0% found this document useful (0 votes)
37 views

Chapter 6

The document discusses elements of grain boundaries in materials. It describes the dislocation model of small-angle grain boundaries, where boundaries can be modeled as arrays of dislocations. It also discusses the five degrees of freedom of a grain boundary, including tilt and twist configurations between crystals. Additionally, it examines the stress field generated by grain boundaries and how grain boundary energy relates to factors like misorientation angle and boundary structure.

Uploaded by

蘇翊愷
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
37 views

Chapter 6

The document discusses elements of grain boundaries in materials. It describes the dislocation model of small-angle grain boundaries, where boundaries can be modeled as arrays of dislocations. It also discusses the five degrees of freedom of a grain boundary, including tilt and twist configurations between crystals. Additionally, it examines the stress field generated by grain boundaries and how grain boundary energy relates to factors like misorientation angle and boundary structure.

Uploaded by

蘇翊愷
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 65

物理冶金 Chapter 6 魏茂國

Elements of Grain Boundaries


 Grain boundaries
 Dislocation model of a small-angle grain boundary
 The five degrees of freedom of a grain boundary
 The stress field of a grain boundary
 Grain-boundary energy
 Low-energy dislocation structures, LEDS
 Dynamic recovery
 Surface tension of the grain boundary
 Boundaries between crystals of different phases
 The grain size
 The effect of grain boundaries on mechanical properties: Hall-Petch relation
 Grain size effects in nanocrystalline materials
 Coincidence site boundaries
 The density of coincidence sites
1
 The Ranganathan relations
物理冶金 Chapter 6 魏茂國

Elements of Grain Boundaries


 Examples involving twist boundaries
 Tilt boundaries

2
物理冶金 Grain Boundaries 魏茂國

 Grain boundaries (晶界)


- Fig. 6.1 shows the crystalline structure of a typical polycrystalline (many crystal)
specimen. The average diameter of the crystals is ~0.05 mm, and each crystal is
seen to be separated from its neighbors by dark lines, the grain boundaries.
- The grain boundaries appear to have an appreciable width, but this is only because
the surface has been etched in an acid solution in order to reveal their presence. A
highly polished polycrystalline specimen of a pure metal that has not been etched
will appear entirely “white” under the microscope; that is, it will not show grain
boundaries.

Fig. 6.1 A polycrystalline zirconium specimen


photographed with polarized light. In this photograph,
individual crystals can be distinguished by a difference
in shading, as well as by the thin dark lines representing
grain boundaries. 350 . Note that most of the triple 3
junctions form 120 angles.
物理冶金 Grain Boundaries 魏茂國

- At low temperatures the grain boundaries are quite strong and do not weaken metals.
In fact, heavily strained pure metals, and most alloys, fail at low temperatures by
cracks that pass through the crystals and not the boundaries. Fractures of this type are
called transgranular fractures (穿晶斷裂).
- At high temperatures and slow strain rates, the grain boundaries lose their strength
more rapidly than do the crystals, with the result that fractures no longer traverse the
crystals but run along the grain boundaries. Fractures of the latter type are called
intergranular fractures (沿晶斷裂).

4
Materials Science and Engineering Brittle Fracture 魏茂國

 Transgranular fracture (穿晶斷裂)


- For most brittle crystalline materials, crack
propagation corresponds to the successive and
repeated breaking of atomic bonds along specific
crystallographic planes (Fig. 8.6a); such a process is
termed cleavage. This type of fracture is said to be
transgranular (or transcrystalline), because the
fracture cracks pass through the grains.
- Macroscopically, the fracture surface may have a
grainy or faceted texture (Fig. 8.3b),
as a result of changes in orientation
of the cleavage planes from grain to
grain.
Figure 8.6 (a) Schematic cross-section profile showing
crack propagation through the interior of grains for
transgranular fracture. (b) Scanning electron fractograph
of ductile cast iron showing a transgranular fracture 5
surface.
Materials Science and Engineering Brittle Fracture 魏茂國

 Intergranular fracture (沿晶斷裂)


- In some alloys, crack propagation is along grain
boundaries (Fig. 8.7a); this fracture is termed
intergranular.
- Fig. 8.7b is a scanning electron micrograph showing a
typical intergranular fracture. This type of fracture
normally results subsequent to the occurrence of
processes that weaken or embrittle grain boundary
regions.

