engineering Materials,
3rd edition. Copyright © 1989 by John
Wiley & Sons, New York. Reprinted by
permission of John Wiley & Sons, Inc.)
8.11 ENVIRONMENTAL EFFECTS
thermal fatigue
Thermal stress—
dependence on
coefficient of
thermal expansion,
modulus of elasticity,
and temperature
change
corrosion fatigue
Environmental factors may also affect the fatigue behavior of materials. A few brief
comments will be given relative to two types of environment-assisted fatigue failure: thermal fatigue and corrosion fatigue.
Thermal fatigue is normally induced at elevated temperatures by fluctuating
thermal stresses; mechanical stresses from an external source need not be present.
The origin of these thermal stresses is the restraint to the dimensional expansion
and/or contraction that would normally occur in a structural member with variations in temperature. The magnitude of a thermal stress developed by a temperature change ¢T is dependent on the coefficient of thermal expansion al and the
modulus of elasticity E according to
s alE ¢T
(8.18)
(The topics of thermal expansion and thermal stresses are discussed in Sections 19.3
and 19.5.) Of course, thermal stresses will not arise if this mechanical restraint is
absent. Therefore, one obvious way to prevent this type of fatigue is to eliminate,
or at least reduce, the restraint source, thus allowing unhindered dimensional
changes with temperature variations, or to choose materials with appropriate physical properties.
Failure that occurs by the simultaneous action of a cyclic stress and chemical
attack is termed corrosion fatigue. Corrosive environments have a deleterious influence and produce shorter fatigue lives. Even the normal ambient atmosphere will
affect the fatigue behavior of some materials. Small pits may form as a result of
chemical reactions between the environment and material, which serve as points of
stress concentration and therefore as crack nucleation sites. In addition, crack propagation rate is enhanced as a result of the corrosive environment. The nature of the
stress cycles will influence the fatigue behavior; for example, lowering the load application frequency leads to longer periods during which the opened crack is in contact with the environment and to a reduction in the fatigue life.
Several approaches to corrosion fatigue prevention exist. On one hand, we
can take measures to reduce the rate of corrosion by some of the techniques
discussed in Chapter 17—for example, apply protective surface coatings, select a
1496T_c08_207-251 11/18/05 11:50 Page 238
REVISED PAGES
238 • Chapter 8 / Failure
more corrosion-resistant material, and reduce the corrosiveness of the environment. And/or it might be advisable to take actions to minimize the probability
of normal fatigue failure, as outlined above—for example, reduce the applied
tensile stress level and impose residual compressive stresses on the surface of
the member.
Creep
creep
Materials are often placed in service at elevated temperatures and exposed to static
mechanical stresses (e.g., turbine rotors in jet engines and steam generators that
experience centrifugal stresses, and high-pressure steam lines). Deformation under
such circumstances is termed creep. Defined as the time-dependent and permanent
deformation of materials when subjected to a constant load or stress, creep is normally an undesirable phenomenon and is often the limiting factor in the lifetime of
a part. It is observed in all materials types; for metals it becomes important only
for temperatures greater than about 0.4Tm (Tm absolute melting temperature).
Amorphous polymers, which include plastics and rubbers, are especially sensitive
to creep deformation as discussed in Section 15.4.
8.12 GENERALIZED CREEP BEHAVIOR
A typical creep test6 consists of subjecting a specimen to a constant load or stress
while maintaining the temperature constant; deformation or strain is measured and
plotted as a function of elapsed time. Most tests are the constant load type, which
yield information of an engineering nature; constant stress tests are employed to
provide a better understanding of the mechanisms of creep.
Figure 8.28 is a schematic representation of the typical constant load creep behavior of metals. Upon application of the load there is an instantaneous deformation, as indicated in the figure, which is mostly elastic. The resulting creep curve
consists of three regions, each of which has its own distinctive strain–time feature.
Primary or transient creep occurs first, typified by a continuously decreasing creep
rate; that is, the slope of the curve diminishes with time. This suggests that the material is experiencing an increase in creep resistance or strain hardening (Section
7.10)—deformation becomes more difficult as the material is strained. For secondary
creep, sometimes termed steady-state creep, the rate is constant; that is, the plot becomes linear. This is often the stage of creep that is of the longest duration. The
constancy of creep rate is explained on the basis of a balance between the competing processes of strain hardening and recovery, recovery (Section 7.11) being the
process whereby a material becomes softer and retains its ability to experience
deformation. Finally, for tertiary creep, there is an acceleration of the rate and ultimate failure. This failure is frequently termed rupture and results from microstructural and/or metallurgical changes; for example, grain boundary separation, and the
formation of internal cracks, cavities, and voids. Also, for tensile loads, a neck may
form at some point within the deformation region. These all lead to a decrease in
the effective cross-sectional area and an increase in strain rate.
6
ASTM Standard E 139, “Standard Practice for Conducting Creep, Creep-Rupture, and
Stress-Rupture Tests of Metallic Materials.”
1496T_c08_207-251 11/18/05 13:43 Page 239
REVISED PAGES
8.13 Stress and Temperature Effects • 239
Creep strain, ⑀
Rupture
Primary
Δt
Δ⑀
Tertiary
Figure 8.28 Typical creep curve of strain
versus time at constant stress and constant
elevated temperature. The minimum creep
rate ¢ ¢t is the slope of the linear
segment in the secondary region. Rupture
lifetime tr is the total time to rupture.
Secondary
Instantaneous deformation
Time, t
tr
For metallic materials most creep tests are conducted in uniaxial tension using
a specimen having the same geometry as for tensile tests (Figure 6.2). On the other
hand, uniaxial compression tests are more appropriate for brittle materials; these
provide a better measure of the intrinsic creep properties inasmuch as there is no
stress amplification and crack propagation, as with tensile loads. Compressive test
specimens are usually right cylinders or parallelepipeds having length-to-diameter
ratios ranging from about 2 to 4. For most materials creep properties are virtually
independent of loading direction.
Possibly the most important parameter from a creep test is the slope of the secondary portion of the creep curve ( ¢ ¢t in Figure 8.28); this is often called the
#
minimum or steady-state creep rate s. It is the engineering design parameter that is
considered for long-life applications, such as a nuclear power plant component that
is scheduled to operate for several decades, and when failure or too much strain
are not options. On the other hand, for many relatively short-life creep situations
(e.g., turbine blades in military aircraft and rocket motor nozzles), time to rupture,
or the rupture lifetime tr , is the dominant design consideration; it is also indicated
in Figure 8.28. Of course, for its determination, creep tests must be conducted to
the point of failure; these are termed creep rupture tests. Thus, a knowledge of these
creep characteristics of a material allows the design engineer to ascertain its suitability for a specific application.
Concept Check 8.5
Superimpose on the same strain-versus-time plot schematic creep curves for both constant tensile stress and constant tensile load, and explain the differences in behavior.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
8.13 STRESS AND TEMPERATURE EFFECTS
Both temperature and the level of the applied stress influence the creep characteristics (Figure 8.29). At a temperature substantially below 0.4Tm, and after the
initial deformation, the strain is virtually independent of time. With either increasing stress or temperature, the following will be noted: (1) the instantaneous strain
at the time of stress application increases, (2) the steady-state creep rate is increased,
and (3) the rupture lifetime is diminished.
1496T_c08_207-251 11/18/05 11:50 Page 240
REVISED PAGES
240 • Chapter 8 / Failure
Figure 8.29 Influence of stress s and
temperature T on creep behavior.
T3 > T2 > T1
3 > 2 > 1
Creep strain
T3 or 3
T2 or 2
T1 or 1
T < 0.4Tm
Time
Dependence of
creep strain rate
on stress
The results of creep rupture tests are most commonly presented as the logarithm of stress versus the logarithm of rupture lifetime. Figure 8.30 is one such plot
for a nickel alloy in which a linear relationship can be seen to exist at each temperature. For some alloys and over relatively large stress ranges, nonlinearity in
these curves is observed.
Empirical relationships have been developed in which the steady-state creep
rate as a function of stress and temperature is expressed. Its dependence on stress
can be written
#
s K1sn
(8.19)
#
where K1 and n are material constants. A plot of the logarithm of s versus the logarithm of s yields a straight line with slope of n; this is shown in Figure 8.31 for a
nickel alloy at three temperatures. Clearly, a straight line segment is drawn at each
temperature.
Now, when the influence of temperature is included,
Dependence of
creep strain rate
on stress and
temperature (in K)
Qc
#
s K2 sn exp a
b
RT
(8.20)
where K2 and Qc are constants; Qc is termed the activation energy for creep.
60
400
300
40
30
200
20
100
80
538°C (1000°F)
10
8
60
649°C (1200°F)
40
6
30
4
20
3
2
102
103
104
Rupture lifetime (h)
105
Stress (103 psi)
427°C (800°F)
Stress (MPa)
Figure 8.30 Stress (logarithmic scale) versus
rupture lifetime (logarithmic scale) for a low
carbon–nickel alloy at three temperatures.
[From Metals Handbook: Properties and
Selection: Stainless Steels, Tool Materials and
Special-Purpose Metals, Vol. 3, 9th edition,
D. Benjamin (Senior Editor), American Society
for Metals, 1980, p. 130.]
1496T_c08_207-251 11/18/05 11:50 Page 241
REVISED PAGES
8.14 Data Extrapolation Methods • 241
30
200
10
8
427°C (800°F)
60
40
6
30
4
538°C (1000°F)
3
20
Stress (103 psi)
20
100
80
Stress (MPa)
Figure 8.31 Stress (logarithmic scale) versus
steady-state creep rate (logarithmic scale) for a
low carbon–nickel alloy at three temperatures.
[From Metals Handbook: Properties and
Selection: Stainless Steels, Tool Materials and
Special-Purpose Metals, Vol. 3, 9th edition,
D. Benjamin (Senior Editor), American Society
for Metals, 1980, p. 131.]
2
10
8
649°C (1200°F)
10–7
10–6
10–5
1
Steady-state creep rate (h–1)
Several theoretical mechanisms have been proposed to explain the creep behavior for various materials; these mechanisms involve stress-induced vacancy diffusion, grain boundary diffusion, dislocation motion, and grain boundary sliding.
Each leads to a different value of the stress exponent n in Equation 8.19. It has
been possible to elucidate the creep mechanism for a particular material by comparing its experimental n value with values predicted for the various mechanisms.
In addition, correlations have been made between the activation energy for creep
(Qc) and the activation energy for diffusion (Qd, Equation 5.8).
Creep data of this nature are represented pictorially for some well-studied systems in the form of stress–temperature diagrams, which are termed deformation
mechanism maps. These maps indicate stress–temperature regimes (or areas) over
which various mechanisms operate. Constant strain rate contours are often also included. Thus, for some creep situation, given the appropriate deformation mechanism map and any two of the three parameters—temperature, stress level, and creep
strain rate—the third parameter may be determined.
8.14 DATA EXTRAPOLATION METHODS
The need often arises for engineering creep data that are impractical to collect from
normal laboratory tests. This is especially true for prolonged exposures (on the order
of years). One solution to this problem involves performing creep and/or creep rupture
tests at temperatures in excess of those required, for shorter time periods, and at a
comparable stress level, and then making a suitable extrapolation to the in-service
condition. A commonly used extrapolation procedure employs the Larson–Miller
parameter, defined as
The Larson–Miller
parameter—in terms
of temperature and
rupture lifetime
T 1C log tr 2
(8.21)
where C is a constant (usually on the order of 20), for T in Kelvin and the rupture
lifetime tr in hours. The rupture lifetime of a given material measured at some specific stress level will vary with temperature such that this parameter remains constant. Or, the data may be plotted as the logarithm of stress versus the Larson–Miller
parameter, as shown in Figure 8.32. Utilization of this technique is demonstrated in
the following design example.
1496T_c08_207-251 11/18/05 11:50 Page 242
REVISED PAGES
242 • Chapter 8 / Failure
103 T(20 + log tr)(°R–h)
25
30
35
40
45
Figure 8.32 Logarithm
stress versus the
Larson–Miller parameter
for an S-590 iron. (From
F. R. Larson and J. Miller,
Trans. ASME, 74, 765, 1952.
Reprinted by permission
of ASME.)
50
1000
100
10
Stress (103 psi)
Stress (MPa)
100
10
1
12
16
20
24
28
103 T(20 + log tr)(K–h)
DESIGN EXAMPLE 8.2
Rupture Lifetime Prediction
Using the Larson–Miller data for S-590 iron shown in Figure 8.32, predict the
time to rupture for a component that is subjected to a stress of 140 MPa (20,000
psi) at 800C 11073 K2.
Solution
From Figure 8.32, at 140 MPa (20,000 psi) the value of the Larson–Miller
parameter is 24.0 103, for T in K and tr in h; therefore,
24.0 103 T 120 log tr 2
1073120 log tr 2
and, solving for the time,
22.37 20 log tr
tr 233 h 19.7 days2
8.15 ALLOYS FOR HIGH-TEMPERATURE USE
There are several factors that affect the creep characteristics of metals. These
include melting temperature, elastic modulus, and grain size. In general, the higher
the melting temperature, the greater the elastic modulus, and the larger the grain
size, the better is a material’s resistance to creep. Relative to grain size, smaller
grains permit more grain-boundary sliding, which results in higher creep rates. This
1496T_c08_207-251 11/18/05 11:50 Page 243
REVISED PAGES
Summary • 243
Figure 8.33 (a) Polycrystalline turbine blade that
was produced by a conventional casting
technique. High-temperature creep resistance is
improved as a result of an oriented columnar
grain structure (b) produced by a sophisticated
directional solidification technique. Creep
resistance is further enhanced when single-crystal
blades (c) are used. (Courtesy of Pratt &
Whitney.)
(a)
Conventional casting
(b)
(c)
Columnar grain
Single crystal
effect may be contrasted to the influence of grain size on the mechanical behavior
at low temperatures [i.e., increase in both strength (Section 7.8) and toughness
(Section 8.6)].
Stainless steels (Section 11.2), the refractory metals, and the superalloys
(Section 11.3) are especially resilient to creep and are commonly employed in hightemperature service applications. The creep resistance of the cobalt and nickel
superalloys is enhanced by solid-solution alloying, and also by the addition of a dispersed phase that is virtually insoluble in the matrix. In addition, advanced processing
techniques have been utilized; one such technique is directional solidification, which
produces either highly elongated grains or single-crystal components (Figure 8.33).
Another is the controlled unidirectional solidification of alloys having specially
designed compositions wherein two-phase composites result.
SUMMARY
Fundamentals of Fracture
Ductile Fracture
Fracture, in response to tensile loading and at relatively low temperatures, may occur
by ductile and brittle modes, both of which involve the formation and propagation
of cracks. For ductile fracture, evidence will exist of gross plastic deformation at the
fracture surface. In tension, highly ductile metals will neck down to essentially a
point fracture; cup-and-cone mating fracture surfaces result for moderate ductility.
Cracks in ductile materials are said to be stable (i.e., resist extension without an
increase in applied stress); and inasmuch as fracture is noncatastrophic, this fracture
mode is almost always preferred.
Brittle Fracture
For brittle fracture, cracks are unstable, and the fracture surface is relatively flat
and perpendicular to the direction of the applied tensile load. Chevron and ridgelike patterns are possible, which indicate the direction of crack propagation. Transgranular (through-grain) and intergranular (between-grain) fractures are found in
brittle polycrystalline materials.
1496T_c08_207-251 11/18/05 11:50 Page 244
REVISED PAGES
244 • Chapter 8 / Failure
Principles of Fracture Mechanics
The significant discrepancy between actual and theoretical fracture strengths of
brittle materials is explained by the existence of small flaws that are capable of
amplifying an applied tensile stress in their vicinity, leading ultimately to crack
formation. Fracture ensues when the theoretical cohesive strength is exceeded at
the tip of one of these flaws.
The fracture toughness of a material is indicative of its resistance to brittle fracture when a crack is present. It depends on specimen thickness, and, for relatively
thick specimens (i.e., conditions of plane strain), is termed the plane strain fracture
toughness. This parameter is the one normally cited for design purposes; its value
is relatively large for ductile materials (and small for brittle ones), and is a function
of microstructure, strain rate, and temperature. With regard to designing against the
possibility of fracture, consideration must be given to material (its fracture toughness), the stress level, and the flaw size detection limit.
Impact Fracture Testing
Qualitatively, the fracture behavior of materials may be determined using Charpy
and Izod impact testing techniques. On the basis of the temperature dependence of
measured impact energy (or appearance of the fracture surface), it is possible to
ascertain whether or not a material experiences a ductile-to-brittle transition and
the temperature range over which such a transition occurs. Low-strength steel
alloys typify this behavior, and, for structural applications, should be used at
temperatures in excess of the transition range. Furthermore, low-strength FCC metals, most HCP metals, and high-strength materials do not experience this ductileto-brittle transition.
Cyclic Stresses (Fatigue)
The S–N Curve
Fatigue is a common type of catastrophic failure wherein the applied stress level
fluctuates with time. Test data are plotted as stress versus the logarithm of the number of cycles to failure. For many metals and alloys, stress diminishes continuously
with increasing number of cycles at failure; fatigue strength and fatigue life are the
parameters used to characterize the fatigue behavior of these materials. On the
other hand, for other metals/alloys, at some point, stress ceases to decrease with,
and becomes independent of, the number of cycles; the fatigue behavior of these
materials is expressed in terms of fatigue limit.
Crack Initiation and Propagation
The processes of fatigue crack initiation and propagation were discussed. Cracks
normally nucleate on the surface of a component at some point of stress concentration. Two characteristic fatigue surface features are beachmarks and striations.
Beachmarks form on components that experience applied stress interruptions; they
normally may be observed with the naked eye. Fatigue striations are of microscopic
dimensions, and each is thought to represent the crack tip advance distance over a
single load cycle.
Factors That Affect Fatigue Life
Measures that may be taken to extend fatigue life include: (1) reducing the mean
stress level; (2) eliminating sharp surface discontinuities; (3) improving the surface
1496T_c08_207-251 11/18/05 11:50 Page 245
REVISED PAGES
Important Terms and Concepts • 245
finish by polishing; (4) imposing surface residual compressive stresses by shot peening; and (5) case hardening by using a carburizing or nitriding process.
Environmental Effects
The fatigue behavior of materials may also be affected by the environment. Thermal stresses may be induced in components that are exposed to elevated temperature fluctuations and when thermal expansion and/or contraction is restrained;
fatigue for these conditions is termed thermal fatigue. The presence of a chemically
active environment may lead to a reduction in fatigue life for corrosion fatigue.
Generalized Creep Behavior
The time-dependent plastic deformation of materials subjected to a constant load
(or stress) and temperatures greater than about 0.4Tm is termed creep. A typical
creep curve (strain versus time) will normally exhibit three distinct regions. For transient (or primary) creep, the rate (or slope) diminishes with time. The plot becomes
linear (i.e., creep rate is constant) in the steady-state (or secondary) region. And
finally, deformation accelerates for tertiary creep, just prior to failure (or rupture).
Important design parameters available from such a plot include the steady-state
creep rate (slope of the linear region) and rupture lifetime.
Stress and Temperature Effects
Both temperature and applied stress level influence creep behavior. Increasing
either of these parameters produces the following effects: (1) an increase in the instantaneous initial deformation; (2) an increase in the steady-state creep rate; and
(3) a diminishment of the rupture lifetime. Analytical expressions were presented
#
which relate s to both temperature and stress.
Data Extrapolation Methods
Extrapolation of creep test data to lower temperature—longer time regimes is possible using the Larson–Miller parameter.
Alloys for High-Temperature Use
Metal alloys that are especially resistant to creep have high elastic moduli and melting temperatures; these include the superalloys, the stainless steels, and the refractory
metals. Various processing techniques are employed to improve the creep properties
of these materials.
I M P O R TA N T T E R M S A N D C O N C E P T S
Brittle fracture
Case hardening
Charpy test
Corrosion fatigue
Creep
Ductile fracture
Ductile-to-brittle transition
Fatigue
Fatigue life
Fatigue limit
Fatigue strength
Fracture mechanics
Fracture toughness
Impact energy
Intergranular fracture
Izod test
Plane strain
Plane strain fracture toughness
Stress raiser
Thermal fatigue
Transgranular fracture
1496T_c08_207-251 11/18/05 11:50 Page 246
REVISED PAGES
246 • Chapter 8 / Failure
REFERENCES
ASM Handbook, Vol. 11, Failure Analysis and Prevention, ASM International, Materials Park,
OH, 1986.
ASM Handbook, Vol. 12, Fractography, ASM
International, Materials Park, OH, 1987.
ASM Handbook, Vol. 19, Fatigue and Fracture,
ASM International, Materials Park, OH, 1996.
Boyer, H. E. (Editor), Atlas of Creep and
Stress–Rupture Curves, ASM International,
Materials Park, OH, 1988.
Boyer, H. E. (Editor), Atlas of Fatigue Curves, ASM
International, Materials Park, OH, 1986.
Colangelo, V. J. and F. A. Heiser, Analysis of Metallurgical Failures, 2nd edition, Wiley, New
York, 1987.
Collins, J.A., Failure of Materials in Mechanical Design, 2nd edition, Wiley, New York, 1993.
Courtney, T. H., Mechanical Behavior of Materials,
2nd edition, McGraw-Hill, New York, 2000.
Dieter, G. E., Mechanical Metallurgy, 3rd edition,
McGraw-Hill, New York, 1986.
Esaklul, K. A., Handbook of Case Histories in Failure Analysis, ASM International, Materials
Park, OH, 1992 and 1993. In two volumes.
Fatigue Data Book: Light Structural Alloys, ASM
International, Materials Park, OH, 1995.
Hertzberg, R. W., Deformation and Fracture Mechanics of Engineering Materials, 4th edition,
Wiley, New York, 1996.
Stevens, R. I., A. Fatemi, R. R. Stevens, and H. O.
Fuchs, Metal Fatigue in Engineering, 2nd edition,
Wiley, New York, 2000.
Tetelman, A. S. and A. J. McEvily, Fracture of Structural Materials, Wiley, New York, 1967. Reprinted
by Books on Demand, Ann Arbor, MI.
Wulpi, D. J., Understanding How Components Fail,
2nd edition, ASM International, Materials
Park, OH, 1999.
QUESTIONS AND PROBLEMS
Principles of Fracture Mechanics
8.1 What is the magnitude of the maximum stress
that exists at the tip of an internal crack having a radius of curvature of 1.9 104 mm
(7.5 106 in.) and a crack length of 3.8
102 mm (1.5 103 in.) when a tensile stress
of 140 MPa (20,000 psi) is applied?
8.2 Estimate the theoretical fracture strength of
a brittle material if it is known that fracture
occurs by the propagation of an elliptically
shaped surface crack of length 0.5 mm (0.02 in.)
and having a tip radius of curvature of 5
103 mm (2 104 in.), when a stress of
1035 MPa (150,000 psi) is applied.
8.3 If the specific surface energy for aluminum
oxide is 0.90 J/m2, using data contained in
Table 12.5, compute the critical stress required
for the propagation of an internal crack of
length 0.40 mm.
8.4 An MgO component must not fail when a tensile stress of 13.5 MPa (1960 psi) is applied.
Determine the maximum allowable surface
crack length if the surface energy of MgO is
1.0 J/m2. Data found in Table 12.5 may prove
helpful.
8.5 A specimen of a 4340 steel alloy with a plane
strain fracture toughness of 54.8 MPa 1m
(50 ksi 1in.) is exposed to a stress of 1030 MPa
(150,000 psi). Will this specimen experience
fracture if it is known that the largest surface
crack is 0.5 mm (0.02 in.) long? Why or why
not? Assume that the parameter Y has a value
of 1.0.
8.6 Some aircraft component is fabricated from
an aluminum alloy that has a plane strain fracture toughness of 40 MPa 1m (36.4 ksi1in.).
It has been determined that fracture results
at a stress of 300 MPa (43,500 psi) when the
maximum (or critical) internal crack length is
4.0 mm (0.16 in.). For this same component
and alloy, will fracture occur at a stress level
of 260 MPa (38,000 psi) when the maximum
internal crack length is 6.0 mm (0.24 in.)? Why
or why not?
8.7 Suppose that a wing component on an aircraft
is fabricated from an aluminum alloy that has
a plane strain fracture toughness of 26 MPa 1m
(23.7 ksi 1in.). It has been determined that
fracture results at a stress of 112 MPa
(16,240 psi) when the maximum internal
crack length is 8.6 mm (0.34 in.). For this same
1496T_c08_207-251 11/18/05 13:43 Page 247
REVISED PAGES
Questions and Problems • 247
8.8
8.9
8.10
8.11
component and alloy, compute the stress level
at which fracture will occur for a critical internal crack length of 6.0 mm (0.24 in.).
A large plate is fabricated from a steel alloy
that has a plane strain fracture toughness of
82.4 MPa1m (75.0 ksi 1in.). If, during service use, the plate is exposed to a tensile stress
of 345 MPa (50,000 psi), determine the minimum length of a surface crack that will lead
to fracture. Assume a value of 1.0 for Y.
Calculate the maximum internal crack length
allowable for a Ti-6Al-4V titanium alloy
(Table 8.1) component that is loaded to a
stress one-half of its yield strength. Assume
that the value of Y is 1.50.
A structural component in the form of a
wide plate is to be fabricated from a steel alloy
that has a plane strain fracture toughness of
98.9 MPa1m (90 ksi 1in.) and a yield strength
of 860 MPa (125,000 psi). The flaw size resolution limit of the flaw detection apparatus is
3.0 mm (0.12 in.). If the design stress is onehalf of the yield strength and the value of Y
is 1.0, determine whether or not a critical flaw
for this plate is subject to detection.
After consultation of other references, write
a brief report on one or two nondestructive
test techniques that are used to detect and
measure internal and/or surface flaws in metal
alloys.
Impact Fracture Testing
8.12 Following is tabulated data that were gathered from a series of Charpy impact tests on
a tempered 4340 steel alloy.
Temperature (C)
Impact Energy (J)
0
25
50
75
100
113
125
150
175
200
105
104
103
97
63
40
34
28
25
24
(a) Plot the data as impact energy versus
temperature.
(b) Determine a ductile-to-brittle transition
temperature as that temperature corresponding to the average of the maximum and
minimum impact energies.
(c) Determine a ductile-to-brittle transition
temperature as that temperature at which the
impact energy is 50 J.
8.13 Following is tabulated data that were gathered from a series of Charpy impact tests on
a commercial low-carbon steel alloy.
Temperature (C)
50
40
30
20
10
0
10
20
30
40
Impact Energy (J)
76
76
71
58
38
23
14
9
5
1.5
(a) Plot the data as impact energy versus
temperature.
(b) Determine a ductile-to-brittle transition
temperature as that temperature corresponding to the average of the maximum and
minimum impact energies.
(c) Determine a ductile-to-brittle transition
temperature as that temperature at which the
impact energy is 20 J.
Cyclic Stresses (Fatigue)
The S–N Curve
8.14 A fatigue test was conducted in which the
mean stress was 70 MPa (10,000 psi), and the
stress amplitude was 210 MPa (30,000 psi).
(a) Compute the maximum and minimum
stress levels.
(b) Compute the stress ratio.
(c) Compute the magnitude of the stress range.
8.15 A cylindrical 1045 steel bar (Figure 8.34) is
subjected to repeated compression-tension
stress cycling along its axis. If the load amplitude is 66,700 N (15,000 lbf), compute the minimum allowable bar diameter to ensure that
fatigue failure will not occur. Assume a safety
factor of 2.0.
1496T_c08_207-251 11/18/05 11:50 Page 248
REVISED PAGES
248 • Chapter 8 / Failure
80
500
70
60
400
50
1045 steel
300
40
2014–T6 aluminum alloy
30
200
20
Stress amplitude (10 3 psi)
Stress amplitude, S (MPa)
Figure 8.34 Stress magnitude S versus the
logarithm of the number N of cycles to fatigue
failure for red brass, an aluminum alloy, and a
plain carbon steel. (Adapted from H. W. Hayden,
W. G. Moffatt, and J. Wulff, The Structure and
Properties of Materials, Vol. III, Mechanical
Behavior, p. 15. Copyright © 1965 by John Wiley
& Sons, New York. Reprinted by permission of
John Wiley & Sons, Inc. Also adapted from ASM
Handbook, Vol. 2, Properties and Selection:
Nonferrous Alloys and Special-Purpose Materials,
1990. Reprinted by permission of ASM
International.)
