Deformation of Earth Materials An Introduction To The Rheology of Solid Earth Shunichiro Karato Instant Download
Deformation of Earth Materials An Introduction To The Rheology of Solid Earth Shunichiro Karato Instant Download
https://ebookbell.com/product/deformation-of-earth-materials-an-
introduction-to-the-rheology-of-solid-earth-shunichiro-
karato-4408448
https://ebookbell.com/product/earth-dynamics-deformations-and-
oscillations-of-the-rotating-earth-smylie-de-4584288
https://ebookbell.com/product/deformation-of-the-continental-crust-
the-legacy-of-mike-coward-geological-society-special-publication-
no-272-a-c-ries-r-w-h-butler-2142134
https://ebookbell.com/product/testing-of-the-plastic-deformation-of-
metals-t-w-clyne-j-e-campbell-49138750
https://ebookbell.com/product/plastic-deformation-of-minerals-and-
rocks-shunichiro-karato-50924138
Plastic Deformation Of Minerals And Rocks Shunichir Karato
https://ebookbell.com/product/plastic-deformation-of-minerals-and-
rocks-shunichir-karato-2347650
https://ebookbell.com/product/constrained-deformation-of-materials-
devices-heterogeneous-structures-and-thermomechanical-modeling-1st-
edition-yl-shen-auth-2530346
https://ebookbell.com/product/continuum-deformation-of-multiagent-
systems-1st-edition-hossein-rastgoftar-auth-5696284
https://ebookbell.com/product/plastic-deformation-of-nanocrystalline-
materials-1st-edition-glezer-6750314
https://ebookbell.com/product/the-physics-of-the-deformation-of-
densely-packed-granular-materials-maarten-anton-koenders-34725598
This page intentionally left blank
Deformation of Earth Materials
Shun-ichiro Karato
Yale University, Department of Geology &
Geophysics, New Haven, CT, USA
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, São Paulo
© S. Karato 2008
Cambridge University Press has no responsibility for the persistence or accuracy of urls
for external or third-party internet websites referred to in this publication, and does not
guarantee that any content on such websites is, or will remain, accurate or appropriate.
Contents
Preface page ix
2 Thermodynamics 13
2.1 Thermodynamics of reversible processes 13
2.2 Some comments on the thermodynamics of a stressed system 28
2.3 Thermodynamics of irreversible processes 29
2.4 Thermally activated processes 32
4 Elasticity 51
4.1 Introduction 51
4.2 Elastic constants 52
4.3 Isothermal versus adiabatic elastic constants 55
4.4 Experimental techniques 57
4.5 Some general trends in elasticity: Birch’s law 59
4.6 Effects of chemical composition 67
4.7 Elastic constants in several crystal structures 70
4.8 Effects of phase transformations 72
5 Crystalline defects 75
5.1 Defects and plastic deformation: general introduction 75
5.2 Point defects 76
5.3 Dislocations 82
5.4 Grain boundaries 94
References 412
Materials index 452
Subject index 454
Understanding the microscopic physics of deformation materials science of deformation of minerals and
is critical in many branches of solid Earth science. rocks over various time-scales are described in addition
Long-term geological processes such as plate tectonics to the applications of these results to important geo-
and mantle convection involve plastic deformation of logical and geophysical problems. Properties of materi-
Earth materials, and hence understanding the plastic als discussed include elastic, anelastic (viscoelastic),
properties of Earth materials is key to the study of and plastic properties. The emphasis is on an interdis-
these geological processes. Interpretation of seismolog- ciplinary approach, and, consequently, I have included
ical observations such as tomographic images or seis- discussions on some advanced, controversial issues
mic anisotropy requires knowledge of elastic, anelastic where they are highly relevant to Earth science prob-
properties of Earth materials and the processes of plas- lems. They include the role of hydrogen, effects of
tic deformation that cause anisotropic structures. pressure, deformation of two-phase materials, local-
Therefore there is an obvious need for understanding ization of deformation and the link between viscoelas-
a range of deformation-related properties of Earth tic deformation and plastic flow. This book is intended
materials in solid Earth science. However, learning to serve as a textbook for a course at a graduate level in
about deformation-related properties is challenging an Earth science program, but it may also be useful for
because deformation in various geological processes students in materials science as well as researchers
involves a variety of microscopic processes. Owing to in both areas. No previous knowledge of geology/
the presence of multiple deformation mechanisms, geophysics or of materials science is assumed. The
the results obtained under some conditions may not basics of continuum mechanics and thermodynamics
necessarily be applicable to a geological problem that are presented as far as they are relevant to the main
involves deformation under different conditions. There- topics of this book.
fore in order to conduct experimental or theoretical Significant progress has occurred in the study of
research on deformation, one needs to have a broad deformation of Earth materials during the last 30
knowledge of various mechanisms to define conditions years, mainly through experimental studies. Experi-
under which a study is to be conducted. Similarly, mental studies on synthetic samples under well-defined
when one attempts to use results of experimental or chemical conditions and the theoretical interpretation
theoretical studies to understand a geological problem, of these results have played an important role in under-
one needs to evaluate the validity of applying partic- standing the microscopic mechanisms of deformation.
ular results to a given geological problem. However, Important progress has also been made to expand
there was no single book available in which a broad the pressure range over which plastic deformation can
range of the physics of deformation of materials was be investigated, and the first low-strain anelasticity
treated in a systematic manner that would be useful for measurements have been conducted. In addition,
a student (or a scientist) in solid Earth science. The some large-strain deformation experiments have been
motivation of writing this book was to fulfill this need. performed that have provided important new insights
In this book, I have attempted to provide a unified, into the microstructural evolution during deformation.
interdisciplinary treatment of the science of deforma- However, experimental data are always obtained under
tion of Earth with an emphasis on the materials limited conditions and their applications to the Earth
science (microscopic) approach. Fundamentals of the involve large extrapolation. It is critical to understand ix
x Preface
the scaling laws based on the physics and chemistry of and a geodynamic interpretation of anomalies in seis-
deformation of materials in order to properly apply mic wave propagation. Some of the representative
experimental data to Earth. A number of examples of experimental data are summarized in tables.
such scaling laws are discussed in this book. However, the emphasis of this book is on presenting
This book consists of three parts: Part I basic theoretical concepts and consequently references
(Chapters 1–3) provides a general background includ- to the data are not exhaustive. Many problems (with
ing basic continuum mechanics, thermodynamics and solutions) are provided to make sure a reader under-
phenomenological theory of deformation. Most of this stands the content of this book. Some of them are
part, particularly Chapters 1 and 2 contain material advanced and these are shown by an asterisk.
that can be found in many other textbooks. Therefore The content of this book is largely based on lectures
those who are familiar with basic continuum mechan- that I have given at the University of Minnesota and
ics and thermodynamics can skip this part. Part II Yale University as well as at other institutions. I thank
(Chapters 4–16) presents a detailed account of materi- students and my colleagues at these institutions who
als science of time-dependent deformation, including have given me opportunities to improve my under-
elastic, anelastic and plastic deformation with an standing of the subjects discussed in this book through
emphasis on anelastic and plastic deformation. They inspiring questions. Some parts of this book have
include, not only the basics of properties of materials been read/reviewed by A. S. Argon, D. Bercovici,
characterizing deformation (i.e., elasticity and viscos- H. W. Green, S. Hier-Majumder, G. Hirth, I. Jackson,
ity (creep strength)), but also the physical princi- D. L. Kohlstedt, J. Korenaga, R. C. Liebermann,
ples controlling the microstructural developments J.-P. Montagner, M. Nakada, C. J. Spiers, J. A. Tullis
(grain size and lattice-preferred orientation). Part III and J. A. Van Orman. However, they do not always
(Chapters 17–21) provides some applications of the agree with the ideas presented in this book and any
materials science of deformation to important geolog- mistakes are obviously my own. W. Landuyt, Z. Jiang
ical and geophysical problems, including the rheolog- and P. Skemer helped to prepare the figures. I should
ical structure of solid Earth and the interpretation of also thank the editors at Cambridge University Press
the pattern of material circulation in the mantle and for their patience. Last but not least, I thank my family,
core from geophysical observations. Specific topics particularly my wife, Yoko, for her understanding, for-
covered include the lithosphere–asthenosphere struc- bearance and support during the long gestation of this
ture, rheological stratification of Earth’s deep mantle monograph. Thank you all.
Part I
General background
1 Stress and strain
The concept of stress and strain is key to the understanding of deformation. When a force is applied to
a continuum medium, stress is developed inside it. Stress is the force per unit area acting on a given
plane along a certain direction. For a given applied force, the stress developed in a material depends
on the orientation of the plane considered. Stress can be decomposed into hydrostatic stress (pressure)
and deviatoric stress. Plastic deformation (in non-porous materials) occurs due to deviatoric stress.
Deformation is characterized by the deformation gradient tensor, which can be decomposed into
rigid body rotation and strain. Deformation such as simple shear involves both strain and rigid body
rotation and hence is referred to as rotational deformation whereas pure shear or tri-axial compression
involves only strain and has no rigid body rotation and hence is referred to as irrotational deformation.
In rotational deformation, the principal axes of strain rotate with respect to those of stress whereas
they remain parallel in irrotational deformation. Strain can be decomposed into dilatational
(volumetric) strain and shear strain. Plastic deformation (in a non-porous material) causes shear strain
and not dilatational strain. Both stress and strain are second-rank tensors, and can be characterized by
the orientation of the principal axes and the magnitude of the principal stress and strain and both have
three invariants that do not depend on the coordinate system chosen.
Key words stress, strain, deformation gradient, vorticity, principal strain, principal stress, invariants
of stress, invariants of strain, normal stress, shear stress, Mohr’s circle, the Flinn diagram, foliation,
lineation, coaxial deformation, non-coaxial deformation.
x3 X
3
x~i ¼ aij xj (1:4)
j¼1
T1
T
where aij is the transformation matrix that satisfies the
orthonormality relation,
T2 x2
X
3
aij ajm ¼ im (1:5)
T3 j¼1
x1
where im is the Kronecker delta (im ¼ 1 for i ¼ m,
FIGURE 1.1 Forces acting on a small pyramid.
im ¼ 0 otherwise). Now in this new coordinate system,
are the force exerted by atom A (B) to B (A). we may write a relation similar to equation (1.2) as,
Consequently these forces caused by atomic displace- X
3
ment within a body must cancel. The long-range force T~i ¼ ~ij n~j : (1:6)
is called a body force, but if one takes this region as j¼1
small, then the magnitude of this body force will Noting that the traction (T) transforms as a vector in
become negligible compared to the surface force (i.e., the same way as the coordinate system, equation (1.4),
the third class of force above). Therefore the net force we have,
acting on the small region must be the forces across
the surface of that region from the neighboring mate- X
3
T~i ¼ aij Tj : (1:7)
rials. To characterize this force, let us consider a small j¼1
piece of block that contains a plane with the area of dS
and whose normal is n (n is the unit vector). Let T be Inserting equation (1.2), the relation (1.7) becomes,
the force (per unit area) acting on the surface dS from
X
3
outside this block (positive when the force is compres- T~i ¼ jk aij nk : (1:8)
sive) and consider the force balance (Fig. 1.1). The j;k¼1
force balance should be attained among the force T
Now using the orthonormality relation (1.5), one has,
as well as the forces T1,2,3 that act on the surface
dS1,2,3 respectively (dS1,2,3 are the projected area of X
3
dS on the plane normal to the x1,2,3 axis). Then the ni ¼ aji n~i : (1:9)
j¼1
force balance relation for the block yields,
X
3 Inserting this relation into equation (1.8) and compar-
T dS ¼ T j dSj : (1:1) ing the result with equation (1.6), one obtains,1
j¼1
X
3
Now using the relation dSj ¼ nj dS, one obtains, ~ij ¼ kl aik ajl : (1:10)
k;l¼1
X
3 X
3
Ti ¼ T ij nj ¼ ij nj (1:2)
j¼1 j¼1
The quantity that follows this transformation law is
referred to as a second rank tensor.
where Ti is the ith component of the force T and ij is
the ith component of the traction Tj, namely the ith
component of force acting on a plane whose normal is 1.1.2. Principal stress, stress invariants
the jth direction ðnij ¼ T ij Þ. This is the definition of
stress. From the balance of torque, one can also show, In any material, there must be a certain orientation of a
plane on which the direction of traction (T) is normal
ij ¼ ji : (1:3) to it. For that direction of n, one can write,
The values of stress thus defined depend on the Ti ¼ ni (1:11)
coordinate system chosen. Let us denote quantities in
a new coordinate system by a tilda, then the new coor-
dinate and the old coordinate system are related to
each other by, 1
In the matrix notation, ~ ¼ A AT where A ¼ aij and AT ¼ aji .
