Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

4-1

Lecture 4
Linear Elasticity and Elastic Moduli, Introduction to Crystal
Structures
In this lecture, we will learn about the linear elastic response exhibited by most materials at
sufficiently low loads and short times. We will also learn about elastic moduli, which relate
stresses to elastic strains in specific tests.

Linear Elasticity
Recall that for elastic deformation, the object returns to original shape when you remove the
load. Most materials (not all!) will, at small strains and under the right conditions (temperature,
strain rate), exhibit elastic strains that are approximately proportional to the stress applied and
are time independent. This behavior is referred to as linear elasticity. Materials that exhibit
linear elasticity are said to obey Hooke’s Law (Figure 4.1)
σ

Materials that exhibit linear elasticity


are said to obey "Hooke's Law"
σ

ε
Figure 4.1: Linear elastic deformation in a uniaxial tension test.
As we will see, this behavior is never perfectly linear, but in many engineering applications, it is
close enough that we can treat it as if it were—this very much simplifies calculations! Because it
can be treated as if it were time-independent, it is sometimes also referred to as instantaneous
elasticity. However, we recall that all deformation is time-dependent on some scale and take this
description to mean that the deformation response of the material is sufficiently fast relative to
the time of the particular test being implemented that we can ignore it (i.e. D << 1).

Elastic Moduli
The most common way to characterize linear elasticity is by the proportionality between stress
and strain in a mechanical property test. The proportionality constant for a particular test is
known as the elastic modulus. A number of commonly used elastic moduli are defined as
follows:

Young's Modulus, E
The most commonly used elastic modulus is Young's modulus, E. Young’s modulus is simply
defined as the ratio of the normal stress s to the normal strain e along the tensile axis in a

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-2

uniaxial tensile test. This is the situation shown in Figure 4.1, where the slope of the s-e data
gives E. Consider the sample shown in Fig. 4.2.
1+ε2

1 1+ε3

1
Figure 4.2: Stresses and strains in a uniaxial tension test.
We apply a stress s3, and get a strain e3. The Young modulus is then just:
σ
E= 3 . (4.1)
ε3

Poisson's Ratio, n
We observe in Figure 4.2 that, associated with the normal strain e3 along x3, there are also normal
strains e1 and e2 along x1 and x2, respectively. For an isotorpic material, we define Poisson’s
ratio as
− ε 2 − ε1
ν= = . (4.2)
ε3 ε3
The Poisson ratio is then just the ratio of the transverse contraction to the normal strain. If the
material is anisotropic, then n will vary with azimuthal angle in the x1-x2 plane. Poisson’s ratio is
not usually thought of as a modulus as we have defined it, but by combining Eq’s 4.1 and 4.2 we
have
−$!%
!= (4.3)
&" ,
So v represents a relationship between a uniaxial stress (in this case s3) and a strain in the
transverse direction (in this case e1).

Shear Modulus, G
The shear modulus, G, is defined as the ratio of shear stress, t, to shear strain, g, for the state of
shear illustrated in Figure 4.3.

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-3

γ τ

Figure 4.3: Shear stress and shear strain

The shear modulus is then just


τ
G= . (4.4)
γ
Note that we are using engineering shear strains here (Fig. 4.3 shows simple shear).
Bulk Modulus, K
The bulk modulus, K, gives the elastic response of the material to a hydrostatic stress sH. This
response is a relative change in volume, ∆V/V, which we have seen is the dilatation, D.
#!
!= (4.5)
∆ .

This kind of loading is common in systems containing a fluid under pressure. The pressure P is
often written as a positive quantity (even though it induces compression, which by normal
definitions of stress and strain is negative). In this case the positive pressure leads to a volume
shrinkage as shown in Figure 4.4.

Figure 4.4: Hydrostatic pressure and dilatation.