100
10
Red brass
0
0
103
104
106
105
107
108
109
1010
Cycles to failure, N
8.16 A 6.4 mm (0.25 in.) diameter cylindrical rod
fabricated from a 2014-T6 aluminum alloy
(Figure 8.34) is subjected to reversed tensioncompression load cycling along its axis. If the
maximum tensile and compressive loads are
5340 N (1200 lbf) and 5340 N (1200 lbf),
respectively, determine its fatigue life.Assume
that the stress plotted in Figure 8.34 is stress
amplitude.
8.17 A 15.2 mm (0.60 in.) diameter cylindrical
rod fabricated from a 2014-T6 aluminum alloy
(Figure 8.34) is subjected to a repeated tensioncompression load cycling along its axis. Compute the maximum and minimum loads that
will be applied to yield a fatigue life of
1.0 108 cycles. Assume that the stress plotted on the vertical axis is stress amplitude, and
data were taken for a mean stress of 35 MPa
(5000 psi).
8.18 The fatigue data for a brass alloy are given as
follows:
Stress Amplitude
(MPa)
Cycles to
Failure
170
148
130
114
92
80
74
3.7 104
1.0 105
3.0 105
1.0 106
1.0 107
1.0 108
1.0 109
(a) Make an S–N plot (stress amplitude versus
logarithm cycles to failure) using these data.
(b) Determine the fatigue strength at 4 106
cycles.
(c) Determine the fatigue life for 120 MPa.
8.19 Suppose that the fatigue data for the brass
alloy in Problem 8.18 were taken from
bending-rotating tests, and that a rod of this
alloy is to be used for an automobile axle that
rotates at an average rotational velocity of 1800
revolutions per minute. Give the maximum
torsional stress amplitude possible for each of
the following lifetimes of the rod: (a) 1 year,
(b) 1 month, (c) 1 day, and (d) 1 hour.
8.20 The fatigue data for a steel alloy are given as
follows:
Stress Amplitude
[MPa (ksi)]
Cycles to
Failure
470 (68.0)
440 (63.4)
390 (56.2)
350 (51.0)
310 (45.3)
290 (42.2)
290 (42.2)
290 (42.2)
104
3 104
105
3 105
106
3 106
107
108
(a) Make an S–N plot (stress amplitude versus logarithm cycles to failure) using these
data.
1496T_c08_207-251 11/18/05 13:43 Page 249
REVISED PAGES
Questions and Problems • 249
(b) What is the fatigue limit for this alloy?
(c) Determine fatigue lifetimes at stress amplitudes of 415 MPa (60,000 psi) and 275 MPa
(40,000 psi).
(d) Estimate fatigue strengths at 2 104 and
6 105 cycles.
8.21 Suppose that the fatigue data for the steel alloy in Problem 8.20 were taken for bendingrotating tests, and that a rod of this alloy is to
be used for an automobile axle that rotates at
an average rotational velocity of 600 revolutions per minute. Give the maximum lifetimes of continuous driving that are allowable
for the following stress levels: (a) 450 MPa
(65,000 psi), (b) 380 MPa (55,000 psi),
(c) 310 MPa (45,000 psi), and (d) 275 MPa
(40,000 psi).
8.22 Three identical fatigue specimens (denoted A,
B, and C) are fabricated from a nonferrous
alloy. Each is subjected to one of the maximumminimum stress cycles listed below; the frequency is the same for all three tests.
Specimen
max (MPa)
min (MPa)
A
B
C
450
300
500
150
300
200
(a) Rank the fatigue lifetimes of these three
specimens from the longest to the shortest.
(b) Now justify this ranking using a schematic
S–N plot.
8.23 Cite five factors that may lead to scatter in
fatigue life data.
Crack Initiation and Propagation
Factors That Affect Fatigue Life
8.24 Briefly explain the difference between fatigue
striations and beachmarks both in terms of
(a) size and (b) origin.
8.25 List four measures that may be taken to
increase the resistance to fatigue of a metal
alloy.
Generalized Creep Behavior
8.26 Give the approximate temperature at which
creep deformation becomes an important consideration for each of the following metals: tin,
molybdenum, iron, gold, zinc, and chromium.
8.27 The following creep data were taken on an
aluminum alloy at 480C (900F) and a constant stress of 2.75 MPa (400 psi). Plot the data
as strain versus time, then determine the
steady-state or minimum creep rate. Note: The
initial and instantaneous strain is not included.
Time
(min)
Strain
Time
(min)
Strain
0
2
4
6
8
10
12
14
16
0.00
0.22
0.34
0.41
0.48
0.55
0.62
0.68
0.75
18
20
22
24
26
28
30
32
34
0.82
0.88
0.95
1.03
1.12
1.22
1.36
1.53
1.77
Stress and Temperature Effects
8.28 A specimen 1015 mm (40 in.) long of a low
carbon–nickel alloy (Figure 8.31) is to be
exposed to a tensile stress of 70 MPa
(10,000 psi) at 427C (800F). Determine its
elongation after 10,000 h. Assume that the total of both instantaneous and primary creep
elongations is 1.3 mm (0.05 in.).
8.29 For a cylindrical low carbon–nickel alloy specimen (Figure 8.31) originally 19 mm (0.75 in.)
in diameter and 635 mm (25 in.) long, what
tensile load is necessary to produce a total
elongation of 6.44 mm (0.25 in.) after 5000 h
at 538C (1000F)? Assume that the sum of
instantaneous and primary creep elongations
is 1.8 mm (0.07 in.).
8.30 If a component fabricated from a low
carbon–nickel alloy (Figure 8.30) is to be
exposed to a tensile stress of 31 MPa (4500
psi) at 649C (1200F), estimate its rupture
lifetime.
8.31 A cylindrical component constructed from a
low carbon–nickel alloy (Figure 8.30) has a diameter of 19.1 mm (0.75 in.). Determine the
maximum load that may be applied for it to
survive 10,000 h at 538C (1000F).
#
8.32 From Equation 8.19, if the logarithm of s
is plotted versus the logarithm of s, then
a straight line should result, the slope of
which is the stress exponent n. Using Figure
8.31, determine the value of n for the low
1496T_c08_207-251 11/18/05 11:50 Page 250
REVISED PAGES
250 • Chapter 8 / Failure
carbon–nickel alloy at each of the three
temperatures.
8.33 (a) Estimate the activation energy for creep
(i.e., Qc in Equation 8.20) for the low carbon–
nickel alloy having the steady-state creep behavior shown in Figure 8.31. Use data taken
at a stress level of 55 MPa (8000 psi) and temperatures of 427C and 538C. Assume that
the stress exponent n is independent of tem#
perature. (b) Estimate s at 649C (922 K).
8.34 Steady-state creep rate data are given here for
some alloy taken at 200C (473 K):
#
s (h1)
3
2.5 10
2.4 102
[MPa (psi)]
55 (8000)
69 (10,000)
If it is known that the activation energy for
creep is 140,000 J/mol, compute the steady-
state creep rate at a temperature of 250C
(523 K) and a stress level of 48 MPa (7000 psi).
8.35 Steady-state creep data taken for an iron at a
stress level of 140 MPa (20,000 psi) are given
here:
#
s (h1)
4
6.6 10
8.8 102
T (K)
1090
1200
If it is known that the value of the stress exponent n for this alloy is 8.5, compute the
steady-state creep rate at 1300 K and a stress
level of 83 MPa (12,000 psi).
Alloys for High-Temperature Use
8.36 Cite three metallurgical/processing techniques
that are employed to enhance the creep resistance of metal alloys.
DESIGN PROBLEMS
8.D1 Each student (or group of students) is to
obtain an object/structure/component that
has failed. It may come from your home, an
automobile repair shop, a machine shop,
etc. Conduct an investigation to determine
the cause and type of failure (i.e., simple
fracture, fatigue, creep). In addition, propose measures that can be taken to prevent
future incidents of this type of failure.
Finally, submit a report that addresses the
above issues.
Principles of Fracture Mechanics
8.D2 (a) For the thin-walled spherical tank discussed in Design Example 8.1, on the basis
of critical crack size criterion [as addressed
in part (a)], rank the following polymers
from longest to shortest critical crack
length: nylon 6,6 (50% relative humidity),
polycarbonate, poly(ethylene terephthalate), and poly(methyl methacrylate).
Comment on the magnitude range of the
computed values used in the ranking relative to those tabulated for metal alloys as
provided in Table 8.3. For these computations, use data contained in Tables B.4 and
B.5 in Appendix B.
(b) Now rank these same four polymers
relative to maximum allowable pressure
according to the leak-before-break criterion,
as described in the (b) portion of Design
Example 8.1. As above, comment on these
values in relation to those for the metal
alloys that are tabulated in Table 8.4.
Data Extrapolation Methods
8.D3 An S-590 iron component (Figure 8.32) must
have a creep rupture lifetime of at least
20 days at 650C (923 K). Compute the maximum allowable stress level.
8.D4 Consider an S-590 iron component (Figure
8.32) that is subjected to a stress of 55 MPa
(8000 psi). At what temperature will the rupture lifetime be 200 h?
8.D5 For an 18-8 Mo stainless steel (Figure 8.35),
predict the time to rupture for a component
that is subjected to a stress of 100 MPa
(14,500 psi) at 600C (873 K).
8.D6 Consider an 18-8 Mo stainless steel component (Figure 8.35) that is exposed to a temperature of 650C (923 K). What is the maximum allowable stress level for a rupture
lifetime of 1 year? 15 years?
1496T_c08_207-251 11/18/05 11:50 Page 251
REVISED PAGES
Design Problems • 251
Figure 8.35 Logarithm stress
versus the Larson–Miller
parameter for an 18-8 Mo
stainless steel. (From F. R.
Larson and J. Miller, Trans.
ASME, 74, 765, 1952. Reprinted
by permission of ASME.)
103 T(20 + log tr)(°R–h)
25
30
35
40
45
50
100
10
10
12
16
20
24
103 T(20 + log tr)(K–h)
28
1
Stress (103 psi)
Stress (MPa)
100
1496T_c09_252-310 12/21/05 8:30 Page 252
Chapter
A
9
2nd REVISE PAGES
Phase Diagrams
scanning electron micrograph showing the microstructure of a plain carbon steel that contains 0.44 wt% C. The large dark
areas are proeutectoid ferrite. Regions having the alternating light and dark lamellar structure are pearlite; the dark and light
layers in the pearlite correspond, respectively, to ferrite and cementite phases. During etching of the surface prior to examination, the ferrite phase was preferentially dissolved; thus, the pearlite appears in topographical relief with cementite layers being
elevated above the ferrite layers. 3000. (Micrograph courtesy of Republic Steel Corporation.)
WHY STUDY Phase Diagrams?
One reason that a knowledge and understanding of
phase diagrams is important to the engineer relates to
the design and control of heat-treating procedures;
some properties of materials are functions of their microstructures, and, consequently, of their thermal histories. Even though most phase diagrams represent
stable (or equilibrium) states and microstructures, they
252 •
are nevertheless useful in understanding the development and preservation of nonequilibrium structures
and their attendant properties; it is often the case
that these properties are more desirable than those
associated with the equilibrium state. This is aptly
illustrated by the phenomenon of precipitation
hardening (Section 11.9).
1496T_c09_252-310 12/21/05 7:56 Page 253
2nd REVISE PAGES
Learning Objectives
After studying this chapter you should be able to do the following:
1. (a) Schematically sketch simple isomorphous
(b) write reactions for all these transformations
and eutectic phase diagrams.
for either heating or cooling.
(b) On these diagrams label the various phase
4. Given the composition of an iron–carbon alloy
regions.
containing between 0.022 wt% C and 2.14 wt%
(c) Label liquidus, solidus, and solvus lines.
C, be able to
2. Given a binary phase diagram, the composition
(a) specify whether the alloy is hypoeutectoid or
of an alloy, its temperature, and assuming that
hypereutectoid,
the alloy is at equilibrium, determine
(b) name the proeutectoid phase,
(a) what phase(s) is (are) present,
(c) compute the mass fractions of proeutectoid
(b) the composition(s) of the phase(s), and
phase and pearlite, and
(c) the mass fraction(s) of the phase(s).
(d) make a schematic diagram of the microstruc3. For some given binary phase diagram, do the
ture at a temperature just below the
following:
eutectoid.
(a) locate the temperatures and compositions
of all eutectic, eutectoid, peritectic, and
congruent phase transformations; and
9.1 INTRODUCTION
The understanding of phase diagrams for alloy systems is extremely important because there is a strong correlation between microstructure and mechanical properties,
and the development of microstructure of an alloy is related to the characteristics of
its phase diagram. In addition, phase diagrams provide valuable information about
melting, casting, crystallization, and other phenomena.
This chapter presents and discusses the following topics: (1) terminology associated with phase diagrams and phase transformations; (2) pressure–temperature
phase diagrams for pure materials; (3) the interpretation of phase diagrams; (4) some
of the common and relatively simple binary phase diagrams, including that for the
iron–carbon system; and (5) the development of equilibrium microstructures, upon
cooling, for several situations.
Definitions and Basic Concepts
component
system
It is necessary to establish a foundation of definitions and basic concepts relating
to alloys, phases, and equilibrium before delving into the interpretation and utilization of phase diagrams. The term component is frequently used in this discussion; components are pure metals and/or compounds of which an alloy is composed.
For example, in a copper–zinc brass, the components are Cu and Zn. Solute and
solvent, which are also common terms, were defined in Section 4.3. Another term
used in this context is system, which has two meanings. First, “system” may refer to
a specific body of material under consideration (e.g., a ladle of molten steel). Or it
may relate to the series of possible alloys consisting of the same components, but
without regard to alloy composition (e.g., the iron–carbon system).
The concept of a solid solution was introduced in Section 4.3. By way of review,
a solid solution consists of atoms of at least two different types; the solute atoms
occupy either substitutional or interstitial positions in the solvent lattice, and the
crystal structure of the solvent is maintained.
1496T_c09_252-310 11/29/05 11:33 Page 254
REVISED PAGES
254 • Chapter 9 / Phase Diagrams
9.2 SOLUBILITY LIMIT
solubility limit
For many alloy systems and at some specific temperature, there is a maximum
concentration of solute atoms that may dissolve in the solvent to form a solid solution; this is called a solubility limit. The addition of solute in excess of this solubility limit results in the formation of another solid solution or compound that has
a distinctly different composition. To illustrate this concept, consider the sugar–
water (C12H22O11–H2O) system. Initially, as sugar is added to water, a sugar–water
solution or syrup forms. As more sugar is introduced, the solution becomes more
concentrated, until the solubility limit is reached, or the solution becomes saturated
with sugar. At this time the solution is not capable of dissolving any more sugar,
and further additions simply settle to the bottom of the container. Thus, the system
now consists of two separate substances: a sugar–water syrup liquid solution and
solid crystals of undissolved sugar.
This solubility limit of sugar in water depends on the temperature of the water
and may be represented in graphical form on a plot of temperature along the ordinate and composition (in weight percent sugar) along the abscissa, as shown in
Figure 9.1. Along the composition axis, increasing sugar concentration is from left
to right, and percentage of water is read from right to left. Since only two components are involved (sugar and water), the sum of the concentrations at any composition will equal 100 wt%. The solubility limit is represented as the nearly vertical
line in the figure. For compositions and temperatures to the left of the solubility
line, only the syrup liquid solution exists; to the right of the line, syrup and solid
sugar coexist. The solubility limit at some temperature is the composition that
corresponds to the intersection of the given temperature coordinate and the solubility limit line. For example, at 20C the maximum solubility of sugar in water is
65 wt%. As Figure 9.1 indicates, the solubility limit increases slightly with rising
temperature.
9.3 PHASES
Also critical to the understanding of phase diagrams is the concept of a phase. A
phase may be defined as a homogeneous portion of a system that has uniform physical and chemical characteristics. Every pure material is considered to be a phase;
Figure 9.1 The
solubility of sugar
(C12H22O11) in a
sugar–water syrup.
100
200
150
60
Liquid
solution
+
solid
sugar
Liquid solution (syrup)
40
100
20
50
Sugar
Water
0
0
20
40
60
80
100
100
80
60
40
20
0
Composition (wt%)
Temperature (°F)
Solubility limit
80
Temperature (°C)
phase
1496T_c09_252-310 11/29/05 11:33 Page 255
REVISED PAGES
9.5 Phase Equilibria • 255
so also is every solid, liquid, and gaseous solution. For example, the sugar–water
syrup solution just discussed is one phase, and solid sugar is another. Each has different physical properties (one is a liquid, the other is a solid); furthermore, each is
different chemically (i.e., has a different chemical composition); one is virtually pure
sugar, the other is a solution of H2O and C12H22O11. If more than one phase is present in a given system, each will have its own distinct properties, and a boundary
separating the phases will exist across which there will be a discontinuous and abrupt
change in physical and/or chemical characteristics. When two phases are present in
a system, it is not necessary that there be a difference in both physical and chemical properties; a disparity in one or the other set of properties is sufficient. When
water and ice are present in a container, two separate phases exist; they are physically dissimilar (one is a solid, the other is a liquid) but identical in chemical makeup.
Also, when a substance can exist in two or more polymorphic forms (e.g., having
both FCC and BCC structures), each of these structures is a separate phase because
their respective physical characteristics differ.
Sometimes, a single-phase system is termed “homogeneous.” Systems composed
of two or more phases are termed “mixtures” or “heterogeneous systems.” Most
metallic alloys and, for that matter, ceramic, polymeric, and composite systems are
heterogeneous. Ordinarily, the phases interact in such a way that the property combination of the multiphase system is different from, and more attractive than, either
of the individual phases.
9.4
MICROSTRUCTURE
Many times, the physical properties and, in particular, the mechanical behavior of
a material depend on the microstructure. Microstructure is subject to direct microscopic observation, using optical or electron microscopes; this topic was touched on
in Sections 4.9 and 4.10. In metal alloys, microstructure is characterized by the number of phases present, their proportions, and the manner in which they are distributed or arranged. The microstructure of an alloy depends on such variables as the
alloying elements present, their concentrations, and the heat treatment of the alloy
(i.e., the temperature, the heating time at temperature, and the rate of cooling to
room temperature).
The procedure of specimen preparation for microscopic examination was briefly
outlined in Section 4.10.After appropriate polishing and etching, the different phases
may be distinguished by their appearance. For example, for a two-phase alloy, one
phase may appear light and the other phase dark, as in the chapter-opening photograph for this chapter. When only a single phase or solid solution is present, the texture will be uniform, except for grain boundaries that may be revealed (Figure 4.12b).
9.5 PHASE EQUILIBRIA
equilibrium
free energy
Equilibrium is another essential concept that is best described in terms of a thermodynamic quantity called the free energy. In brief, free energy is a function of the
internal energy of a system, and also the randomness or disorder of the atoms or
molecules (or entropy). A system is at equilibrium if its free energy is at a minimum under some specified combination of temperature, pressure, and composition.
In a macroscopic sense, this means that the characteristics of the system do not
change with time but persist indefinitely; that is, the system is stable. A change in
temperature, pressure, and/or composition for a system in equilibrium will result in
an increase in the free energy and in a possible spontaneous change to another state
whereby the free energy is lowered.
1496T_c09_252-310 11/29/05 12:46 Page 256
REVISED PAGES
256 • Chapter 9 / Phase Diagrams
phase equilibrium
metastable
The term phase equilibrium, often used in the context of this discussion, refers
to equilibrium as it applies to systems in which more than one phase may exist.
Phase equilibrium is reflected by a constancy with time in the phase characteristics
of a system. Perhaps an example best illustrates this concept. Suppose that a
sugar–water syrup is contained in a closed vessel and the solution is in contact with
solid sugar at 20C. If the system is at equilibrium, the composition of the syrup is
65 wt% C12H22O1135 wt% H2O (Figure 9.1), and the amounts and compositions
of the syrup and solid sugar will remain constant with time. If the temperature of
the system is suddenly raised—say, to 100C—this equilibrium or balance is temporarily upset in that the solubility limit has been increased to 80 wt% C12H22O11
(Figure 9.1). Thus, some of the solid sugar will go into solution in the syrup. This
will continue until the new equilibrium syrup concentration is established at the
higher temperature.
This sugar–syrup example illustrates the principle of phase equilibrium using a
liquid–solid system. In many metallurgical and materials systems of interest, phase
equilibrium involves just solid phases. In this regard the state of the system is reflected in the characteristics of the microstructure, which necessarily include not
only the phases present and their compositions but, in addition, the relative phase
amounts and their spatial arrangement or distribution.
Free energy considerations and diagrams similar to Figure 9.1 provide information about the equilibrium characteristics of a particular system, which is important; but they do not indicate the time period necessary for the attainment of a
new equilibrium state. It is often the case, especially in solid systems, that a state of
equilibrium is never completely achieved because the rate of approach to equilibrium is extremely slow; such a system is said to be in a nonequilibrium or metastable
state. A metastable state or microstructure may persist indefinitely, experiencing
only extremely slight and almost imperceptible changes as time progresses. Often,
metastable structures are of more practical significance than equilibrium ones. For
example, some steel and aluminum alloys rely for their strength on the development of metastable microstructures during carefully designed heat treatments
(Sections 10.5 and 11.9).
Thus not only is an understanding of equilibrium states and structures important, but also the speed or rate at which they are established and the factors that
affect the rate must be considered. This chapter is devoted almost exclusively to
equilibrium structures; the treatment of reaction rates and nonequilibrium structures is deferred to Chapter 10 and Section 11.9.
Concept Check 9.1
What is the difference between the states of phase equilibrium and metastability?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
9.6 ONE-COMPONENT (OR UNARY)
PHASE DIAGRAMS
phase diagram
Much of the information about the control of the phase structure of a particular
system is conveniently and concisely displayed in what is called a phase diagram, also
often termed an equilibrium diagram. Now, there are three externally controllable
1496T_c09_252-310 11/29/05 12:46 Page 257
REVISED PAGES
9.6 One-Component (or Unary) Phase Diagrams • 257
parameters that will affect phase structure—viz. temperature, pressure, and composition—and phase diagrams are constructed when various combinations of these
parameters are plotted against one another.
Perhaps the simplest and easiest type of phase diagram to understand is
that for a one-component system, in which composition is held constant (i.e., the
phase diagram is for a pure substance); this means that pressure and temperature
are the variables. This one-component phase diagram (or unary phase diagram)
[sometimes also called a pressure–temperature (or P–T ) diagram] is represented as a two-dimensional plot of pressure (ordinate, or vertical axis) versus
temperature (abscissa, or horizontal axis). Most often, the pressure axis is scaled
logarithmically.
We illustrate this type of phase diagram and demonstrate its interpretation using as an example the one for H2O, which is shown in Figure 9.2. Here it may be
noted that regions for three different phases—solid, liquid, and vapor—are delineated on the plot. Each of the phases will exist under equilibrium conditions over
the temperature–pressure ranges of its corresponding area. Furthermore, the three
curves shown on the plot (labeled aO, bO, and cO) are phase boundaries; at any
point on one of these curves, the two phases on either side of the curve are in equilibrium (or coexist) with one another. That is, equilibrium between solid and vapor
phases is along curve aO—likewise for the solid-liquid, curve bO, and the liquidvapor, curve cO. Also, upon crossing a boundary (as temperature and/or pressure
is altered), one phase transforms to another. For example, at one atmosphere pressure, during heating the solid phase transforms to the liquid phase (i.e., melting
occurs) at the point labeled 2 on Figure 9.2 (i.e., the intersection of the dashed
horizontal line with the solid-liquid phase boundary); this point corresponds to a
temperature of 0C. Of course, the reverse transformation (liquid-to-solid, or solidification) takes place at the same point upon cooling. Similarly, at the intersection
of the dashed line with the liquid-vapor phase boundary [point 3 (Figure 9.2), at
100C] the liquid transforms to the vapor phase (or vaporizes) upon heating;
1,000
b
Pressure (atm)
100
Liquid
(Water)
Solid
(Ice)
10
c
3
2
1.0
0.1
Vapor
(Steam)
O
0.01
a
0.001
20
0
20
40
60
80
100
120
Temperature (°C)
Figure 9.2 Pressure–temperature phase diagram for H2O.
Intersection of the dashed horizontal line at 1 atm pressure with
the solid-liquid phase boundary (point 2) corresponds to the
melting point at this pressure (T 0C). Similarly, point 3, the
intersection with the liquid-vapor boundary, represents the boiling
point (T 100C).
1496T_c09_252-310 11/29/05 12:46 Page 258
REVISED PAGES
258 • Chapter 9 / Phase Diagrams
condensation occurs for cooling. And, finally, solid ice sublimes or vaporizes upon
crossing the curve labeled aO.
As may also be noted from Figure 9.2, all three of the phase boundary curves
intersect at a common point, which is labeled O (and for this H2O system, at a temperature of 273.16 K and a pressure of 6.04 103 atm). This means that at this
point only, all of the solid, liquid, and vapor phases are simultaneously in equilibrium with one another. Appropriately, this, and any other point on a P–T phase diagram where three phases are in equilibrium, is called a triple point; sometimes it
is also termed an invariant point inasmuch as its position is distinct, or fixed by definite values of pressure and temperature. Any deviation from this point by a change
of temperature and/or pressure will cause at least one of the phases to disappear.
Pressure-temperature phase diagrams for a number of substances have been
determined experimentally, which also have solid, liquid, and vapor phase regions.
In those instances when multiple solid phases (i.e., allotropes, Section 3.6) exist,
there will appear a region on the diagram for each solid phase, and also other
triple points. The pressure-temperature phase diagram for carbon (which includes
phase regions for diamond and graphite) is shown on the front and back covers
of this book.
Binary Phase Diagrams
Another type of extremely common phase diagram is one in which temperature
and composition are variable parameters, and pressure is held constant—normally
1 atm. There are several different varieties; in the present discussion, we will concern ourselves with binary alloys—those that contain two components. If more than
two components are present, phase diagrams become extremely complicated and
difficult to represent. An explanation of the principles governing and the interpretation of phase diagrams can be demonstrated using binary alloys even though most
alloys contain more than two components.
Binary phase diagrams are maps that represent the relationships between temperature and the compositions and quantities of phases at equilibrium, which influence the microstructure of an alloy. Many microstructures develop from phase
transformations, the changes that occur when the temperature is altered (ordinarily upon cooling). This may involve the transition from one phase to another, or the
appearance or disappearance of a phase. Binary phase diagrams are helpful in predicting phase transformations and the resulting microstructures, which may have
equilibrium or nonequilibrium character.
9.7 BINARY ISOMORPHOUS SYSTEMS
Possibly the easiest type of binary phase diagram to understand and interpret is the
type that is characterized by the copper–nickel system (Figure 9.3a). Temperature
is plotted along the ordinate, and the abscissa represents the composition of the alloy, in weight percent (bottom) and atom percent (top) of nickel. The composition
ranges from 0 wt% Ni (100 wt% Cu) on the left horizontal extremity to 100 wt%
Ni (0 wt% Cu) on the right. Three different phase regions, or fields, appear on the
diagram, an alpha (a) field, a liquid (L) field, and a two-phase a L field. Each
region is defined by the phase or phases that exist over the range of temperatures
and compositions delimited by the phase boundary lines.
The liquid L is a homogeneous liquid solution composed of both copper and nickel.
The a phase is a substitutional solid solution consisting of both Cu and Ni atoms, and
1496T_c09_252-310 11/29/05 11:33 Page 259
REVISED PAGES
9.7 Binary Isomorphous Systems • 259
Composition (at% Ni)
0
20
40
60
80
100
1600
2800
1500
Liquid
1453°C
2600
Solidus line
Liquidus line
1300
2400
␣ +L
B
1200
2200
␣
A
1100
Temperature (°F)
1400
Temperature (°C)
Figure 9.3 (a) The
copper–nickel phase
diagram. (b) A
portion of the
copper–nickel phase
diagram for which
compositions and
phase amounts are
determined at point
B. (Adapted from
Phase Diagrams of
Binary Nickel Alloys,
P. Nash, Editor, 1991.
Reprinted by
permission of ASM
International,
Materials Park, OH.)
2000
1085°C
1000
0
40
20
(Cu)
60
80
100
Composition (wt% Ni)
(Ni)
(a)
1300
Liquid
Temperature (°C)
Tie line
B
␣ + Liquid
␣ + Liquid
␣
1200
R
S
␣
20
40
30
CL
C0
Composition (wt% Ni)
50
C␣
(b)
isomorphous
having an FCC crystal structure. At temperatures below about 1080C, copper and
nickel are mutually soluble in each other in the solid state for all compositions. This
complete solubility is explained by the fact that both Cu and Ni have the same crystal structure (FCC), nearly identical atomic radii and electronegativities, and similar
valences, as discussed in Section 4.3. The copper–nickel system is termed isomorphous
because of this complete liquid and solid solubility of the two components.