Stress and strain 5
X
3
σ2
ðij ij Þnj ¼ 0: (1:12)
j¼1 σn
x1
For this equation to have a non-trivial solution other σ3 θ σ3
than n ¼ 0, one must have, τ
x3
x2
ij ij ¼ 0 (1:13)
σ2
where Xij is the determinant of a matrix Xij. Writing
equation (1.13) explicitly, one obtains,
11
12 13 σ1
21 22 23 ¼ 3 þ I 2 þ II þ III ¼ 0
31 32 33 FIGURE 1.2 Geometry of normal and shear stress on a plane.
(1:14)
1.1.3. Normal stress, shear stress,
with
Mohr’s circle
I ¼ 11 þ 22 þ 33 (1:15a)
Now let us consider the normal and shear stress on a
given plane subjected to an external force (Fig. 1.2).
II ¼ 11 22 11 33 33 22 þ 212 þ 213 þ 223
Let x1 be the axis parallel to the maximum compres-
(1:15b) sional stress 1 and x2 and x3 be the axes perpendicular
to x1. Consider a plane whose normal is at the angle
III ¼ 11 22 33 þ 212 23 31 11 223 from x3 (positive counterclockwise). Now, we define a
(1:15c)
22 213 33 212 : new coordinate system whose x01 axis is normal to the
plane, but the x02 axis is the same as the x2 axis. Then
Therefore, there are three solutions to equation (1.14), the transformation matrix is,
1 ; 2 ; 3 ð1 42 43 Þ.These are referred to as the
2 3
principal stresses. The corresponding n is the orienta- cos 0 sin
tion of principal stress. If the stress tensor is written ½aij ¼ 4 0 1 0 5 (1:18)
using the coordinate whose orientation coincides with sin 0 cos
the orientation of principal stress, then,
2 3 and hence,
1 0 0
2 3
½ij ¼ 4 0 2 0 5: (1:16) 1 þ 3 1 3
þ cos 2 0
1 3
sin 2
0 0 3 6 2 2 2 7
6 7
6 7
½~
ij ¼ 6 0 2 0 7:
6 7
It is also seen that because equation (1.14) is a scalar 4 1 3 1 þ 3 1 3 5
sin 2 0 cos 2
equation, the values of I, II and III are independ- 2 2 2
ent of the coordinate. These quantities are called the (1:19)
invariants of stress tensor. These quantities play
important roles in the formal theory of plasticity (see
Section 3.3). Equations (1.15a–c) can also be written Problem 1.1
in terms of the principal stress as,
Derive equation (1.19).
I ¼ 1 þ 2 þ 3 (1:17a)
Solution
II ¼ 1 2 2 3 3 1 (1:17b)
2 32 32 3
cos 0 sin 1 0 0 cos 0 sin A = ( 0 , σ1)
6 76 76 7
6 76 76 7 B = ( 0 , σ3)
½~
ij ¼ 6 0 1 0 76 0 2 0 76 0 1 0 7
4 54 54 5 τ C = ( 0 , (σ1 + σ3) / 2 )
sin 0 cos 0 0 3 sin 0 cos R = (σ1 − σ3) / 2
2 1 þ 3 1 3 1 3 3
þ cos 2 0 sin 2
6 2 2 2 7 R
6 7
6 7
¼6 0 2 0 7:
6 7 σn
4 5 A C B
1 3 1 þ 3 1 3
sin 2 0 cos 2
2 2 2
Therefore the shear stress and normal stress n on this FIGURE 1.3 A Mohr circle corresponding to two-dimensional stress
plane are showing the variation of normal, n , and shear stress, , on a plane.
1 3
~13 ¼ sin 2 (1:20)
2
and Problem 1.2
1 þ 3 1 3
~33 n ¼ cos 2 (1:21) Show that the second invariant of deviatoric stress
2 2 1h
can be written as II0 ¼ ð1 2 Þ2 þ ð2 3 Þ2 þ
respectively. It follows that the maximum shear stress i 6
is on the two conjugate planes that are inclined by ð3 1 Þ2 :
p=4 with respect to the x1 axis and its absolute mag-
nitude is ð1 3 Þ=2. Similarly, the maximum com-
Solution
pressional stress is on a plane that is normal to the x1
axis and its value is 1. It is customary to use 1 3 as If one uses a coordinate system parallel to the
(differential (or deviatoric)) stress in rock deformation principal axes of stress, from equation (1.15), one
literature, but the shear stress, ð1 3 Þ=2, is also has II0 ¼ 01 02 01 03 03 02 . Using I0 ¼ 01 þ 02 þ
often used. Eliminating from equations (1.20) and
03 ¼ 0; one finds I2 ¼ 02 02 02
1 þ 2 þ 3 þ 2ð1 2 þ
0 0
(1.21), one has,
2 3 þ 3 1 Þ ¼ 0. Therefore II0 ¼ 2 ð1 þ 2 þ02
0 0 0 0 1 02 02
3 Þ:
1 þ 3 2 1
2 þ n ¼ ð1 3 Þ2 : (1:22) Now, inserting 01 ¼ 1 13 ð1 þ 2 þ 3 Þ etc., one
2 4 h i
obtains II0 ¼ 16 ð1 2 Þ2 þ ð2 3 Þ2 þð3 1 Þ2 :
Thus, the normal and shear stress on planes with var-
ious orientations can be visualized on a two-dimensional
plane (–n space) as a circle whose center is located Problem 1.3
at ð0; ð1 þ 3 Þ=2Þ and the radius ð1 3 Þ=2
(Fig. 1.3). This is called a Mohr’s circle and plays an Show that when the stress has axial symmetry with
important role in studying the brittle fracture that is respect to the x1 axis (i.e., 2 ¼ 3 ), then n ¼ Pþ
controlled by the stress state (shear–normal stress ratio; ð1 3 Þðcos2 13Þ.
see Section 7.3).
When 1 ¼ 2 ¼ 3 ð¼ PÞ, then the stress is isotro- Solution
pic (hydrostatic). The hydrostatic component of stress
does not cause plastic flow (this is not true for porous From (1.21), one obtains, n ¼ ð1 þ 3 Þ=2 þ
materials, but we do not discuss porous materials ðð1 3 Þ=2Þ cos 2. Now cos 2 ¼ 2 cos2 1 and
here), so it is useful to define deviatoric stress P ¼ 13 ð1 þ 2 þ 3 Þ ¼ 13 ð1 þ 23 Þ ¼ 1 23 ð1 3 Þ.
Therefore n ¼ P þ ð1 3 Þðcos2 13Þ.
0ij ij ij P: (1:23)
hence 2
Note that in some literature, another definition of shear strain is used in
which "ij ¼ @ ui =@xj þ @ uj =@xi for i 6¼ j and "ii ¼ @ ui =@xi ; e.g., Hobbs
@ ui @Xi
¼ ij : (1:30) et al. (1976). In such a case, the symbol ij is often used for the non-
@ xj @xj diagonal (i 6¼ j) strain component instead of "ij .
8 Deformation of Earth Materials
Let us first consider the physical meaning of the Obviously, normal strain can be present in defor-
@u
second part, 12ð@@ xuji @ xji Þ. The second part is an anti- mation without a volume change. For example,
0 1
symmetric tensor, namely, " 0 0
@
"ij ¼ 0 2 " 1
0 A represents an elongation
1 @ ui @ uj
oij ¼ ¼ oji ðoii ¼ 0Þ: (1:35) 0 0 12 "
2 @ xj @ xi
along the 1-axis and contraction along the 2 and
The displacement of a small vector duj due to the 3 axes without volume change.
operation of this matrix is given by, Now let us consider the off-diagonal components
X
3 of strain tensor. From equation (1.39), it is clear that
uo
d~i ¼ oij duj : (1:36) when all the diagonal components are zero, then all the
j¼1
displacement vectors must be normal to the direction
Since oii ¼ 0, the displacement occurs only to the direc- of the initial vector. Therefore, there is no change in
tions that are normal to the initial orientation. Therefore length due to the off-diagonal component of strain.
the operation of this matrix causes the rotation of mate- Note, also, that since strain is a symmetric tensor,
rial points with the axis that is normal to both ith and jth "ij ¼ "ji , the directions of rotation of two orthogonal
directions with the magnitude (positive clockwise), axes are toward the opposite direction with the same
uo
d~ magnitude (Fig. 1.5). Consequently, the angle of two
i
tan ij ¼ ¼ oji ¼ oij : (1:37) orthogonal axes change from p=2 to (see Problem 1.4),
dui
p
(Again this rotation tensor is defined using the defor- tan1 2"ij : (1:43)
med state. So it is referred to as the Eulerian rotation 2
tensor.) To represent this, a rotation vector is often Therefore, the off-diagonal components of strain ten-
used that is defined as, sor (i.e., "ij with i 6¼ j) represent the shape change with-
wð¼ ðo1 ; o2 ; o3 ÞÞ ðo23 ; o31 ; o12 Þ: (1:38) out volume change, namely shear strain.
X
3
u"i ¼
d~ "ij duj : (1:39) Solution
j¼1
Let the small angle of rotation of the i axis to the j axis
From equation (1.39), it follows that the length of a
due to the operation of strain tensor be ij (positive
component of vector u0i changes to,
clockwise), then (Fig. 1.5),
u~i ¼ ð1 þ "ii Þu0i : (1:40)
du~j
Therefore the diagonal component of strain tensor tan ij ¼ ij ¼ ð"ji þ oji Þ ¼ "ij þ oij :
dui
represents the change in length, so that this component
of strain, "ii , is called normal strain. Consequently, Similarly, if the rotation of the j axis relative to the i
axis is ji , one obtains,
V
¼ ð1 þ "11 Þð1 þ "22 Þð1 þ "33 Þ 1 þ "11 þ "22 þ "33
V0 du~i
tan ji ¼ ji ¼ ð"ij þ oij Þ ¼ "ij oij :
(1:41) duj
where V0 is initial volume and V is the final volume and (Note that the rigid-body rotations of the two axes are
the strain is assumed to be small (this assumption can opposite with the same magnitude.) Therefore, the net
be relaxed and the same argument can be applied to a change in the angle between i and j axes is given by
finite strain, see e.g., M ASE (1970)). Thus,
4ij ¼ ij þ ji ¼ 2"ij tan 4ij :
X
3
4V
"kk ¼ : (1:42) Hence 4ij ¼ tan1 2"ij .
k¼1
V
Stress and strain 9
x2′ "11 " "12 "13
x2 "21 "22 " "32 ¼ "3 þ I" "2 þ II" " þ III"
"31 "32 "33 "
¼0
(1:48)
with
x1′ II" ¼ "11 "22 "11 "33 "33 "22 þ "212 þ "213 þ "223
(1:49b)
x1 III" ¼ "11 "22 "33 þ 2"12 "23 "31 "11 "223 "22 "213
FIGURE 1.5 Geometry of shear deformation. "33 "212 : (1:49c)
1.2.3. Principal strain, strain ellipsoid Therefore, there are three solutions of equation (1.48),
"1 ; "2 ; "3 ð"1 4"2 4"3 Þ. These are referred to as the prin-
We have seen two different cases for strain, one in which cipal strain. The corresponding u0 are the orientations
the displacement caused by the strain tensor is normal to of principal strain. If the strain tensor is written using
the original direction of the material line and another the coordinate whose orientation coincides with the
where the displacement is normal to the original mate- orientation of principal strain, then,
rial line. In this section, we will learn that in any material 2 3
and in any geometry of strain, there are three directions "1 0 0
½"ij ¼ 0 "2 0 5:
4 (1:50)
along which the displacement is normal to the direction
0 0 "3
of original line segment. These are referred to as the
orientation of principal strain, and the magnitude of A strain ellipsoid is a useful way to visualize the
strain along these orientations are called principal strain. geometry of strain. Let us consider a spherical body
One can define the principal strains ð"1 ; "2 ; "3 ; in a space and deform it. The shape of a sphere is
"1 4"2 4"3 Þ in the following way. Recall that the nor- described by,
mal displacement along the direction i, u~i , along the
vector u is given by, ðu1 Þ2 þ ðu2 Þ2 þ ðu3 Þ2 ¼ 1: (1:51)
X
3 The shape of the sphere will change due to deforma-
u~i ¼ "ij uj : (1:44) tion. Let us choose a coordinate system such that the
j¼1
directions of 1, 2 and 3 axes coincide with the directions
Now, let u be the direction in space along which the of principal strain. Then the length of each axis of the
displacement is parallel to the direction u. Then, original sphere along each direction of the coordinate
system should change to u~i ¼ ð1 þ "ii Þui , and therefore
u~i ¼ "ui (1:45)
the sphere will change to an ellipsoid,
where " is a scalar quantity to be determined. From
u 1 Þ2
ð~ u2 Þ2
ð~ u3 Þ2
ð~
equations (1.44) and (1.45), 2
þ 2
þ ¼ 1: (1:52)
ð1 þ "1 Þ ð1 þ "2 Þ ð1 þ "3 Þ2
X
3
ð"ij "ij Þuj ¼ 0: (1:46) A three-dimensional ellipsoid defined by this equa-
j¼1
tion is called a strain ellipsoid. For example, if the
For this equation to have a non-trivial solution other shape of grains is initially spherical, then the shape of
than u ¼ 0, one must have, grains after deformation represents the strain ellip-
soid. The strain of a rock specimen can be deter-
j"ij "ij j ¼ 0 (1:47)
mined by the measurements of the shape of grains
where Xij is the determinant of a matrix Xij. Writing or some objects whose initial shape is inferred to be
equation (1.47) explicitly, one gets, nearly spherical.