Figure 8.7 (a) Schematic cross-section profile showing


crack propagation along grain boundaries for intergranular
fracture. (b) Scanning electron fractograph showing an 6
intergranular fracture surface.
物理冶金 Dislocation Model of a Small-Angle Grain Boundary 魏茂國

 Dislocation model of a small-angle grain boundary


- In 1940, both Bragg and Burgers introduced the idea that boundaries between
crystals of the same structure might to be considered as arrays of dislocations.
- Fig. 6.2A shows an example of such an elementary boundary in a simple cubic
lattice in which the crystal on the right is rotated with respect to that on the left
about the [100] direction (the direction perpendicular to the plane of the paper). The
line between points a and b corresponds to the boundary.
- The lattice on both sides of the boundary is seen to be inclines downward toward
the boundary, with the result that certain nearly vertical lattice planes of both
crystals terminate at the boundary as positive-edge dislocations.
- The greater the angular rotation of one crystal relative to the other, the greater is the
inclination of the planes that terminate as dislocations at the boundary, and the
closer is the spacing of the dislocations in the vertical boundary.

7
物理冶金 Dislocation Model of a Small-Angle Grain Boundary 魏茂國

- Referring to Fig. 6.2B, it can be seen that


sin  / 2   b / 2d (6.1)
where b is the Burgers vector of a dislocation in the boundary and d is the spacing
between the dislocations.
- If the angle of rotation of the crystal structure across the boundary is assumed to be
small, then sin(/2) may be replaced by /2, and the equation relating the angle of tilt
across a boundary composed of simple edge dislocations becomes
 b/d (6.2)
- Small-angle boundaries commonly have less (A) (B)
than 10-15 of misorientation and can be
described by a dislocation model such as the
one shown in Fig. 6.2.

Fig. 6.2 (A) Dislocation model of a small-angle grain boundary.


(B) The geometrical relationship between , the angle of tilt, 8
and d, the spacing between the dislocations.
物理冶金 Dislocation Model of a Small-Angle Grain Boundary 魏茂國

- Fig. 6.3 shows a photograph of low-angle boundaries in a magnesium crystal. Low-


angle boundaries may also be observed by transmission electron microscopy; an
example is shown in Fig. 6.4.

Fig. 6.3 Low-angle boundaries in a magnesium Fig. 6.4 A low-angle boundary in a copper-13.2
specimen. The rows of etch pits correspond to atomic percent aluminum specimen deformed
positions where dislocations intersect the 0.7% in tension as observed in the transmission
surface. electron microscope. This boundary has both a
tilt and a twist character. Magnification: 32,000.
9
物理冶金 Five Degrees of Freedom of a Grain Boundary 魏茂國

 Five degrees of freedom of a grain boundary


- There are 5 degrees of freedom of a grain boundary, as illustrated in Fig. 6.5.
- In Fig. 6.5A, note that it is symmetrically positioned between the 2 crystals that are
tilted with respect to each other about a horizontal axis that runs out of the plane of
the paper.
- In Fig. 6.5B, a symmetrical tilt boundary with a vertical tilt axis is shown.
- Fig. 6.5C shows a basically different way of orienting 2 crystals relative to each
other. They are rotated about an axis normal to the boundary, instead of about an
axis lying in the boundary, as in the tilt boundaries in Figs. 6.5A and 6.5B. This is
called a twist boundary.
- In addition, the boundary itself has 2 degrees of freedom, as indicated in Figs. 6.5D
and 6.5E.
- In Fig. 6.5, there are 3 ways that we can tilt or twist one crystal to another, and 2
ways that we can align the boundary between the crystals.