And the bulk modulus relates pressure to volume change as


ΔV = 1
V K P . (4.6)

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-4

Elastic Moduli vs. Elastic Constants


It is common to call E, n, G, and K, “elastic constants” and indeed they are typically treated as
“materials properties.” By far the most common reference values for the elastic stiffness of
materials are E and n. However, treating elastic moduli as materials constants requires that the
materials be isotropic and homogeneous. Since real materials are anisotropic and
inhomogeneous, elastic moduli cannot represent true materials properties. Instead, each one
represents the behavior of a particular material in a particular test.
For example, due to the anisotropy of copper, it is possible to make pure copper tensile test
samples having Young’s moduli that vary by about a factor of 3! Clearly, the Young moduli for
those samples are specific to the internal structure of each one. There is no such thing as a
Young’s modulus for Cu that applies to all Cu samples! Yet, you can look up “Young’s modulus
of Cu” online or in reputable textbooks and find a value near E = 130 GPa listed as a “materials
property” or even as an “elastic constant” (which it definitely is not!). How can this be? As we
will see, by convention, people assume a sample with a very specific internal structure to
determine the elastic moduli for a certain material. The Young’s modulus for pure metals is
determined for a particular structure known as a random polycrystal, which we will study
shortly. As it turns out, this structure is very difficult to come by in real life, but it is a well-
defined quantity that produces an elastic modulus near the middle of the range. Thus, the
textbook value of Young’s modulus for Cu is that of a random polycrystal of Cu and is a
reasonable representative for the class of samples that could be made of pure Cu. However, it
does not apply to any other Cu sample, therefore it cannot represent a “property” that is intrinsic
to Cu in general.
As we will see, for homogenous materials, it is possible to come up with elastic constants
that would apply to all instances of that material, even if it is very anisotropic. This approach is
very important to understanding materials properties and we will develop it in the next lecture.
Meanwhile, it is important to realize that in our uniaxial-tension-test-centric world, E and n
are considered basic elastic properties and are used in lieu of real materials elastic constants
because of their convenience and simplicity. If you do not need to predict what the modulus of a
particular material might be, or if you don’t care why a particular material has a particular
modulus, you can in principle just measure the particular sample that you are interested in and
report the measured modulus values. It does not matter how anisotropic the sample is, all other
samples measured exactly the same way in exactly the same orientation will give the same
values. From this perspective, the obtained values are “materials properties,” but only for those
specific samples.
In some cases, materials actually are isotropic. As we will see, it turns out that only two
independent constants are needed to describe the elastic behavior of a linearly elastic, isotropic,
homogeneous solid. Thus, for such cases, any of the four elastic moduli presented above can be
described terms of two of the others. If E and n are known for instance, G and K can be
computed simply from
E E
G= and κ= .
2(1+ ν ) 3(1 − 2ν )
However, if the material is anisotropic, these conversions don’t apply.

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-5

Hooke’s Law in 3-D


Of course, engineering structures are usually not loaded in any of the simple configurations
shown above; rather they are subjected to complex 3-D loadings. However, Young’s modulus
and Poisson’s ratio are defined for uniaxial tests only. How can we use these simple “properties”
to describe the elastic response of a material subjected to a bi- or triaxial stress state?
If the material is isotropic, and we are in the principal stress coordinates, we can find the
relationships among stress and strain along the principal axes for biaxial or triaxial loading at a
single point using the principle of superposition.
Suppose that a state of triaxial stress exists at a point in an isotropic homogeneous solid and
is expressed in principal coordinates as shown in Fig. 4.5.
σz

σy

σx

Figure 4.5: Principal stresses in a triaxial stress state.

What are the principal strains in this case? Let’s look along the x-axis first. The strain along the
x-axis due to a stress sx imposed along the x-axis in a uniaxial test would just be

σx
εx σ =
x E
The strain along the x-axis due to a stress sy imposed along the y-axis in a uniaxial test would
just be
σy
ε x σ = −νε y = −ν .
y E
Similarly,
σz
ε x σ = −νε z = −ν .
z E
By superposition, we can write for the 3-D stress state,

εx = εx σ + εx σ + εx σ ,
x y z
or
1"
εx = #σ x − ν (σ y + σ z )$% . (4.7)
E
Similarly, we can find

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-6

1"
εy = #σ y − ν (σ x + σ z )$% , (4.8)
E
and
1"
εz = σ z − ν (σ x + σ y )$% . (4.9)
E#
Equations 4.7-4.9 constitute Hooke’s Law in 3-D.