10 Deformation of Earth Materials
k=∞
Problem 1.5*
a = (ε1 + 1) / (ε2 + 1)
Calculate the strain ellipsoid, and find how the
principal axes of the strain ellipsoid rotate with strain.
Also find the relation between the angle of tilt of the
initially vertical line and the angle of the maximum
elongation direction relative to the horizontal axis.
Solution
1 k=0
For simplicity, let us analyze the geometry in the x–y plane
1 b = (ε2 + 1) / (ε3 + 1)
(normal to the shear plane) where shear occurs. Consider
a circle defined by x2 þ y2 ¼ 1: By deformation, this FIGURE 1.6 The Flinn diagram (after H OBBS et al., 1976).
circle changes to an ellipsoid, ðx þ yÞ2 þ y2 ¼ 1, i.e.,
as ! 1 and "2 ¼ A1=2
yy 1 changes from 0 at ¼ 0
2 2 2 to –1 at ! 1.
x þ 2xy þ ð þ 1Þy ¼ 1: (1)
Combined with the relation ð"1 þ 1Þ=ð"2 þ 1Þ ¼ other possible causes for lineation including the pref-
ð"2 þ 1Þ=ð"3 þ 1Þ, we obtain ð"2 þ 1Þ3 ¼ 1 and hence erential growth of minerals (e.g., H OBBS et al., 1976).
"2 ¼ 0. Therefore deformation is plane strain. Consequently, the interpretation of the significance
of these reference frames (foliation/lineation) in natu-
ral rocks is not always unique. In particular, the ques-
tion of growth origin versus deformation origin, and
1.2.5. Foliation, lineation (Fig. 1.7)
the strain ellipsoid versus the shear plane/shear direc-
When the anisotropic microstructure of a rock is tion can be elusive in some cases. Interpretation and
studied, it is critical to define the reference frame of identification of foliation/lineation become more diffi-
the coordinate. Once one identifies a plane of reference cult if the deformation geometry is not constant with
and the reference direction on that plane, then the three time. Consequently, it is important to state clearly how
orthogonal axes (parallel to lineation (X direction), one defines foliation/lineation in the structural analysis
normal to lineation on the foliation plane (Y direction), of a deformed rock. For more details on foliation and
normal to foliation (Z direction)) define the reference lineation, a reader is referred to a structural geology
frame. textbook such as H OBBS et al. (1976).
Foliation is usually used to define a reference plane
and lineation is used define a reference direction on the 1.2.6. Various deformation geometries
foliation plane. Foliation is a planar feature in a given
rock, but its origin can be various (H OBBS et al., 1976). The geometry of strain is completely characterized by the
The foliation plane may be defined by a plane normal principal strain, and therefore a diagram such as the Flinn
to the maximum shortening strain (Fig. 1.7). Foliation diagram (Fig. 1.6) can be used to define strain. However,
can also be caused by compositional layering, grain-size in order to characterize the geometry of deformation
variation and the orientation of platy minerals such as completely, it is necessary to characterize the deformation
mica. When deformation is heterogeneous, such as the gradient tensor ðdij ð¼ "ij þ oij ÞÞ. Therefore the rota-
case for S-C mylonite (L ISTER and S NOKE , 1984), one tional component (vorticity tensor), oij, must also be
can identify two planar structures, one corresponds to characterized. In this connection, it is important to dis-
the strain ellipsoid (a plane normal to maximum short- tinguish between irrotational and rotational deformation
ening, "3 ) and another to the shear plane. geometry. Rotational deformation geometry refers to
Lineation is a linear feature that occurs repetitively deformation in which oij ¼ 6 0, and irrotational deforma-
in a rock. In most cases, the lineation is found on the tion geometry corresponds to oij ¼ 0. The distinction
foliation plane, although there are some exceptions. between them is important at finite strain. To illustrate
The most common is mineral lineation, which is defined this point, let us consider two-dimensional deformation
by the alignment of non-spherical minerals such as (Fig. 1.8). For irrotational deformation, the orientations
clay minerals. The alignment of spinel grains in a spinel of the principal axes of strain are always parallel to those
lherzolite and recrystallized orthopyroxene in a garnet of principal stress. Therefore such a deformation is called
lherzolite are often used to define the lineation in peri- coaxial deformation. In contrast, when deformation is
dotites. One cause of lineation is strain, and in this case, rotational, such as simple shear, the orientations of
the direction of lineation is parallel to the maximum principal axes of strain rotate progressively with respect
elongation direction. However, there are a number of to those of the stress (see Problem 1.5). This type of
12 Deformation of Earth Materials
The nature of the deformation of materials depends on the physical and chemical state of the materials.
Thermodynamics provides a rigorous way by which the physical and chemical state of materials can
be characterized. A brief account is made of the concepts of thermodynamics of reversible as well
as irreversible processes that are needed to understand the plastic deformation of materials and
related processes. The principles governing the chemical equilibrium are outlined including the
concept of chemical potential, the law of mass action, and the Clapeyron slope (i.e., the slope of a
phase boundary in the pressure-temperature space). When a system is out of equilibrium, a flow of
materials and/or energy occurs. The principles governing the irreversible processes are outlined.
Irreversible processes often occur through thermally activated processes. The basic concepts of
thermally activated processes are summarized based on the statistical physics.
Key words entropy, chemical potential, Gibbs free energy, fugacity, activity, Clapeyron slope,
phase diagrams, rate theory, generalized force, the Onsager reciprocal relation.
2.1. Thermodynamics of reversible 2.1.1. The first and the second principles
processes of thermodynamics
Thermodynamics provides a framework by which the The first principle of thermodynamics is the law of conser-
nature of thermochemical equilibrium is defined, and, vation of energy, which states that the change in the inter-
in cases where a system is out of equilibrium, it defines nal energy, dE, is the sum of the mechanical work done to
the direction to which a given material will change. It the system, the change in the energy due to the addition of
gives a basis for analyzing the composition and struc- materials and the heat added to the system, namely,
ture of geological materials, experimental data and the
dE ¼ W þ Z þ Q (2:1)
way in which the experimental results should be
extrapolated to Earth’s interior where necessary. This where W ¼ P dV (the symbol is used to indicate a
chapter provides a succinct review of some of the change in some quantity that depends on the path) is
important concepts in thermodynamics that play sig- the mechanical work done to the system where P is the
nificant roles in understanding the deformation of pressure, dV is the volume change, Z is the change in
materials in Earth’s interior. More complete discus- internal energy due to the change in the number of
sions on thermodynamics can be found in the text- atomic species, i.e.,
books such as C ALLEN (1960), DE G ROOT and M AZUR X @E
(1962), L ANDAU and L IFSHITZ (1964) and P RIGOGINE Z ¼ dni (2:2)
i
@ni S;V;nj
and D EFAY (1950).
13
14 Deformation of Earth Materials
@E @F If we consider a process in which the quantity Z is kept
P¼ ¼ (2:14c)
@V S;ni @V T;ni constant, then, dZ ¼ 0 and
@Z @Y @Z
@H @G ¼ : (2:17)
V¼ ¼ : (2:14d) @X Y;ni @X Z;ni @Y X;ni
@P S;ni @P T;ni
It can be seen that the thermodynamic quantities such Examples of such a relation include
as T, P, S and V (and i) can be derived from E, H, F
@S @P @S
and G. Therefore these quantities (E, H, F and G) ¼ (2:18a)
@T P;ni @T S;ni @P T;ni
are called the thermodynamic potentials. The thermo-
dynamic potentials assume the minimum value at
@T @S @T
thermochemical equilibrium. Because we will mostly ¼ (2:18b)
@V S;ni @V T;ni @S V;ni
consider a system at constant temperature and pres-
sure, the most frequently used thermodynamic and
potential is the Gibbs free energy. i is the thermo-
dynamic potential of the ith species (per unit mole). To @V @V @P
¼ : (2:18c)
emphasize the fact that i is the thermodynamic poten- @T P;ni @P T;ni @T V;ni
tial of the ith species per mole, it is often called the
partial molar thermodynamic potential (partial molar These thermodynamic identities (the Maxwell relations
Gibbs free energy when the independent variables are T and the relations (2.18)) are often used in manipulating
and P). thermodynamic relationships (e.g., Chapter 4).
Using the rule of calculus, it follows from (2.13) Now let us rewrite (2.13d) as,
and (2.14), X i
1 P
2 2 dS ¼ dE þ dV þ dni : (2:19)
@ E @ E @T @P T T i
T
¼ ) ¼
@S @V ni @V @S ni @V S;ni @S V;ni
At equilibrium, the entropy is a maximum, i.e., dS ¼ 0.
(2:15a)
Consider a case where two systems (1 and 2) are in
2 contact. In this case the condition for equilibrium can
@2 H @ H @T @V
¼ ) ¼ be written as
@S @P ni @P @S ni @P S;ni @S V;ni
(2:15b) 1 1 P1 P2
dS ¼ dE1 þ dE2 þ dV1 þ dV2
T1 T2 T1 T2
2 X (2:20)
@2 F @ F @S @P i i
¼ ) ¼ þ 1 dni1 þ 2 dni2 ¼ 0:
@T @V ni @V @T ni @V T;ni @T V;ni i
T1 T2
P1 ¼ P2 (2:22b) 1.0
and
Sconfig /R
of the system. These variables are called intensive quan-
tities. In contrast, quantities such as entropy, internal
energy and Gibbs free energy increase linearly with the 0.4
size of the system. They are called extensive quantities.
It follows that,
0.2
SðlE; lV; lni Þ ¼ lSðE; V; ni Þ (2:23)
where U is the energy of a static lattice (at T ¼ 0 K). Derive equation (2.32).
Therefore from S ¼ ð@F=@TÞ,
(
X
hoi
Solution
Svib ¼ kB log 1 exp
i
2pkB T From (2.10) and (2.12), noting that E, V and S are the
) extensive variables, one obtains,
X hoi 1
2pkB T expðhoi =2pkB TÞ 1 @G @E @V @S
i ¼ ¼ þP T
@nmol
T;P @nmol T;P @nmol T;P @nmol T;P
X hoi
kB log : @Sconfig
2pkB T ¼ e þ P Tsvib T
i @nmol
0 @Sconfig
The approximation is for high temperature, i.e., T :
@nmol
hoi =2pkB T 1.
where 0 e þ P Tsvib and e, and svib are molar
internal energy, molar volume and molar (vibrational)
vibration has a lower entropy. When the vibrational entropy respectively.
frequency changes between two phases (A and B), then Now, noting that dx ¼ dnmol =Nmol ðfrom
the change in entropy is given by, x ¼ n=N ¼ nmol =Nmol Þ, it follows from (2.28),
X @Sconfig =@nmol T;P ¼ R logðx=ð1 xÞÞ R log x.
oBi oBD
Svib SA B
vib Svib ¼ kB log R log : Therefore one obtains ðT; P; xÞ ¼ 0 ðT; PÞþ
i
oA
i oAD
RT log x.