1496T_c09_252-310 11/29/05 11:33 Page 260
REVISED PAGES
260 • Chapter 9 / Phase Diagrams
A couple of comments are in order regarding nomenclature. First, for metallic
alloys, solid solutions are commonly designated by lowercase Greek letters (a, b, g,
etc.). Furthermore, with regard to phase boundaries, the line separating the L and
a L phase fields is termed the liquidus line, as indicated in Figure 9.3a; the liquid phase is present at all temperatures and compositions above this line.The solidus
line is located between the a and a L regions, below which only the solid a phase
exists.
For Figure 9.3a, the solidus and liquidus lines intersect at the two composition
extremities; these correspond to the melting temperatures of the pure components.
For example, the melting temperatures of pure copper and nickel are 1085C and
1453C, respectively. Heating pure copper corresponds to moving vertically up the
left-hand temperature axis. Copper remains solid until its melting temperature is
reached. The solid-to-liquid transformation takes place at the melting temperature,
and no further heating is possible until this transformation has been completed.
For any composition other than pure components, this melting phenomenon
will occur over the range of temperatures between the solidus and liquidus lines; both
solid a and liquid phases will be in equilibrium within this temperature range. For
example, upon heating an alloy of composition 50 wt% Ni–50 wt% Cu (Figure 9.3a),
melting begins at approximately 1280C (2340F); the amount of liquid phase continuously increases with temperature until about 1320C (2410F), at which the alloy
is completely liquid.
9.8 INTERPRETATION OF PHASE DIAGRAMS
For a binary system of known composition and temperature that is at equilibrium,
at least three kinds of information are available: (1) the phases that are present, (2)
the compositions of these phases, and (3) the percentages or fractions of the phases.
The procedures for making these determinations will be demonstrated using the
copper–nickel system.
Phases Present
Bi-Sb
Ge-Si
The establishment of what phases are present is relatively simple. One just locates
the temperature–composition point on the diagram and notes the phase(s) with
which the corresponding phase field is labeled. For example, an alloy of composition 60 wt% Ni–40 wt% Cu at 1100C would be located at point A in Figure 9.3a;
since this is within the a region, only the single a phase will be present. On the
other hand, a 35 wt% Ni–65 wt% Cu alloy at 1250C (point B) will consist of both
a and liquid phases at equilibrium.
Determination of Phase Compositions
Bi-Sb
Ge-Si
tie line
The first step in the determination of phase compositions (in terms of the concentrations of the components) is to locate the temperature–composition point on the
phase diagram. Different methods are used for single- and two-phase regions. If
only one phase is present, the procedure is trivial: the composition of this phase
is simply the same as the overall composition of the alloy. For example, consider
the 60 wt% Ni–40 wt% Cu alloy at 1100C (point A, Figure 9.3a). At this composition and temperature, only the a phase is present, having a composition of 60 wt%
Ni–40 wt% Cu.
For an alloy having composition and temperature located in a two-phase region, the situation is more complicated. In all two-phase regions (and in two-phase
regions only), one may imagine a series of horizontal lines, one at every temperature; each of these is known as a tie line, or sometimes as an isotherm. These tie
1496T_c09_252-310 11/29/05 11:33 Page 261
REVISED PAGES
9.8 Interpretation of Phase Diagrams • 261
lines extend across the two-phase region and terminate at the phase boundary lines
on either side. To compute the equilibrium concentrations of the two phases, the
following procedure is used:
1. A tie line is constructed across the two-phase region at the temperature of
the alloy.
2. The intersections of the tie line and the phase boundaries on either side are
noted.
3. Perpendiculars are dropped from these intersections to the horizontal composition axis, from which the composition of each of the respective phases is read.
For example, consider again the 35 wt% Ni–65 wt% Cu alloy at 1250C, located
at point B in Figure 9.3b and lying within the a L region. Thus, the problem is to
determine the composition (in wt% Ni and Cu) for both the a and liquid phases. The
tie line has been constructed across the a L phase region, as shown in Figure 9.3b.
The perpendicular from the intersection of the tie line with the liquidus boundary
meets the composition axis at 31.5 wt% Ni–68.5 wt% Cu, which is the composition
of the liquid phase, CL. Likewise, for the solidus–tie line intersection, we find a composition for the a solid-solution phase, Ca, of 42.5 wt% Ni–57.5 wt% Cu.
Determination of Phase Amounts
Bi-Sb
Ge-Si
lever rule
The relative amounts (as fraction or as percentage) of the phases present at equilibrium may also be computed with the aid of phase diagrams. Again, the single- and
two-phase situations must be treated separately.The solution is obvious in the singlephase region: Since only one phase is present, the alloy is composed entirely of that
phase; that is, the phase fraction is 1.0 or, alternatively, the percentage is 100%. From
the previous example for the 60 wt% Ni–40 wt% Cu alloy at 1100C (point A in
Figure 9.3a), only the a phase is present; hence, the alloy is completely or 100% a.
If the composition and temperature position is located within a two-phase region,
things are more complex. The tie line must be utilized in conjunction with a procedure that is often called the lever rule (or the inverse lever rule), which is applied
as follows:
1. The tie line is constructed across the two-phase region at the temperature of
the alloy.
2. The overall alloy composition is located on the tie line.
3. The fraction of one phase is computed by taking the length of tie line from
the overall alloy composition to the phase boundary for the other phase, and
dividing by the total tie line length.
4. The fraction of the other phase is determined in the same manner.
5. If phase percentages are desired, each phase fraction is multiplied by 100.
When the composition axis is scaled in weight percent, the phase fractions
computed using the lever rule are mass fractions—the mass (or weight) of a
specific phase divided by the total alloy mass (or weight). The mass of each
phase is computed from the product of each phase fraction and the total
alloy mass.
In the employment of the lever rule, tie line segment lengths may be determined either by direct measurement from the phase diagram using a linear scale,
preferably graduated in millimeters, or by subtracting compositions as taken from
the composition axis.
1496T_c09_252-310 11/29/05 11:33 Page 262
REVISED PAGES
262 • Chapter 9 / Phase Diagrams
Consider again the example shown in Figure 9.3b, in which at 1250C both a
and liquid phases are present for a 35 wt% Ni–65 wt% Cu alloy. The problem is to
compute the fraction of each of the a and liquid phases. The tie line has been constructed that was used for the determination of a and L phase compositions. Let
the overall alloy composition be located along the tie line and denoted as C0, and
mass fractions be represented by WL and Wa for the respective phases. From the
lever rule, WL may be computed according to
S
RS
(9.1a)
Ca C0
Ca CL
(9.1b)
WL
or, by subtracting compositions,
Lever rule
expression for
computation of
liquid mass fraction
(per Figure 9.3b)
WL
Composition need be specified in terms of only one of the constituents for a binary
alloy; for the computation above, weight percent nickel will be used (i.e., C0 35
wt% Ni, Ca 42.5 wt% Ni, and CL 31.5 wt% Ni), and
WL
42.5 35
0.68
42.5 31.5
Wa
R
RS
(9.2a)
C0 CL
Ca CL
(9.2b)
35 31.5
0.32
42.5 31.5
Similarly, for the a phase,
Lever rule
expression for
computation of
a-phase mass
fraction (per
Figure 9.3b)
Of course, identical answers are obtained if compositions are expressed in weight
percent copper instead of nickel.
Thus, the lever rule may be employed to determine the relative amounts or
fractions of phases in any two-phase region for a binary alloy if the temperature
and composition are known and if equilibrium has been established. Its derivation
is presented as an example problem.
It is easy to confuse the foregoing procedures for the determination of phase
compositions and fractional phase amounts; thus, a brief summary is warranted. Compositions of phases are expressed in terms of weight percents of the components (e.g.,
wt% Cu, wt% Ni). For any alloy consisting of a single phase, the composition of
that phase is the same as the total alloy composition. If two phases are present, the
tie line must be employed, the extremities of which determine the compositions of
the respective phases. With regard to fractional phase amounts (e.g., mass fraction
of the a or liquid phase), when a single phase exists, the alloy is completely that
phase. For a two-phase alloy, on the other hand, the lever rule is utilized, in which
a ratio of tie line segment lengths is taken.
1496T_c09_252-310 11/29/05 11:33 Page 263
REVISED PAGES
9.8 Interpretation of Phase Diagrams • 263
Concept Check 9.2
A copper–nickel alloy of composition 70 wt% Ni–30 wt% Cu is slowly heated from
a temperature of 1300C (2370F).
(a)
(b)
(c)
(d)
At what temperature does the first liquid phase form?
What is the composition of this liquid phase?
At what temperature does complete melting of the alloy occur?
What is the composition of the last solid remaining prior to complete
melting?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Concept Check 9.3
Is it possible to have a copper–nickel alloy that, at equilibrium, consists of an a
phase of composition 37 wt% Ni–63 wt% Cu, and also a liquid phase of composition 20 wt% Ni–80 wt% Cu? If so, what will be the approximate temperature of
the alloy? If this is not possible, explain why.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
EXAMPLE PROBLEM 9.1
Lever Rule Derivation
Derive the lever rule.
Solution
Consider the phase diagram for copper and nickel (Figure 9.3b) and alloy of
composition C0 at 1250C, and let Ca, CL, Wa, and WL represent the same parameters as above. This derivation is accomplished through two conservationof-mass expressions. With the first, since only two phases are present, the sum
of their mass fractions must be equal to unity; that is,
Wa WL 1
(9.3)
For the second, the mass of one of the components (either Cu or Ni) that is
present in both of the phases must be equal to the mass of that component in
the total alloy, or
WaCa WLCL C0
(9.4)
Simultaneous solution of these two equations leads to the lever rule expressions for this particular situation, Equations 9.1b and 9.2b:
Ca C0
Ca CL
C0 CL
Wa
Ca CL
WL
(9.1b)
(9.2b)
1496T_c09_252-310 11/29/05 12:46 Page 264
REVISED PAGES
264 • Chapter 9 / Phase Diagrams
For multiphase alloys, it is often more convenient to specify relative phase
amount in terms of volume fraction rather than mass fraction. Phase volume
fractions are preferred because they (rather than mass fractions) may be determined from examination of the microstructure; furthermore, the properties of a
multiphase alloy may be estimated on the basis of volume fractions.
For an alloy consisting of a and b phases, the volume fraction of the a phase,
Va, is defined as
a phase volume
fraction—
dependence on
volumes of a and b
phases
Va
va
va vb
(9.5)
where va and vb denote the volumes of the respective phases in the alloy. Of course,
an analogous expression exists for Vb, and, for an alloy consisting of just two phases,
it is the case that Va Vb 1.
On occasion conversion from mass fraction to volume fraction (or vice versa)
is desired. Equations that facilitate these conversions are as follows:
Va
Conversion of mass
fractions of a and b
phases to volume
fractions
Wa
ra
Wb
Wa
ra
rb
(9.6a)
Wb
Vb
rb
Wb
Wa
ra
rb
(9.6b)
and
Wa
Conversion of
volume fractions of
a and b phases to
mass fractions
Wb
Vara
Vara Vbrb
Vbrb
Vara Vbrb
(9.7a)
(9.7b)
In these expressions, ra and rb are the densities of the respective phases; these may
be determined approximately using Equations 4.10a and 4.10b.
When the densities of the phases in a two-phase alloy differ significantly, there
will be quite a disparity between mass and volume fractions; conversely, if the phase
densities are the same, mass and volume fractions are identical.
9.9 DEVELOPMENT OF MICROSTRUCTURE IN
ISOMORPHOUS ALLOYS
Equilibrium Cooling
At this point it is instructive to examine the development of microstructure that
occurs for isomorphous alloys during solidification. We first treat the situation in
which the cooling occurs very slowly, in that phase equilibrium is continuously
maintained.
1496T_c09_252-310 11/29/05 11:33 Page 265
REVISED PAGES
9.9 Development of Microstructure in Isomorphous Alloys • 265
Figure 9.4
Schematic
representation of the
development of
microstructure
during the
equilibrium
solidification of a
35 wt% Ni–65 wt%
Cu alloy.
L
L
(35 Ni)
L
(35 Ni)
␣ (46 Ni)
1300
a
␣
+
L
L (32 Ni)
b
␣ (46 Ni)
Temperature (°C)
c
␣ (43 Ni)
␣ (43 Ni)
L (24 Ni)
d
␣
␣
L (32 Ni)
␣
␣
1200
L (24 Ni)
e
␣ (35 Ni)
␣
␣
␣
␣ ␣
␣ ␣ ␣
␣ ␣
␣
␣
␣
␣
␣ (35 Ni)
1100
20
␣
30
40
␣
␣
50
Composition (wt% Ni)
Bi-Sb
Ge-Si
Let us consider the copper–nickel system (Figure 9.3a), specifically an alloy
of composition 35 wt% Ni–65 wt% Cu as it is cooled from 1300C. The region of
the Cu–Ni phase diagram in the vicinity of this composition is shown in Figure 9.4.
Cooling of an alloy of the above composition corresponds to moving down the
vertical dashed line. At 1300C, point a, the alloy is completely liquid (of composition 35 wt% Ni–65 wt% Cu) and has the microstructure represented by the circle inset in the figure. As cooling begins, no microstructural or compositional
changes will be realized until we reach the liquidus line (point b, 1260C). At this
point, the first solid a begins to form, which has a composition dictated by the tie
line drawn at this temperature [i.e., 46 wt% Ni–54 wt% Cu, noted as a(46 Ni)];
the composition of liquid is still approximately 35 wt% Ni–65 wt% Cu [L(35 Ni)],
which is different from that of the solid a. With continued cooling, both compositions and relative amounts of each of the phases will change. The compositions
of the liquid and a phases will follow the liquidus and solidus lines, respectively.
Furthermore, the fraction of the a phase will increase with continued cooling.
Note that the overall alloy composition (35 wt% Ni–65 wt% Cu) remains unchanged during cooling even though there is a redistribution of copper and nickel
between the phases.
At 1250C, point c in Figure 9.4, the compositions of the liquid and a phases
are 32 wt% Ni–68 wt% Cu [L(32 Ni)] and 43 wt% Ni–57 wt% Cu [a(43 Ni)],
respectively.
1496T_c09_252-310 11/29/05 11:33 Page 266
REVISED PAGES
266 • Chapter 9 / Phase Diagrams
The solidification process is virtually complete at about 1220C, point d; the
composition of the solid a is approximately 35 wt% Ni–65 wt% Cu (the overall
alloy composition) while that of the last remaining liquid is 24 wt% Ni–76 wt%
Cu. Upon crossing the solidus line, this remaining liquid solidifies; the final product then is a polycrystalline a-phase solid solution that has a uniform 35 wt% Ni–65
wt% Cu composition (point e, Figure 9.4). Subsequent cooling will produce no
microstructural or compositional alterations.
Nonequilibrium Cooling
Conditions of equilibrium solidification and the development of microstructures,
as described in the previous section, are realized only for extremely slow cooling
rates. The reason for this is that with changes in temperature, there must be readjustments in the compositions of the liquid and solid phases in accordance with
the phase diagram (i.e., with the liquidus and solidus lines), as discussed. These
readjustments are accomplished by diffusional processes—that is, diffusion in both
solid and liquid phases and also across the solid–liquid interface. Inasmuch as diffusion is a time-dependent phenomenon (Section 5.3), to maintain equilibrium
during cooling, sufficient time must be allowed at each temperature for the appropriate compositional readjustments. Diffusion rates (i.e., the magnitudes of the
diffusion coefficients) are especially low for the solid phase and, for both phases,
decrease with diminishing temperature. In virtually all practical solidification situations, cooling rates are much too rapid to allow these compositional readjustments and maintenance of equilibrium; consequently, microstructures other than
those previously described develop.
Some of the consequences of nonequilibrium solidification for isomorphous alloys will now be discussed by considering a 35 wt% Ni–65 wt% Cu alloy, the same
composition that was used for equilibrium cooling in the previous section. The portion of the phase diagram near this composition is shown in Figure 9.5; in addition,
microstructures and associated phase compositions at various temperatures upon
cooling are noted in the circular insets. To simplify this discussion it will be assumed
that diffusion rates in the liquid phase are sufficiently rapid such that equilibrium
is maintained in the liquid.
Let us begin cooling from a temperature of about 1300C; this is indicated by
point a¿ in the liquid region. This liquid has a composition of 35 wt% Ni–65 wt%
Cu [noted as L(35 Ni) in the figure], and no changes occur while cooling through
the liquid phase region (moving down vertically from point a¿ ). At point b¿ (approximately 1260C), a-phase particles begin to form, which, from the tie line constructed, have a composition of 46 wt% Ni–54 wt% Cu [a(46 Ni)].
Upon further cooling to point c¿ (about 1240C), the liquid composition has
shifted to 29 wt% Ni–71 wt% Cu; furthermore, at this temperature the composition of the a phase that solidified is 40 wt% Ni–60 wt% Cu [a(40 Ni)]. However,
since diffusion in the solid a phase is relatively slow, the a phase that formed at
point b has not changed composition appreciably—that is, it is still about 46 wt%
Ni—and the composition of the a grains has continuously changed with radial position, from 46 wt% Ni at grain centers to 40 wt% Ni at the outer grain perimeters. Thus, at point c¿, the average composition of the solid a grains that have formed
would be some volume weighted average composition, lying between 46 and
40 wt% Ni. For the sake of argument, let us take this average composition to be
42 wt% Ni–58 wt% Cu [a(42 Ni)]. Furthermore, we would also find that, on the
basis of lever-rule computations, a greater proportion of liquid is present for these
1496T_c09_252-310 1/10/06 23:53 Page 267
2nd REVISE PAGES
9.9 Development of Microstructure in Isomorphous Alloys • 267
Figure 9.5
Schematic
representation of the
development of
microstructure
during the
nonequilibrium
solidification of a
35 wt% Ni–65 wt%
Cu alloy.
L
L
(35 Ni)
L (35 Ni)
1300
␣ (46 Ni)
␣ +L
a
␣
b
␣ (40 Ni)
L (29 Ni)
␣ (46 Ni)
Temperature (°C)
c
L (24 Ni)
␣ (42 Ni)
d
L (21 Ni)
␣ (35 Ni)
e
1200
L (29 Ni)
␣ (46 Ni)
␣ (40 Ni)
␣ (38 Ni)
␣ (31 Ni)
L (24 Ni)
␣ (46 Ni)
␣ (40 Ni)
␣ (35 Ni)
f
␣ (46 Ni)
␣ (40 Ni)
␣ (35 Ni)
␣ (31 Ni)
L (21 Ni)
␣ (46 Ni)
␣ (40 Ni)
␣ (35 Ni)
␣ (31 Ni)
1100
20
30
40
Composition (wt% Ni)
50
60
nonequilibrium conditions than for equilibrium cooling. The implication of this
nonequilibrium solidification phenomenon is that the solidus line on the phase
diagram has been shifted to higher Ni contents—to the average compositions of
the a phase (e.g., 42 wt% Ni at 1240C)—and is represented by the dashed line in
Figure 9.5. There is no comparable alteration of the liquidus line inasmuch as it is
assumed that equilibrium is maintained in the liquid phase during cooling because
of sufficiently rapid diffusion rates.
At point d¿ (1220C) and for equilibrium cooling rates, solidification should
be completed. However, for this nonequilibrium situation, there is still an appreciable proportion of liquid remaining, and the a phase that is forming has a composition of 35 wt% Ni [a(35 Ni)]; also the average a-phase composition at this point
is 38 wt% Ni [a(38 Ni)].
Nonequilibrium solidification finally reaches completion at point e¿ (1205C).
The composition of the last a phase to solidify at this point is about 31 wt% Ni; the
average composition of the a phase at complete solidification is 35 wt% Ni. The
inset at point f ¿ shows the microstructure of the totally solid material.
The degree of displacement of the nonequilibrium solidus curve from the equilibrium one will depend on rate of cooling. The slower the cooling rate, the smaller
1496T_c09_252-310 11/29/05 11:33 Page 268
REVISED PAGES
268 • Chapter 9 / Phase Diagrams
this displacement; that is, the difference between the equilibrium solidus and average solid composition is lower. Furthermore, if the diffusion rate in the solid phase
is increased, this displacement will be diminished.
There are some important consequences for isomorphous alloys that have solidified under nonequilibrium conditions. As discussed above, the distribution of
the two elements within the grains is nonuniform, a phenomenon termed segregation; that is, concentration gradients are established across the grains that are represented by the insets of Figure 9.5. The center of each grain, which is the first
part to freeze, is rich in the high-melting element (e.g., nickel for this Cu–Ni system), whereas the concentration of the low-melting element increases with position from this region to the grain boundary. This is termed a cored structure, which
gives rise to less than the optimal properties. As a casting having a cored structure is reheated, grain boundary regions will melt first inasmuch as they are richer
in the low-melting component. This produces a sudden loss in mechanical integrity
due to the thin liquid film that separates the grains. Furthermore, this melting may
begin at a temperature below the equilibrium solidus temperature of the alloy.
Coring may be eliminated by a homogenization heat treatment carried out at a
temperature below the solidus point for the particular alloy composition. During
this process, atomic diffusion occurs, which produces compositionally homogeneous grains.
9.10 MECHANICAL PROPERTIES OF
ISOMORPHOUS ALLOYS
We shall now briefly explore how the mechanical properties of solid isomorphous
alloys are affected by composition as other structural variables (e.g., grain size) are
held constant. For all temperatures and compositions below the melting temperature of the lowest-melting component, only a single solid phase will exist. Therefore, each component will experience solid-solution strengthening (Section 7.9), or
an increase in strength and hardness by additions of the other component. This
effect is demonstrated in Figure 9.6a as tensile strength versus composition for the
50
300
40
30
200
0
(Cu)
20
40
60
Composition (wt% Ni)
(a)
80
100
(Ni)
Tensile strength (ksi)
Tensile strength (MPa)
60
400
Elongation (% in 50 mm [2 in.])
60
50
40
30
20
0
(Cu)
20
40
60
Composition (wt% Ni)
80
100
(Ni)
(b)
Figure 9.6 For the copper–nickel system, (a) tensile strength versus composition, and
(b) ductility (%EL) versus composition at room temperature. A solid solution exists over
all compositions for this system.
1496T_c09_252-310 11/29/05 11:33 Page 269
REVISED PAGES
9.11 Binary Eutectic Systems • 269
copper–nickel system at room temperature; at some intermediate composition, the
curve necessarily passes through a maximum. Plotted in Figure 9.6b is the ductility
(%EL)–composition behavior, which is just the opposite of tensile strength; that is,
ductility decreases with additions of the second component, and the curve exhibits
a minimum.
9.11 BINARY EUTECTIC SYSTEMS
Another type of common and relatively simple phase diagram found for binary alloys is shown in Figure 9.7 for the copper–silver system; this is known as a binary
eutectic phase diagram. A number of features of this phase diagram are important
and worth noting. First, three single-phase regions are found on the diagram: a, b,
and liquid. The a phase is a solid solution rich in copper; it has silver as the solute
component and an FCC crystal structure. The -phase solid solution also has an
FCC structure, but copper is the solute. Pure copper and pure silver are also considered to be a and b phases, respectively.
Thus, the solubility in each of these solid phases is limited, in that at any temperature below line BEG only a limited concentration of silver will dissolve in
copper (for the a phase), and similarly for copper in silver (for the b phase). The
solubility limit for the a phase corresponds to the boundary line, labeled CBA, between the a(a b) and a(a L) phase regions; it increases with temperature to
Composition (at% Ag)
0
20
40
60
80
100
2200
1200
A
2000
Liquidus
1000
Liquid
1800
F
␣ +L
␣
800
1600
779°C (TE)
B
+L
E
8.0
(C␣ E)
71.9
(CE)
91.2
(C E)
G
1400

1200
600
1000
Solvus
␣ +
800
400
C
600
H
200
0
(Cu)
20
40
60
Composition (wt% Ag)
80
400
100
(Ag)
Figure 9.7 The copper–silver phase diagram. [Adapted from Binary Alloy Phase
Diagrams, 2nd edition, Vol. 1, T. B. Massalski (Editor-in-Chief), 1990. Reprinted by
permission of ASM International, Materials Park, OH.]
Temperature (°F)
Temperature (°C)
Solidus
1496T_c09_252-310 11/29/05 11:33 Page 270
REVISED PAGES
270 • Chapter 9 / Phase Diagrams
solvus line
solidus line
liquidus line
invariant point
The eutectic reaction
(per Figure 9.7)
a maximum [8.0 wt% Ag at 779C (1434F)] at point B, and decreases back to zero
at the melting temperature of pure copper, point A [1085C (1985F)]. At temperatures below 779C (1434F), the solid solubility limit line separating the a and
a b phase regions is termed a solvus line; the boundary AB between the a and
a L fields is the solidus line, as indicated in Figure 9.7. For the b phase, both
solvus and solidus lines also exist, HG and GF, respectively, as shown. The maximum solubility of copper in the b phase, point G (8.8 wt% Cu), also occurs at 779C
(1434F). This horizontal line BEG, which is parallel to the composition axis and
extends between these maximum solubility positions, may also be considered a
solidus line; it represents the lowest temperature at which a liquid phase may exist
for any copper–silver alloy that is at equilibrium.
There are also three two-phase regions found for the copper–silver system (Figure 9.7): a L, b L, and a b. The a- and b-phase solid solutions coexist for
all compositions and temperatures within the a b phase field; the a liquid and
b liquid phases also coexist in their respective phase regions. Furthermore, compositions and relative amounts for the phases may be determined using tie lines
and the lever rule as outlined previously.
As silver is added to copper, the temperature at which the alloys become totally liquid decreases along the liquidus line, line AE; thus, the melting temperature of copper is lowered by silver additions. The same may be said for silver: the
introduction of copper reduces the temperature of complete melting along the other
liquidus line, FE. These liquidus lines meet at the point E on the phase diagram,
through which also passes the horizontal isotherm line BEG. Point E is called an
invariant point, which is designated by the composition CE and temperature TE; for
the copper–silver system, the values of CE and TE are 71.9 wt% Ag and 779C
(1434F), respectively.
An important reaction occurs for an alloy of composition CE as it changes temperature in passing through TE; this reaction may be written as follows:
L1CE 2 Δ a 1CaE 2 b1CbE 2
cooling
heating
eutectic reaction
(9.8)
Or, upon cooling, a liquid phase is transformed into the two solid a and b phases
at the temperature TE; the opposite reaction occurs upon heating. This is called a
eutectic reaction (eutectic means easily melted), and CE and TE represent the eutectic composition and temperature, respectively; CaE and CbE are the respective
compositions of the a and b phases at TE. Thus, for the copper–silver system, the
eutectic reaction, Equation 9.8, may be written as follows:
cooling
L171.9 wt% Ag2 Δ a18.0 wt%Ag2 b191.2 wt% Ag2
heating
Often, the horizontal solidus line at TE is called the eutectic isotherm.
The eutectic reaction, upon cooling, is similar to solidification for pure components in that the reaction proceeds to completion at a constant temperature, or
isothermally, at TE. However, the solid product of eutectic solidification is always
two solid phases, whereas for a pure component only a single phase forms. Because
of this eutectic reaction, phase diagrams similar to that in Figure 9.7 are termed
eutectic phase diagrams; components exhibiting this behavior comprise a eutectic
system.
1496T_c09_252-310 11/29/05 11:33 Page 271
REVISED PAGES
9.11 Binary Eutectic Systems • 271
In the construction of binary phase diagrams, it is important to understand that
one or at most two phases may be in equilibrium within a phase field. This holds
true for the phase diagrams in Figures 9.3a and 9.7. For a eutectic system, three
phases (a, b, and L) may be in equilibrium, but only at points along the eutectic
isotherm. Another general rule is that single-phase regions are always separated
from each other by a two-phase region that consists of the two single phases that
it separates. For example, the a b field is situated between the a and b singlephase regions in Figure 9.7.
Another common eutectic system is that for lead and tin; the phase diagram
(Figure 9.8) has a general shape similar to that for copper–silver. For the lead–tin
system the solid solution phases are also designated by a and b; in this case, a represents a solid solution of tin in lead and, for b, tin is the solvent and lead is the
solute. The eutectic invariant point is located at 61.9 wt% Sn and 183C (361F). Of
course, maximum solid solubility compositions as well as component melting temperatures will be different for the copper–silver and lead–tin systems, as may be
observed by comparing their phase diagrams.
On occasion, low-melting-temperature alloys are prepared having near-eutectic
compositions. A familiar example is the 60–40 solder, containing 60 wt% Sn and
40 wt% Pb. Figure 9.8 indicates that an alloy of this composition is completely
molten at about 185C (365F), which makes this material especially attractive as a
low-temperature solder, since it is easily melted.