Deformation Mechanism of Linear Elasticity


To understand the mechanical behavior of materials, we must understand the deformation
mechanisms—that is, how atoms move relative to each other to accomplish the macroscopic
shape changes that we observe. In this section, we consider how atomic bonding determines
elastic properties.
When a material is subjected to loads, those loads are distributed across the bonds between
atoms in the material. In response those bonds can get longer or shorter and change their bond
angles. We will refer to these changes in length and angle generically as bond stretching. If the
deformation is not too severe such that the atoms in the material don’t change their nearest
neighbors, then when the load is removed, the bonds will return to their original lengths and
angles and the material will return to its original undeformed shape. This type of bond stretching
is the mechanism of linear elasticity. While virtually all solids exhibit linear elastic behavior at
low enough strains (and temperatures!) metals and ceramics generally show linear elastic
behavior to very high stresses so we will focus on them in this development.
To simplify our arguments, we will start with a simple cubic single crystal; i.e., a material in
which the atoms are arranged in a 3-D cubic array with equilibrium spacing ao between atoms as
shown in Figure 4.6. We pick a pair of atoms in the solid and draw a cube of side ao around them
so that the two atoms are at the centers of opposing faces of the cube as shown by the dotted
lines. We imagine that the solid is loaded in uniaxial tension parallel to the axis of this bond.

Figure 4.6: Stretching of bonds across an atomic plane.


If the applied normal stress is s, the force F on this one bond must be F = ao2σ . If the resulting
macroscopic strain is e, then the strained length of our single bond is a = ao(1 + e). Thus, if we
know the F vs. a behavior of a single bond in the solid, we can find the macroscopic s vs. e
relation.

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-7

This is relatively straightforward. Interatomic bonds are typically described in terms of the
potential energy vs. distance curve for the two atoms. A schematic is shown in Figure 4.7.

Figure 4.7: Potential energy and force between atoms as a function of separation.
At very large separations, the potential energy is defined to be zero. As the separation is
decreased, the potential energy decreases as the bond forms, then goes through a minimum and
then increases. The very sharp increase in energy at low atomic separations is due to the forced
interpenetration of the inner electron shells of the atoms. To a good approximation, the shape of
this curve for many bonds can be described by a curve of the forma
−c b
φ (a) = n + m . (4.10)
a a
The curve shown in Fig. 4.7 is the characteristic “energy well” of the bond. Two features of
this curve may be easily measured. The depth of the well, fb, represents the energy that must be
supplied to separate the atoms to infinity. It is called the binding energy. It is proportional to the
energy of sublimation (vaporization) of the solid, which can be determined using calorimetry.
The interatomic separation, ao, of the crystal when no stress is applied is just the value of a
corresponding to the minimum potential energy. In the absence of an externally applied force,
the bond adopts this spacing, which, for this crystal, is the equilibrium lattice parameter. The
value of ao is readily measured by x-ray diffraction.
To obtain the F vs. a curve, one must only differentiate the f vs. a curve since F = df/da. A
plot of F vs. a for the same bond is also shown in Fig. 4.7. Note that F = 0 at a = ao, increases at
a slowly decreasing rate, but then bends over, goes through a maximum and decreases again as a
increases.
When we convert the F vs. a data to s vs. e following the prescription outlined above, we get
the curve shown in Figure 4.8.

a
The specific case where n = 6 and m = 12 is known as the Lennard-Jones potential; a very common form used in
materials science and many other disciplines.

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-8

Figure 4.8: Expected stress vs. strain behavior for the stretching of a single bond
Note that the s vs. e curve has the same shape as the F vs. a curve, only scaled linearly and
shifted on the a axis. We are interested in the slope of the s vs. e curve at ao. This represents
Young’s modulus for a single bond and is given by
1 dF 1 d 2φ
Slope = E = = . (4.11)
ao da ao ao da2 a
o