(2:30)
where oA;B
D is a characteristic frequency of lattice vibra-
tion (the Debye frequency; see Box 4.3 in Chapter 4) of
a phase A or B. Activity
In a solid, the micro-state may be defined either by In deriving (2.32), we made an assumption that the
small displacements of atomic positions from their component under consideration has a small quantity
18 Deformation of Earth Materials
(dilute solution) so that atoms in the component do not increases logarithmically with pressure. For a non-
interact with each other or with other species. Such a ideal gas, one can assume a similar relation, i.e.,
material is called an ideal solution. In a real material
fðP; TÞ
where the interaction of atoms of a given component is ðP; TÞ ¼ ðP0 ; TÞ þ RT log (2:41)
P0
not negligible, a modification of these relations is
needed. A useful way to do this is to introduce the where ðT; P0 Þ is identical to the ideal gas. This is the
concept of activity (of the ith component), ai, which is definition of fugacity, f. The fugacity coefficient, , is
defined by, often used to characterize the deviation from ideal gas,
102
H2O
100
10–2
10–4
0 2 4 6 8 10 12 14
Pressure (GPa)
(b) 20
H2O
15 3
1573 K
Fugacity (GPa)
10 2 l/lm
ity
ac
fug
l/lm
5 1
s)
(ideal ga
fugacity
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Pressure (GPa)
20 Deformation of Earth Materials
When a fluid behaves like an ideal gas whose equa- TABLE 2.1 Equation of state parameters, a(T), b(T), c(T) and d(T)
tion of state is P ¼ RT, then its fugacity defined by for water and carbon dioxide.
equation (2.41) is equal to its (partial) pressure. All parameters are assumed to be parabolic function of
However, as fluids are compressed, their resistance temperature: m ¼ m0 þ m1 T þ m2 T 2 (where m is a, b, c or d).
to compression increases and the molar volume Units listed in the table are for m0. The unit for m1 is
does not change with pressure as much as an equation ½m1 ¼ ½m0 =T, and for m2 is ½m2 ¼ ½m0 =T 2 . Units:
of state of an ideal gas would imply. If the molar a (m6 Pa K1/2 mol1), b (m3), c (m3 Pa1/2), d (m3 Pa1).
volume does not change with pressure, for example,
then the fugacity will be an exponential function of CO2 H2O
pressure, a a0 ¼ 5.373 a0 ¼ 5.395 10
ðTÞ P a1¼ 5.6829 103 a1¼ 6.362 102
fðP; T Þ P0 exp : (2:47) a2¼ 4.045 106 a2¼ 2.368 105
RT
b b0¼ 4.288 105 b0¼ 2.7732 105
Important examples are water and carbon dioxide. – b1¼ 2.0179 108
The fugacities of water and carbon dioxide can be – b2 ¼ 9.2125 1012
calculated from the equations of state (Fig. 2.2). c c0 ¼ 7.526 1010 c0 ¼ 3.934 1010
Water behaves like a nearly ideal gas up to 0.3 GPa c1 ¼ 1.1440 1013 c1 ¼ 5.66 1013
(at T 41000 K), but its property starts to deviate – c2 ¼ 2.485 1016
from ideal gas behavior above 0.5 GPa. At
d d0 ¼ 3.707 1015 d0 ¼ 2.186 1015
P ¼ 2 GPa ðT ¼ 1500 KÞ, for example, the fugacity d1 ¼ 1.198 1020 d1 ¼ 3.6836 1018
of water is 13 GPa and at P ¼ 3 GPa ðT ¼ 1500 KÞ, d2 ¼ 1.0464 1022 d2 ¼ 1.6127 1021
it is 55 GPa. The large fugacity of water under high
pressures means that water is chemically highly reac-
tive under deep Earth conditions. The behavior of
carbon dioxide is similar. When extrapolating labo-
Solution
ratory data involving these fluids obtained at low Using equation (2.46), one obtains
pressures to higher pressures, one must take into
account the non-ideal gas behavior of these fluids Z P pffiffiffiffi
f 1 aR T pffiffiffi
(see Chapter 10). log ¼ þbþc þ d d
P RT 0 ðRTþbÞðRTþ2bÞ
Z P
1 a 1 1
¼ pffiffiffiffi
Problem 2.2
RT 0 T ðRTþbÞ ðRT=2þbÞ
pffiffiffi o
þbþc þ d d
The equations of state of water and carbon dioxide are
approximately given by the following formula (F ROST
and performing elementary integration and remember-
and W OOD , 1997b),
ing that the parameters a, b, c and d are functions of
pffiffiffiffi temperature, T, one obtains
RT aðTÞR T
ðP; TÞ ¼ þ bðTÞ
P ðRT
pffiffiffiffi þ bðTÞPÞðRT þ 2bðTÞPÞ fðP; TÞ aðTÞ RTþbðTÞP bðTÞP
þ cðTÞ P þ dðTÞP: log ¼ pffiffiffiffi log þ
P bðTÞRT T RTþ2bðTÞP RT
pffiffiffiffi 2
Where parameters (a, b, c and d) are functions of 2 cðTÞP P dðTÞP
þ þ :
temperature, but not of pressure (see Table 2.1). 3 RT 2RT
Show that the fugacity of these fluids is given by
Note that these gases behave like an ideal gas (i.e.,
fðP; TÞ aðTÞ RT þ bðTÞP bðTÞP f ! P) as P ! 0 as they should. At intermediate
log ¼ pffiffiffiffi log þ
P bðTÞRT T RT þ 2bðTÞP RT pressures (P 5–20 GPa for water or carbon dioxide),
pffiffiffiffi 2
2 cðTÞP P dðTÞP the third term (bðTÞP=RT) dominates and f=P
þ þ
3 RT 2RT expðbðTÞP=RTÞ whereas at extreme pressures (i.e.,
and using the parameter values shown below calculate P ! 1), f=P exp d ðTÞP2 =2RT . The results of the
the fugacities of water and carbon dioxide for the con- fugacity calculation are shown in Fig. 2.2.
ditions 0 5 P 5 20 GPa and 1000 5 T 52000 K.
Thermodynamics 21
This is called the law of mass action that relates the Discuss how the fugacity of water is controlled in a
concentration of chemical species with their chemical system that contains water as well as other materials such
potential. When the solution is not ideal (a case where as olivine, a metal that modifies the oxygen fugacity
solute atoms have a strong interaction with others), (H OBBS , 1984; K ARATO et al., 1986). For simplicity,
then equation (2.51a) must be modified to, assume that all the gaseous phases behave like an ideal gas.
22 Deformation of Earth Materials
TABLE 2.2 Thermodynamic properties of various oxides and fH2 O þ fH2 þ fO2 ¼ P: (3)
metals relevant to the oxygen fugacity buffer.
3
In the case where only water is present, then the dis-
( 106 m /mol): molar volume, h0 ðkJ/molÞ: molar enthalpy sociation of one mole of water produces one mole of
of formation from elements, s (J/mol K): molar entropy.
All quantities are at room pressure and T ¼ 298 K. Molar
hydrogen and 1/2 mole of oxygen, so fH2 ¼ 2fO2 .
volumes of some materials change with temperature and Inserting this into the equation for the law of mass
pressure as well as with phase transformations. However, action, and noting that one has
these changes are small relative to the difference in molar
volume of metals and their oxides. 3 2=3 1=3 2=3
fH2 O þ f P K ðT; PÞ ¼ P (4)
22=3 H2 O 0 1
h0 s where for simplicity, we assume that all the gasses are
O2 24 798 0 205.15 ideal, so that all the fugacity coefficients are 1.This
Fe 7.09 0 27.28 equation gives the fugacity of water when only water
FeO 12.00 272.04 59.80 is present. At high pressures, exceeding 1 GPa, the
Ni 6.59 0 29.87 second term in this equation is small (confirm this
NiO 10.97 239.74 37.99 yourself), so that fH2 O P, but when significant disso-
Mo 9.39 0 28.66 ciation occurs (at lower pressures), then the water
MoO2 19.58 587.85 50.02 fugacity will be lower.
Now consider a case where some other species are
present that also react with oxygen, hydrogen etc. For
example, let us consider a case where material A (e.g.,
1010 Fe) reacts with oxygen to form another mineral Ax Oy
(e.g., Fe2O3), namely,
T = 1373 K
y
105 xA þ O2 ¼ Ax Oy : (5)
2
Oxygen Fugacity (Pa)
Now the total pressure of the gas must be the same as the For a given chemical composition, a stable phase at a
given pressure, P, so that (assuming ideal gas behavior) given pressure and temperature is the phase for which
Thermodynamics 23
the Gibbs free energy is the minimum. When a material This type of phase transformation does not involve
with a given chemical composition can assume several changes in density or in entropy (hence no latent
phases, then as the P, T conditions change, the phase heat). Note that although there is no change in density
with the minimum Gibbs free energy may change from in these types of transformation, there is a change in
one to another. In these cases, the stable phase for a the elastic constants and thermal expansion (the
material changes with these variables, and a phase second derivatives of Gibbs free energy), and therefore
transformation occurs. They include to transforma- there must be a change in seismic wave velocities asso-
tion in quartz, order–disorder transformation in pla- ciated with a second-order phase transformation.
gioclase, (olivine) to (wadsleyite) transformation Schematic diagrams showing the change in free
in (Mg, Fe)2SiO4 and (bcc) to " (hcp) transformation energy associated with a first- and a second-order
in iron. transformation are shown in Fig. 2.4. In the case
A phase transformation may be classified into two where a first-order transformation is considered, a
groups. In some cases, a phase transformation involves material can assume two possible states. When the
a change in the first derivatives of Gibbs free energy free energy of one phase is lower than the other, then
(e.g., ð@G=@TÞP;ni ¼ S or ð@G=@PÞT;ni ¼ V, where S is a phase with lower free energy is more stable. Therefore
entropy and V is volume). This type of phase trans- if the transition from one state to the other is kineti-
formations is called the first order phase transforma- cally possible, then all the materials will transform to a
tion. Many phase transformations in silicates and phase with the lowest free energy. Note, however, that
metals are of this type. In these cases, there is a change this transition involves kinetic processes over a local
in density (molar volume) and heat is either released or maximum of free energy, and therefore the transfor-
absorbed upon the phase transformation (due to the mation takes a certain time to be completed.
change in molar entropy; recall that T dS is the latent Consequently, a metastable phase can exist in the case
heat). Another is the case where there is no change in of a first-order transformation when the kinetics
the molar volume or entropy (the first derivatives of involved are sluggish for a given time-scale. Examples
Gibbs free energy), but changes occur only in the include the presence of diamond at the Earth’s surface
second derivatives. This type of phase transformations (the stable phase for carbon at the Earth’s surface is
is referred to as the second order phase transformation. graphite, so we would not have diamond if the presence
Many of the structural phase transformations belong of everything on Earth were controlled by thermody-
to this class. The to transformation of quartz is namic stability), and the possible presence of metasta-
close to this type and many structural transformations ble olivine in cold regions of subducting slabs (see
of perovskite belong to this type (e.g., G HOSE , 1985). Chapters 17 and 20). The situation is different for a
(a) (P, T ) = (P1, T1) (P, T ) = (Pc, Tc) (P, T ) = (P2, T2)
G G G
phase 2 phase 1
phase 1 phase 2 phase 2
phase 1
(b)
(P, T ) = (P1, T1) (P, T ) = (Pc, Tc) (P, T ) = (P2, T2)
G G G
FIGURE 2.4 Free energy of a system having (a) first-order and (b) second-order phase transformations. The horizontal axis represents atomic
configuration (atomic positions, crystal structure).