Composition (at% Sn)
0
20
40
60
80
100
327°C
600
300
Liquid
500
Temperature (°C)
␣ +L
␣
200
 +L
183°C
400

18.3
61.9
97.8
300
100
␣ + 
200
100
0
0
(Pb)
20
40
60
Composition (wt% Sn)
80
100
(Sn)
Figure 9.8 The lead–tin phase diagram. [Adapted from Binary Alloy Phase Diagrams,
2nd edition, Vol. 3, T. B. Massalski (Editor-in-Chief), 1990. Reprinted by permission of
ASM International, Materials Park, OH.]
Temperature (°F)
232°C
1496T_c09_252-310 11/29/05 11:33 Page 272
REVISED PAGES
272 • Chapter 9 / Phase Diagrams
Concept Check 9.4
At 700C (1290F), what is the maximum solubility (a) of Cu in Ag? (b) Of Ag
in Cu?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Concept Check 9.5
Below is a portion of the H2O–NaCl phase diagram:
10
50
Liquid
(brine)
40
10
30
Salt
Liquid
(brine)
Ice
Liquid
(brine)
20
10
Temperature (°C)
Temperature (°C)
0
0
20
10
Ice Salt
NaCl
H2O
30
0
100
20
10
90
20
80
30
70
Composition (wt%)
(a) Using this diagram, briefly explain how spreading salt on ice that is at a temperature below 0C (32F) can cause the ice to melt.
(b) At what temperature is salt no longer useful in causing ice to melt?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
EXAMPLE PROBLEM 9.2
Determination of Phases Present and Computation
of Phase Compositions
For a 40 wt% Sn–60 wt% Pb alloy at 150C (300F), (a) What phase(s) is (are)
present? (b) What is (are) the composition(s) of the phase(s)?
Solution
(a) Locate this temperature–composition point on the phase diagram (point
B in Figure 9.9). Inasmuch as it is within the a b region, both a and b phases
will coexist.
1496T_c09_252-310 12/21/05 7:56 Page 273
2nd REVISE PAGES
9.11 Binary Eutectic Systems • 273
600
300
Liquid
+L
200
+L
B
300
+
100
400
Temperature (°F)
Temperature (°C)
500
200
C
100
0
0
(Pb)
20
60
C
80
C1
100
(Sn)
Composition (wt% Sn)
Figure 9.9 The lead–tin phase diagram. For a 40 wt% Sn–60 wt% Pb alloy at 150C
(point B), phase compositions and relative amounts are computed in Example
Problems 9.2 and 9.3.
(b) Since two phases are present, it becomes necessary to construct a tie line
across the a b phase field at 150C, as indicated in Figure 9.9. The composition of the a phase corresponds to the tie line intersection with the a (a b)
solvus phase boundary—about 10 wt% Sn–90 wt% Pb, denoted as Ca. Similarly for the b phase, which will have a composition of approximately 98 wt%
Sn–2 wt% Pb (Cb).
EXAMPLE PROBLEM 9.3
Relative Phase Amount Determinations—Mass and
Volume Fractions
For the lead–tin alloy in Example Problem 9.2, calculate the relative amount of
each phase present in terms of (a) mass fraction and (b) volume fraction. At
150C take the densities of Pb and Sn to be 11.23 and 7.24 g/cm3, respectively.
Solution
(a) Since the alloy consists of two phases, it is necessary to employ the lever
rule. If C1 denotes the overall alloy composition, mass fractions may be computed by subtracting compositions, in terms of weight percent tin, as follows:
Cb C1
98 40
0.66
Cb Ca
98 10
C1 Ca
40 10
Wb
0.34
Cb Ca
98 10
Wa
1496T_c09_252-310 11/29/05 11:33 Page 274
REVISED PAGES
274 • Chapter 9 / Phase Diagrams
(b) To compute volume fractions it is first necessary to determine the density
of each phase using Equation 4.10a. Thus
ra
CSn1a2
rSn
100
CPb1a2
rPb
where CSn(a) and CPb(a) denote the concentrations in weight percent of tin and
lead, respectively, in the a phase. From Example Problem 9.2, these values are
10 wt% and 90 wt%. Incorporation of these values along with the densities
of the two components lead to
ra
100
10.64 g/cm3
10
90
7.24 g/cm3
11.23 g/cm3
Similarly for the b phase:
rb C
Sn1b2
rSn
100
CPb1b2
rPb
100
98
2
7.24 g/cm3
11.23 g/cm3
7.29 g/cm3
Now it becomes necessary to employ Equations 9.6a and 9.6b to determine
Va and Vb as
Va
Wa
ra
Wb
Wa
ra
rb
0.66
10.64 g/cm3
0.57
0.66
0.34
10.64 g/cm3
7.29 g/cm3
Wb
Vb
rb
Wa
Wb
ra
rb
0.34
7.29 g/cm3
0.43
0.66
0.34
10.64 g/cm3
7.29 g/cm3
1496T_c09_252-310 12/21/05 7:56 Page 275
2nd REVISE PAGES
9.11 Binary Eutectic Systems • 275
MATERIALS OF IMPORTANCE
Lead-Free Solders
olders are metal alloys that are used to bond
or join two or more components (usually other
metal alloys). They are used extensively in the
electronics industry to physically hold assemblies
together; furthermore, they must allow expansion
and contraction of the various components, must
transmit electrical signals, and also dissipate any
heat that is generated. The bonding action is accomplished by melting the solder material, allowing it to flow among and make contact with the
components to be joined (which do not melt),
and, finally, upon solidification, forming a physical
bond with all of these components.
In the past, the vast majority of solders have
been lead-tin alloys. These materials are reliable,
inexpensive, and have relatively low melting temperatures. The most common lead–tin solder has a
composition of 63 wt% Sn–37 wt% Pb. According
to the lead–tin phase diagram, Figure 9.8, this composition is near the eutectic and has melting temperature of about 183C, the lowest temperature
possible with the existence of a liquid phase (at
equilibrium) for the lead–tin system. It follows that
this alloy is often called a “eutectic lead-tin solder.”
Unfortunately, lead is a mildly toxic metal, and
there is serious concern about the environmental
impact of discarded lead-containing products that
can leach into groundwater from landfills or pollute the air if incinerated. Consequently, in some
countries legislation has been enacted that bans
the use of lead-containing solders. This has forced
the development of lead-free solders that, among
other things, must have relatively low melting temperatures (or temperature ranges). Some of these
are ternary alloys (i.e., composed of three metals),
to include tin–silver–copper and tin–silver–bismuth
solders. The compositions of several lead-free
solders are listed in Table 9.1.
Of course, melting temperatures (or temperature ranges) are important in the development
and selection of these new solder alloys, information that is available from phase diagrams. For example, the tin-bismuth phase diagram is presented
in Figure 9.10. Here it may be noted that a eutectic
Table 9.1 Compositions, Solidus Temperatures, and Liquidus Temperatures
for Five Lead-Free Solders
Composition
(wt%)
Solidus
Temperature (C )
Liquidus
Temperature (C )
118
139
211
118
139
213
217
217
227
227
52 In/48 Sn*
57 Bi/43 Sn*
91.8 Sn/3.4 Ag/
4.8 Bi
95.5 Sn/3.8 Ag/
0.7 Cu*
99.3 Sn/0.7 Cu*
*The compositions of these alloys are eutectic compositions; therefore, their solidus and liquidus temperatures are identical.
Source: Adapted from E. Bastow, “Solder Families and
How They Work,”Advanced Materials & Processes,
Vol. 161, No. 12, M. W. Hunt (Editor-in-chief), ASM
International, 2003, p. 28. Reprinted by permission of
ASM International, Materials Park, OH.
exists at 57 wt% Bi and 139C, which are indeed
the composition and melting temperature of the
Bi–Sn soldier in Table 9.1
Composition (at% Bi)
300
0
20
40
60
80
100
271°C
L
232°C
Temperature (°C)
S
200
Bi + L
+L
139°C
Sn
21
57
100
Sn + Bi
Bi
Sn
13°C
0
0
20
40
60
(Sn)
80
100
(Bi)
Composition (wt% Bi)
Figure 9.10 The tin–bismuth phase diagram.
[Adapted from ASM Handbook, Vol. 3, Alloy Phase
Diagrams, H. Baker (Editor), ASM International,
1992, p. 2.106. Reprinted by permission of ASM
International, Materials Park, OH.]
1496T_c09_252-310 11/29/05 11:33 Page 276
REVISED PAGES
276 • Chapter 9 / Phase Diagrams
9.12 DEVELOPMENT OF MICROSTRUCTURE IN
EUTECTIC ALLOYS
Depending on composition, several different types of microstructures are possible
for the slow cooling of alloys belonging to binary eutectic systems. These possibilities will be considered in terms of the lead–tin phase diagram, Figure 9.8.
The first case is for compositions ranging between a pure component and the
maximum solid solubility for that component at room temperature [20C (70F)].
For the lead–tin system, this includes lead-rich alloys containing between 0 and
about 2 wt% Sn (for the a phase solid solution), and also between approximately
99 wt% Sn and pure tin (for the b phase). For example, consider an alloy of composition C1 (Figure 9.11) as it is slowly cooled from a temperature within the liquid-phase region, say, 350C; this corresponds to moving down the dashed vertical
line ww¿ in the figure. The alloy remains totally liquid and of composition C1 until
we cross the liquidus line at approximately 330C, at which time the solid a phase
begins to form. While passing through this narrow a L phase region, solidification proceeds in the same manner as was described for the copper–nickel alloy in the
preceding section; that is, with continued cooling more of the solid a forms. Furthermore, liquid- and solid-phase compositions are different, which follow along the
liquidus and solidus phase boundaries, respectively. Solidification reaches completion at the point where ww¿ crosses the solidus line. The resulting alloy is polycrystalline with a uniform composition of C1, and no subsequent changes will occur
Figure 9.11 Schematic
representations of the
equilibrium microstructures for
a lead–tin alloy of composition
C1 as it is cooled from the
liquid-phase region.
400
L
w
(C1 wt% Sn)
L
a
b
L
300
Liquidus
c
Temperature (°C)
+L
Solidus
200
(C1 wt% Sn)
100
+
w
0
10
C1
20
Composition (wt% Sn)
30
1496T_c09_252-310 11/29/05 15:20 Page 277
REVISED PAGES
9.12 Development of Microstructure in Eutectic Alloys • 277
upon cooling to room temperature. This microstructure is represented schematically
by the inset at point c in Figure 9.11.
The second case considered is for compositions that range between the room
temperature solubility limit and the maximum solid solubility at the eutectic temperature. For the lead–tin system (Figure 9.8), these compositions extend from about
2 wt% Sn to 18.3 wt% Sn (for lead-rich alloys) and from 97.8 wt% Sn to approximately 99 wt% Sn (for tin-rich alloys). Let us examine an alloy of composition C2
as it is cooled along the vertical line xx¿ in Figure 9.12. Down to the intersection of
xx¿ and the solvus line, changes that occur are similar to the previous case, as we
pass through the corresponding phase regions (as demonstrated by the insets at
points d, e, and f ). Just above the solvus intersection, point f, the microstructure
consists of a grains of composition C2. Upon crossing the solvus line, the a solid
solubility is exceeded, which results in the formation of small b-phase particles;
these are indicated in the microstructure inset at point g. With continued cooling,
these particles will grow in size because the mass fraction of the b phase increases
slightly with decreasing temperature.
The third case involves solidification of the eutectic composition, 61.9 wt% Sn
(C3 in Figure 9.13). Consider an alloy having this composition that is cooled from
a temperature within the liquid-phase region (e.g., 250C) down the vertical line yy¿
x
d
Figure 9.12 Schematic
representations of the
equilibrium microstructures
for a lead–tin alloy of
composition C2 as it is cooled
from the liquid-phase region.
L
L
(C2 wt% Sn)
300
␣
L
e
Temperature (°C)
␣
␣
␣ +L
␣
200
␣
␣
C2 wt% Sn
f

Solvus
line
g
␣
100
␣ +
x
0
10
20
30
C2
Composition (wt% Sn)
40
50
1496T_c09_252-310 1/9/06 13:00 Page 278
2nd REVISE PAGES
278 • Chapter 9 / Phase Diagrams
300
600
y
L
L
(61.9 wt%
Sn)
500
h
200
␣
+L
183°C
97.8
i
18.3
 400
300
␣+
100
Temperature (°F)
␣ +L
Temperature (°C)
Figure 9.13
Schematic
representations of
the equilibrium
microstructures for a
lead–tin alloy of
eutectic composition
C3 above and below
the eutectic
temperature.
200
␣ (18.3 wt%
Sn)
 (97.8 wt%
Sn)
100
y
0
0
20
40
60
C3
(61.9)
(Pb)
80
100
(Sn)
Composition (wt%Sn)
Eutectic, Pb-Sn
in Figure 9.13. As the temperature is lowered, no changes occur until we reach the
eutectic temperature, 183C. Upon crossing the eutectic isotherm, the liquid transforms to the two a and b phases. This transformation may be represented by the
reaction
cooling
L161.9 wt% Sn2 Δ a118.3 wt% Sn2 b197.8 wt% Sn2
heating
eutectic structure
(9.9)
in which the a- and b-phase compositions are dictated by the eutectic isotherm end
points.
During this transformation, there must necessarily be a redistribution of the
lead and tin components, inasmuch as the a and b phases have different compositions neither of which is the same as that of the liquid (as indicated in Equation 9.9).
This redistribution is accomplished by atomic diffusion. The microstructure of the
solid that results from this transformation consists of alternating layers (sometimes
called lamellae) of the a and b phases that form simultaneously during the transformation. This microstructure, represented schematically in Figure 9.13, point i, is
called a eutectic structure and is characteristic of this reaction. A photomicrograph
of this structure for the lead–tin eutectic is shown in Figure 9.14. Subsequent cooling of the alloy from just below the eutectic to room temperature will result in only
minor microstructural alterations.
The microstructural change that accompanies this eutectic transformation is
represented schematically in Figure 9.15; here is shown the a-b layered eutectic
growing into and replacing the liquid phase. The process of the redistribution of
lead and tin occurs by diffusion in the liquid just ahead of the eutectic–liquid interface. The arrows indicate the directions of diffusion of lead and tin atoms; lead
atoms diffuse toward the a-phase layers since this a phase is lead-rich (18.3 wt%
Sn–81.7 wt% Pb); conversely, the direction of diffusion of tin is in the direction of
the b, tin-rich (97.8 wt% Sn–2.2 wt% Pb) layers. The eutectic structure forms in
1496T_c09_252-310 1/9/06 13:00 Page 279
2nd REVISE PAGES
9.12 Development of Microstructure in Eutectic Alloys • 279
Figure 9.14 Photomicrograph showing
the microstructure of a lead–tin alloy of
eutectic composition. This microstructure
consists of alternating layers of a leadrich a-phase solid solution (dark layers),
and a tin-rich b-phase solid solution (light
layers). 375. (Reproduced with
permission from Metals Handbook, 9th
edition, Vol. 9, Metallography and
Microstructures, American Society for
Metals, Materials Park, OH, 1985.)
eutectic phase
primary phase
these alternating layers because, for this lamellar configuration, atomic diffusion of
lead and tin need only occur over relatively short distances.
The fourth and final microstructural case for this system includes all compositions other than the eutectic that, when cooled, cross the eutectic isotherm. Consider, for example, the composition C4, Figure 9.16, which lies to the left of the
eutectic; as the temperature is lowered, we move down the line zz¿, beginning at
point j. The microstructural development between points j and l is similar to that
for the second case, such that just prior to crossing the eutectic isotherm (point l),
the a and liquid phases are present having compositions of approximately 18.3 and
61.9 wt% Sn, respectively, as determined from the appropriate tie line. As the temperature is lowered to just below the eutectic, the liquid phase, which is of the eutectic composition, will transform to the eutectic structure (i.e., alternating a and b
lamellae); insignificant changes will occur with the a phase that formed during cooling through the a L region. This microstructure is represented schematically by
the inset at point m in Figure 9.16. Thus, the a phase will be present both in the eutectic structure and also as the phase that formed while cooling through the a L
phase field. To distinguish one a from the other, that which resides in the eutectic
structure is called eutectic a, while the other that formed prior to crossing the
eutectic isotherm is termed primary a; both are labeled in Figure 9.16. The
photomicrograph in Figure 9.17 is of a lead–tin alloy in which both primary a and
eutectic structures are shown.

␣
Pb
Sn

␣

Figure 9.15 Schematic representation of
the formation of the eutectic structure for
the lead–tin system. Directions of diffusion
of tin and lead atoms are indicated by
blue and red arrows, respectively.
Liquid
Pb
Sn
Pb
Eutectic
growth
direction
1496T_c09_252-310 1/9/06 13:00 Page 280
2nd REVISE PAGES
280 • Chapter 9 / Phase Diagrams
L
(C4 wt% Sn)
z
␣
j
300
600
L
␣ +L
L
200
k
␣
 +L
l
400

m
L (61.9 wt% Sn)
100
Eutectic
structure
300
Temperature (°F)
Temperature (°C)
␣ (18.3
wt% Sn)
500
Primary ␣
(18.3 wt% Sn)
␣ + 
200
Eutectic ␣
(18.3 wt% Sn)
 (97.8 wt% Sn)
100
z
0
0
(Pb)
20
60
C4
(40)
80
100
(Sn)
Composition (wt% Sn)
Figure 9.16 Schematic representations of the equilibrium microstructures for a lead–tin
alloy of composition C4 as it is cooled from the liquid-phase region.
microconstituent
In dealing with microstructures, it is sometimes convenient to use the term
microconstituent—that is, an element of the microstructure having an identifiable
and characteristic structure. For example, in the point m inset, Figure 9.16, there are
two microconstituents—namely, primary a and the eutectic structure. Thus, the
eutectic structure is a microconstituent even though it is a mixture of two phases,
because it has a distinct lamellar structure, with a fixed ratio of the two phases.
Figure 9.17 Photomicrograph
showing the microstructure of a
lead–tin alloy of composition 50 wt%
Sn–50 wt% Pb. This microstructure is
composed of a primary lead-rich a
phase (large dark regions) within a
lamellar eutectic structure consisting
of a tin-rich b phase (light layers)
and a lead-rich a phase (dark layers).
400. (Reproduced with permission
from Metals Handbook, 9th edition,
Vol. 9, Metallography and
Microstructures, American Society for
Metals, Materials Park, OH, 1985.)
1496T_c09_252-310 11/29/05 11:33 Page 281
REVISED PAGES
9.12 Development of Microstructure in Eutectic Alloys • 281
300
Temperature (°C)
L
+L
200
+L
Q
P
Figure 9.18 The lead–tin
phase diagram used in
computations for relative
amounts of primary a and
eutectic microconstituents
for an alloy of composition
C¿4.
R
100
0
(Pb)
(Sn)
18.3
C4
61.9
97.8
Composition (wt% Sn)
It is possible to compute the relative amounts of both eutectic and primary a
microconstituents. Since the eutectic microconstituent always forms from the liquid
having the eutectic composition, this microconstituent may be assumed to have a
composition of 61.9 wt% Sn. Hence, the lever rule is applied using a tie line between the a–(a b) phase boundary (18.3 wt% Sn) and the eutectic composition.
For example, consider the alloy of composition C¿4 in Figure 9.18. The fraction of
the eutectic microconstituent We is just the same as the fraction of liquid WL from
which it transforms, or
Lever rule
expression for
computation
of eutectic
microconstituent
and liquid phase
mass fractions
(composition C¿4,
Figure 9.18)
Lever rule
expression for
computation of
primary a phase
mass fraction
We WL
P
PQ
C¿4 18.3
C¿4 18.3
61.9 18.3
43.6
(9.10)
Furthermore, the fraction of primary a, Wa¿, is just the fraction of the a phase
that existed prior to the eutectic transformation or, from Figure 9.18,
Wa¿
Q
PQ
61.9 C¿4
61.9 C¿4
61.9 18.3
43.6
(9.11)
The fractions of total a, Wa (both eutectic and primary), and also of total b, Wb,
are determined by use of the lever rule and a tie line that extends entirely across
the a b phase field. Again, for an alloy having composition C¿4,
Lever rule
expression for
computation of
total a phase
mass fraction
Wa
QR
PQR
97.8 C¿4
97.8 C¿4
97.8 18.3
79.5
(9.12)
1496T_c09_252-310 11/29/05 11:33 Page 282
REVISED PAGES
282 • Chapter 9 / Phase Diagrams
and
Lever rule
expression for
computation of
total b phase
mass fraction
Wb
P
PQR
C¿4 18.3
C¿4 18.3
97.8 18.3
79.5
(9.13)
Analogous transformations and microstructures result for alloys having compositions to the right of the eutectic (i.e., between 61.9 and 97.8 wt% Sn). However,
below the eutectic temperature, the microstructure will consist of the eutectic and
primary b microconstituents because, upon cooling from the liquid, we pass through
the b liquid phase field.
When, for case 4 (represented in Figure 9.16), conditions of equilibrium are
not maintained while passing through the a (or b) liquid phase region, the following consequences will be realized for the microstructure upon crossing the eutectic isotherm: (1) grains of the primary microconstituent will be cored, that is,
have a nonuniform distribution of solute across the grains; and (2) the fraction of
the eutectic microconstituent formed will be greater than for the equilibrium
situation.
9.13 EQUILIBRIUM DIAGRAMS HAVING
INTERMEDIATE PHASES OR COMPOUNDS
terminal solid
solution
intermediate solid
solution
intermetallic
compound
The isomorphous and eutectic phase diagrams discussed thus far are relatively
simple, but those for many binary alloy systems are much more complex. The eutectic copper–silver and lead–tin phase diagrams (Figures 9.7 and 9.8) have only
two solid phases, a and b; these are sometimes termed terminal solid solutions,
because they exist over composition ranges near the concentration extremities of
the phase diagram. For other alloy systems, intermediate solid solutions (or intermediate phases) may be found at other than the two composition extremes. Such
is the case for the copper–zinc system. Its phase diagram (Figure 9.19) may at first
appear formidable because there are some invariant points and reactions similar
to the eutectic that have not yet been discussed. In addition, there are six different solid solutions—two terminal (a and h) and four intermediate (b, g, d, and ).
(The b¿ phase is termed an ordered solid solution, one in which the copper and
zinc atoms are situated in a specific and ordered arrangement within each unit
cell.) Some phase boundary lines near the bottom of Figure 9.19 are dashed to
indicate that their positions have not been exactly determined. The reason for this
is that at low temperatures, diffusion rates are very slow and inordinately long
times are required for the attainment of equilibrium. Again, only single- and twophase regions are found on the diagram, and the same rules outlined in Section
9.8 are utilized for computing phase compositions and relative amounts. The commercial brasses are copper-rich copper–zinc alloys; for example, cartridge brass
has a composition of 70 wt% Cu–30 wt% Zn and a microstructure consisting of
a single a phase.
For some systems, discrete intermediate compounds rather than solid solutions
may be found on the phase diagram, and these compounds have distinct chemical
formulas; for metal–metal systems, they are called intermetallic compounds. For example, consider the magnesium–lead system (Figure 9.20). The compound Mg2Pb
0
(Cu)
200
400
600
800
1000
20
20
+L
40
+
40
+
+
Composition (wt% Zn)
+
+L
Composition (at% Zn)
60
60
+
+
+
+
L
80
+L
Liquid
80
+
+
L
600
800
1000
(Zn)
400
100
+L
1200
1400
1600
1800
2000
100
2200
Figure 9.19 The copper–zinc phase diagram. [Adapted from Binary Alloy Phase Diagrams, 2nd edition, Vol. 2, T. B. Massalski
(Editor-in-Chief), 1990. Reprinted by permission of ASM International, Materials Park, OH.]
Temperature (°C)
1200
0
1496T_c09_252-310 11/29/05 11:33 Page 283
REVISED PAGES
9.13 Equilibrium Diagrams Having Intermediate Phases or Compounds • 283
Temperature (°F)
1496T_c09_252-310 11/29/05 11:33 Page 284
REVISED PAGES
284 • Chapter 9 / Phase Diagrams
Composition (at% Pb)
0
5
10
700
20
30
40
70 100
L
L
+
Mg2Pb
600
+L
1200
M
1000
400
+
L
L
+
Mg2Pb
600
300
200
+ Mg2Pb
+
Mg2Pb
100
Mg2Pb
0
0
(Mg)
20
40
60
80
Composition (wt% Pb)
800
Temperature (°F)
Temperature (°C)
500
400
200
100
(Pb)
Figure 9.20 The magnesium–lead phase diagram. [Adapted from Phase Diagrams of
Binary Magnesium Alloys, A. A. Nayeb-Hashemi and J. B. Clark (Editors), 1988. Reprinted
by permission of ASM International, Materials Park, OH.]
has a composition of 19 wt% Mg–81 wt% Pb (33 at% Pb), and is represented as a
vertical line on the diagram, rather than as a phase region of finite width; hence,
Mg2Pb can exist by itself only at this precise composition.
Several other characteristics are worth noting for this magnesium–lead system. First, the compound Mg2Pb melts at approximately 550C (1020F), as indicated by point M in Figure 9.20. Also, the solubility of lead in magnesium is rather
extensive, as indicated by the relatively large composition span for the a-phase
field. On the other hand, the solubility of magnesium in lead is extremely limited. This is evident from the very narrow b terminal solid-solution region on the
right or lead-rich side of the diagram. Finally, this phase diagram may be thought
of as two simple eutectic diagrams joined back to back, one for the Mg–Mg2Pb
system and the other for Mg2Pb–Pb; as such, the compound Mg2Pb is really
considered to be a component. This separation of complex phase diagrams into
smaller-component units may simplify them and, furthermore, expedite their
interpretation.
9.14 EUTECTOID AND PERITECTIC REACTIONS
In addition to the eutectic, other invariant points involving three different phases
are found for some alloy systems. One of these occurs for the copper–zinc system
(Figure 9.19) at 560C (1040F) and 74 wt% Zn–26 wt% Cu. A portion of the
1496T_c09_252-310 11/29/05 11:33 Page 285
REVISED PAGES
9.14 Eutectoid and Peritectic Reactions • 285
+L
+
+L
600
P
L
598°C
+
560°C
E
60
1000
+
500
1200
Temperature (°F)
700
Temperature (°C)
Figure 9.21 A region of the
copper–zinc phase diagram that
has been enlarged to show
eutectoid and peritectic invariant
points, labeled E (560C, 74 wt%
Zn) and P (598C, 78.6 wt% Zn),
respectively. [Adapted from
Binary Alloy Phase Diagrams,
2nd edition, Vol. 2, T. B. Massalski
(Editor-in-Chief), 1990. Reprinted
by permission of ASM
International, Materials
Park, OH.]
70
80
+L
90
Composition (wt% Zn)
phase diagram in this vicinity appears enlarged in Figure 9.21. Upon cooling, a
solid d phase transforms into two other solid phases (g and ) according to the
reaction
The eutectoid
reaction (per point
E, Figure 9.21)
eutectoid reaction
peritectic reaction
The peritectic
reaction (per point
P, Figure 9.21)
cooling
d Δ g
heating
(9.14)
The reverse reaction occurs upon heating. It is called a eutectoid (or eutectic-like)
reaction, and the invariant point (point E, Figure 9.21) and the horizontal tie line
at 560C are termed the eutectoid and eutectoid isotherm, respectively. The feature
distinguishing “eutectoid” from “eutectic” is that one solid phase instead of a liquid transforms into two other solid phases at a single temperature. A eutectoid reaction is found in the iron–carbon system (Section 9.18) that is very important in
the heat treating of steels.
The peritectic reaction is yet another invariant reaction involving three phases
at equilibrium. With this reaction, upon heating, one solid phase transforms into a
liquid phase and another solid phase. A peritectic exists for the copper–zinc system
(Figure 9.21, point P) at 598C (1108F) and 78.6 wt% Zn–21.4 wt% Cu; this reaction is as follows:
cooling
dL Δ
heating
(9.15)
The low-temperature solid phase may be an intermediate solid solution (e.g.,
in the above reaction), or it may be a terminal solid solution. One of the latter
peritectics exists at about 97 wt% Zn and 435C (815F) (see Figure 9.19), wherein
the h phase, when heated, transforms to and liquid phases. Three other peritectics are found for the Cu–Zn system, the reactions of which involve b, d, and
g intermediate solid solutions as the low-temperature phases that transform upon
heating.