There are three interesting and fundamental things to note at this point. First, the s vs. e
curve is never truly linear—the slope decreases continuously with increasing e. That is, there is
no true linearly-elastic regime! However, it is close enough to linear over a small range (e < 0.01
or so) for many common calculations. In practical terms, we are able to ignore the non-linearity
in most metals and ceramics because the stress required to generate strains beyond e = .001 or so
becomes so high that plastic deformation or fracture occurs firstb. However, if the material is
exceptionally strong (or compliant) it might get into a regime where the non-linearity is
important. In addition, in any case where high accuracy is required, the non-linearity of elastic
stiffness must be accounted for. Second, the stiffness of the material is proportional to the second
derivative, i.e. the curvature, of the atomic bond potential. This is why, for a given ao, a stronger
bond (as determined by fb) is associated with a stiffer material. For example, since materials
with higher fb have higher melting temperature, Tm, we can see that materials with higher Tm are
also, on average, stiffer. Third, similarly, for a given bond strength the stiffness increases as ao
decreases. In fact, using Equation 4.10, it is possible to show that 𝐸!"#$ ∝ 1⁄𝑎%& for bonds that
can be treated as forming by Coulomb attraction at a fixed charge level (strictly speaking, ionic
bonds, but applies to others as well.)
For a high modulus, we want to find materials with high binding energies and small
interatomic spacings. The binding energy depends strongly on the type of bonding between
atoms. In general binding energies are highest for covalent solids and somewhat smaller for ionic
and metallic solids.

Ebond vs. E
How do we get from the modulus of a single bond, Ebond, to the elastic modulus of a bulk
material? In general, this is very difficult! Consider the simple square 2-D crystal shown in
Figure 4.9a. Obviously, if we pull it along the lines of the primary first-nearest-neighbor (1NN)
bonds, we will get a different result than if we pull at an angle q to those bonds. In fact, we might

b
This is because metal and ceramic crystals are very stiff!

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-9

expect to see E and n vary as shown in Figure 4.9b. The answer is further complicated by
second-nearest neighbor (2NN) bonds, shown by the dotted lines in Figure 4.9c. If the 2NN
bonds are significant, then they will contribute to the Poisson ratio for tension along the primary
1NN bonds. It is evident that, even if the shapes of the primary interatomic bond potentials are
well known, the Young modulus of the material will depend on the details of the arrangements of
the atoms and all of the 2NN, 3NN, etc. interactions.
(a) (b) (c)

Young's modulus
𝜃 0˚ 45˚ 90˚ 𝜃
Figure 4.9: Stiffness in various directions in a 2-D crystal.
We will address this problem in three ways in this course:
• We will learn about the elastic constants of single crystals. This is the simplest case
for which we can obtain a complete elastic description of an anisotropic linear-elastic
material. Such materials can have homogeneous stress distributions that are very
anisotropic.
• We will learn about the average elastic constants of materials consisting of many
differently oriented anisotropic components (individual single crystals, phases in
different morphologies, molecules, etc.) and for which the orientation distribution is
not random. Such materials are macroscopically anisotropic but have very
inhomogeneous stress distributions.
• We will learn some of the conditions under which materials are macroscopically
isotropic.
In the next lecture, we will learn that simple homogeneous single crystals are generally
anisotropic. In other solids, the types of bonding may change with direction. For example, a
polymer can be formed from long linear chain molecules formed with primary covalent carbon-
carbon bonds all aligned in one direction with only secondary (weak) bonding between chains.
Such a material will be very stiff in the direction of the carbon “backbone” and very compliant in
the perpendicular direction. Only homogeneous amorphous materials, such as silicate glasses,
are generally isotropic both microscopicallyc and macroscopically.

Introduction to Crystallography
A surprisingly large fraction of materials and the vast majority of engineering materials are
essentially crystalline. In addition, single crystals represent one of the rare cases in which we can
understand the properties of materials as the level of basic physics, without the need to average
c
At least down to the nm scale. On the scale of atoms (sub nm), the structure itself is inhomogeneous (atoms and the
gaps between them).

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-10

over many different structures. Thus, we will dedicate significant effort to understanding the
mechanical properties of crystals. In this section, we briefly review how crystal structures are
defined as well as crystal defects.

Crystal Structure
A crystal is a three-dimensional periodic array of atoms. We describe a crystal structure in terms
of a crystal lattice and an atom basis. A lattice is a periodic 3D arrangement of points in space.
Each lattice point has identical surroundings. There are 14 possible crystal lattices, the Bravais
lattices, as shown in Figure 4.10.