24 Deformation of Earth Materials
dP @V1 =@T @V2 =@T @S1 =@T @S2 =@T
¼ ¼
Problem 2.6* dT eq @V1 =@P @V2 =@P @S1 =@P @S2 =@P
@S1 =@T @S2 =@T 1 2
Derive equation (2.55). ¼ ¼
@V1 =@P @V2 =@T 1=K1 1=K2
C1 C2
¼ : (6)
Solution Tð1 2 Þ
Now solving equations (1a) and (1b), the slope of the eA eB sA sB
P¼ þ T: (2:58)
phase boundary in the T–P space is given by, A B A B
26 Deformation of Earth Materials
XA XB XA XB
(c) (d)
T T
Liquid
Tc
T1 A B
A+B A+B
solvus
XA X XB XA X XB
Solid-solution, eutectic melting a modification to the phase diagram toward the end-
When there are two or more components in the system, member component representing the effects of finite
there are additional degrees of freedom by which the solubility (Fig. 2.5c). A phase diagram for a silicate and
chemical potential is controlled. Consequently, the water system at high T is an important example.
phase diagram depends on how the chemical potential Consider the equilibrium at temperature T1 below the
of each phase varies with the composition. For sim- eutectic point. When the amount of B is small, then the
plicity let us consider a two-component system. The only phase that exists is a phase A that contains a small
component i ¼ 1 and 2 may assume various phases amount of B. According to the Gibbs phase rule, in
such as solid and liquid. Two cases may be distin- such a case we have f ¼ c p þ 2 ¼ 3, that is this
guished. One is the case in which the two components phase, i.e., phase A with a small amount of B can
mix well in both the solid and liquid phases. In this exist for a range of T, P and composition. When the
case, the contribution from the configurational amount of B in the system increases, then at a certain
entropy is similar for both the solid and liquid phases, point, the phase A can no longer dissolve all the com-
and the free energy of each phase changes with compo- ponent B and there will be two phases ðX2 4X 4X1 Þ.
sition similarly following the compositional depend- The same thing happens from another side, namely the
ence of internal energy, entropy and the molar B-phase side. Consequently the domain is divided into
volume. The phase diagram corresponding to this one-phase domains in each side of the phase diagram
case is shown in Fig. 2.5a. In such a case, solid A and (A-rich or B-rich, X5X1 ; X 4X2 ) and a two-phase
B are said to form a solid-solution. Another is the case domain ðX2 4X 4X1 Þ. In the latter domain, there are
where mixing occurs only in the liquid phase. In this two phases that co-exist, and therefore the degree of
case, the contribution from the configurational freedom is f ¼ 2. Consequently, if temperature and
entropy is important only in the liquid phase. pressure are prescribed, then the chemical composi-
Consequently, the free energy of the liquid becomes tions of a material must be fixed. The boundaries
low in the intermediate concentration of a given spe- between the one- and two-phase regions correspond
cies, and therefore the solidus of the system is reduced to the solubility of each species into another.
significantly at intermediate compositions (Fig. 2.5b). Usually the solubility of another phase into a given
Melting behavior due to this type of mixing property is phase increases with temperature, so the boundaries
called eutectic melting. separating two one-phase domains will become closer
The solid-solution type behavior is observed when as temperature rises. These boundaries are often
the solid phases involved have similar properties (crys- referred to as a solvus. When mutual solubility is
tal structure and chemical bonding). The examples large, then at a certain temperature below the melting
include magnesiowüstite (MgO and FeO), olivine temperature, the two solvus curves merge. Above this
(fayalite Fe2SiO4 and forsterite Mg2SiO4), plagioclase critical temperature (Tc) the two phases mix com-
feldspar (albite NaAlSi3O8 and anorthite CaAl2Si2O8). pletely. Above this temperature mixing occurs both in
In all of these cases, ions that have similar ionic radii solid-state and liquid-state, and therefore the phase
are incorporated as a solid-solution in the solid phase. diagram above this temperature should look like that
If the ionic radii are largely different then the solubility of a solid-solution (Fig. 2.5d). Obviously the solvus
in the solid phase is limited and the eutectic behavior curves or any of the boundaries on a phase diagram
will occur. This is the case for the MgO–CaO, also depend on pressure. The temperature and pressure
MgSiO3–Mg2SiO4 systems. dependence of solvus curves for various combinations
of minerals is used as petrological barometers and/or
thermometers (e.g., W OOD and F RASER , 1976).
Solvus
Let us now consider a two-component system in which Effects of non-stoichiometry: a phase diagram
there is a finite solubility of each phase into another in for an open system
the solid state as well as in the liquid state. First, con- The phase diagram considered above assumes that the
sider a system in which mixing is complete in the liquid chemical composition of each phase is independent of
state and a small degree of mixing also occurs in the T and P except in cases where finite solubility of one
solid state. In such a case, a phase diagram needs to be component occurs in each phase. For example, a phase
modified. A solid phase always contains, in this case, a diagram for (Mg, Fe)O is usually constructed assuming
finite amount of secondary component so that there is that this is a two-component system (MgO and FeO)
28 Deformation of Earth Materials
10 000 / T(K)
O, oxygen), the ratio of the number of atoms of X and
O (stoichiometry) can deviate from what the chemical magnetite
formula would indicate. The deviation from the formal
chemical formula is referred to as non-stoichiometry. γ-iron wüstite
When non-stoichiometry occurs in an ionic crystal, 6.5
then charge balance must be maintained by creating
another type of charged species. This is usually done by
creating point defects or by incorporating another spe-
cies. One example is an Fe-bearing mineral such as
6.0
olivine ((Mg, Fe)2SiO4) that can have non-stoichiometry –10 –5 0 5
caused by a change of valence state of iron log f O (MPa)
2
ðFe2þ , Fe3þ Þ. In this case the charge balance is main- FIGURE 2.6 A phase diagram of Fe-O at 0.1 MPa.
tained by the change in the concentration of M-site
vacancies that have a negative effective charge (see at the intermediate oxygen fugacity and hematite
Chapter 5). Another example is a combined substitu- (Fe2O3) at high oxygen fugacity. Iron in wüstite is
tion such as Al3þ þ Fe3þ , Si4þ þ Fe2þ . In these mostly ferrous iron (Fe2 þ), whereas in magnetite
cases, an additional variable such as oxygen fugacity there are both ferrous iron (Fe2 þ) and ferric iron
or the activity of Al2 O3 is needed to specify the degree (Fe3 þ) and finally at high oxygen fugacity all iron
of non-stoichiometry. The degree of non-stoichiometry changes to ferric iron (Fe3 þ). The stability of iron-
in the former type of processes is usually small (10 4 bearing olivine can be analyzed in a similar way.
or less in olivine) but can be large in an Fe-rich Olivine accepts ferrous iron but not ferric iron (ferric
compound such as FeO (in FeO the non-stoichiometry iron is present in olivine but only with a very small
is 8%, i.e., Fe0.92O). Even in cases where the degree of amount, 1–10 ppm, as point defects) and therefore it
non-stoichiometry is small, its effects on physical prop- is stable only within a certain range of oxygen fugacity
erties can be important. In a binary material (such as that is determined by the stability of wüstite (FeO).
XO), the oxygen fugacity is used as an additional var- A somewhat different phase diagram applies when a
iable in constructing a phase diagram (N ITSAN , 1974). given mineral favors ferric iron more than ferrous iron.
(In a ternary system such as Mg2SiO4, the stoichiom- In such a case, even at an oxygen fugacity in which iron
etry is defined by two ratios (i.e., Mg=O; Mg=Si), and would occur as FeO, iron in that mineral can be ferric
hence one needs two additional parameters to completely iron. In some cases, the stability field of the ferric iron-
describe the chemical state of the system. Both oxygen bearing phase expands to a much lower oxygen fugac-
fugacity and the oxide activity must be specified in such ity, and in such a case a mineral containing ferric iron
a case.) could co-exist with metallic iron. An important case is
To illustrate this point, let us consider a phase dia- silicate perovskite that favors ferric iron, and the for-
gram of Fe–O. Iron (Fe) can assume three different mation of silicate perovskite from ringwoodite leads to
valence states dependent on oxygen fugacity, fO2: met- the formation of metallic iron (e.g., F ROST et al., 2004).
allic iron Fe0 at low oxygen fugacity, ferrous iron Fe2 þ
at intermediate oxygen fugacity and ferric iron Fe3 þ at 2.2. Some comments on the
high oxygen fugacity, see Fig. 2.6. Each species (Fe0, thermodynamics of a stressed system
Fe2 þ and Fe3 þ) has a different chemical character and
therefore the stable phases at different conditions will In the usual treatment of thermodynamics, the energy
depend on the oxygen fugacity. Consequently in an change of a system due to mechanical work is treated
Fe–O system, four compounds may be present depen- assuming hydrostatic stress. That is dW ¼ P dV, i.e.,
dent upon the oxygen fugacity, i.e., metallic iron at low the work done against pressure. An extension of such a
oxygen fugacity, wüstite (FeO) and magnetite (Fe3O4) treatment to non-hydrostatic stress conditions is
Thermodynamics 29
certain probability (see section 2.3). This process can natural to define the gradients of these quantities as
be assisted by the applied stress. Once this happens, generalized forces that drive the motion of correspond-
then upon the removal of the stress, the atomic config- ing (extensive) quantities. Therefore we define a gener-
uration of the material would not revert to the original alized force by
configuration instantaneously: thus the deformation
X r (2:62)
becomes irreversible (and time dependent because the
process now depends on the probability of atomic where ¼ 1=T; P=T; =T and define a linear rela-
jumps). tionship between the force and the flux, J, of an exten-
Plastic deformation occurs due to irreversible pro- sive quantity,
cesses. Therefore an understanding of irreversible pro-
cesses is critical in the study of plastic deformation. The J ¼ L X ¼ L r (2:63)
important issues here are the relationship between where L is a material constant that is often referred to
mass flux and various driving forces for mass flux as a phenomenological coefficient. For instance, the
and the interaction of various (generalized) forces. extensive quantity corresponding to an intensive
More detailed discussions of the thermodynamics of parameter 1=T is the internal energy, so the flux corre-
irreversible processes can be found in C ALLEN , 1960, sponding to rð1=TÞ will be the flux of energy, i.e., heat
DE G ROOT and M AZUR (1962) and G LANSDORFF and
flow. In this case, the empirical law is Fourier’s law,
P RIGOGINE (1971).
JT ¼ krT (2:64)
2.3.1. Flux and the generalized forces
where JT is the energy flux and k is thermal conductiv-
Irreversible motion or the flux of atoms (or energy) ity. Comparing Fourier’s law with equations (2.62) and
occurs when a system is out of equilibrium. One can (2.63) ( ¼ 1=T), one obtains
surmise that some forces cause such a flux.
L ¼ kT2 . (2:65)
Experimental observations show that in most cases
the rate of motion of atoms or the flux of atoms (or Similarly, the extensive quantity corresponding to an
energy) is linearly proportional to the force. An exam- intensive parameter =T is the number of atoms, n,
ple of such an empirical law is Fick’s law of diffusion so the corresponding flux is the flux of atoms. The
(of atoms) (also see Chapter 8), empirical relation for this case is therefore Fick’s law
J ¼ Drc: (2:61) of atomic diffusion,
From the above discussion, two methods of general- the fact that the entropy production rate (equation 2.66)
ization of the above linear relationship are obvious. First, is a scalar, so that upon the transformation of the coor-
Fick’s law describes the flux of atoms due to the concen- dinate this quantity must not change.
tration gradient, but more generally, atomic flux due to Among the off-diagonal components that represent
the gradient of chemical potential X ¼ rð=TÞ can be the coupling of different fluxes, the following symme-
considered. The chemical potential may include not only try relation, the Onsager reciprocal relation, must be
the gradient of concentration (gradient of configura- satisfied,
tional entropy) but also the gradient of free energy due
Lij ¼ Lji (2:70)
to other effects such as electrostatic field. Electrostatic
interaction among different diffusing species plays an if the independent fluxes are written as linear functions
important role in ionic crystals (see Chapter 8 for more of the independent generalized forces. The Onsager
details). In such a case, the electrostatic energy must be reciprocal relation is a consequence of the symmetry
included in the chemical potential thus of fluctuation with respect to time (for details see e.g.,
C ALLEN , 1960; DE G ROOT and M AZUR , 1962; L ANDAU
ð0 þ RT log ðc=c0 Þ þ q Þ
X ¼ r ¼ r and L IFSHITZ , 1964).
T T
q rc q rc
¼ r R ¼ ER (2:68)
T c T c 2.3.2. Some notes on the driving forces for
plastic deformation
where q is the electrostatic charge of the species, is the
electrostatic potential and E is the electrostatic field The driving forces considered in the previous section
(E ¼ r ). may be called thermodynamic forces (such as a force
Second, in writing equation (2.62), we considered due to the gradient of concentration of a given species).
only one force and one flux. A natural generalization However, under a deviatoric stress, there are forces that
of this relation is directly arise from the applied macroscopic (stress) field.