1496T_c09_252-310 11/29/05 11:33 Page 286
REVISED PAGES
286 • Chapter 9 / Phase Diagrams
Composition (at% Ti)
30
40
50
60
70
1500
2600
L
1400
1310°C
44.9 wt% Ti
Temperature (°C)
 +L
␥ +L
1200
2200
␥ +L
1100
1000
Temperature (°F)
2400
1300
Figure 9.22 A portion of
the nickel–titanium phase
diagram on which is shown
a congruent melting point
for the g-phase solid
solution at 1310C and 44.9
wt% Ti. [Adapted from
Phase Diagrams of Binary
Nickel Alloys, P. Nash
(Editor), 1991. Reprinted
by permission of ASM
International, Materials
Park, OH.]
2000
␥
 +␥
1800
␥ +␦
␦
900
30
40
50
60
70
Composition (wt% Ti)
9.15 CONGRUENT PHASE TRANSFORMATIONS
congruent
transformation
Phase transformations may be classified according to whether or not there is any
change in composition for the phases involved. Those for which there are no compositional alterations are said to be congruent transformations. Conversely, for incongruent transformations, at least one of the phases will experience a change in
composition. Examples of congruent transformations include allotropic transformations (Section 3.6) and melting of pure materials. Eutectic and eutectoid reactions, as well as the melting of an alloy that belongs to an isomorphous system, all
represent incongruent transformations.
Intermediate phases are sometimes classified on the basis of whether they
melt congruently or incongruently. The intermetallic compound Mg2Pb melts congruently at the point designated M on the magnesium–lead phase diagram, Figure 9.20. Also, for the nickel–titanium system, Figure 9.22, there is a congruent
melting point for the g solid solution that corresponds to the point of tangency
for the pairs of liquidus and solidus lines, at 1310C and 44.9 wt% Ti. Furthermore, the peritectic reaction is an example of incongruent melting for an intermediate phase.
Concept Check 9.6
The figure on the next page is the hafnium–vanadium phase diagram, for which
only single-phase regions are labeled. Specify temperature–composition points
at which all eutectics, eutectoids, peritectics, and congruent phase transformations
occur. Also, for each, write the reaction upon cooling. [Phase diagram from ASM
Handbook, Vol. 3, Alloy Phase Diagrams, H. Baker (Editor), 1992, p. 2.244.
Reprinted by permission of ASM International, Materials Park, OH.]
1496T_c09_252-310 12/21/05 7:56 Page 287
2nd REVISE PAGES
9.17 The Gibbs Phase Rule • 287
2200
Hf
L
Temperature (°C)
2000
1800
1600
V
1400
HfV2
1200
Hf
1000
0
20
(Hf)
40
60
Composition (wt% V)
80
100
(V)
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
9.16 CERAMIC AND TERNARY PHASE DIAGRAMS
It need not be assumed that phase diagrams exist only for metal–metal systems; in
fact, phase diagrams that are very useful in the design and processing of ceramic
systems have been experimentally determined for quite a number of these materials. Ceramic phase diagrams are discussed in Section 12.7.
Phase diagrams have also been determined for metallic (as well as ceramic)
systems containing more than two components; however, their representation
and interpretation may be exceedingly complex. For example, a ternary, or threecomponent, composition–temperature phase diagram in its entirety is depicted by
a three-dimensional model. Portrayal of features of the diagram or model in two
dimensions is possible but somewhat difficult.
9.17 THE GIBBS PHASE RULE
Gibbs phase rule
General form of the
Gibbs phase rule
The construction of phase diagrams as well as some of the principles governing the
conditions for phase equilibria are dictated by laws of thermodynamics. One of these
is the Gibbs phase rule, proposed by the nineteenth-century physicist J. Willard
Gibbs. This rule represents a criterion for the number of phases that will coexist
within a system at equilibrium, and is expressed by the simple equation
PFCN
(9.16)
where P is the number of phases present (the phase concept is discussed in Section
9.3). The parameter F is termed the number of degrees of freedom or the number
of externally controlled variables (e.g., temperature, pressure, composition) which
must be specified to completely define the state of the system. Expressed another
1496T_c09_252-310 11/29/05 11:33 Page 288
REVISED PAGES
288 • Chapter 9 / Phase Diagrams
way, F is the number of these variables that can be changed independently without
altering the number of phases that coexist at equilibrium. The parameter C in Equation 9.16 represents the number of components in the system. Components are normally elements or stable compounds and, in the case of phase diagrams, are the materials at the two extremities of the horizontal compositional axis (e.g., H2O and
C12H22O11, and Cu and Ni for the phase diagrams shown in Figures 9.1 and 9.3a,
respectively). Finally, N in Equation 9.16 is the number of noncompositional variables (e.g., temperature and pressure).
Let us demonstrate the phase rule by applying it to binary temperature–
composition phase diagrams, specifically the copper–silver system, Figure 9.7. Since
pressure is constant (1 atm), the parameter N is 1—temperature is the only noncompositional variable. Equation 9.16 now takes the form
PFC1
(9.17)
Furthermore, the number of components C is 2 (viz. Cu and Ag), and
PF213
or
F3P
Consider the case of single-phase fields on the phase diagram (e.g., a, b, and
liquid regions). Since only one phase is present, P 1 and
F3P
312
This means that to completely describe the characteristics of any alloy that exists
within one of these phase fields, we must specify two parameters; these are composition and temperature, which locate, respectively, the horizontal and vertical
positions of the alloy on the phase diagram.
For the situation wherein two phases coexist, for example, a L, b L, and
a b phase regions, Figure 9.7, the phase rule stipulates that we have but one
degree of freedom since
F3P
321
Thus, it is necessary to specify either temperature or the composition of one of
the phases to completely define the system. For example, suppose that we decide
to specify temperature for the a L phase region, say, T1 in Figure 9.23. The
compositions of the a and liquid phases (Ca and CL) are thus dictated by the
extremities of the tie line constructed at T1 across the a L field. Note that only
the nature of the phases is important in this treatment and not the relative phase
amounts. This is to say that the overall alloy composition could lie anywhere along
this tie line constructed at temperature T1 and still give Ca and CL compositions for
the respective a and liquid phases.
The second alternative is to stipulate the composition of one of the phases for
this two-phase situation, which thereby fixes completely the state of the system. For
example, if we specified Ca as the composition of the a phase that is in equilibrium
with the liquid (Figure 9.23), then both the temperature of the alloy (T1) and the
composition of the liquid phase (CL) are established, again by the tie line drawn
across the a L phase field so as to give this Ca composition.
1496T_c09_252-310 11/29/05 11:33 Page 289
REVISED PAGES
9.17 The Gibbs Phase Rule • 289
Figure 9.23 Enlarged
copper-rich section of the
Cu–Ag phase diagram in
which the Gibbs phase rule
for the coexistence of two
phases (i.e., a and L) is
demonstrated. Once the
composition of either phase
(i.e., Ca or CL) or the
temperature (i.e., T1) is
specified, values for the two
remaining parameters are
established by construction
of the appropriate tie line.
L
1000
T1
Temperature (°C)
C
800
␣+L
CL
600
400
0
(Cu)
20
40
Composition (wt% Ag)
60
For binary systems, when three phases are present, there are no degrees of freedom, since
F3P
330
This means that the compositions of all three phases as well as the temperature are
fixed. This condition is met for a eutectic system by the eutectic isotherm; for the
Cu–Ag system (Figure 9.7), it is the horizontal line that extends between points B
and G. At this temperature, 779C, the points at which each of the a, L, and b phase
fields touch the isotherm line correspond to the respective phase compositions;
namely, the composition of the a phase is fixed at 8.0 wt% Ag, that of the liquid at
71.9 wt% Ag, and that of the b phase at 91.2 wt% Ag. Thus, three-phase equilibrium will not be represented by a phase field, but rather by the unique horizontal
isotherm line. Furthermore, all three phases will be in equilibrium for any alloy composition that lies along the length of the eutectic isotherm (e.g., for the Cu–Ag
system at 779C and compositions between 8.0 and 91.2 wt% Ag).
One use of the Gibbs phase rule is in analyzing for nonequilibrium conditions.
For example, a microstructure for a binary alloy that developed over a range of
temperatures and consisting of three phases is a nonequilibrium one; under these
circumstances, three phases will exist only at a single temperature.
Concept Check 9.7
For a ternary system, three components are present; temperature is also a variable.
What is the maximum number of phases that may be present for a ternary system,
assuming that pressure is held constant?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
1496T_c09_252-310 11/29/05 11:33 Page 290
REVISED PAGES
290 • Chapter 9 / Phase Diagrams
T h e I ro n – C a r b o n S ys t e m
Of all binary alloy systems, the one that is possibly the most important is that
for iron and carbon. Both steels and cast irons, primary structural materials in
every technologically advanced culture, are essentially iron–carbon alloys. This
section is devoted to a study of the phase diagram for this system and the development of several of the possible microstructures. The relationships among
heat treatment, microstructure, and mechanical properties are explored in Chapters 10 and 11.
9.18 THE IRON–IRON CARBIDE (Fe–Fe3C)
PHASE DIAGRAM
A portion of the iron–carbon phase diagram is presented in Figure 9.24. Pure iron,
upon heating, experiences two changes in crystal structure before it melts. At room
temperature the stable form, called ferrite, or a iron, has a BCC crystal structure.
Ferrite experiences a polymorphic transformation to FCC austenite, or g iron, at
912C (1674F). This austenite persists to 1394C (2541F), at which temperature
the FCC austenite reverts back to a BCC phase known as d ferrite, which finally
ferrite
austenite
Composition (at% C)
1600
0
5
10
15
20
25
1538°C
1493°C
L
1400
2500
+L
1394°C
1147°C
2.14
, Austenite
4.30
2000
1000
+ Fe3C
912°C
800
+
Temperature (°F)
Temperature (°C)
1200
1500
727°C
0.76
0.022
600
+ Fe3C
, Ferrite
Cementite (Fe3C)
400
0
(Fe)
1
2
3
4
Composition (wt% C)
5
6
1000
6.70
Figure 9.24 The iron–iron carbide phase diagram. [Adapted from Binary Alloy Phase
Diagrams, 2nd edition, Vol. 1, T. B. Massalski (Editor-in-Chief), 1990. Reprinted by
permission of ASM International, Materials Park, OH.]
1496T_c09_252-310 11/29/05 11:33 Page 291
REVISED PAGES
9.18 The Iron–Iron Carbide (Fe–Fe3C) Phase Diagram • 291
Figure 9.25
Photomicrographs of
(a) a ferrite (90)
and (b) austenite
(325). (Copyright
1971 by United
States Steel
Corporation.)
(a)
cementite
(b)
melts at 1538C (2800F). All these changes are apparent along the left vertical axis
of the phase diagram.1
The composition axis in Figure 9.24 extends only to 6.70 wt% C; at this concentration the intermediate compound iron carbide, or cementite (Fe3C), is formed, which
is represented by a vertical line on the phase diagram. Thus, the iron–carbon system
may be divided into two parts: an iron-rich portion, as in Figure 9.24, and the other
(not shown) for compositions between 6.70 and 100 wt% C (pure graphite). In practice, all steels and cast irons have carbon contents less than 6.70 wt% C; therefore,
we consider only the iron–iron carbide system. Figure 9.24 would be more appropriately labeled the Fe–Fe3C phase diagram, since Fe3C is now considered to be a component. Convention and convenience dictate that composition still be expressed in
“wt% C” rather than “wt% Fe3C”; 6.70 wt% C corresponds to 100 wt% Fe3C.
Carbon is an interstitial impurity in iron and forms a solid solution with each
of a and d ferrites, and also with austenite, as indicated by the a, d, and g singlephase fields in Figure 9.24. In the BCC a ferrite, only small concentrations of carbon are soluble; the maximum solubility is 0.022 wt% at 727C (1341F). The limited
solubility is explained by the shape and size of the BCC interstitial positions, which
make it difficult to accommodate the carbon atoms. Even though present in relatively low concentrations, carbon significantly influences the mechanical properties
of ferrite. This particular iron–carbon phase is relatively soft, may be made magnetic at temperatures below 768C (1414F), and has a density of 7.88 g/cm3. Figure 9.25a is a photomicrograph of a ferrite.
1
The reader may wonder why no b phase is found on the Fe–Fe3C phase diagram, Figure
9.24 (consistent with the a, b, g, etc. labeling scheme described previously). Early investigators observed that the ferromagnetic behavior of iron disappears at 768C and attributed this
phenomenon to a phase transformation; the “b” label was assigned to the high-temperature
phase. Later it was discovered that this loss of magnetism did not result from a phase transformation (see Section 20.6) and, therefore, the presumed b phase did not exist.
1496T_c09_252-310 11/29/05 11:33 Page 292
REVISED PAGES
292 • Chapter 9 / Phase Diagrams
The austenite, or g phase of iron, when alloyed with carbon alone, is not stable
below 727C (1341F), as indicated in Figure 9.24. The maximum solubility of carbon
in austenite, 2.14 wt%, occurs at 1147C (2097F). This solubility is approximately
100 times greater than the maximum for BCC ferrite, since the FCC interstitial
positions are larger (see the results of Problem 4.5), and, therefore, the strains imposed on the surrounding iron atoms are much lower. As the discussions that follow demonstrate, phase transformations involving austenite are very important in
the heat treating of steels. In passing, it should be mentioned that austenite is nonmagnetic. Figure 9.25b shows a photomicrograph of this austenite phase.
The d ferrite is virtually the same as a ferrite, except for the range of temperatures over which each exists. Since the d ferrite is stable only at relatively high
temperatures, it is of no technological importance and is not discussed further.
Cementite (Fe3C) forms when the solubility limit of carbon in a ferrite is exceeded below 727C (1341F) (for compositions within the a Fe3C phase region).
As indicated in Figure 9.24, Fe3C will also coexist with the g phase between 727
and 1147C (1341 and 2097F). Mechanically, cementite is very hard and brittle; the
strength of some steels is greatly enhanced by its presence.
Strictly speaking, cementite is only metastable; that is, it will remain as a compound indefinitely at room temperature. However, if heated to between 650 and 700C
(1200 and 1300F) for several years, it will gradually change or transform into a iron
and carbon, in the form of graphite, which will remain upon subsequent cooling to
room temperature. Thus, the phase diagram in Figure 9.24 is not a true equilibrium
one because cementite is not an equilibrium compound. However, inasmuch as the decomposition rate of cementite is extremely sluggish, virtually all the carbon in steel
will be as Fe3C instead of graphite, and the iron–iron carbide phase diagram is, for all
practical purposes, valid.As will be seen in Section 11.2, addition of silicon to cast irons
greatly accelerates this cementite decomposition reaction to form graphite.
The two-phase regions are labeled in Figure 9.24. It may be noted that one eutectic exists for the iron–iron carbide system, at 4.30 wt% C and 1147C (2097F);
for this eutectic reaction,
Eutectic reaction for
the iron-iron carbide
system
cooling
L Δ g Fe3C
heating
(9.18)
the liquid solidifies to form austenite and cementite phases. Of course, subsequent
cooling to room temperature will promote additional phase changes.
It may be noted that a eutectoid invariant point exists at a composition of 0.76 wt%
C and a temperature of 727C (1341F).This eutectoid reaction may be represented by
Eutectoid reaction
for the iron-iron
carbide system
g10.76 wt% C2 Δ a10.022 wt% C2 Fe3C 16.7 wt%C2
cooling
heating
(9.19)
or, upon cooling, the solid g phase is transformed into a iron and cementite. (Eutectoid phase transformations were addressed in Section 9.14.) The eutectoid phase
changes described by Equation 9.19 are very important, being fundamental to the
heat treatment of steels, as explained in subsequent discussions.
Ferrous alloys are those in which iron is the prime component, but carbon as
well as other alloying elements may be present. In the classification scheme of ferrous alloys based on carbon content, there are three types: iron, steel, and cast iron.
Commercially pure iron contains less than 0.008 wt% C and, from the phase diagram, is composed almost exclusively of the ferrite phase at room temperature. The
1496T_c09_252-310 11/29/05 11:33 Page 293
REVISED PAGES
9.19 Development of Microstructure in Iron–Carbon Alloys • 293
iron–carbon alloys that contain between 0.008 and 2.14 wt% C are classified as
steels. In most steels the microstructure consists of both a and Fe3C phases. Upon
cooling to room temperature, an alloy within this composition range must pass
through at least a portion of the g-phase field; distinctive microstructures are subsequently produced, as discussed below. Although a steel alloy may contain as much
as 2.14 wt% C, in practice, carbon concentrations rarely exceed 1.0 wt%. The properties and various classifications of steels are treated in Section 11.2. Cast irons are
classified as ferrous alloys that contain between 2.14 and 6.70 wt% C. However,
commercial cast irons normally contain less than 4.5 wt% C. These alloys are discussed further also in Section 11.2.
9.19 DEVELOPMENT OF MICROSTRUCTURE IN
IRON–CARBON ALLOYS
Several of the various microstructures that may be produced in steel alloys and their
relationships to the iron–iron carbon phase diagram are now discussed, and it is
shown that the microstructure that develops depends on both the carbon content
and heat treatment. This discussion is confined to very slow cooling of steel alloys,
in which equilibrium is continuously maintained. A more detailed exploration of
the influence of heat treatment on microstructure, and ultimately on the mechanical properties of steels, is contained in Chapter 10.
Phase changes that occur upon passing from the g region into the a Fe3C
phase field (Figure 9.24) are relatively complex and similar to those described for
the eutectic systems in Section 9.12. Consider, for example, an alloy of eutectoid
composition (0.76 wt% C) as it is cooled from a temperature within the g phase region, say, 800C—that is, beginning at point a in Figure 9.26 and moving down the
1100
␥
1000
␥ + Fe3C
900
␥
Temperature (°C)
x
␥
800
a
␣ +␥
␥
727°C
␣
b
700
␥
␣
600
Fe3C
␣ + Fe3C
500
x
400
0
1.0
Composition (wt% C)
2.0
Figure 9.26 Schematic
representations of the
microstructures for an
iron–carbon alloy of eutectoid
composition (0.76 wt% C) above
and below the eutectoid
temperature.
1496T_c09_252-310 11/29/05 11:33 Page 294
REVISED PAGES
294 • Chapter 9 / Phase Diagrams
Figure 9.27 Photomicrograph of a eutectoid
steel showing the pearlite microstructure
consisting of alternating layers of a ferrite
(the light phase) and Fe3C (thin layers most
of which appear dark). 500. (Reproduced
with permission from Metals Handbook,
9th edition, Vol. 9, Metallography and
Microstructures, American Society for
Metals, Materials Park, OH, 1985.)
pearlite
vertical line xx¿. Initially, the alloy is composed entirely of the austenite phase having
a composition of 0.76 wt% C and corresponding microstructure, also indicated in
Figure 9.26. As the alloy is cooled, there will occur no changes until the eutectoid
temperature (727C) is reached. Upon crossing this temperature to point b, the
austenite transforms according to Equation 9.19.
The microstructure for this eutectoid steel that is slowly cooled through the
eutectoid temperature consists of alternating layers or lamellae of the two phases
(a and Fe3C) that form simultaneously during the transformation. In this case, the
relative layer thickness is approximately 8 to 1. This microstructure, represented
schematically in Figure 9.26, point b, is called pearlite because it has the appearance of mother of pearl when viewed under the microscope at low magnifications.
Figure 9.27 is a photomicrograph of a eutectoid steel showing the pearlite. The
pearlite exists as grains, often termed “colonies”; within each colony the layers
are oriented in essentially the same direction, which varies from one colony to
another. The thick light layers are the ferrite phase, and the cementite phase appears as thin lamellae most of which appear dark. Many cementite layers are so
thin that adjacent phase boundaries are so close together that they are indistinguishable at this magnification, and, therefore, appear dark. Mechanically, pearlite
has properties intermediate between the soft, ductile ferrite and the hard, brittle
cementite.
The alternating a and Fe3C layers in pearlite form as such for the same reason that the eutectic structure (Figures 9.13 and 9.14) forms—because the composition of the parent phase [in this case austenite (0.76 wt% C)] is different
from either of the product phases [ferrite (0.022 wt% C) and cementite (6.7 wt%
C)], and the phase transformation requires that there be a redistribution of the
carbon by diffusion. Figure 9.28 illustrates schematically microstructural changes
that accompany this eutectoid reaction; here the directions of carbon diffusion
are indicated by arrows. Carbon atoms diffuse away from the 0.022 wt% ferrite
regions and to the 6.7 wt% cementite layers, as the pearlite extends from the
grain boundary into the unreacted austenite grain. The layered pearlite forms because carbon atoms need diffuse only minimal distances with the formation of
this structure.
Furthermore, subsequent cooling of the pearlite from point b in Figure 9.26 will
produce relatively insignificant microstructural changes.
1496T_c09_252-310 12/21/05 7:56 Page 295
2nd REVISE PAGES
9.19 Development of Microstructure in Iron–Carbon Alloys • 295
Figure 9.28 Schematic
representation of the
formation of pearlite from
austenite; direction of carbon
diffusion indicated by arrows.
Austenite grain
boundary
Ferrite ()
Austenite
( )
Ferrite ()
Austenite
( )
Ferrite ()
Cementite
(Fe3C)
Growth direction
of pearlite
Ferrite ()
Carbon diffusion
Hypoeutectoid Alloys
1100
1000
+ Fe3C
y
M
900
c
Temperature (°C)
hypoeutectoid alloy
Microstructures for iron–iron carbide alloys having other than the eutectoid composition are now explored; these are analogous to the fourth case described in
Section 9.12 and illustrated in Figure 9.16 for the eutectic system. Consider a composition C0 to the left of the eutectoid, between 0.022 and 0.76 wt% C; this is termed
a hypoeutectoid (less than eutectoid) alloy. Cooling an alloy of this composition is
represented by moving down the vertical line yy¿ in Figure 9.29. At about 875C,
point c, the microstructure will consist entirely of grains of the g phase, as shown
800
d
e
N
Te
O
f
700
Pearlite
600
Fe3C
Proeutectoid
Eutectoid
+ Fe3C
500
y
400
0
1.0
C0
Composition (wt% C)
2.0
Figure 9.29 Schematic
representations of the
microstructures for an
iron–carbon alloy of
hypoeutectoid composition C0
(containing less than 0.76 wt%
C) as it is cooled from within
the austenite phase region to
below the eutectoid
temperature.
1496T_c09_252-310 11/29/05 11:33 Page 296
REVISED PAGES
296 • Chapter 9 / Phase Diagrams
proeutectoid ferrite
Figure 9.30
Photomicrograph
of a 0.38 wt% C
steel having a
microstructure
consisting of pearlite
and proeutectoid
ferrite. 635.
(Photomicrograph
courtesy of Republic
Steel Corporation.)
schematically in the figure. In cooling to point d, about 775C, which is within the
a g phase region, both these phases will coexist as in the schematic microstructure. Most of the small a particles will form along the original g grain boundaries.
The compositions of both a and g phases may be determined using the appropriate
tie line; these compositions correspond, respectively, to about 0.020 and 0.40 wt% C.
While cooling an alloy through the a g phase region, the composition of the
ferrite phase changes with temperature along the a 1a g2 phase boundary, line
MN, becoming slightly richer in carbon. On the other hand, the change in composition of the austenite is more dramatic, proceeding along the 1a g2 g boundary,
line MO, as the temperature is reduced.
Cooling from point d to e, just above the eutectoid but still in the a g region,
will produce an increased fraction of the a phase and a microstructure similar to
that also shown: the a particles will have grown larger. At this point, the compositions of the a and g phases are determined by constructing a tie line at the temperature Te; the a phase will contain 0.022 wt% C, while the g phase will be of the
eutectoid composition, 0.76 wt% C.
As the temperature is lowered just below the eutectoid, to point f, all the g
phase that was present at temperature Te (and having the eutectoid composition)
will transform to pearlite, according to the reaction in Equation 9.19. There will be
virtually no change in the a phase that existed at point e in crossing the eutectoid
temperature—it will normally be present as a continuous matrix phase surrounding the isolated pearlite colonies. The microstructure at point f will appear as the
corresponding schematic inset of Figure 9.29. Thus the ferrite phase will be present
both in the pearlite and also as the phase that formed while cooling through the
a g phase region. The ferrite that is present in the pearlite is called eutectoid
ferrite, whereas the other, that formed above Te, is termed proeutectoid (meaning
pre- or before eutectoid) ferrite, as labeled in Figure 9.29. Figure 9.30 is a photomicrograph of a 0.38 wt% C steel; large, white regions correspond to the proeutectoid ferrite. For pearlite, the spacing between the a and Fe3C layers varies
from grain to grain; some of the pearlite appears dark because the many closespaced layers are unresolved at the magnification of the photomicrograph. The
Proeutectoid
ferrite
Pearlite
1496T_c09_252-310 11/29/05 11:33 Page 297
REVISED PAGES
9.19 Development of Microstructure in Iron–Carbon Alloys • 297
chapter-opening photograph for this chapter is a scanning electron micrograph of
a hypoeutectoid (0.44 wt% C) steel in which may also be seen both pearlite and
proeutectoid ferrite, only at a higher magnification. Note also that two microconstituents are present in these micrographs—proeutectoid ferrite and pearlite—
which will appear in all hypoeutectoid iron–carbon alloys that are slowly cooled to
a temperature below the eutectoid.
The relative amounts of the proeutectoid a and pearlite may be determined in
a manner similar to that described in Section 9.12 for primary and eutectic microconstituents. We use the lever rule in conjunction with a tie line that extends from
the a (a Fe3C) phase boundary (0.022 wt% C) to the eutectoid composition
(0.76 wt% C), inasmuch as pearlite is the transformation product of austenite having
this composition. For example, let us consider an alloy of composition C0¿ in Figure 9.31. Thus, the fraction of pearlite, Wp, may be determined according to
Lever rule
expression for
computation of
pearlite mass
fraction
(composition C0¿,
Figure 9.31)
Wp
T
TU
C0¿ 0.022
C0¿ 0.022
0.76 0.022
0.74
(9.20)
Furthermore, the fraction of proeutectoid a, Wa¿, is computed as follows:
Lever rule
expression for
computation of
proeutectoid ferrite
mass fraction
Wa¿
U
TU
0.76 C0¿
0.76 C0¿
0.76 0.022
0.74
(9.21)
Of course, fractions of both total a (eutectoid and proeutectoid) and cementite are
determined using the lever rule and a tie line that extends across the entirety of the
a Fe3C phase region, from 0.022 to 6.7 wt% C.
␥
␥ + Fe3C
Temperature
Figure 9.31 A
portion of the
Fe–Fe3C phase
diagram used in
computations for
relative amounts
of proeutectoid
and pearlite
microconstituents for
hypoeutectoid (C0¿)
and hypereutectoid
(C1¿) compositions.
␣
T
U
V
X
␣ + Fe3C
6.70
0.022 C0
0.76
C1
Composition (wt% C)
1496T_c09_252-310 12/21/05 7:56 Page 298
2nd REVISE PAGES
298 • Chapter 9 / Phase Diagrams
1100
Figure 9.32 Schematic
representations of the
microstructures for an
iron–carbon alloy of
hypereutectoid composition
C1 (containing between 0.76
and 2.14 wt% C), as it is
cooled from within the
austenite phase region to
below the eutectoid
temperature.
P
+ Fe3C
1000
z
g
900
Fe3C
Temperature (°C)
800
h
+
O
700
i
Pearlite
600
Proeutectoid
Eutectoid Fe3C
Fe3C
500
+ Fe3C
z'
400
0
1.0
2.0
C1
Composition (wt% C)
Hypereutectoid Alloys
hypereutectoid
alloy
proeutectoid
cementite
Analogous transformations and microstructures result for hypereutectoid alloys,
those containing between 0.76 and 2.14 wt% C, which are cooled from temperatures
within the g phase field. Consider an alloy of composition C1 in Figure 9.32 that,
upon cooling, moves down the line zz¿. At point g only the g phase will be present
with a composition of C1; the microstructure will appear as shown, having only g
grains. Upon cooling into the g Fe3C phase field—say, to point h—the cementite
phase will begin to form along the initial g grain boundaries, similar to the a phase
in Figure 9.29, point d. This cementite is called proeutectoid cementite—that which
forms before the eutectoid reaction. Of course, the cementite composition remains
constant (6.70 wt% C) as the temperature changes. However, the composition of
the austenite phase will move along line PO toward the eutectoid. As the temperature is lowered through the eutectoid to point i, all remaining austenite of eutectoid composition is converted into pearlite; thus, the resulting microstructure consists of pearlite and proeutectoid cementite as microconstituents (Figure 9.32). In
the photomicrograph of a 1.4 wt% C steel (Figure 9.33), note that the proeutectoid
cementite appears light. Since it has much the same appearance as proeutectoid ferrite (Figure 9.30), there is some difficulty in distinguishing between hypoeutectoid
and hypereutectoid steels on the basis of microstructure.