Figure 4.10: The 14 Bravais latticesd

d
From B.D. Cullity and S.R. Stock, Elements of X-ray Diffraction, 3rd Ed. Prentice Hall 2001.

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-11

One can think of constructing a lattice by translating an initial point by any integer
combination of a set of lattice translation vectors. The primitive translation vectors are the
shortest set of 3 vectors which can be combined as integer multiples to describe every lattice
point. That is, in a three-dimensional lattice, all lattice points must be described as end points of
the vector
r = la1+ma2+na3 (4.12)
where a1, a2, and a3 are the independent primitive lattice translation vectors and l, m, and n are
(positive and negative) integers. We see that we can reach any lattice point from any other by a
vector r as described above.
Since a lattice repeats in a regular fashion, we can choose a small element of the lattice,
called a unit cell, which can be duplicated and translated to generate the entire lattice. Unit cells
can be thought of as identical units that can be stacked up together to fill space. We can always
define the primitive unit cell as the parallelopiped with sides defined by the primitive lattice
translation vectors for a particular lattice. The primitive unit cell contains only one lattice point,
but may not necessarily contain all symmetry elements of the infinite lattice. It is common to
define a conventional unit cell as the smallest unit cell which contains all the symmetry
elements of the infinite lattice. This unit cell may contain more than one lattice point. We can
understand this as follows:
A symmetry operation is one which leaves the lattice unchanged. Lattice translations are
symmetry operations. Another important type of symmetry operation is rotation. To illustrate this
type of symmetry, view a lattice along a certain direction or axis. Now rotate the lattice through a
full circle about that axis. If during that rotation, the lattice coincides with itself n times, we call
that axis an n-fold rotational symmetry axis, or we say that the lattice has n-fold rotational
symmetry about that axis. Figure 4.11a shows the primitive unit cell from a hexagonal crystal
from Figure 4.10. We see that if we rotate the crystal by 180˚ around the c-axis, the crystal is
unchanged. Thus, n = 2 and the c-axis is a 2-fold symmetry axis. We can also find 2-fold
symmetry axes in this unit cell perpendicular to the c-axis along lines parallel to the diagonals of
the top and bottom faces as shown in Figure 2b.

(a) (b) (c) (d)

Figure 4.11: For hexagonal crystals: (a) Primitive unit cell, (b) symmetry axes in primitive unit
cell (3 two-fold axes), (c) conventional unit cell showing relationship with primitive unit cell,
and (d) symmetry axes in conventional hexagonal unit cell (1 six-fold axis and 6 two-fold axes).
(Ovals indicate two-fold axes and hexagon indicates six-fold axis.)

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-12

However, if we start to fill space with this primitive unit cell, we see that the lattice that is
generated has much more symmetry. Figure 4.11c shows the conventional unit cell that is used
for a hexagonal crystal. It has 6-fold symmetry along the c-axis and 6 different 2-fold axes
perpendicular to the c-axis as shown in Figure 4.11d. There is quite a difference between the
symmetries of the primitive and conventional unit cells! Because the symmetry of crystals is
important, we virtually always use the conventional unit cell for any crystal.
A basis is a grouping of atoms about each lattice point. The complete crystal structure is
created by placing the basis group of atoms at every lattice point. A few examples are shown in
Figure 4.12.

Figure 4.12: Examples of crystal structures decomposed into lattices and bases.
While there are many crystal structures, a surprising number of engineering materials are
based on just three structures, face-centered cubic (FCC), body centered cubic (BCC) and
hexagonal close packed (HCP). FCC and BCC unit cells are shown in Figure 4.10. The HCP
lattice is a simple hexagonal lattice with 2 atoms per lattice point as shown in Figure 4.13. For
this class, you should become familiar with these crystal structures. This will provide you with a
solid foundation for understanding more complex crystal structures if needed.

Figure 4.13: Conventional (left) and primitive (right) unit cells for the HCP structure.

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-13

Miller Notation
Miller notation is a consistent way to describe directions and planes in crystalline materials.