X In a solid under deviatoric stress, what forces drive
Ji ¼ Lij Xj (2:69)
j
large-scale atomic motion leading to plastic flow? Two
cases can be distinguished. First, for an atomic species
where we include various types of forces and corre- such as individual atoms or isolated point defects, there
sponding fluxes. The phenomenological coefficients dis- is no direct interaction between applied stress and these
cussed above correspond to the diagonal components, atomic species. In a perfect crystal, atoms are in their
Lkk, namely a material parameter describing the flow of lattice sites without any stress or strain. Similarly for
intensive variables (matter (or energy)) due to the corre- point defects, most of them are also isotropic and are
sponding gradient of extensive parameters (generalized not associated with deviatoric strain so that they do not
force). Off-diagonal components of phenomenological interact with deviatoric stress (exceptions are defect
coefficients (Lij with i ¼
6 j) express the coupling of differ- complexes, which can have anisotropic strain field and
ent fluxes. Physically these terms represent the flux of the interact with shear stress causing anelasticity; see
ith variable caused by the generalized force conjugate to Chapter 11). Consequently applied stress provides no
the jth variable. For example, the temperature gradient direct driving forces for atomic species. The most impor-
can cause the diffusion flux of atoms (Soret effect), and tant driving force for the motion of these atomic species
conversely the concentration gradient can cause the (e.g., point defects) under deviatoric stress is the ther-
temperature gradient (Dufour effect). Diffusion in a modynamic force, X ¼ rð=TÞ ¼ Rðrc=cÞ, caused
multi-component system is an important example by the heterogeneity of defect concentration due to the
where coupling among different species can cause heterogeneous microstructure of a polycrystalline mate-
important effects on mass transport (e.g., L ASAGA , rial (see Chapter 8). Heterogeneity of defect concentra-
1997, see also Chapter 8). The entropy production rate tion can also occur due to the local stress heterogeneity
contains various terms including the scalar (PV term), due to the presence of a dislocation or a grain boundary
vector (heat flow, diffusional mass flow) as well as tensor (see Chapters 8 and 9). In these cases, dislocation
(energy dissipation due to plastic flow). A coupling of motion or motion of atoms across a grain boundary
flux of various variables can occur only when these occurs due to (indirect) thermodynamic force. Second
variables have the same transformation with respect to is the case for dislocation motion, which is different
the change of the coordinate system. This comes from from the case for point defects. Due to the long-range
32 Deformation of Earth Materials
X
displacement field associated with it, a dislocation has a ¼ Lij Xi Xj (2:73)
direct mechanical interaction with the applied stress i;j
(Peach–Koehler force: see Chapter 5). A dislocation
where we used the Onsager reciprocal relation. Now con-
moves both by the direct effect of applied stress as well
sider a case where one force (Xk) is fixed but others are free
as by the thermodynamic forces caused by the concen-
to vary. The entropy production rate is minimum with
tration gradient of point defects around it.
respect to Xi ði ¼
6 kÞ when the following condition is met,
In some literature, the gradient in chemical potential
that drives processes such as diffusion under stress, , is @ X ij j
¼ L X ¼ Ji ¼ 0 ði 6¼ kÞ: (2:74)
written as (e.g., S CHMALZRIED , 1995) @Xi j
O
r ¼ Rr log a r (2:71) At the stationary state, all the fluxes other than the one
T T
corresponding to a fixed force (Xk) should vanish. Now
where a is the activity and O is the molar volume of the from the conservation equation for an extensive
species involved. This expression is misleading and parameter i ; ð@i =@tÞ ¼ r Ji , it is seen that the
should be used with care. First, this equation would state of minimum entropy production rate (i.e.,
imply that the applied stress has a direct effect on defect Ji ¼ 0) corresponds to the stationary state
motion. This is not true in many cases. For example, ð@i =@t ¼ 0Þ. This is referred to as the principle of
diffusion flux in a polycrystalline material under devia- minimum entropy production rate.
toric stress is caused by the indirect thermodynamic
force due to the local (grain scale) variation in the
normal component of stress at the place where crea- 2.4. Thermally activated processes
tion/destruction of defects occur (Chapter 8). In other
2.4.1. Absolute rate theory
words, stress in this equation should not be the macro-
scopic stress.3 The use of equation (2.71) will obscure Basic theory
this basic physics of stress-induced kinetic processes An irreversible process involves an atomic jump from
and is not recommended. Second, in the case of diffu- one stable position to the next. The atomic jump is
sion flux in a polycrystalline material, the two terms in often caused by a large fluctuation of atomic positions
equation (2.71) are redundant if the stress dependence due to lattice vibration. Such a process is referred to as a
of concentration is included in the activity term thermally activated process. The energy of an atom in a
because in this case, the driving force is a thermody- crystal has a minimum value at the lattice sites. As one
namic driving force, rð=TÞ ¼ Rr log a, caused by measures the energy of an atom from one lattice site to
the grain-scale variation in stress, and is given the neighboring site, the energy has a local maximum
by rð=TÞ ¼ Rr log a ¼ ðO=TÞ ð =LÞ ðO=TÞ somewhere in the middle of the two lattice sites. The
r ða ¼ a0 expð O=RTÞÞ. value of the energy maximum depends on the pass
along which one moves an atom. The position at which
2.3.3. Stationary (steady-state) state: the principle the energy at the peak is minimum is called a saddle point
of minimum entropy production rate (Fig. 2.8). Through thermal vibration, atomic positions
fluctuate. Therefore, with some non-zero probability, a
The flow of materials (or energy) causes entropy pro- system can assume an atomic configuration in which an
duction (energy dissipation). The rate of local entropy
production is given by H
dS X i i X i S
¼ JX ¼ J ri : (2:72)
dt i i
H*
Inserting the relation (2.69) into (2.72), one gets
A
A′
FIGURE 2.8 A plot of enthalpy as a function of atomic position
3
If equation (2.71) is used, with being the macroscopic stress, then there associated with an atomic jump from position A to position A0 . S is the
would be no diffusion or deformation for a homogeneous deviatoric saddle point. H ¼ E þ PV is the activation enthalpy (E* is the
stress. activation energy and V* is the activation volume).
Thermodynamics 33
atom goes to a saddle point. At a saddle point, the ES EA is always positive whereas VS VA can be
potential energy of an atom has a maximum with respect positive or negative (see Chapter 10). Consequently,
to the spatial coordinate, and consequently, this config- an atomic jump is always enhanced by higher temper-
uration is unstable and an atom moves to the next stable ature, whereas higher pressure usually suppresses an
position (or goes back to the original position). When the atomic jump (Chapter 10). Also note that the temper-
velocity of an atom at that saddle position is ‘‘positive,’’ ature dependence of thermally activated processes
then the atom goes to the next stable position (a lattice results in the time dependence of the process because
site). Since this new configuration is stable, that atom these processes occur as a result of fluctuation.
does not go back to the original position instantaneously.
Therefore the motion over a saddle point is irreversible. Effects of stress on thermal activation
The rate at which this occurs may be calculated by The above formula shows that the rate at which a ther-
applying the principles of statistical physics and the fol- mally activated process occurs is strongly dependent on
lowing theory is called the absolute rate theory, which temperature and pressure through activation enthalpy,
was developed by E YRING (1935). The application of H*. When an atom (or defect) has an interaction with
this theory to atomic migration in a solid was made by stress, the activation enthalpy becomes dependent on the
F LYNN (1968) and V INEYARD (1957) (see also F LYNN stress. This is a case for dislocations (see Chapters 5 and
(1972)). Noting that the activated state has different 9) which have an anisotropic strain field and hence have
chemical bonding and hence the frequencies of lattice strong interactions with applied (deviatoric) stress. In
vibration are different, one obtains (e.g., F LYNN (1972)) such a case, the activation enthalpy becomes stress
dependent, H*( ). In these cases, applied stress changes
H
w ¼ ~ exp (2:75) the energy of a dislocation when it moves from the
RT ground state to the saddle point configuration. The
where stress works on a piece of dislocation when a piece of
dislocation moves to the direction of the Peach–Koehler
Q
3N force (Chapter 5) and hence the enthalpy will be
oi
i¼1 reduced. On the other hand, when a dislocation moves
~ ¼ (2:76)
Q
3N in the opposite direction, the enthalpy for the barrier will
oi0
i¼2 increase. The net rate of motion will then be given by
The formal theory of deformation plays an important role in formulating energy dissipation
(i.e., seismic wave attenuation) and non-linear rheological relationships. This chapter presents a
brief summary of the phenomenological theory of plastic deformation. This includes the classification
of deformation (elastic, viscous, plastic etc.), the mathematical formula for constitutive relations,
formulation of transient creep and the mathematical formula appropriate for non-linear rheology.
Key words elasticity, visco-elasticity, anelasticity, constitutive relations, Levy–von Mises equation,
mechanical equation of state, transient creep, creep response function, creep compliance function,
Kramers–Kronig relation, Maxwell model, Voigt model, Zener model, Burgers model.
Strain
transient steady state
be non-recoverable. Fracture or low-temperature
yielding at high stress is nearly time-independent yet
deformation is non-recoverable. This type of behavior
is, however, better considered as an end-member εe
behavior of plastic deformation (e.g., H ART (1970)).
The mode of non-elastic deformation is conveniently Time
classified into brittle and plastic (ductile) deformation.
FIGURE 3.2 A schematic diagram of a strain versus time relation for
In more casual terms, brittle deformation is fracture and a constant stress test showing various stages of deformation, where "e
ductile deformation is flow. These two processes are is the elastic strain. This figure corresponds to a case where work-
nearly independent and therefore the non-elastic hardening occurs in the transient stage.
response of a material to the external force is either
brittle fracture or ductile flow whichever is easier.
Fracture involves the macroscopic breaking of chemical elastic response yielding elastic strain, but we focus on
bonds (at the scale of cracks), which occurs in most the non-elastic component. Non-elastic deformation
cases in a localized fashion. In most cases, ductile flow usually starts with a transient period in which the strain
involves the microscopic motion of atoms and in many rate (at constant stress) or stress (at constant strain
cases occurs homogeneously. Consequently, micro- rate) changes with time. Eventually a material will
scopic, thermally activated motion of atoms (defects) usually assume steady-state deformation in which the
controls ductile deformation, but thermally activated strain rate (for constant stress) or stress (for constant
processes play a less important role in brittle fracture. strain rate) remains constant with time. A few points
In this book, we focus on plastic flow, but a brief should be noted. First, the steady state is only approx-
summary of brittle deformation is given in Chapter 7. imately defined, and in fact, there are several quasi-
Extensive reviews of brittle deformation are given by steady-states of deformation for a given material.
P ATERSON and W ONG (2005) and S CHOLZ (2002). This is due to the fact that plastic deformation is sensi-
tive to microstructures and there is a hierarchy
3.2. Some general features of plastic of microstructures. As microstructures evolve, plastic
deformation properties also evolve. Consequently, when different
microstructures assume quasi-steady-state at different
Fig. 3.2 illustrates a typical result of mechanical tests in time-scales, there will be several quasi-steady-states of
the plastic regime. In all cases, there is an instantaneous deformation for a given material. One example is the
deformation of a polycrystalline material in the dislo-
cation creep regime (see Chapter 9). Dislocation micro-
1
In the literature, the term ‘‘plastic deformation’’ is often used to describe
structures at each grain level will evolve during
nearly time-independent yielding, and the other type of strongly time-
dependent deformation is called viscous flow. However, in this book I use
deformation, and quasi-steady-state deformation will
‘‘plastic deformation’’ for any time-dependent non-recoverable be reached when dislocation microstructures achieve
deformation. steady state (constant dislocation density, constant
36 Deformation of Earth Materials
subgrain structures etc.). At the same time, orienta- steady state, the plastic properties of a material are
tions of individual grains will evolve with deformation characterized by a single scalar parameter, viscosity,2
to result in a lattice-preferred orientation (see , defined by,
Chapter 14) that will also affect plastic deformation.
ij 2"_ ij : (3:2)
Consequently, another quasi-steady-state will be
achieved when steady-state lattice-preferred orienta- Consequently, the plastic properties in such a case
tion is established. Grain size may also evolve during can be determined by measuring the one component
deformation that affects the plastic flow (see of strain rate corresponding to one type of stress.
Chapter 13). In most cases, these grain-scale micro- For example, one can apply normal stress on the x1
structures evolve with longer time constants than plane, 1 , and observe the corresponding strain
dislocation microstructures. Therefore quasi-steady- rate "_ 1 . The viscosity can be determined by
states corresponding to steady-state lattice-preferred ¼ ð1 3 Þ=2ð"_ 1 "_ 3 Þ ¼ ð1 3 Þ=3"_ 1 (note that
orientation and steady-state grain size will be achieved "_ 3 ¼ 12"_ 1 ) where 3 is the lateral stress (confining
in later stages than the quasi-steady-state correspond- pressure). Given this effective viscosity, strain rates
ing to steady-state dislocation microstructures. It is corresponding to any other stress components can
important to define the scale that one is talking about be calculated from (3.2). For a more general aniso-
when one discusses the time-dependent deformation of tropic rheology, one has to determine the rheological
materials. Second, under some conditions, steady-state properties for a large number of orientations.
deformation becomes unstable and localized unstable In the case of non-linear rheology, one may deter-
deformation occurs (Chapter 16). mine a relationship such as "_ 1 ¼ Að1 3 Þn (n > 1).