Relative amounts of both pearlite and proeutectoid Fe3C microconstituents may
be computed for hypereutectoid steel alloys in a manner analogous to that for
hypoeutectoid materials; the appropriate tie line extends between 0.76 and 6.70
1496T_c09_252-310 11/29/05 11:33 Page 299
REVISED PAGES
9.19 Development of Microstructure in Iron–Carbon Alloys • 299
Figure 9.33
Photomicrograph
of a 1.4 wt% C
steel having a
microstructure
consisting of a white
proeutectoid
cementite network
surrounding the
pearlite colonies.
1000. (Copyright
1971 by United
States Steel
Corporation.)
Proeutectoid
cementite
Pearlite
wt% C. Thus, for an alloy having composition C¿1 in Figure 9.31, fractions of pearlite
Wp and proeutectoid cementite WFe3C¿ are determined from the following lever rule
expressions:
6.70 C1¿
6.70 C1¿
X
(9.22)
Wp
VX
6.70 0.76
5.94
and
C1¿ 0.76
C1¿ 0.76
V
(9.23)
WFe3C¿
VX
6.70 0.76
5.94
Concept Check 9.8
Briefly explain why a proeutectoid phase (ferrite or cementite) forms along austenite grain boundaries. Hint: Consult Section 4.6.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
EXAMPLE PROBLEM 9.4
Determination of Relative Amounts of Ferrite, Cementite,
and Pearlite Microconstituents
For a 99.65 wt% Fe–0.35 wt% C alloy at a temperature just below the eutectoid, determine the following:
(a) The fractions of total ferrite and cementite phases
(b) The fractions of the proeutectoid ferrite and pearlite
(c) The fraction of eutectoid ferrite
1496T_c09_252-310 11/29/05 11:33 Page 300
REVISED PAGES
300 • Chapter 9 / Phase Diagrams
Solution
(a) This part of the problem is solved by application of the lever rule expressions employing a tie line that extends all the way across the a Fe3C
phase field. Thus, C¿0 is 0.35 wt% C, and
Wa
6.70 0.35
0.95
6.70 0.022
and
WFe3C
0.35 0.022
0.05
6.70 0.022
(b) The fractions of proeutectoid ferrite and pearlite are determined by using the lever rule and a tie line that extends only to the eutectoid composition (i.e., Equations 9.20 and 9.21). Or
Wp
0.35 0.022
0.44
0.76 0.022
Wa¿
0.76 0.35
0.56
0.76 0.022
and
(c) All ferrite is either as proeutectoid or eutectoid (in the pearlite). Therefore, the sum of these two ferrite fractions will equal the fraction of total ferrite; that is,
Wa¿ Wae Wa
where Wae denotes the fraction of the total alloy that is eutectoid ferrite.Values
for Wa and Wa¿ were determined in parts (a) and (b) as 0.95 and 0.56,
respectively. Therefore,
Wae Wa Wa¿ 0.95 0.56 0.39
Nonequilibrium Cooling
In this discussion on the microstructural development of iron–carbon alloys it has
been assumed that, upon cooling, conditions of metastable equilibrium2 have been
continuously maintained; that is, sufficient time has been allowed at each new temperature for any necessary adjustment in phase compositions and relative amounts
as predicted from the Fe–Fe3C phase diagram. In most situations these cooling rates
are impractically slow and really unnecessary; in fact, on many occasions nonequilibrium conditions are desirable. Two nonequilibrium effects of practical importance
are (1) the occurrence of phase changes or transformations at temperatures other
than those predicted by phase boundary lines on the phase diagram, and (2) the
2
The term “metastable equilibrium” is used in this discussion inasmuch as Fe3C is only a
metastable compound.
1496T_c09_252-310 11/29/05 11:33 Page 301
REVISED PAGES
9.20 The Influence of Other Alloying Elements • 301
Eutectoid temperature (°C)
Mo
1200
W
2200
Si
2000
1000
1800
Cr
1600
800
1400
Mn
Eutectoid temperature (°F)
2400
Ti
Figure 9.34 The dependence of
eutectoid temperature on alloy
concentration for several alloying
elements in steel. (From Edgar C.
Bain, Functions of the Alloying
Elements in Steel, American Society
for Metals, 1939, p. 127.)
1200
600
Ni
0
2
4
6
8
10
12
1000
14
Concentration of alloying elements (wt%)
existence at room temperature of nonequilibrium phases that do not appear on the
phase diagram. Both are discussed in the next chapter.
9.20 THE INFLUENCE OF OTHER ALLOYING
ELEMENTS
Additions of other alloying elements (Cr, Ni, Ti, etc.) bring about rather dramatic
changes in the binary iron–iron carbide phase diagram, Figure 9.24. The extent
of these alterations of the positions of phase boundaries and the shapes of the
phase fields depends on the particular alloying element and its concentration.
One of the important changes is the shift in position of the eutectoid with respect to temperature and to carbon concentration. These effects are illustrated
in Figures 9.34 and 9.35, which plot the eutectoid temperature and eutectoid
composition (in wt% C) as a function of concentration for several other alloying elements. Thus, other alloy additions alter not only the temperature of the
eutectoid reaction but also the relative fractions of pearlite and the proeutectoid
phase that form. Steels are normally alloyed for other reasons, however—usually
either to improve their corrosion resistance or to render them amenable to heat
treatment (see Section 11.8).
Figure 9.35 The dependence of eutectoid
composition (wt% C) on alloy concentration
for several alloying elements in steel. (From
Edgar C. Bain, Functions of the Alloying
Elements in Steel, American Society for Metals,
1939, p. 127.)
Eutectoid composition (wt% C)
0.8
Ni
0.6
Cr
0.4
Si
Mo
W
0.2
0
Mn
Ti
0
2
4
6
8
10 12 14
Concentration of alloying elements (wt%)
1496T_c09_252-310 11/29/05 11:33 Page 302
REVISED PAGES
302 • Chapter 9 / Phase Diagrams
SUMMARY
Phase Equilibria
One-Component (or Unary) Phase Diagrams
Binary Phase Diagrams
Interpretation of Phase Diagrams
Equilibrium phase diagrams are a convenient and concise way of representing the
most stable relationships between phases in alloy systems. This discussion began
by considering the unary (or pressure–temperature) phase diagram for a onecomponent system. Solid-, liquid-, and vapor-phase regions are found on this type
of phase diagram. For binary systems, temperature and composition are variables,
whereas external pressure is held constant. Areas, or phase regions, are defined
on these temperature-versus-composition plots within which either one or two
phases exist. For an alloy of specified composition and at a known temperature, the phases present, their compositions, and relative amounts under equilibrium conditions may be determined. Within two-phase regions, tie lines and the
lever rule must be used for phase composition and mass fraction computations,
respectively.
Binary Isomorphous Systems
Development of Microstructure in Isomorphous Alloys
Mechanical Properties of Isomorphous Alloys
Several different kinds of phase diagram were discussed for metallic systems. Isomorphous diagrams are those for which there is complete solubility in the solid
phase; the copper–nickel system displays this behavior. Also discussed for alloys belonging to isomorphous systems were the development of microstructure for both
cases of equilibrium and nonequilibrium cooling, and the dependence of mechanical characteristics on composition.
Binary Eutectic Systems
Development of Microstructure in Eutectic Alloys
In a eutectic reaction, as found in some alloy systems, a liquid phase transforms
isothermally to two different solid phases upon cooling. Such a reaction is noted on
the copper–silver and lead–tin phase diagrams. Complete solid solubility for all compositions does not exist; instead, solid solutions are terminal—there is only a limited solubility of each component in the other. Four different kinds of microstructures that may develop for the equilibrium cooling of alloys belonging to eutectic
systems were discussed.
Equilibrium Diagrams Having Intermediate Phases or Compounds
Eutectoid and Peritectic Reactions
Congruent Phase Transformations
Other equilibrium phase diagrams are more complex, having intermediate compounds and/or phases, possibly more than a single eutectic, and other reactions including eutectoid, peritectic, and congruent phase transformations. These are found
for copper–zinc and magnesium–lead systems.
1496T_c09_252-310 11/29/05 11:33 Page 303
REVISED PAGES
References • 303
The Gibbs Phase Rule
The Gibbs phase rule was introduced; it is a simple equation that relates the number of phases present in a system at equilibrium with the number of degrees of freedom, the number of components, and the number of noncompositional variables.
The Iron–Iron Carbide (Fe–Fe3C) Phase Diagram
Development of Microstructure in Iron–Carbon Alloys
Considerable attention was given to the iron–carbon system, and specifically, the
iron–iron carbide phase diagram, which technologically is one of the most important. The development of microstructure in many iron–carbon alloys and steels
depends on the eutectoid reaction in which the FCC austenite phase of composition 0.76 wt% C transforms isothermally to the BCC a ferrite phase (0.022 wt%
C) and the intermetallic compound, cementite (Fe3C). The microstructural product of an iron–carbon alloy of eutectoid composition is pearlite, a microconstituent consisting of alternating layers of ferrite and cementite. The microstructures of alloys having carbon contents less than the eutectoid (hypoeutectoid) are
comprised of a proeutectoid ferrite phase in addition to pearlite. On the other
hand, pearlite and proeutectoid cementite constitute the microconstituents for
hypereutectoid alloys—those with carbon contents in excess of the eutectoid
composition.
I M P O R TA N T T E R M S A N D C O N C E P T S
Austenite
Cementite
Component
Congruent transformation
Equilibrium
Eutectic phase
Eutectic reaction
Eutectic structure
Eutectoid reaction
Ferrite
Free energy
Gibbs phase rule
Hypereutectoid alloy
Hypoeutectoid alloy
Intermediate solid solution
Intermetallic compound
Invariant point
Isomorphous
Lever rule
Liquidus line
Metastable
Microconstituent
Pearlite
Peritectic reaction
Phase
Phase diagram
Phase equilibrium
Primary phase
Proeutectoid cementite
Proeutectoid ferrite
Solidus line
Solubility limit
Solvus line
System
Terminal solid solution
Tie line
REFERENCES
ASM Handbook, Vol. 3, Alloy Phase Diagrams,
ASM International, Materials Park, OH,
1992.
ASM Handbook, Vol. 9, Metallography and Microstructures, ASM International, Materials
Park, OH, 2004.
Hansen, M. and K. Anderko, Constitution of Binary
Alloys, 2nd edition, McGraw-Hill, New York,
1958. First Supplement (R. P. Elliott), 1965.
Second Supplement (F. A. Shunk), 1969.
Reprinted by Genium Publishing Corp., Schenectady, NY.
Massalski,T. B., H. Okamoto, P. R. Subramanian, and
L. Kacprzak (Editors), Binary Phase Diagrams,
2nd edition, ASM International, Materials
Park, OH, 1990. Three volumes. Also on CD
with updates.
Okamoto, H., Desk Handbook: Phase Diagrams for
Binary Alloys, ASM International, Materials
Park, OH, 2000.
1496T_c09_252-310 11/29/05 11:33 Page 304
REVISED PAGES
304 • Chapter 9 / Phase Diagrams
Petzow, G., Ternary Alloys, A Comprehensive Compendium of Evaluated Constitutional Data and
Phase Diagrams, Wiley, New York, 1988–1995.
Fifteen volumes.
Villars, P., A. Prince, and H. Okamoto (Editors),
Handbook of Ternary Alloy Phase Diagrams,
ASM International, Materials Park, OH, 1995.
Ten volumes. Also on CD.
QUESTIONS AND PROBLEMS
Solubility Limit
9.1 Consider the sugar–water phase diagram of
Figure 9.1.
(a) How much sugar will dissolve in 1000 g
of water at 80C (176F)?
(b) If the saturated liquid solution in part (a)
is cooled to 20C (68F), some of the sugar
will precipitate out as a solid. What will be
the composition of the saturated liquid solution (in wt% sugar) at 20C?
(c) How much of the solid sugar will come
out of solution upon cooling to 20C?
9.2 At 100C, what is the maximum solubility (a)
of Pb in Sn? (b) of Sn in Pb?
Microstructure
9.3 Cite three variables that determine the microstructure of an alloy.
Phase Equilibria
9.4 What thermodynamic condition must be met
for a state of equilibrium to exist?
One-Component (or Unary)
Phase Diagrams
9.5 Consider a specimen of ice that is at 25C
and 10 atm pressure. Using Figure 9.2, the
pressure–temperature phase diagram for
H2O, determine the pressure to which the
specimen must be raised or lowered to cause
it (a) to melt, and (b) to sublime.
9.6 At a pressure of 0.1 atm, determine (a) the
melting temperature for ice, and (b) the
boiling temperature for water.
Binary Isomorphous Systems
9.7 Given here are the solidus and liquidus temperatures for the copper–gold system. Construct the phase diagram for this system and
label each region.
Composition
(wt% Au)
Solidus
Temperature (C)
Liquidus
Temperature (C)
0
20
40
60
80
90
95
100
1085
1019
972
934
911
928
974
1064
1085
1042
996
946
911
942
984
1064
Interpretation of Phase Diagrams
(Binary Isomorphous Systems)
(Binary Eutectic Systems)
(Equilibrium Diagrams Having Intermediate Phases
or Compounds)
9.8 Cite the phases that are present and the phase
compositions for the following alloys:
(a) 15 wt% Sn–85 wt% Pb at 100C (212F)
(b) 25 wt% Pb–75 wt% Mg at 425C (800F)
(c) 85 wt% Ag–15 wt% Cu at 800C (1470F)
(d) 55 wt% Zn–45 wt% Cu at 600C (1110F)
(e) 1.25 kg Sn and 14 kg Pb at 200C (390F)
(f) 7.6 lbm Cu and 144.4 lbm Zn at 600C
(1110F)
(g) 21.7 mol Mg and 35.4 mol Pb at 350C
(660F)
(h) 4.2 mol Cu and 1.1 mol Ag at 900C
(1650F)
9.9 Is it possible to have a copper–silver alloy that,
at equilibrium, consists of a b phase of composition 92 wt% Ag–8 wt% Cu, and also a liquid
phase of composition 76 wt% Ag–24 wt% Cu?
If so, what will be the approximate temperature
of the alloy? If this is not possible, explain why.
9.10 Is it possible to have a copper–silver alloy
that, at equilibrium, consists of an a phase of
composition 4 wt% Ag–96 wt% Cu, and also
a b phase of composition 95 wt% Ag–5 wt%
1496T_c09_252-310 11/29/05 11:33 Page 305
REVISED PAGES
Questions and Problems • 305
Cu? If so, what will be the approximate temperature of the alloy? If this is not possible,
explain why.
9.11 A lead–tin alloy of composition 30 wt%
Sn–70 wt% Pb is slowly heated from a temperature of 150C (300F).
(a) At what temperature does the first liquid
phase form?
(b) What is the composition of this liquid
phase?
(c) At what temperature does complete
melting of the alloy occur?
(d) What is the composition of the last solid
remaining prior to complete melting?
9.12 A 50 wt% Ni–50 wt% Cu alloy is slowly
cooled from 1400C (2550F) to 1200C
(2190F).
(a) At what temperature does the first solid
phase form?
(b) What is the composition of this solid
phase?
(c) At what temperature does the liquid
solidify?
(d) What is the composition of this last
remaining liquid phase?
9.13 For an alloy of composition 52 wt% Zn–48
wt% Cu, cite the phases present and their
mass fractions at the following temperatures:
1000C, 800C, 500C, and 300C.
9.14 Determine the relative amounts (in terms of
mass fractions) of the phases for the alloys
and temperatures given in Problem 9.8.
9.15 A 2.0-kg specimen of an 85 wt% Pb–15 wt%
Sn alloy is heated to 200C (390F); at this
temperature it is entirely an a-phase solid
solution (Figure 9.8). The alloy is to be melted
to the extent that 50% of the specimen is
liquid, the remainder being the a phase. This
may be accomplished by heating the alloy or
changing its composition while holding the
temperature constant.
(a) To what temperature must the specimen
be heated?
(b) How much tin must be added to the 2.0-kg
specimen at 200C to achieve this state?
9.16 A magnesium–lead alloy of mass 7.5 kg consists of a solid a phase that has a composition
just slightly below the solubility limit at 300C
(570F).
(a) What mass of lead is in the alloy?
(b) If the alloy is heated to 400C (750F),
how much more lead may be dissolved in the
a phase without exceeding the solubility limit
of this phase?
9.17 A 65 wt% Ni–35 wt% Cu alloy is heated to a
temperature within the liquid-phase
region. If the composition of the a phase is
70 wt% Ni, determine:
(a) The temperature of the alloy
(b) The composition of the liquid phase
(c) The mass fractions of both phases
9.18 A 40 wt% Pb–60 wt% Mg alloy is heated to
a temperature within the liquid-phase
region. If the mass fraction of each phase is
0.5, then estimate:
(a) The temperature of the alloy
(b) The compositions of the two phases
9.19 For alloys of two hypothetical metals A and
B, there exist an a, A-rich phase and a b,
B-rich phase. From the mass fractions of both
phases for two different alloys provided in the
table below, (which are at the same temperature), determine the composition of the phase
boundary (or solubility limit) for both a and
b phases at this temperature.
Alloy Composition
70 wt% A–30 wt% B
35 wt% A–65 wt% B
Fraction
Phase
Fraction
Phase
0.78
0.36
0.22
0.64
9.20 A hypothetical A–B alloy of composition
40 wt% B–60 wt% A at some temperature is
found to consist of mass fractions of 0.66 and
0.34 for the a and b phases, respectively. If
the composition of the a phase is 13 wt%
B–87 wt% A, what is the composition of the
b phase?
9.21 Is it possible to have a copper–silver alloy of
composition 20 wt% Ag–80 wt% Cu that, at
equilibrium, consists of a and liquid phases
having mass fractions Wa 0.80 and WL
0.20? If so, what will be the approximate
temperature of the alloy? If such an alloy is
not possible, explain why.
1496T_c09_252-310 11/29/05 11:33 Page 306
REVISED PAGES
306 • Chapter 9 / Phase Diagrams
9.22 For 5.7 kg of a magnesium–lead alloy of composition 50 wt% Pb–50 wt% Mg, is it possible, at equilibrium, to have a and Mg2Pb
phases with respective masses of 5.13 and
0.57 kg? If so, what will be the approximate
temperature of the alloy? If such an alloy is
not possible, then explain why.
9.23 Derive Equations 9.6a and 9.7a, which may be
used to convert mass fraction to volume fraction, and vice versa.
9.24 Determine the relative amounts (in terms of
volume fractions) of the phases for the alloys
and temperatures given in Problems 9.8a, b, and
d. Given here are the approximate densities of
the various metals at the alloy temperatures:
Metal
Temperature (C)
Density (g/cm3)
Cu
Mg
Pb
Pb
Sn
Zn
600
425
100
425
100
600
8.68
1.68
11.27
10.96
7.29
6.67
Development of Microstructure in Eutectic Alloys
9.28 Briefly explain why, upon solidification, an
alloy of eutectic composition forms a microstructure consisting of alternating layers of
the two solid phases.
9.29 What is the difference between a phase and
a microconstituent?
9.30 Is it possible to have a magnesium–lead alloy
in which the mass fractions of primary a and
total a are 0.60 and 0.85, respectively, at 460C
(860F)? Why or why not?
9.31 For 2.8 kg of a lead–tin alloy, is it possible to
have the masses of primary b and total b of
2.21 kg and 2.53 kg, respectively, at 180C
(355F)? Why or why not?
9.32 For a lead–tin alloy of composition 80 wt%
Sn–20 wt% Pb and at 180C (355F) do the
following:
(a) Determine the mass fractions of a and
b phases.
(b) Determine the mass fractions of primary
b and eutectic microconstituents.
(c) Determine the mass fraction of eutectic b.
Development of Microstructure in
Isomorphous Alloys
9.25 (a) Briefly describe the phenomenon of
coring and why it occurs.
(b) Cite one undesirable consequence of
coring.
Mechanical Properties of Isomorphous Alloys
9.26 It is desirable to produce a copper–nickel
alloy that has a minimum noncold-worked
tensile strength of 380 MPa (55,000 psi) and
a ductility of at least 45%EL. Is such an alloy
possible? If so, what must be its composition?
If this is not possible, then explain why.
Binary Eutectic Systems
9.27 A 60 wt% Pb–40 wt% Mg alloy is rapidly
quenched to room temperature from an elevated temperature in such a way that the hightemperature microstructure is preserved. This
microstructure is found to consist of the a
phase and Mg2Pb, having respective mass
fractions of 0.42 and 0.58. Determine the approximate temperature from which the alloy
was quenched.
9.33 The microstructure of a copper–silver alloy at
775C (1425F) consists of primary a and eutectic structures. If the mass fractions of these
two microconstituents are 0.73 and 0.27, respectively, determine the composition of the
alloy.
9.34 Consider the hypothetical eutectic phase diagram for metals A and B, which is similar to
that for the lead–tin system, Figure 9.8. Assume that: (l) a and b phases exist at the A
and B extremities of the phase diagram,
respectively; (2) the eutectic composition is
36 wt% A–64 wt% B; and (3) the composition of the a phase at the eutectic temperature is 88 wt% A–12 wt% B. Determine the
composition of an alloy that will yield primary
b and total b mass fractions of 0.367 and 0.768,
respectively.
9.35 For a 64 wt% Zn–36 wt% Cu alloy, make
schematic sketches of the microstructure that
would be observed for conditions of very slow
cooling at the following temperatures: 900C
(1650F), 820C (1510F), 750C (1380F), and
600C (1100F). Label all phases and indicate
their approximate compositions.
1496T_c09_252-310 11/29/05 11:33 Page 307
REVISED PAGES
Questions and Problems • 307
9.36 For a 76 wt% Pb–24 wt% Mg alloy, make
schematic sketches of the microstructure that
would be observed for conditions of very slow
cooling at the following temperatures: 575C
(1070F), 500C (930F), 450C (840F), and
300C (570F). Label all phases and indicate
their approximate compositions.
9.37 For a 52 wt% Zn–48 wt% Cu alloy, make
schematic sketches of the microstructure that
would be observed for conditions of very slow
cooling at the following temperatures: 950C
(1740F), 860C (1580F), 800C (1470F), and
600C (1100F). Label all phases and indicate
their approximate compositions.
9.38 On the basis of the photomicrograph (i.e.,
the relative amounts of the microconstituents) for the lead–tin alloy shown in
Figure 9.17 and the Pb–Sn phase diagram
(Figure 9.8), estimate the composition of the
alloy, and then compare this estimate with
the composition given in the figure legend of
Figure 9.17. Make the following assumptions:
(1) the area fraction of each phase and microconstituent in the photomicrograph is
equal to its volume fraction; (2) the densities
of the a and b phases as well as the eutectic
structure are 11.2, 7.3, and 8.7 g/cm3, respectively; and (3) this photomicrograph represents the equilibrium microstructure at
180C (355F).
9.39 The room-temperature tensile strengths of
pure copper and pure silver are 209 MPa and
125 MPa, respectively.
(a) Make a schematic graph of the roomtemperature tensile strength versus composition for all compositions between pure copper
and pure silver. (Hint: you may want to consult Sections 9.10 and 9.11, as well as Equation 9.24 in Problem 9.64.)
(b) On this same graph schematically plot
tensile strength versus composition at 600C.
(c) Explain the shapes of these two curves, as
well as any differences between them.
Equilibrium Diagrams Having Intermediate
Phases or Compounds
9.40 Two intermetallic compounds, A3B and
AB3, exist for elements A and B. If the compositions for A3B and AB3 are 91.0 wt%
A–9.0 wt% B and 53.0 wt% A–47.0 wt% B,
respectively, and element A is zirconium,
identify element B.
Congruent Phase Transformations
Binary Eutectic Systems
Equilibrium Diagrams Having Intermediate Phases
or Compounds
Eutectoid and Peritectic Reactions
9.41 What is the principal difference between congruent and incongruent phase transformations?
9.42 Figure 9.36 is the tin–gold phase diagram, for
which only single-phase regions are labeled.
Specify temperature–composition points at
which all eutectics, eutectoids, peritectics,
and congruent phase transformations occur. Also, for each, write the reaction upon
cooling.
9.43 Figure 9.37 is a portion of the copper–
aluminum phase diagram for which only
single-phase regions are labeled. Specify
temperature–composition points at which all
eutectics, eutectoids, peritectics, and congruent phase transformations occur. Also, for
each, write the reaction upon cooling.
9.44 Construct the hypothetical phase diagram
for metals A and B between room temperature (20C) and 700C given the following
information:
• The melting temperature of metal A is 480C.
• The maximum solubility of B in A is 4 wt%
B, which occurs at 420C.
• The solubility of B in A at room temperature is 0 wt% B.
• One eutectic occurs at 420C and 18 wt%
B–82 wt% A.
• A second eutectic occurs at 475C and
42 wt% B–58 wt% A.
• The intermetallic compound AB exists at a
composition of 30 wt% B–70 wt% A, and
melts congruently at 525C.
• The melting temperature of metal B is
600C.
• The maximum solubility of A in B is
13 wt% A, which occurs at 475C.
• The solubility of A in B at room temperature is 3 wt% A.
1496T_c09_252-310 11/29/05 11:33 Page 308
REVISED PAGES
308 • Chapter 9 / Phase Diagrams
1000
800
L
Temperature (°C)
Figure 9.36 The
tin–gold phase
diagram. (Adapted
with permission
from Metals
Handbook, 8th
edition, Vol. 8,
Metallography,
Structures and Phase
Diagrams, American
Society for Metals,
Metals Park, OH,
1973.)
600
400
200
0
(Sn)
40
60
Composition (wt% Au)
80
100
(Au)
1100
1000
L
1
900
1
800
Temperature (°C)
Figure 9.37 The
copper–aluminum
phase diagram.
(Adapted with
permission from
Metals Handbook,
8th edition, Vol. 8,
Metallography
Structures and Phase
Diagrams, American
Society for Metals,
Metals Park, OH,
1973.)
20
700
2
2
600
1
1
500
2
2
400
0
(Cu)
4
8
12
16
Composition (wt% Al)
20
24
28
1496T_c09_252-310 11/29/05 11:33 Page 309
REVISED PAGES
Questions and Problems • 309
Figure 9.38
Logarithm pressureversus-temperature
phase diagram
for H2O.
10,000
Ice III
1,000
Pressure (atm)
100
Liquid
10
1.0
A
Ice I
0.1
Vapor
B
0.01
C
0.001
–20
0
20
40
60
Temperature (°C)
The Gibbs Phase Rule
9.45 In Figure 9.38 is shown the pressure–
temperature phase diagram for H2O. Apply
the Gibbs phase rule at points A, B, and C; that
is, specify the number of degrees of freedom at
each of the points—that is, the number of externally controllable variables that need be
specified to completely define the system.
9.51
The Iron–Iron Carbide (Fe–Fe3C) Phase Diagram
Development of Microstructure in
Iron–Carbon Alloys
9.46 Compute the mass fractions of a ferrite and
cementite in pearlite.
9.47 (a) What is the distinction between hypoeutectoid and hypereutectoid steels?
(b) In a hypoeutectoid steel, both eutectoid
and proeutectoid ferrite exist. Explain the difference between them. What will be the
carbon concentration in each?
9.48 What is the carbon concentration of an iron–
carbon alloy for which the fraction of total
cementite is 0.10?
9.49 What is the proeutectoid phase for an iron–
carbon alloy in which the mass fractions of
total ferrite and total cementite are 0.86 and
0.14, respectively? Why?
9.50 Consider 3.5 kg of austenite containing
0.95 wt% C, cooled to below 727C (1341F).
(a) What is the proeutectoid phase?
9.52
9.53
9.54
9.55
80
100
120
(b) How many kilograms each of total ferrite
and cementite form?
(c) How many kilograms each of pearlite and
the proeutectoid phase form?
(d) Schematically sketch and label the resulting microstructure.
Consider 6.0 kg of austenite containing
0.45 wt% C, cooled to below 727C (1341F).
(a) What is the proeutectoid phase?
(b) How many kilograms each of total ferrite
and cementite form?
(c) How many kilograms each of pearlite and
the proeutectoid phase form?
(d) Schematically sketch and label the
resulting microstructure.
Compute the mass fractions of proeutectoid
ferrite and pearlite that form in an iron–
carbon alloy containing 0.35 wt% C.
The microstructure of an iron–carbon alloy
consists of proeutectoid ferrite and pearlite; the
mass fractions of these two microconstituents
are 0.174 and 0.826, respectively. Determine the
concentration of carbon in this alloy.