Directions
The procedure for determining the Miller indices of a direction are:
1. Find the coefficients of a vector parallel to the direction of interest in unit cell
coordinates.
We consider the line that intercepts the unit cell at 1 ½ 0 in Figure 4.14. (Recall that
unit cell coordinates do not have to be orthogonal nor the same length on all axes!)
2. Scale these three numbers to the smallest integers having the same ratio.
In our example, 1 ½ 0 becomes 2 1 0.
3. Enclose in appropriate brackets.
a. Use square brackets to indicate a specific direction. The specific direction in
the example in Fig. 4.14 is then [210]. direction as shown in Figure 4.12b. The
Miller indices of several other specific directions are shown in Fig. 4.14 as
well. Note that overbars are used to indicate negative directions, e.g. [2(1(0].
b. Use angle brackets to specify a family of crystallographically equivalent
directions. For example, if the unit cell were cubic, then all of the directions
where the indices were some combination of 2, 1, and 0 would be
crystallographically equivalent and we would say that they are all <210>
directions. In Figure 4.14, this would include [210], [120], and [12(0]. Many
other <210> directions could be drawn. However, if a and b are different
lengths, then [210] and [120] are not crystallographically equivalent.

Figure 4.14: Miller indices of various directions in a crystal.

Planes
The procedure for determining the Miller indices of a plane are:
1. Find intercepts of the plane with unit cell axes.

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-14

For example, in Figure 4.15, the intercepts are ¼ ½ 1


2. Take reciprocals
¼, ½, 1 becomes 4 2 1. If a plane is parallel to an axis, its fractional intercept on that axis
is infinity, and the Miller index is zero.
3. Find the lowest common denominator
In the example shown, no change
4. Use appropriate brackets:
a. Parentheses to specify a specific plane, e.g., (4 2 1)
b. Curly brackets to indicate a family of crystallographically equivalent planes, e.g.
{421}.

Figure 4.15: Miller indices of a plane

Miller-Bravis Indices for Hexagonal Cystals


A conventional unit cell for a hexagonal crystal is shown in Figure 4.16.

Figure 4.16: Three crystallographically equivalent planes in a hexagonal crystal.

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-15

If we use the lattice translation vectors of the primitive hexagonal unit cell to find the Miller
indices of the planes on the faces of the conventional hexagonal unit cell as shown, we will find:
Plane A: (100)
Plane B: (010)
Plane C: (11(0)
The planes (100), (010) and (11(0) are actually crystallographically equivalent, yet using our
standard process for Miller indices, their Miller indices have different forms. Furthermore, the
plane perpendicular to the c axis is (001), even though it is not crystallographically equivalent to
(100) and (010). This makes it difficult to describe which planes and directions are
crystallographically equivalent using this notation! To solve this problem, a 4-axis system is used
to describe the hexagonal structure. In this system, the vectors a1, a2, a3, and a4 (also written as
a1, a2, a3, and c) as shown in Figure 4.17, are used.

Figure 4.17: The four-axis system used to describe hexagonal crystals.


If we apply the procedure for Miller indices described above, but using all four indices, we find
that the indices for planes A, B, and C are then given by
Plane A: (101(0)
Plane B: (011(0)
Plane C: (11(00)
In this notation it is clear that planes A, B, and C are all members of the {11(00} family.
Furthermore, the plane perpendicular to the c axis is now (0001), so it is clear that it belongs to a
different family.
Because we are trying to use 4 numbers to describe a 3-dimensional quantity, our solutions
following the procedures above are indeterminate. To restrict our choices to one unique result,
we apply the additional rule that the first three indices must sum to zero. That is, a1 + a2 + a3 = 0.
The same 4-index approach applies to directions. An example showing that the close-packed
directions in HCP belong to the < 21 ( 1(0 > family is provided in Figure 4. 16.

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker


4-16

- - -
[1 1 2 0] [2 1 1 0] The close packed directions are the
a
~3 - -
<2 1 1 0> type.

- a2 [1- 2 1- 0]
[1 2 1 0] ~

a
- - ~1 -
[2 1 1 0] [1 1 2 0]

Figure 4.16: The close-packed directions in HCP


Now that we understand the physical basis for linear elasticity as well as the basics of
crystallography, we are ready to proceed to study elastic constants in single crystals.

Cornell MS&E 2610 Lecture 4 Ó2024 S.P. Baker

You might also like