In this chapter we will discuss some general (phe- But this equation cannot be generalized to "_ ij ¼ Anij ,
nomenological) aspects of the mathematical descrip- because such a relation violates the rule of transforma-
tion of mechanical behavior in the plastic regime. The tion of tensors (if one rotates the coordinate system,
issues that we discuss in this chapter include: (i) the the left- and the right-hand sides would change differ-
formulation of non-linear flow laws (section 3.3), ently). If a material is isotropic, stress and strain rate
(ii) the formulation of transient deformation behavior must be related by a scalar quantity. Therefore, for
(section 3.4) and (iii) the influence of ductile deforma- non-linear rheology, one must have,
tion on energy loss and on apparent elastic moduli
"_ ij ¼ Bðij Þ ij (3:3)
(section 3.5).
where B is a scalar function of stress. Because B is a
3.3. Constitutive relationships for scalar, it must contain only the invariants of (devia-
non-linear rheology toric) stress tensor (see Chapter 1). Because the first
invariant of stress tensor is hydrostatic pressure, the
The plastic properties of a material are expressed in first invariant is usually assumed not to contribute to
terms of a relationship between stress and strain or deformation (other than its effect on a constant
strain-rate. Therefore a material parameter (or para- term related to the rate of thermal activation, see
meters) that characterizes plastic deformation must Chapter 10). The third invariant is also assumed not
express a relation between two second rank tensors to contribute to assure that energy dissipation is always
(deviatoric stress ij and deviatoric strain rate "_ ij ). In positive (see equation (3.11)). In this case, the simplest
the simplest case of linear rheology at steady state, the form will be,
relation should be
X B ¼ BðII Þ ¼ B IIðn1Þ=2 : (3:4)
ij 2 ijkl "_ kl (3:1)
k;l Therefore equation (3.3) becomes
where ijkl is the viscosity tensor. It follows that a large "_ ij ¼ B IIðn1Þ=2
ij : (3:5)
number of measurements along various orientations
are needed to fully characterize the plastic properties
2
Similar to an elastic material (see Chapter 4), there are two viscosities for
of a material. However, in many cases, a material is
an isotropic material, one for shear and another for bulk deformation.
approximately isotropic. This would be the case where However, for a perfectly dense solid, the bulk viscosity is infinite.
one considers deformation of a polycrystal whose Therefore we consider only shear viscosity here (for the bulk viscosity, see
microstructure is nearly isotropic. In such a case, at also Chapter 12).
Phenomenological theory of deformation 37
This is referred to as the Levy–von Mises equation. Therefore II ¼ 13ð1 3 Þ2 (note that only deviatoric
In this case (i.e., isotropic rheology), one can trans- stress contributes to plastic flow in the ductile regime).
late the results from one type of deformation geom- The flow law for the tri-axial test assumes the form,
etry to another deformation geometry. Note that
when large internal stress exists in addition to the "_ 1 ¼ "_ ¼ Cð1 3 Þn
applied stress, the BðII Þ term will be largely modi-
fied by the internal stress. Consequently, the plastic Comparing this relation with equation (3.5),
flow of a material will be affected by the presence of B ¼ ðC=2Þ3ðnþ1Þ=2 .
internal stress. An application of this notion to plas- Now for simple shear,
tic deformation during a phase transformation is 2 3
1
discussed in Chapter 15. 0 2_ 0
"_ ij ¼ 4 12_ 0 05
0 0 0
Problem 3.1*
and
Let us consider two cases. In uni-axial (tri-axial) 2 3
compression tests, one applies uni-axial stress, and 0 0
0ij ¼ 4 0 0 5:
measures uni-axial strain rate. Usually, the results of
0 0 0
such tests are presented as a relationship between
1 3 and "_ 1 as,
Therefore II ¼ 2 and using (3.5), one gets
C ðnþ1Þ=2 n
1
2_ ¼ 2 3 . Hence D ¼ 3ðnþ1Þ=2 C.
"_ 1 ¼ Cð1 3 Þn (1)
Plastic deformation dissipates energy. The energy pressure are attained). If the mechanical equation of
dissipation per unit volume due to plastic deformation state indeed exists, then one can predict the mechanical
is given by properties of a material including transient behavior
X without knowing its detailed history. H ART (1970)
¼ ij "_ ij (3:11) showed that if a mechanical equation of state exists,
i;j
then y must be a unique function of strain, yð"Þ, and
Using the equations (3.6) and (3.7), one obtains should not explicitly depend on time (see also
1=n ðnþ1Þ=n M C C ARTNEY (1976)). Constitutive relations consistent
¼ B1 nþ1
e ¼ B "_ e : (3:12) with this requirement include
Note that the energy dissipation by plastic flow is a
¼ A"q "_ r þ 0 (3:14)
strong function of stress (or strain rate).
(e.g., H ART , 1970; P OIRIER , 1980) or
3.4. Constitutive relation for transient s
"
creep ¼ B 1 exp "_ r þ 0 (3:15)
"0
Transient creep is a phenomenon in which the rate of (e.g., C HINH et al., 2004; V OCE , 1948) (q, r, s > 0) which
deformation at a given stress (temperature and pres- have been used in the literature to describe strain
sure) changes with time (or strain). Since all of these (work) hardening behavior. Note that equation (3.14)
variables (stress, temperature, pressure) are kept con- does not describe steady-state flow (at infinite strain,
stant, this temporal variation of mechanical properties strength would become infinite which is physically
must be attributed to the evolution of internal structure meaningless), whereas (3.15) does.
(or internal mechanical state such as stress distribu- However, there is no reason to believe that a
tion). Therefore a general equation for transient creep mechanical equation of state exists for a range of
is given by materials or conditions, and in fact, in many cases,
¼ Fð";
_ T; P; yÞ (3:13) mechanical properties of materials indeed depend on
their history and hence a mechanical equation of state
where y is a parameter that defines the internal struc- does not exist. For example, when deformation occurs
ture or state of a material such as dislocation density or by grain-size sensitive creep and when grain size is
grain size, or the distribution of internal stress. As controlled by grain growth, then the strength of the
deformation proceeds, y evolves that leads to transient material depends on the entire history of that material
behavior. Annealing may also cause a change in y (the initial grain size and the temperature–pressure
without a change in strain that leads to time-dependent history). Similarly, deformation superposed with
(transient) behavior. The specific form of such a func- pre-existing deformation depends on the pre-existing
tion depends on the microscopic mechanisms of defor- conditions such as the initial stress (this is the case for
mation that are discussed in Chapters 8 and 9. In this post-glacial rebound, in which deformation occurs
section I will summarize some of the general aspects of due to stress caused by the variation of surface load
the formulation of time-dependent deformation. in addition to the pre-existing stress due to long-term
There have been some discussions as to how to convection, see Chapter 18).
formulate time-dependent deformation. The question In these cases, the transient creep behavior can be
is closely related to the issue of whether a mechanical described explicitly in terms of time variation of strain
equation of state exists or not (H ART , 1970; H OLLOMON , (at a constant stress). In many cases the experimental
1947; Z ENER and H OLLOMON , 1946). When the stress data can be fitted to the logarithmic constitutive equa-
required for plastic deformation depends on the instan- tion, namely,
taneous values of the strain, strain rate, temperature
"ðtÞ ¼ "ela þ B logð1 þ Atn Þ þ "_ s t ð05 n 1Þ (3:16)
and pressure and not on their past histories, it is said
that a mechanical equation of state exists (similar to where "ela is elastic strain A and B are constants, and "_ s is
the equation of state of material, by which the density the steady-state strain-rate. This equation is reduced to
of a given material is defined only by the state variables the following ‘‘parabolic’’ flow law for Atn 1,
such as temperature and pressure without any depend-
ence on the path through which the temperature and "ðtÞ ¼ "ela þ ABtn þ "_ s t (3:17)
Another Random Document on
Scribd Without Any Related Topics
The next step (F) consists in laying out the form of the propeller blade upon
all four sides and ends of the block, and Step G is the final one of cutting out
the propeller, scooping out its blades concave on one side, and carving them
convex on the opposite side. A very sharp knife must be used for cutting; and
the work must be done slowly and carefully, because the least slip is likely to
ruin the propeller. The entering-edge of each blade is the almost straight
edge, and should be cut very thin. The ends of the blades should also be cut
thin, while the hub should be cut away as much as can safely be done without
weakening the propeller.
When you have completed cutting the propellers, place them at their centers
across the edge of a knife-blade, and if they do not balance perfectly, locate
the trouble and correct it. Finish the work with fine emery-paper, and then
shellac it. Some boys glue silk over the ends of their propeller blades, for a
distance of ½ inch or so, to reinforce them and make them less likely to split.
The Propeller-shafts are made of heavy piano-wire, bent into a hook at one
end (Fig. 38) to receive the rubber strands of the motor, and cut of the right
length to extend through the hole in the bearing, through a glass bead,
through the propeller, and then to bend over the side of the hub (Figs. 37 and
38). By bending over the end of the shaft against the hub, it is held securely
in place.
The Motors consist of twelve strands of 1/8-inch flat rubber, each, and as
these are 1 yard in length, exactly 24 yards of rubber are required. The
rubber is not connected direct to the hooks on the bow and propeller-shafts,
as the wire would quickly cut through the strands. Instead, small rings are
bent out of wire, with pieces of small rubber-tubing slipped over the wire, and
the ends of the rubber strands are looped through these rings and bound in
place with thread (Fig. 39). The wire rings are then slipped on and off the
hooks quickly. As light and heat cause rubber to deteriorate, you must remove
the motors from the machine after use, pack away in a covered box, and keep
in a cool place, in order to get the longest life possible out of the rubber.
It has been found that rubber motors can be wound much farther by
lubricating them with glycerine. It is only necessary to put a few drops of the
glycerine upon a clean cloth, and rub it over the outside strands; then wind
the motors, and it will work over the surface of the inner strands until all parts
are covered.
Fig. 45.—A Home-made Motor Winder.
Fig. 46.—The Kind of Egg-beater to Use.
Fig. 47.—How the Motors are Connected to Winder for Winding.
To Complete the Boat, go over the work carefully, trim off all projecting
edges, drive nail heads beneath the surfaces, putty nail holes and cracks, and
give the wood two coats of paint of whatever color you want to have the
motor-boat.
The Propeller (E, Fig. 54) is cut from the side of a tin can. Cut a piece 3
inches long and ¾ inch wide, round its ends, and with the point of a nail
pierce a hole through it each side of the center of the length of the piece (Fig.
55). To finish the propeller, it is only necessary to take hold of the two ends
and twist the piece into the shape shown in Fig. 56.
The Propeller-shaft requires a short piece of wire with one end bent into a
hook (F, Fig. 56). Stick the straight end of this shaft through one hole in the
propeller, and the hooked end through the other hole, then twist the hooked
end over on to the main part of the shaft, as shown in Fig. 57. Make a tight
twist so the propeller will be held perfectly rigid on the shaft.
The Bearing Plate G (Figs. 54 and 58) supports the propeller. Cut it out of a
piece of tin 1½ inches wide by 3 inches long, bend it in half crosswise to give
it stiffness, and then bend it lengthwise to the angle shown so it will fit over
the slanted stern of the boat. Punch two holes through the upper end for
nailing the plate to the stern, and a hole at the lower end for the propeller-
shaft to run through.
For a Thrust Bearing, slip a couple of beads over the propeller-shaft,
between the propeller and bearing plate G. Probably you can find glass beads
in your mother's button bag.
Fig. 54.—Longitudinal Section of Assembled Motor-boat.
Figs. 55-59.—Details of Propeller.
Fig. 60.—Rubber-band Motor.
After slipping the beads on to the shaft, and sticking the shaft through the
hole in bearing plate G, bend the end of the shaft into a hook; then screw a
small screw-hook into the bottom of the hull, at the bow end (I, Fig. 54), and
you will be ready for
The Rubber-band Motor. Rubber-bands about 1½ inches in length are best
for the purpose. Loop these together end to end (Fig. 60) to form a strand
that will reach from hook I to the hook on the propeller-shaft; then form three
more strands of this same length, and slip the end loops of all four strands
over the hooks.