The mass fractions of total ferrite and total
cementite in an iron–carbon alloy are 0.91 and
0.09, respectively. Is this a hypoeutectoid or
hypereutectoid alloy? Why?
The microstructure of an iron–carbon alloy
consists of proeutectoid cementite and pearlite;
the mass fractions of these microconstituents
1496T_c09_252-310 11/29/05 11:33 Page 310
REVISED PAGES
310 • Chapter 9 / Phase Diagrams
9.56
9.57
9.58
9.59
9.60
9.61
9.62
are 0.11 and 0.89, respectively. Determine the
concentration of carbon in this alloy.
Consider 1.5 kg of a 99.7 wt% Fe–0.3 wt% C
alloy that is cooled to a temperature just
below the eutectoid.
(a) How many kilograms of proeutectoid
ferrite form?
(b) How many kilograms of eutectoid ferrite
form?
(c) How many kilograms of cementite form?
Compute the maximum mass fraction of
proeutectoid cementite possible for a hypereutectoid iron–carbon alloy.
Is it possible to have an iron–carbon alloy for
which the mass fractions of total cementite
and proeutectoid ferrite are 0.057 and 0.36,
respectively? Why or why not?
Is it possible to have an iron–carbon alloy for
which the mass fractions of total ferrite and
pearlite are 0.860 and 0.969, respectively?
Why or why not?
Compute the mass fraction of eutectoid cementite in an iron–carbon alloy that contains
1.00 wt% C.
The mass fraction of eutectoid cementite in an
iron–carbon alloy is 0.109. On the basis of this
information, is it possible to determine the
composition of the alloy? If so, what is its
composition? If this is not possible, explain
why.
The mass fraction of eutectoid ferrite in an
iron–carbon alloy is 0.71. On the basis of this
information, is it possible to determine the
composition of the alloy? If so, what is its
composition? If this is not possible, explain
why.
9.63 For an iron–carbon alloy of composition 3 wt%
C–97 wt% Fe, make schematic sketches of
the microstructure that would be observed
for conditions of very slow cooling at the
following temperatures: 1250C (2280F),
1145C (2095F), and 700C (1290F). Label
the phases and indicate their compositions
(approximate).
9.64 Often, the properties of multiphase alloys
may be approximated by the relationship
E 1alloy2 EaVa EbVb
(9.24)
where E represents a specific property (modulus of elasticity, hardness, etc.), and V is the
volume fraction.The subscripts a and b denote
the existing phases or microconstituents. Employ the relationship above to determine the
approximate Brinell hardness of a 99.75 wt%
Fe–0.25 wt% C alloy. Assume Brinell hardnesses of 80 and 280 for ferrite and pearlite,
respectively, and that volume fractions may
be approximated by mass fractions.
The Influence of Other Alloying Elements
9.65 A steel alloy contains 95.7 wt% Fe, 4.0 wt%
W, and 0.3 wt% C.
(a) What is the eutectoid temperature of this
alloy?
(b) What is the eutectoid composition?
(c) What is the proeutectoid phase?
Assume that there are no changes in the
positions of other phase boundaries with the
addition of W.
9.66 A steel alloy is known to contain 93.65 wt%
Fe, 6.0 wt% Mn, and 0.35 wt% C.
(a) What is the approximate eutectoid temperature of this alloy?
(b) What is the proeutectoid phase when this
alloy is cooled to a temperature just below the
eutectoid?
(c) Compute the relative amounts of the
proeutectoid phase and pearlite. Assume that
there are no alterations in the positions of other
phase boundaries with the addition of Mn.
1496T_c10_311-357 12/31/05 9:55 Page 311
Chapter
2nd REVISE PAGES
10
Phase Transformations
in Metals:
Development of Microstructure and Alteration of Mechanical Properties
A
photomicrograph of a pearlitic steel that has partially transformed to spheroidite. 2000.
(Courtesy of United States Steel Corporation.)
WHY STUDY Phase Transformations in Metals?
The development of a set of desirable mechanical
characteristics for a material often results from a phase
transformation that is wrought by a heat treatment.
The time and temperature dependencies of some
phase transformations are conveniently represented on
modified phase diagrams. It is important to know how
to use these diagrams in order to design a heat treat-
ment for some alloy that will yield the desired roomtemperature mechanical properties. For example, the
tensile strength of an iron–carbon alloy of eutectoid
composition (0.76 wt% C) can be varied between
approximately 700 MPa (100,000 psi) and 2000 MPa
(300,000 psi) depending on the heat treatment
employed.
• 311
1496T_c10_311-357 12/31/05 9:55 Page 312
2nd REVISE PAGES
Learning Objectives
After careful study of this chapter you should be able to do the following:
1. Make a schematic fraction transformationpearlite, coarse pearlite, spheroidite, bainite,
versus-logarithm of time plot for a typical
martensite, and tempered martensite. Now, in
solid–solid transformation; cite the equation
terms of microstructure (or crystal structure),
that describes this behavior.
briefly explain these behaviors.
2. Briefly describe the microstructure for each of
4. Given the isothermal transformation (or continthe following microconstituents that are found
uous cooling transformation) diagram for some
in steel alloys: fine pearlite, coarse pearlite,
iron–carbon alloy, design a heat treatment that
spheroidite, bainite, martensite, and tempered
will produce a specified microstructure.
martensite.
3. Cite the general mechanical characteristics for
each of the following microconstituents: fine
10.1 INTRODUCTION
transformation rate
One reason for the versatility of metallic materials lies in the wide range of mechanical properties they possess, which are accessible to management by various
means.Three strengthening mechanisms were discussed in Chapter 7—namely, grain
size refinement, solid-solution strengthening, and strain hardening. Additional techniques are available wherein the mechanical properties are reliant on the characteristics of the microstructure.
The development of microstructure in both single- and two-phase alloys ordinarily involves some type of phase transformation—an alteration in the number
and/or character of the phases. The first portion of this chapter is devoted to a brief
discussion of some of the basic principles relating to transformations involving solid
phases. Inasmuch as most phase transformations do not occur instantaneously,
consideration is given to the dependence of reaction progress on time, or the transformation rate. This is followed by a discussion of the development of two-phase
microstructures for iron–carbon alloys. Modified phase diagrams are introduced that
permit determination of the microstructure that results from a specific heat treatment. Finally, other microconstituents in addition to pearlite are presented, and, for
each, the mechanical properties are discussed.
P h a s e Tr a n s f o r m a t i o n s
10.2 BASIC CONCEPTS
phase transformation
A variety of phase transformations are important in the processing of materials,
and usually they involve some alteration of the microstructure. For purposes of this
discussion, these transformations are divided into three classifications. In one group
are simple diffusion-dependent transformations in which there is no change in either the number or composition of the phases present. These include solidification
of a pure metal, allotropic transformations, and, recrystallization and grain growth
(see Sections 7.12 and 7.13).
In another type of diffusion-dependent transformation, there is some alteration
in phase compositions and often in the number of phases present; the final microstructure ordinarily consists of two phases. The eutectoid reaction, described by
Equation 9.19, is of this type; it receives further attention in Section 10.5.
1496T_c10_311-357 11/30/05 7:37 Page 313
REVISED PAGES
10.3 The Kinetics of Phase Transformations • 313
The third kind of transformation is diffusionless, wherein a metastable phase is
produced. As discussed in Section 10.5, a martensitic transformation, which may be
induced in some steel alloys, falls into this category.
10.3 THE KINETICS OF PHASE TRANSFORMATIONS
nucleation, growth
With phase transformations, normally at least one new phase is formed that has
different physical/chemical characteristics and/or a different structure than the
parent phase. Furthermore, most phase transformations do not occur instantaneously. Rather, they begin by the formation of numerous small particles of the
new phase(s), which increase in size until the transformation has reached completion. The progress of a phase transformation may be broken down into two distinct
stages: nucleation and growth. Nucleation involves the appearance of very small
particles, or nuclei of the new phase (often consisting of only a few hundred atoms),
which are capable of growing. During the growth stage these nuclei increase in size,
which results in the disappearance of some (or all) of the parent phase. The transformation reaches completion if the growth of these new phase particles is allowed
to proceed until the equilibrium fraction is attained. We now discuss the mechanics of these two processes, and how they relate to solid-state transformations.
Nucleation
There are two types of nucleation: homogeneous and heterogeneous. The distinction
between them is made according to the site at which nucleating events occur. For
the homogeneous type, nuclei of the new phase form uniformly throughout the
parent phase, whereas for the heterogeneous type, nuclei form preferentially at
structural inhomogeneities, such as container surfaces, insoluble impurities, grain
boundaries, dislocations, and so on. We begin by discussing homogeneous nucleation because its description and theory are simpler to treat. These principles are
then extended to a discussion of the heterogeneous type.
Homogeneous Nucleation
free energy
A discussion of the theory of nucleation involves a thermodynamic parameter called
free energy (or Gibbs free energy), G. In brief, free energy is a function of other
thermodynamic parameters, of which one is the internal energy of the system (i.e.,
the enthalpy, H), and another is a measurement of the randomness or disorder of
the atoms or molecules (i.e., the entropy, S). It is not our purpose here to provide
a detailed discussion of the principles of thermodynamics as they apply to materials systems. However, relative to phase transformations, an important thermodynamic
parameter is the change in free energy ¢G; a transformation will occur spontaneously only when ¢G has a negative value.
For the sake of simplicity, let us first consider the solidification of a pure material, assuming that nuclei of the solid phase form in the interior of the liquid as atoms
cluster together so as to form a packing arrangement similar to that found in the
solid phase. Furthermore, it will be assumed that each nucleus is spherical in geometry and has a radius r. This situation is represented schematically in Figure 10.1.
There are two contributions to the total free energy change that accompany a
solidification transformation. The first is the free energy difference between the
solid and liquid phases, or the volume free energy, ¢Gv. Its value will be negative
if the temperature is below the equilibrium solidification temperature, and the magnitude of its contribution is the product of ¢Gv and the volume of the spherical nucleus (i.e., 43 pr 3). The second energy contribution results from the formation of the
1496T_c10_311-357 12/31/05 9:55 Page 314
2nd REVISE PAGES
314 • Chapter 10 / Phase Transformations in Metals
Volume =
4 3
r
3
Liquid
Figure 10.1 Schematic diagram showing the
nucleation of a spherical solid particle in a liquid.
r
Solid
Area = 4r2
Solid-liquid
interface
solid–liquid phase boundary during the solidification transformation. Associated
with this boundary is a surface free energy, g, which is positive; furthermore, the
magnitude of this contribution is the product of g and the surface area of the nucleus (i.e., 4pr2). Finally, the total free energy change is equal to the sum of these
two contributions—that is,
¢G 43 pr3 ¢Gv 4pr2g
(10.1)
These volume, surface, and total free energy contributions are plotted schematically as a function of nucleus radius in Figures 10.2a and 10.2b. Here (Figure 10.2a)
it will be noted that for the curve corresponding to the first term on the right-hand
side of Equation 10.1, the free energy (which is negative) decreases with the third
power of r. Furthermore, for the curve resulting from the second term in Equation 10.1, energy values are positive and increase with the square of the radius.
Consequently, the curve associated with the sum of both terms (Figure 10.2b) first
increases, passes through a maximum, and finally decreases. In a physical sense,
this means that as a solid particle begins to form as atoms in the liquid cluster together, its free energy first increases. If this cluster reaches a size corresponding
to the critical radius r*, then growth will continue with the accompaniment of a
4 r2
0
+
radius, r
4 3
r Gv
3
–
(a)
Free energy change, G
+
Free energy change, G
Total free energy
change for a
solidification
transformation
r*
G*
0
radius, r
–
(b)
Figure 10.2 (a) Schematic curves for volume free energy and surface free energy
contributions to the total free energy change attending the formation of a spherical
embryo/nucleus during solidification. (b) Schematic plot of free energy versus
embryo/nucleus radius, on which is shown the critical free energy change (¢G*) and
the critical nucleus radius (r*).
1496T_c10_311-357 11/30/05 7:37 Page 315
REVISED PAGES
10.3 The Kinetics of Phase Transformations • 315
decrease in free energy. On the other hand, a cluster of radius less than the critical will shrink and redissolve. This subcritical particle is an embryo, whereas the
particle of radius greater than r* is termed a nucleus. A critical free energy, ¢G*,
occurs at the critical radius and, consequently, at the maximum of the curve in
Figure 10.2b. This ¢G* corresponds to an activation free energy, which is the free
energy required for the formation of a stable nucleus. Equivalently, it may be considered an energy barrier to the nucleation process.
Since r* and ¢G* appear at the maximum on the free energy-versus-radius curve
of Figure 10.2b, derivation of expressions for these two parameters is a simple matter. For r*, we differentiate the ¢G equation (Equation 10.1) with respect to r, set
the resulting expression equal to zero, and then solve for r ( r*). That is,
d1¢G2
43 p¢Gv 13r2 2 4pg12r2 0
dr
(10.2)
which leads to the result
For homogeneous
nucleation, critical
radius of a stable
solid particle nucleus
r*
2g
¢Gv
(10.3)
Now, substitution of this expression for r* into Equation 10.1 yields the following
expression for ¢G*:
For homogeneous
nucleation,
activation free
energy required for
the formation of a
stable nucleus
¢G*
16pg3
31¢Gv 2 2
(10.4)
This volume free energy change ¢Gv is the driving force for the solidification
transformation, and its magnitude is a function of temperature. At the equilibrium
solidification temperature Tm, the value of ¢Gv is zero, and with diminishing temperature its value becomes increasingly more negative.
It can be shown that ¢Gv is a function of temperature as
¢Gv
¢Hf 1Tm T 2
Tm
(10.5)
where ¢Hf is the latent heat of fusion (i.e., the heat given up during solidification),
and Tm and the temperature T are in Kelvin. Substitution of this expression for ¢Gv
into Equations 10.3 and 10.4 yields
Dependence of
critical radius on
surface free energy,
latent heat of fusion,
melting temperature,
and transformation
temperature
Activation free
energy expression
r* a
2gTm
1
ba
b
¢Hf
Tm T
(10.6)
16pg3T m2
1
b
3¢H 2f
1Tm T 2 2
(10.7)
and
¢G* a
Thus, from these two equations, both the critical radius r* and the activation free energy ¢G* decrease as temperature T decreases. (The g and ¢Hf parameters in these
expressions are relatively insensitive to temperature changes.) Figure 10.3, a schematic
¢G-versus-r plot that shows curves for two different temperatures, illustrates these
relationships. Physically, this means that with a lowering of temperature at temperatures below the equilibrium solidification temperature (Tm), nucleation occurs more
1496T_c10_311-357 11/30/05 7:37 Page 316
REVISED PAGES
316 • Chapter 10 / Phase Transformations in Metals
+
r*1
T 2 > T1
r*2
G
G*1
at T1
G*
2
0
radius, r
Figure 10.3 Schematic freeenergy-versus-embryo/nucleusradius curves for two different
temperatures. The critical free
energy change (¢G*) and critical
nucleus radius (r*) are indicated
for each temperature.
at T2
–
readily. Furthermore, the number of stable nuclei n* (having radii greater than r*) is
a function of temperature as
n* K1 exp a
¢G*
b
kT
(10.8)
where the constant K1 is related to the total number of nuclei of the solid phase.
For the exponential term of this expression, changes in temperature have a greater
effect on the magnitude of the ¢G* term in the numerator than the T term in the
denominator. Consequently, as the temperature is lowered below Tm the exponential term in Equation 10.8 also decreases such that the magnitude of n* increases.
This temperature dependence (n* versus T) is represented in the schematic plot of
Figure 10.4a.
There is another important temperature-dependent step that is involved in and
also influences nucleation: the clustering of atoms by short-range diffusion during
the formation of nuclei. The influence of temperature on the rate of diffusion (i.e.,
magnitude of the diffusion coefficient, D) is given in Equation 5.8. Furthermore,
this diffusion effect is related to the frequency at which atoms from the liquid attach themselves to the solid nucleus, nd. Or, the dependence of nd on temperature
is the same as for the diffusion coefficient—namely,
nd K2 exp a
(10.9)
Tm
Tm
Tm
G*
kT
exp –
Qd
kT
d
Temperature
exp –
Temperature
T
Temperature
Figure 10.4 For
solidification,
schematic plots of
(a) number of stable
nuclei versus
temperature,
(b) frequency of
atomic attachment
versus temperature,
and (c) nucleation
rate versus
temperature (also
shown are curves for
parts a and b).
Qd
b
kT
.
N
n*
.
Number of stable nuclei, n*
Frequency of attachment, d
n*, d, N
(a)
(b)
(c)
1496T_c10_311-357 11/30/05 7:37 Page 317
REVISED PAGES
10.3 The Kinetics of Phase Transformations • 317
where Qd is a temperature-independent parameter—the activation energy for
diffusion—and K2 is a temperature-independent constant. Thus, from Equation 10.9
a diminishment of temperature results in a reduction in nd. This effect, represented
by the curve shown in Figure 10.4b, is just the reverse of that for n* as discussed
above.
The principles and concepts just developed are now extended
# to a discussion
of another important nucleation parameter, the nucleation rate N (which has units
of nuclei per unit volume per second). This rate is simply proportional to the product of n* (Equation 10.8) and nd (Equation 10.9); that is,
Nucleation rate
expression for
homogeneous
nucleation
#
Qd
¢G*
N K3n*nd K1K2K3 c exp a
b exp a b d
kT
kT
(10.10)
Here K3 is the number of atoms on a nucleus surface. Figure 10.4c schematically
plots nucleation rate as a function of temperature
and, in addition, the curves of
#
Figures 10.4a and 10.4b from which the N curve is derived. Note (Figure 10.4c) that,
with a lowering of temperature from below Tm, the nucleation rate first increases,
achieves a maximum, and
# subsequently diminishes.
The shape of this N curve is explained as
# follows: for the upper region of the
curve (a sudden and dramatic increase in N with decreasing T ), ¢G* is greater
than Qd, which means that the exp(¢G*kT ) term of Equation 10.10 is much
smaller than exp (Qd kT). In other words, the nucleation rate is suppressed at high
temperatures due to a small activation driving force. With continued diminishment
of temperature, there comes a point at which ¢G* becomes smaller than the
temperature-independent Qd with the result that exp(QdkT ) 6 exp(¢G*kT),
or that, at lower temperatures, a low atomic mobility suppresses the nucleation
rate. This# accounts for the shape of the lower curve segment (a precipitous reduc#
tion of N with a continued diminishment of temperature). Furthermore, the N
curve of Figure 10.4c necessarily passes through a maximum over the intermediate temperature range where values for ¢G* and Qd are of approximately the
same magnitude.
Several qualifying comments are in order regarding the above discussion. First,
although we assumed a spherical shape for nuclei, this method may be applied to
any shape with the same final result. Furthermore, this treatment may be utilized
for types of transformations other than solidification (i.e., liquid–solid)—for example, solid–vapor and solid–solid. However, magnitudes of ¢Gv and g, in addition to
diffusion rates of the atomic species, will undoubtedly differ among the various
transformation types. In addition, for solid–solid transformations, there may be
volume changes attendant to the formation of new phases. These changes may lead
to the introduction of microscopic strains, which must be taken into account in the
¢G expression of Equation 10.1, and, consequently, will affect the magnitudes of
r* and ¢G*.
From Figure 10.4c it is apparent that during the cooling of a liquid, an
appreciable nucleation rate (i.e., solidification) will begin only after the temperature has been lowered to below the equilibrium solidification (or melting) temperature (Tm). This phenomenon is termed supercooling (or undercooling), and
the degree of supercooling for homogeneous nucleation may be significant (on
the order of several hundred degrees Kelvin) for some systems. In Table 10.1 is
tabulated, for several materials, typical degrees of supercooling for homogeneous
nucleation.
1496T_c10_311-357 11/30/05 7:37 Page 318
REVISED PAGES
318 • Chapter 10 / Phase Transformations in Metals
Table 10.1 Degree of Supercooling (T ) Values
(Homogeneous Nucleation)
for Several Metals
Metal
Antimony
Germanium
Silver
Gold
Copper
Iron
Nickel
Cobalt
Palladium
T (C )
135
227
227
230
236
295
319
330
332
Source: D. Turnbull and R. E. Cech, “Microscopic
Observation of the Solidification of Small Metal
Droplets,” J. Appl. Phys., 21, 808 (1950).
EXAMPLE PROBLEM 10.1
Computation of Critical Nucleus Radius and
Activation Free Energy
(a) For the solidification of pure gold, calculate the critical radius r* and the
activation free energy ¢G* if nucleation is homogeneous. Values for the latent
heat of fusion and surface free energy are 1.16 109 J/m3 and 0.132 J/m2,
respectively. Use the supercooling value found in Table 10.1.
(b) Now calculate the number of atoms found in a nucleus of critical size. Assume a lattice parameter of 0.413 nm for solid gold at its melting temperature.
Solution
(a) In order to compute the critical radius, we employ Equation 10.6, using
the melting temperature of 1064C for gold, assuming a supercooling value of
230C (Table 10.1), and realizing that ¢Hf is negative. Hence
r* a
c
2gTm
1
ba
b
¢Hf
Tm T
12210.132 J/m2 211064 273 K2
1.16 10 J/m
9
3
da
1
b
230 K
1.32 109 m 1.32 nm
For computation of the activation free energy, Equation 10.7 is employed.
Thus
16pg3Tm2
1
b
¢G* a
2
3¢Hf
1Tm T 2 2
11621p210.132 J/m2 2 3 11064 273 K2 2
1
c
dc
d
13211.16 109 J/m3 2 2
1230 K2 2
9.64 1019 J
1496T_c10_311-357 11/30/05 7:37 Page 319
REVISED PAGES
10.3 The Kinetics of Phase Transformations • 319
(b) In order to compute the number of atoms in a nucleus of critical size (assuming a spherical nucleus of radius r*), it is first necessary to determine the
number of unit cells, which we then multiply by the number of atoms per unit
cell. The number of unit cells found in this critical nucleus is just the ratio of
critical nucleus and unit cell volumes. Inasmuch as gold has the FCC crystal
structure (and a cubic unit cell), its unit cell volume is just a3, where a is the lattice parameter (i.e., unit cell edge length); its value is 0.413 nm, as cited in the
problem statement. Therefore, the number of unit cells found in a radius of critical size is just
4
pr*3
critical nucleus volume
3
# unit cells/particle
(10.11)
unit cell volume
a3
4
a b1p211.32 nm2 3
3
137 unit cells
10.413 nm2 3
Inasmuch as there is the equivalence of four atoms per FCC unit cell (Section 3.4), the total number of atoms per critical nucleus is just
(137 unit cells/critical nucleus)(4 atoms/unit cell) 548 atoms/critical nucleus
Heterogeneous Nucleation
For heterogeneous
nucleation of a solid
particle, relationship
among solid-surface,
solid-liquid, and
liquid-surface
interfacial energies
and the wetting
angle
Although levels of supercooling for homogeneous nucleation may be significant (on
occasion several hundred degrees Celsius), in practical situations they are often on
the order of only several degrees Celsius. The reason for this is that the activation
energy (i.e., energy barrier) for nucleation ( ¢G* of Equation 10.4) is lowered when
nuclei form on preexisting surfaces or interfaces, since the surface free energy (g
of Equation 10.4) is reduced. In other words, it is easier for nucleation to occur at
surfaces and interfaces than at other sites. Again, this type of nucleation is termed
heterogeneous.
In order to understand this phenomenon, let us consider the nucleation, on a
flat surface, of a solid particle from a liquid phase. It is assumed that both the liquid and solid phases “wet” this flat surface, that is, both of these phases spread out
and cover the surface; this configuration is depicted schematically in Figure 10.5.
Also noted in the figure are three interfacial energies (represented as vectors) that
exist at two-phase boundaries—gSL, gSI, and gIL—as well as the wetting angle u (the
angle between the gSI and gSL vectors). Taking a surface tension force balance in
the plane of the flat surface leads to the following expression:
gIL gSI gSL cos u
Liquid
Solid
SL
IL
SI
Surface or interface
(10.12)
Figure 10.5 Heterogeneous
nucleation of a solid from a liquid.
The solid–surface (gSI), solid–liquid
(gSL), and liquid–surface (gIL)
interfacial energies are represented
by vectors. The wetting angle (u) is
also shown.
1496T_c10_311-357 11/30/05 7:37 Page 320
REVISED PAGES
320 • Chapter 10 / Phase Transformations in Metals
Figure 10.6 Schematic freeenergy-versus-embryo/nucleusradius plot on which are
presented curves for both
homogeneous and
heterogeneous nucleation.
Critical free energies and the
critical radius are also shown.
r*
G*hom
G
G*het
0
r
Now, using a somewhat involved procedure similar to the one presented above
for homogeneous nucleation (which we have chosen to omit), it is possible to derive
equations for r* and ¢G*; these are as follows:
For heterogeneous
nucleation, critical
radius of a stable
solid particle nucleus
For heterogeneous
nucleation,
activation free
energy required for
the formation of a
stable nucleus
r*
¢G* a
2gSL
¢Gv
16pg3SL
b S1u2
3¢Gv2
(10.13)
(10.14)
The S(u) term of this last equation is a function only of u (i.e., the shape of the nucleus), which will have a numerical value between zero and unity.1
From Equation 10.13, it is important to note that the critical radius r* for heterogeneous nucleation is the same as for homogeneous, inasmuch as gSL is the same surface energy as g in Equation 10.3. It is also evident that the activation energy barrier
for heterogeneous nucleation (Equation 10.14) is smaller than the homogeneous barrier (Equation 10.4) by an amount corresponding to the value of this S(u) function, or
* ¢Ghom
¢Ghet
* S1u2
(10.15)
Figure 10.6, a schematic graph of ¢G versus nucleus radius, plots curves for both
types of nucleation, and indicates the difference in the magnitudes of ¢G*het and
¢G*hom, in addition to the constancy of r*. This lower ¢G* for heterogeneous means
that a smaller energy must be overcome during the nucleation process (than for homogeneous), and, therefore, heterogeneous nucleation
occurs more readily (Equa#
tion 10.10). In terms of the nucleation rate, the N versus T curve (Figure 10.4c) is
shifted to higher temperatures for heterogeneous. This effect is represented in Figure 10.7, which also shows that a much smaller degree of supercooling (¢T) is
required for heterogeneous nucleation.
Growth
The growth step in a phase transformation begins once an embryo has exceeded
the critical size, r*, and becomes a stable nucleus. Note that nucleation will continue
to occur simultaneously with growth of the new phase particles; of course, nucleation
1
For example, for u angles of 30 and 90, values of S(u) are approximately 0.01 and 0.5,
respectively.
1496T_c10_311-357 12/31/05 9:55 Page 321
2nd REVISE PAGES
10.3 The Kinetics of Phase Transformations • 321
Figure 10.7 Nucleation rate versus temperature for
both homogeneous and heterogeneous nucleation.
Degree of supercooling (¢T ) for each is also shown.
Tm
T het
T hom
Temperature
.
Nhet
.
Nhom
Nucleation rate
cannot occur in regions that have already transformed to the new phase. Furthermore, the growth process will cease in any region where particles of the new phase
meet, since here the transformation will have reached completion.
Particle growth occurs by long-range atomic diffusion, which normally involves
several steps—for example, diffusion through the parent phase, across
a phase
#
boundary, and then into the nucleus. Consequently, the growth rate G is determined
by the rate of diffusion, and its temperature dependence is the same as for the diffusion coefficient (Equation 5.8)—namely,
Dependence of
particle growth rate
on the activation
energy for diffusion
and temperature
#
Q
G C exp a b
kT
(10.16)
where Q (the activation energy) and C (a preexponential)
are independent of tem#
perature.2 The temperature dependence of G is represented# by one of the curves
in Figure 10.8; also shown is a curve for the nucleation rate, N (again, almost always
the rate for heterogeneous nucleation). Now, at a specific
temperature,
the overall
#
#
transformation rate is equal to some product of N and G. The third curve of
Tm
.
Temperature
Growth rate, G
Figure 10.8 Schematic plot# showing
curves
for nucleation rate (N ), growth rate
#
(G ), and overall transformation rate
versus temperature.
Overall
transformation
rate
.
Nucleation rate, N
Rate
#
Processes the rates of which depend on temperature as G in Equation (10.16) are sometimes termed thermally activated. Also, a rate equation of this form (i.e., having the exponential temperature dependence) is termed an Arrhenius rate equation.
2
thermally activated
transformation
1496T_c10_311-357 12/31/05 9:55 Page 322
2nd REVISE PAGES
322 • Chapter 10 / Phase Transformations in Metals
Figure 10.8, which is for the total rate, represents this combined effect. The general
shape of this curve is the same as for the nucleation rate,
# in that it has a peak or
maximum that has been shifted upward relative to the N curve.
Whereas this treatment on transformations has been developed for solidification,
the same general principles also apply to solid–solid and solid–gas transformations.
As we shall see below, the rate of transformation and the time required for the
transformation to proceed to some degree of completion (e.g., time to 50% reaction
completion, t0.5) are inversely proportional to one another (Equation 10.18). Thus, if
the logarithm of this transformation time (i.e., log t0.5) is plotted versus temperature,
a curve having the general shape shown in Figure 10.9b results. This “C-shaped”
curve is a virtual mirror image (through a vertical plane) of the transformation rate
curve of Figure 10.8, as demonstrated in Figure 10.9. It is often the case that the
kinetics of phase transformations are represented using logarithm time- (to some
degree of transformation) versus-temperature plots (for example, see Section 10.5).
Several physical phenomena may be explained in terms of the transformation
rate-versus-temperature curve of Figure 10.8. First, the size of the product phase
particles will depend on transformation temperature. For example, for transformations that occur at temperatures near to Tm, corresponding to low nucleation and
high growth rates, few nuclei form that grow rapidly. Thus, the resulting microstructure will consist of few and relatively large phase particles (e.g., coarse grains).
Conversely, for transformations at lower temperatures, nucleation rates are high
and growth rates low, which results in many small particles (e.g., fine grains).
Also, from Figure 10.8, when a material is cooled very rapidly through the temperature range encompassed by the transformation rate curve to a relatively low
temperature where the rate is extremely low, it is possible to produce nonequilibrium phase structures (for example, see Sections 10.5 and 11.9).
Kinetic Considerations of Solid-State Transformations
Te
Te
Temperature
Figure 10.9 Schematic plots of (a)
transformation rate versus temperature,
and (b) logarithm time [to some degree
(e.g., 0.5 fraction) of transformation] versus
temperature. The curves in both (a) and
(b) are generated from the same set of
data—i.e., for horizontal axes, the time
[scaled logarithmically in the (b) plot] is
just the reciprocal of the rate from plot (a).
Temperature
kinetics
The previous discussion of this section has centered on the temperature dependences
of nucleation, growth, and transformation rates. The time dependence of rate (which
is often termed the kinetics of a transformation) is also an important consideration,
often in the heat treatment of materials. Also, since many transformations of interest to materials scientists and engineers involve only solid phases, we have decided
to devote the following discussion to the kinetics of solid-state transformations.
With many kinetic investigations, the fraction of reaction that has occurred is
measured as a function of time while the temperature is maintained constant. Transformation progress is usually ascertained by either microscopic examination or measurement of some physical property (such as electrical conductivity) the magnitude
Rate
1
t0.5
(a)
Time (t0.5)
(logarithmic scale)
(b)
1496T_c10_311-357 11/30/05 7:37 Page 323
REVISED PAGES
10.3 The Kinetics of Phase Transformations • 323
Fraction of transformation, y
1.0
Figure 10.10 Plot of fraction
reacted versus the logarithm of
time typical of many solid-state
transformations in which
temperature is held constant.
0.5
t0.5
0
Nucleation
Growth
Logarithm of heating time, t
of which is distinctive of the new phase. Data are plotted as the fraction of transformed material versus the logarithm of time; an S-shaped curve similar to that in
Figure 10.10 represents the typical kinetic behavior for most solid-state reactions.
Nucleation and growth stages are also indicated in the figure.
For solid-state transformations displaying the kinetic behavior in Figure 10.10,
the fraction of transformation y is a function of time t as follows:
Transformation
rate—reciprocal of
the halfway-tocompletion
transformation time
y 1 exp1kt n 2
(10.17)
where k and n are time-independent constants for the particular reaction. The above
expression is often referred to as the Avrami equation.
By convention, the rate of a transformation is taken as the reciprocal of time
required for the transformation to proceed halfway to completion, t0.5, or
rate
1
(10.18)
t0.5
Temperature will have a profound influence on the kinetics and thus on the
rate of a transformation. This is demonstrated in Figure 10.11, where y-versus-log t
100
Percent recrystallized
Avrami equation—
dependence of
fraction of
transformation
on time
80
135C
60
119C
113C 102C
88C
43C
40
20
0
1
10
102
Time (min)
(Logarithmic scale)
104
Figure 10.11 Percent recrystallization as a function of time and at constant
temperature for pure copper. (Reprinted with permission from Metallurgical
Transactions, Vol. 188, 1950, a publication of The Metallurgical Society of AIME,
Warrendale, PA. Adapted from B. F. Decker and D. Harker, “Recrystallization in Rolled
Copper,” Trans. AIME, 188, 1950, p. 888.)
1496T_c10_311-357 11/30/05 7:37 Page 324
REVISED PAGES
324 • Chapter 10 / Phase Transformations in Metals
S-shaped curves at several temperatures for the recrystallization of copper are
shown.
A detailed discussion on the influence of both temperature and time on phase
transformations is provided in Section 10.5.
10.4 METASTABLE VERSUS EQUILIBRIUM STATES
supercooling
superheating
Phase transformations may be wrought in metal alloy systems by varying temperature, composition, and the external pressure; however, temperature changes by means
of heat treatments are most conveniently utilized to induce phase transformations.
This corresponds to crossing a phase boundary on the composition-temperature
phase diagram as an alloy of given composition is heated or cooled.
During a phase transformation, an alloy proceeds toward an equilibrium state
that is characterized by the phase diagram in terms of the product phases, their
compositions, and relative amounts.As the previous section noted, most phase transformations require some finite time to go to completion, and the speed or rate is
often important in the relationship between the heat treatment and the development of microstructure. One limitation of phase diagrams is their inability to indicate the time period required for the attainment of equilibrium.
The rate of approach to equilibrium for solid systems is so slow that true
equilibrium structures are rarely achieved. When phase transformations are induced
by temperature changes, equilibrium conditions are maintained only if heating or
cooling is carried out at extremely slow and unpractical rates. For other than
equilibrium cooling, transformations are shifted to lower temperatures than indicated by the phase diagram; for heating, the shift is to higher temperatures. These
phenomena are termed supercooling and superheating, respectively. The degree of
each depends on the rate of temperature change; the more rapid the cooling or
heating, the greater the supercooling or superheating. For example, for normal cooling rates the iron–carbon eutectoid reaction is typically displaced 10 to 20C (18 to
36F) below the equilibrium transformation temperature.3
For many technologically important alloys, the preferred state or microstructure is a metastable one, intermediate between the initial and equilibrium states;
on occasion, a structure far removed from the equilibrium one is desired. It thus
becomes imperative to investigate the influence of time on phase transformations.
This kinetic information is, in many instances, of greater value than a knowledge
of the final equilibrium state.
M i c ro s t r u c t u r a l a n d P ro p e r t y C h a n ge s i n
I ro n – C a r b o n A l l oys
Some of the basic kinetic principles of solid-state transformations are now extended
and applied specifically to iron–carbon alloys in terms of the relationships among
heat treatment, the development of microstructure, and mechanical properties. This
3
It is important to note that the treatments relating to the kinetics of phase transformations in Section 10.3 are constrained to the condition of constant temperature. By way of
contrast, the discussion of this section pertains to phase transformations that occur with
changing temperature. This same distinction exists between Sections 10.5 (Isothermal
Transformation Diagrams) and 10.6 (Continuous Cooling Transformation Diagrams).
1496T_c10_311-357 11/30/05 7:37 Page 325
REVISED PAGES
10.5 Isothermal Transformation Diagrams • 325
system has been chosen because it is familiar and because a wide variety of microstructures and mechanical properties are possible for iron–carbon (or steel) alloys.
10.5 ISOTHERMAL TRANSFORMATION DIAGRAMS
Pearlite
Consider again the iron–iron carbide eutectoid reaction
Eutectoid reaction
for the iron-iron
carbide system
cooling
g10.76 wt% C2 Δ a10.022 wt% C2 Fe3C16.70 wt% C2
heating
(10.19)
which is fundamental to the development of microstructure in steel alloys. Upon
cooling, austenite, having an intermediate carbon concentration, transforms to a ferrite phase, having a much lower carbon content, and also cementite, with a much
higher carbon concentration. Pearlite is one microstructural product of this transformation (Figure 9.27), and the mechanism of pearlite formation was discussed
previously (Section 9.19) and demonstrated in Figure 9.28.
Temperature plays an important role in the rate of the austenite-to-pearlite
transformation. The temperature dependence for an iron–carbon alloy of eutectoid
composition is indicated in Figure 10.12, which plots S-shaped curves of the percentage transformation versus the logarithm of time at three different temperatures.
For each curve, data were collected after rapidly cooling a specimen composed of
100% austenite to the temperature indicated; that temperature was maintained constant throughout the course of the reaction.
A more convenient way of representing both the time and temperature dependence of this transformation is in the bottom portion of Figure 10.13. Here, the
vertical and horizontal axes are, respectively, temperature and the logarithm of time.
Two solid curves are plotted; one represents the time required at each temperature
for the initiation or start of the transformation; the other is for the transformation
conclusion. The dashed curve corresponds to 50% of transformation completion.
These curves were generated from a series of plots of the percentage transformation versus the logarithm of time taken over a range of temperatures. The S-shaped
curve [for 675C (1247F)], in the upper portion of Figure 10.13, illustrates how the
data transfer is made.
In interpreting this diagram, note first that the eutectoid temperature [727C
(1341F)] is indicated by a horizontal line; at temperatures above the eutectoid and
Percent pearlite
600°C
50
650°C
0
1
10
Time (s)
675°C
102
50
100
103
Percent austenite
0
100
Figure 10.12 For an
iron–carbon alloy of
eutectoid composition
(0.76 wt% C),
isothermal fraction
reacted versus the
logarithm of time for
the austenite-to-pearlite
transformation.
1496T_c10_311-357 11/30/05 7:37 Page 326
REVISED PAGES
Percent of austenite
transformed to pearlite
326 • Chapter 10 / Phase Transformations in Metals
100
Transformation
ends
Transformation
temperature 675°C
50
Transformation
begins
0
1
102
10
103
104
105
Time (s)
Eutectoid temperature
Austenite (stable)
Austenite
(unstable)
1200
Pearlite
600
50% Completion curve
1000
Completion curve
(~100% pearlite)
500
Begin curve
(~ 0% pearlite)
400
1
10
Temperature (°F)
Temperature (°C)
700
1400
Figure 10.13
Demonstration of how
an isothermal
transformation diagram
(bottom) is generated
from percentage
transformation-versuslogarithm of time
measurements (top).
[Adapted from H.
Boyer, (Editor), Atlas
of Isothermal
Transformation and
Cooling Transformation
Diagrams, American
Society for Metals, 1977,
p. 369.]
800
102
103
104
105
Time (s)
isothermal
transformation
diagram
for all times, only austenite will exist, as indicated in the figure.The austenite-to-pearlite
transformation will occur only if an alloy is supercooled to below the eutectoid; as indicated by the curves, the time necessary for the transformation to begin and then end
depends on temperature. The start and finish curves are nearly parallel, and they approach the eutectoid line asymptotically. To the left of the transformation start curve,
only austenite (which is unstable) will be present, whereas to the right of the finish
curve, only pearlite will exist. In between, the austenite is in the process of transforming to pearlite, and thus both microconstituents will be present.
According to Equation 10.18, the transformation rate at some particular temperature is inversely proportional to the time required for the reaction to proceed to 50%
completion (to the dashed line in Figure 10.13). That is, the shorter this time, the
higher is the rate. Thus, from Figure 10.13, at temperatures just below the eutectoid
(corresponding to just a slight degree of undercooling) very long times (on the order
of 105 s) are required for the 50% transformation, and therefore the reaction rate is
very slow. The transformation rate increases with decreasing temperature such that at
540C (1000F) only about 3 s is required for the reaction to go to 50% completion.
Several constraints are imposed on using diagrams like Figure 10.13. First,
this particular plot is valid only for an iron–carbon alloy of eutectoid composition; for other compositions, the curves will have different configurations. In
addition, these plots are accurate only for transformations in which the temperature of the alloy is held constant throughout the duration of the reaction. Conditions of constant temperature are termed isothermal; thus, plots such as Figure
10.13 are referred to as isothermal transformation diagrams, or sometimes as
time–temperature–transformation (or T–T–T ) plots.
1496T_c10_311-357 12/31/05 9:55 Page 327
2nd REVISE PAGES
10.5 Isothermal Transformation Diagrams • 327
1s
700
␥
Temperature (°C)
␥
␥
␥
1 day
Eutectoid
temperature
Austenite (stable)
␥
727°C
1h
1400
␥
␣ Ferrite
␥
Coarse pearlite
1200
C
B
D
600
Fe3C
Fine pearlite
Temperature (°F)
A
1 min
1000
500
Austenite → pearlite
transformation
Denotes that a transformation
is occurring
800
1
10
102
103
104
105
Time (s)
Figure 10.14 Isothermal transformation diagram for a eutectoid iron–carbon alloy, with
superimposed isothermal heat treatment curve (ABCD). Microstructures before, during,
and after the austenite-to-pearlite transformation are shown. [Adapted from H. Boyer
(Editor), Atlas of Isothermal Transformation and Cooling Transformation Diagrams,
American Society for Metals, 1977, p. 28.]
coarse pearlite
fine pearlite
An actual isothermal heat treatment curve (ABCD) is superimposed on the
isothermal transformation diagram for a eutectoid iron–carbon alloy in Figure 10.14.
Very rapid cooling of austenite to a temperature is indicated by the near-vertical
line AB, and the isothermal treatment at this temperature is represented by the
horizontal segment BCD. Of course, time increases from left to right along this line.
The transformation of austenite to pearlite begins at the intersection, point C (after approximately 3.5 s), and has reached completion by about 15 s, corresponding
to point D. Figure 10.14 also shows schematic microstructures at various times during the progression of the reaction.
The thickness ratio of the ferrite and cementite layers in pearlite is approximately 8 to 1. However, the absolute layer thickness depends on the temperature
at which the isothermal transformation is allowed to occur. At temperatures just
below the eutectoid, relatively thick layers of both the a-ferrite and Fe3C phases
are produced; this microstructure is called coarse pearlite, and the region at which
it forms is indicated to the right of the completion curve on Figure 10.14. At these
temperatures, diffusion rates are relatively high, such that during the transformation illustrated in Figure 9.28 carbon atoms can diffuse relatively long distances,
which results in the formation of thick lamellae. With decreasing temperature, the
carbon diffusion rate decreases, and the layers become progressively thinner. The
thin-layered structure produced in the vicinity of 540C is termed fine pearlite; this
is also indicated in Figure 10.14. To be discussed in Section 10.7 is the dependence
of mechanical properties on lamellar thickness. Photomicrographs of coarse and
fine pearlite for a eutectoid composition are shown in Figure 10.15.
1496T_c10_311-357 11/30/05 7:37 Page 328
REVISED PAGES
328 • Chapter 10 / Phase Transformations in Metals
Figure 10.15
Photomicrographs of
(a) coarse pearlite
and (b) fine pearlite.
3000. (From K. M.
Ralls et al., An
Introduction to
Materials Science and
Engineering, p. 361.
Copyright © 1976 by
John Wiley & Sons,
New York. Reprinted
by permission of
John Wiley & Sons,
Inc.)
For iron–carbon alloys of other compositions, a proeutectoid phase (either
ferrite or cementite) will coexist with pearlite, as discussed in Section 9.19. Thus
additional curves corresponding to a proeutectoid transformation also must be included on the isothermal transformation diagram. A portion of one such diagram
for a 1.13 wt% C alloy is shown in Figure 10.16.
Bainite
In addition to pearlite, other microconstituents that are products of the austenitic
transformation exist; one of these is called bainite. The microstructure of bainite
consists of ferrite and cementite phases, and thus diffusional processes are involved
900
1600
A
800
Eutectoid temperature
+
700
C
A
A
+
600
1400
1200
P
P
1000
500
1
10
102
Time (s)
103
104
Temperature (°F)
A
Temperature (°C)
bainite
Figure 10.16
Isothermal
transformation diagram
for a 1.13 wt% C
iron–carbon alloy:
A, austenite; C,
proeutectoid cementite;
P, pearlite. [Adapted
from H. Boyer (Editor),
Atlas of Isothermal
Transformation and
Cooling Transformation
Diagrams, American
Society for Metals, 1977,
p. 33.]
1496T_c10_311-357 11/30/05 7:37 Page 329
REVISED PAGES
10.5 Isothermal Transformation Diagrams • 329
Figure 10.17 Transmission electron
micrograph showing the structure of bainite.
A grain of bainite passes from lower left to
upper right-hand corners, which consists of
elongated and needle-shaped particles of
Fe3C within a ferrite matrix. The phase
surrounding the bainite is martensite.
(Reproduced with permission from Metals
Handbook, 8th edition, Vol. 8, Metallography,
Structures and Phase Diagrams, American
Society for Metals, Materials Park, OH,
1973.)
Martensite
Cementite
Ferrite
in its formation. Bainite forms as needles or plates, depending on the temperature
of the transformation; the microstructural details of bainite are so fine that their
resolution is possible only using electron microscopy. Figure 10.17 is an electron
micrograph that shows a grain of bainite (positioned diagonally from lower left to
upper right); it is composed of a ferrite matrix and elongated particles of Fe3C; the
various phases in this micrograph have been labeled. In addition, the phase that
surrounds the needle is martensite, the topic to which a subsequent section is
addressed. Furthermore, no proeutectoid phase forms with bainite.
The time–temperature dependence of the bainite transformation may also be
represented on the isothermal transformation diagram. It occurs at temperatures
below those at which pearlite forms; begin-, end-, and half-reaction curves are just
extensions of those for the pearlitic transformation, as shown in Figure 10.18, the
isothermal transformation diagram for an iron–carbon alloy of eutectoid composition that has been extended to lower temperatures. All three curves are C-shaped
and have a “nose” at point N, where the rate of transformation is a maximum. As
may be noted, whereas pearlite forms above the nose [i.e., over the temperature
range of about 540 to 727C (1000 to 1341F)], at temperatures between about 215
and 540C (420 and 1000F), bainite is the transformation product.
It should also be noted that pearlitic and bainitic transformations are really
competitive with each other, and once some portion of an alloy has transformed
to either pearlite or bainite, transformation to the other microconstituent is not
possible without reheating to form austenite.
Spheroidite
spheroidite
If a steel alloy having either pearlitic or bainitic microstructures is heated to,
and left at, a temperature below the eutectoid for a sufficiently long period of
time—for example, at about 700C (1300F) for between 18 and 24 h—yet another
microstructure will form. It is called spheroidite (Figure 10.19). Instead of the alternating ferrite and cementite lamellae (pearlite), or the microstructure observed
for bainite, the Fe3C phase appears as sphere-like particles embedded in a continuous phase matrix. This transformation has occurred by additional carbon diffusion
with no change in the compositions or relative amounts of ferrite and cementite
1496T_c10_311-357 11/30/05 7:37 Page 330
REVISED PAGES
330 • Chapter 10 / Phase Transformations in Metals
800
A
Eutectoid temperature
1400
700
A
1200
A
+
P
1000
N
500
A+B
800
B
400
A
600
300
50%
200
100
10–1
1
10
102
103
400
104
Temperature (°F)
Temperature (°C)
600
P
Figure 10.18
Isothermal
transformation diagram
for an iron–carbon
alloy of eutectoid
composition, including
austenite-to-pearlite
(A–P) and austeniteto-bainite (A–B)
transformations.
[Adapted from H.
Boyer (Editor), Atlas
of Isothermal
Transformation
and Cooling
Transformation
Diagrams, American
Society for Metals,
1977, p. 28.]
105
Time (s)
phases. The chapter-opening photograph for this chapter is a photomicrograph that
shows a pearlitic steel that has partially transformed to spheroidite. The driving force
for this transformation is the reduction in a–Fe3C phase boundary area. The kinetics of spheroidite formation are not included on isothermal transformation diagrams.
Concept Check 10.1
Which is the more stable, the pearlitic or the spheroiditic microstructure? Why?
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
Figure 10.19 Photomicrograph of a steel
having a spheroidite microstructure. The
small particles are cementite; the continuous
phase is a ferrite. 1000. (Copyright 1971
by United States Steel Corporation.)
1496T_c10_311-357 11/30/05 7:37 Page 331
REVISED PAGES
10.5 Isothermal Transformation Diagrams • 331
Martensite
martensite
Yet another microconstituent or phase called martensite is formed when austenitized iron–carbon alloys are rapidly cooled (or quenched) to a relatively low temperature (in the vicinity of the ambient). Martensite is a nonequilibrium single-phase
structure that results from a diffusionless transformation of austenite. It may be thought
of as a transformation product that is competitive with pearlite and bainite. The
martensitic transformation occurs when the quenching rate is rapid enough to prevent
carbon diffusion. Any diffusion whatsoever will result in the formation of ferrite
and cementite phases.
The martensitic transformation is not well understood. However, large numbers of atoms experience cooperative movements, in that there is only a slight displacement of each atom relative to its neighbors. This occurs in such a way that
the FCC austenite experiences a polymorphic transformation to a body-centered
tetragonal (BCT) martensite. A unit cell of this crystal structure (Figure 10.20) is
simply a body-centered cube that has been elongated along one of its dimensions;
this structure is distinctly different from that for BCC ferrite. All the carbon atoms
remain as interstitial impurities in martensite; as such, they constitute a supersaturated solid solution that is capable of rapidly transforming to other structures if
heated to temperatures at which diffusion rates become appreciable. Many steels,
however, retain their martensitic structure almost indefinitely at room temperature.
The martensitic transformation is not, however, unique to iron–carbon alloys.
It is found in other systems and is characterized, in part, by the diffusionless transformation.
Since the martensitic transformation does not involve diffusion, it occurs almost
instantaneously; the martensite grains nucleate and grow at a very rapid rate—the
velocity of sound within the austenite matrix. Thus the martensitic transformation
rate, for all practical purposes, is time independent.
Martensite grains take on a plate-like or needle-like appearance, as indicated
in Figure 10.21. The white phase in the micrograph is austenite (retained austenite)
that did not transform during the rapid quench. As already mentioned, martensite
as well as other microconstituents (e.g., pearlite) can coexist.
Being a nonequilibrium phase, martensite does not appear on the iron–iron carbide phase diagram (Figure 9.24).The austenite-to-martensite transformation is, however, represented on the isothermal transformation diagram. Since the martensitic
transformation is diffusionless and instantaneous, it is not depicted in this diagram
as the pearlitic and bainitic reactions are. The beginning of this transformation is
represented by a horizontal line designated M(start) (Figure 10.22). Two other horizontal and dashed lines, labeled M(50%) and M(90%), indicate percentages of the
austenite-to-martensite transformation. The temperatures at which these lines are
Figure 10.20 The body-centered tetragonal unit cell for
martensitic steel showing iron atoms (circles) and sites that
may be occupied by carbon atoms (crosses). For this
tetragonal unit cell, c 7 a.
c
a
a
1496T_c10_311-357 11/30/05 7:37 Page 332
REVISED PAGES
332 • Chapter 10 / Phase Transformations in Metals
Figure 10.21 Photomicrograph showing the
martensitic microstructure. The needleshaped grains are the martensite phase, and
the white regions are austenite that failed to
transform during the rapid quench. 1220.
(Photomicrograph courtesy of United States
Steel Corporation.)
athermal
transformation
plain carbon steel
alloy steel
located vary with alloy composition but, nevertheless, must be relatively low because
carbon diffusion must be virtually nonexistent.4 The horizontal and linear character
of these lines indicates that the martensitic transformation is independent of time;
it is a function only of the temperature to which the alloy is quenched or rapidly
cooled. A transformation of this type is termed an athermal transformation.
Consider an alloy of eutectoid composition that is very rapidly cooled from a
temperature above 727C (1341F) to, say, 165C (330F). From the isothermal
transformation diagram (Figure 10.22) it may be noted that 50% of the austenite
will immediately transform to martensite; and as long as this temperature is maintained, there will be no further transformation.
The presence of alloying elements other than carbon (e.g., Cr, Ni, Mo, and W)
may cause significant changes in the positions and shapes of the curves in the isothermal transformation diagrams. These include (1) shifting to longer times the nose of
the austenite-to-pearlite transformation (and also a proeutectoid phase nose, if such
exists), and (2) the formation of a separate bainite nose. These alterations may be
observed by comparing Figures 10.22 and 10.23, which are isothermal transformation diagrams for carbon and alloy steels, respectively.
Steels in which carbon is the prime alloying element are termed plain carbon
steels, whereas alloy steels contain appreciable concentrations of other elements,
including those cited in the preceding paragraph. Section 11.2 tells more about the
classification and properties of ferrous alloys.
Concept Check 10.2
Cite two major differences between martensitic and pearlitic transformations.
[The answer may be found at www.wiley.com/college/callister (Student Companion Site).]
4
The alloy that is the subject of Figure 10.21 is not an iron-carbon alloy of eutectoid composition; furthermore, its 100% martensite transformation temperature lies below the
ambient. Since the photomicrograph was taken at room temperature, some austenite
(i.e., the retained austenite) is present, having not transformed to martensite.
1496T_c10_311-357 11/30/05 7:37 Page 333
REVISED PAGES
10.5 Isothermal Transformation Diagrams • 333
Figure 10.22 The
complete isothermal
transformation diagram
for an iron–carbon
alloy of eutectoid
composition: A,
austenite; B, bainite; M,
martensite; P, pearlite.
800
A
1400
Eutectoid temperature
700
A
1200
A
+
600
P
P
B
800
A
400
+
B
A
300
Temperature (°F)
Temperature (°C)
1000
500
600
M(start)
200
50%
M+A
M(50%)
400
M(90%)
100
200
0
10–1
1
102
10
103
104
105
Time (s)
800
A
1400
Eutectoid temperature
700
A+F
F+P
1200
A+F
+P
A
600
1000
800
400
A+B
50%
B
600
M(start)
300
M+A
M(50%)
M(90%)
400
200
M
100
200
0
1
10
102
103
Time (s)
104
105
106
Temperature (°F)
500
Temperature (°C)
Figure 10.23
Isothermal
transformation
diagram for an alloy
steel (type 4340): A,
austenite; B, bainite;
P, pearlite; M,
martensite; F,
proeutectoid ferrite.
[Adapted from H.
Boyer (Editor), Atlas
of Isothermal
Transformation
and Cooling
Transformation
Diagrams, American
Society for Metals,
1977, p. 181.]
1496T_c10_311-357 11/30/05 7:37 Page 334
REVISED PAGES
334 • Chapter 10 / Phase Transformations in Metals
EXAMPLE PROBLEM 10.2
Microstructural Determinations for Three
Isothermal Heat Treatments
Using the isothermal transformation diagram for an iron–carbon alloy of eutectoid composition (Figure 10.22), specify the nature of the final microstructure (in terms of microconstituents present and approximate percentages) of
a small specimen that has been subjected to the following time–temperature
treatments. In each case assume that the specimen begins at 760C (1400F)
and that it has been held at this temperature long enough to have achieved a
complete and homogeneous austenitic structure.
(a) Rapidly cool to 350C (660F), hold for 104 s, and quench to room temperature.
(b) Rapidly cool to 250C (480F), hold for 100 s, and quench to room temperature.
(c) Rapidly cool to 650C (1200F), hold for 20 s, rapidly cool to 400C (750F),
hold for 103 s, and quench to room temperature.
Solution
The time–temperature paths for all three treatments are shown in Figure 10.24.
In each case the initial cooling is rapid enough to prevent any transformation
from occurring.
(a) At 350C austenite isothermally transforms to bainite; this reaction
begins after about 10 s and reaches completion at about 500 s elapsed
time. Therefore, by 104 s, as stipulated in this problem, 100% of the
specimen is bainite, and no further transformation is possible, even though
the final quenching line passes through the martensite region of the
diagram.
(b) In this case it takes about 150 s at 250C for the bainite transformation
to begin, so that at 100 s the specimen is still 100% austenite. As the specimen is cooled through the martensite region, beginning at about 215C, progressively more of the austenite instantaneously transforms to martensite. This
transformation is complete by the time room temperature is reached, such
that the final microstructure is 100% martensite.
(c) For the isothermal line at 650C, pearlite begins to form after about 7 s;
by the time 20 s has elapsed, only approximately 50% of the specimen has
transformed to pearlite. The rapid cool to 400C is indicated by the vertical
line; during this cooling, very little, if any, remaining