To Wind the Motor, give the propeller about one hundred turns with your
finger; then, keep hold of the propeller until you launch the boat.
There are many ways of elaborating upon the design and construction of this
toy motor-boat, but, having given the necessary instructions for building a
simple model, I am going to leave further development for you to work out.
Here is an opportunity for you to use your ingenuity. Devise an adjustable
rudder, add a keel, finish off the cockpit with a coaming, install a headlight
made from a pocket flashlight—in fact, see just how complete a motor-boat
model you can build.
CHAPTER V
HOME-MADE TOY WATER-MOTORS
You can own a water-motor like the one shown in Fig. 61, because its
construction requires nothing but easily obtained materials.
Prepare Eight Paddles 1¾ inches wide and 2½ inches long, out of cigar-
box wood. Locate the positions for the ends of the paddles, upon the side
pieces, by drawing a horizontal line, a vertical line, and two diagonal lines at
angles of 45 degrees, through their centers. This will simplify the matter of
spacing the paddles equidistant from one another (Fig. 67). Use brads for
fastening the side pieces to the paddle ends. Those removed from the cigar
boxes will do.
The Wheel Shaft should be a trifle shorter than the inside width of the can,
and enough smaller than the ¼-inch hole in the wheel side pieces so the
wheel will turn freely. Locate the centers for the axle upon the two sides of
the can, in the proper position so there will be the same margin above and at
the ends of the wheel. Drive a nail through each side of the can into the axle
end.
Fig. 65.—The Completed Water-motor Wheel.
Figs. 66 and 67.—Details of Water-motor Wheel.
An Outlet for the water after it has passed over the wheel paddles must be
provided, and the best way is to fasten a strip to two opposite sides of the
can so as to raise the bottom about an inch, as shown in Figs. 62, 63, and 64.
For a Pulley-belt use a piece of heavy cord. Cut a slot through the front of
the can for the belt to run through, and make this slot large enough so the
cord will not rub against the sides (Fig. 63).
Pulley-wheels for attaining different speeds can be made of spools of
various sizes. A bicycle wheel with the tire removed, mounted in a frame, is
excellent for a large wheel.
Connecting up the Water-motor. If
you operate the water-motor in the
kitchen sink, you can either build a
platform as shown in Fig. 61, to bring the
spout of the varnish-can case up to the
level of the faucet, or you can set the
water-motor in the sink and lead a piece
of rubber tubing from the spout to the
faucet, as shown in Fig. 68. If you use the
latter arrangement, slip the lower end of
the rubber tubing over a short piece of
glass, brass, or tin tubing, and stick the
short tubing through a hole in a cork
large enough to fit the spout of the
Figs. 68-69.—How to Make a Water-
tight Connection between Faucet varnish-can case (Fig. 69). If you raise
and Water-motor. the water-motor high enough so the
faucet will set down into the spout, you
can cut a large enough hole for the faucet, through a cork, and then fit the
cork in the spout as shown in Fig. 64.
Another Water-motor. The little water-motor in Fig. 70 will furnish
sufficient power to operate simple mechanical toys.
The Water-motor Wheel. Procure two baking-powder can covers for the
ends of the water-motor wheel (A, Fig. 72), a cigar-box out of which to make
the wheel paddles, and a stick ¼ inch square and 5 inches long for the wheel
axle (B, Fig. 72).
Cut eight paddles from the
cigar-box wood 1 inch wide
and 5 inches long. Take a pair
of these strips and fasten
them to one can cover, in line
with each other, and close
against the sides of the cover
(C, Fig. 73). Fasten with tacks
or brads driven through the
cover into the ends of the
strips. Take another pair of
strips and fasten them to the
same cover, in a similar
manner, at right angles to pair
C (D, Fig. 72). Then tack the
pairs of strips E and F to the
cover halfway between pairs C
and D. With the paddles in
position, locate the exact Fig. 70.—A Small Water-motor that can be
Operated in a Wash-Basin.
center of the end of the can
cover, and drive a nail through
at this point into the end of axle B. Slip the free ends of the paddles into the
other can cover, and carefully drive tacks or brads through the cover into
them. Drive a nail through the center of the cover into the end of axle B.
The Wheel Supports. Figure 74 shows the supports for the wheel. Cut the
end pieces G 4 inches wide and 6 inches high, and the cross strips H 1¾
inches wide and 5½ inches long. Nail pieces G to H, as shown, allowing the
lower ends of G to extend ½ inch below strips H, and leaving a space of ½
inch between strips H. The axle holes in pieces G (Fig. 74) should be located
in the center of the width of these pieces, and halfway between their tops and
strips H. Bore the holes with a
gimlet, or make them by
driving a large nail through the
pieces, and then withdrawing
it.
To Mount the Wheel upon
the supports, withdraw the
nails driven into the ends of
axle B, slip the wheel between
uprights G, and drive the nails
through the holes in G back
into the holes in the axle ends
(Fig. 71).
The Pulley Wheel. One can
cover should be converted into
a pulley by winding several
turns of string around it, near
each edge, leaving a groove
between the string. Coat the Fig. 71.—The Water-motor Wheel.
string with glue to make it
stick fast to the cover.
The Water-motor Case.
Figure 70 shows how the
water-motor case is
constructed by fastening
boards N, I, J, K, L, and M to
the wheel supports G. There
must be a slot through I and
another through J, for the
string belt to pass through,
and a hole through K for the
intake of water from a faucet. These can be cut out of the edges of the
boards, as shown, before they are nailed in place. Leave an opening between
boards N and M, and the bottom of ends G, for an outlet for waste water.
Figs. 72 and 73.—Details of Water-motor Wheel.
Fig. 75.—Upright.
The window openings may be cut in each end, but it will make a stronger car
if they are simply drawn upon it. Cut four cardboard steps similar to Fig. 93
and tack them to the sides of the front and rear platforms. When the car has
been put together, replace the wire in the tops of uprights I and J (Fig. 81),
running the ends through the roof (see Fig. 84). Paint the sides and ends of
the car yellow with brown trimmings, and paint the roof a light gray. Water
colors can be used for the purpose. Letter the name of your car-line upon the
sides and the number of the car upon each end and side. The route should be
lettered upon strips of cardboard with pins run through them as shown in Fig.
94, these strips to stick in the roof of the car (see Figs. 84 and 85).
Having seen how the car is made, you will find it a simple matter to make
designs for
Other Cars, using the same scheme for the trucks, and altering the patterns
for the sides, ends, and roof, to suit the design.
Nothing has, as yet, been said about the
Operation of the Railway, and though Fig. 76 probably shows sufficiently
clear how it is run, a few words may be helpful. The car or cars are placed
between the wooden tracks, and the trolley (or cord attached to the ring on
top of the car) is tied to the trolley-line as in the illustration. Upon starting
your engine, water-motor, or whatever motive-power you have, the car will
run from one end of the track to the other. When it has reached the support
of the trolley-line, it will stop long enough for the cord trolley to pass around
the wooden wheel, and then run in the opposite direction until the other
support is reached. It will thus be seen that the trolley hangs to the upper
part of the cable, or trolley-line, in running one way, and to the lower part on
the return run. In changing the direction of the run, the ring to which the
trolley is attached slides to the other end of the car.
Fig. 99.—
Partitions.
Fig. 100.—Front View of Elevator Shaft.
Figs. 101 and 102.—Elevator Car Details.
To Make the Car Rise to the top of the shaft again, remove bottle Q.
Replace the bottle upon the end of hook R, and it will be in position for the
next trip downwards.
Cut the holes Y and Z (Fig. 100) through the outside wall of the shaft for hand
holes through which to reach bottle Q and hook R.
Figures 97, 100, and 103 show
A Simple Control for stopping the car at the
different floor levels. Stick S may be a piece of
broom-handle, curtain-pole, or flagstaff. Bore a
hole through the bottom of the shaft, directly
below holes C in the floors (T, Fig. 100), and
slip the stick through hole T and into slots C.
Then locate on one side of stick S points just
below the under face of each floor, and upon
the opposite side locate points just above
where the back edge of the elevator will come
when the car is raised to each floor level (Fig.
97). Remove the stick, and drill or bore a small
hole at each point marked; then replace it, nail
a small block (U, Fig. 100) across the top end
to hold it in place, and drive a nail, with its
head filed off, into each of the holes.
When stick S is turned to the position shown in
Fig. 97, while the car is going down, the first
nail below the car will project beneath it and
bring it to a stop; and if the stick is turned in
the opposite direction while the elevator is
going up, the first nail above the car will project
over the back edge of portion E and bring the
car to a stop.
Two Levers operate the brakes (W, Figs. 97
Fig. 103.—Detail of Brake and
Controlling Levers. and 100). Cut these of the shape shown in Fig.
103, and screw one to each side wall. Then
tack a piece of cord to stick S, wrap the ends of
the cord once around, slip them through screw-eyes V screwed into the side
walls, and tie to tacks driven into levers W.
One series of brakes can now be set by pulling forward one lever, and the
other series by pulling forward the other lever. By driving a nail into stick S at
X (Fig. 100), and a nail into the bottom of the shaft, each side of stick S, the
levers will turn the stick just far enough in either direction to bring the brakes
into operation.
If there is a kitchen porch to your house, construct
The Outdoor Elevator shown in Fig. 104 to run from the ground up to that
porch. If you live in an upper story of an apartment building, your elevator
can be made to run to a greater
height, which, of course, will make
more fun.
It will save considerable work to
use the porch, because for one
thing you will not have to build an
upper platform to stand upon to
reach the elevator car when it runs
to the top, and for another thing
the supports for the guides and
cable can be fastened directly to
one of the porch posts.
Figure 105 shows a large detail of
The Guide Supports. Cross strips
A, B, and C should be 18 or 20
inches long, about 2 inches wide,
and 1 inch thick. At a distance of
about 1 inch from one end of strips
A and B screw a screw-eye into one
edge, and 8 inches from those eyes
screw a second screw-eye (D, Fig.
105). Screw-eyes with ½ inch eyes
are large enough. A dozen will cost
about 5 cents at the hardward
store. The elevator guides are
fastened to these.
Besides the screw-eyes there must
be two clothes-line pulleys for the
cable to run over. These cost 5
Fig. 104.—An Outdoor Elevator. cents apiece. Screw one pulley into
the edge of strip B, halfway
between the two screw-eyes D (E, Fig. 105), the other into an edge of strip C
at the same distance from the end that you have placed the pulley in strip B
(F, Fig. 105).
Nail strip A to the porch post as close to the ground as you can get it, strip B
to the same face of the same post, about 18 inches above the porch railing,
and strip C to the opposite face of the post at the same height as strip B. Nail
these strips securely in place.
If you cannot find a starch-box
or other small box out of
which to make
The Car, go to a grocery
store. You will be sure to find
just what you want there. It is
not likely that the grocer will
charge you anything for a
small box like this. If you have
placed screw-eyes D 8 inches
apart, as directed, the width of
the box should be a trifle less
than this measurement, but if
the box you pick up is wider
the screw-eyes can be spaced
as much farther apart as is
necessary to accommodate it.
Figure 106 shows how the box
is converted into the car.
Screw two screw-eyes into
each side of the box, one
above the other, as shown at
G, for the elevator guides to
run through, screw another Fig. 105.—Supports for Elevator Guides and
into the exact center of the Cables.
top (H) to tie the hoisting
cable to, and screw another
into the exact center of the bottom to tie the lowering cable to. Nail a narrow
strip across the open front of the car, at the bottom, to keep things from
falling out.
Get heavy wrapping-twine or stovepipe wire for
The Guides. Attach these guides to screw-eyes D in strip B, first, drop them
through screw-eyes G in the sides of the car, and then fasten to screw-eyes D
in strip A.
The Counter-balance is a large can filled with earth, sand, or small stones.
Its weight must be equal to about three times that of the empty car. Fasten
the lifting cable through holes punched in opposite sides of the can (Fig. 107).
Use a strong wrapping-twine for
Welcome to our website – the perfect destination for book lovers and
knowledge seekers. We believe that every book holds a new world,
offering opportunities for learning, discovery, and personal growth.
That’s why we are dedicated to bringing you a diverse collection of
books, ranging from classic literature and specialized publications to
self-development guides and children's books.
ebookbell.com