10
物理冶金 Five Degrees of Freedom of a Grain Boundary 魏茂國

(A) (B) (C)

(D) (E)

11
Fig. 6.5 The 5 degrees of freedom of a grain boundary.
物理冶金 Stress Field of a Grain Boundary 魏茂國

 雙曲函數
在數學中,雙曲函數是一類與常見的三角函數(圓函數)類似的函數。最基本的
雙曲函數是雙曲正弦函數和雙曲餘弦函數,從它們可以導出雙曲正切函數等
,其推導也類似於三角函數的推導。雙曲函數的反函數稱為反雙曲函數。雙曲
函數的定義域是實數,其自變量的值叫做雙曲角。
e x  e x
sinh x 
2
e x  e x
cosh x 
2
sinh x e x  e  x
tanh x   x
cosh x e  e  x
1 cosh x e x  e  x
coth x    x
tanh x sinh x e  e  x
1 2
sec hx   x
cosh x e  e  x
1 2
csc hx   x
sinh x e  e  x 12
Source: https://zh.wikipedia.org/zh-tw/%E5%8F%8C%E6%9B%B2%E5%87%BD%E6%95%B0
物理冶金 Stress Field of a Grain Boundary 魏茂國

 Stress field of a grain boundary


- The boundaries do not possess long-range stress fields.
- In Fig. 6.6, a slip plane is shown that passes through one of the boundary
dislocations. Now consider the shear stress on this slip plane at a distance, x, to the
right of the boundary.
- The contribution to the shear stress at this point due to a specific boundary
dislocation, such as that at point a, may be
obtained using the shear-stress equation of
Eqs. 4.9 or
b x x 2  y 2 
 xy  
2 1  v  x 2
y 
2 2
(6.3)

where xy is the shear stress on the slip plane,


 the shear modulus, and v Poisson’s ratio.

Fig. 6.6 A diagram defining the parameters used in computing the


stress due to a simple tilt boundary. The y scale is in units of d, the 13
distance between adjacent boundary dislocations, and n in the number.
物理冶金 Stress Field of a Grain Boundary 魏茂國

- It is convenient to assume that y = -nd, where d is the distance between adjacent tilt
boundary dislocations and n is an ordinal number defining the positions of the
dislocations in the boundary. Thus, for the dislocation at point a in Fig. 6.6, n = 2 and
Eq. 6.3 becomes
 xy 
b


x x 2  4d 2 

2 1  v  x 2  4d 2 
2 (6.4)

- The total shear stress at a distance x from the tilt boundary is now obtained by
summing the stress contributions from all the dislocations in the boundary.
b x[ x 2   nd  ]
n   2
 xy  
2 1  v  n   [ x 2   nd 2 ]2
(6.5)

The solution to this equation may be found in standard texts on dislocation theory
b x
 xy   2
21  v  d [sinh 2 (x / d )]
(6.6)

where xy is the shear stress on the slip plane,  the shear modulus, x the distance on
the slip plane from the boundary, v Poisson’s ratio, and d the distance between edge
14
dislocations in the boundary.
物理冶金 Stress Field of a Grain Boundary 魏茂國

- A plot of xy as a function of x is shown in Fig. 6.7A in which the metal is assumed to
be iron, with  = 86 GPa, v = 0.3, and b = 0.248 nm. It is also assumed that the
spacing between tilt boundary dislocations, d, is 22b.
Note that the shear stress due to the boundary falls very rapidly with increasing x. A
second curve is drawn that shows the magnitude of the shear stress due to a single
edge dislocation located at the intersection of the boundary with the slip plane.

(A) (B)
Fig. 6.7 (A) The stress, xy, due to a tilt boundary and due to a single edge dislocation as functions of
15the
distance measured in Burgers vectors. (B) Same as in (A) but at an expanded stress d = 22b.
物理冶金 Stress Field of a Grain Boundary 魏茂國

- While the shear stress due to a single dislocation exceeds the critical resolved shear
stress (CRSS) at all values of x in this diagram, note that the boundary stress equals
the CRSS at a distance of only ~25b. Further note that the boundary stress is
negligible when x is greater than ~50b.

16
物理冶金 Grain Boundary Energy 魏茂國

 Grain boundary energy


- It is generally true that any dislocation, no matter what is orientation, possesses a
strain energy.
- Since a grain boundary can consist of an array of dislocations, grain boundaries
should have strain energies. Its energy is normally expressed in terms of an energy
per unit area (J/m2).
- Following the derivation in Hirth and Lothe, it will first be assumed that there are 2
infinitely long parallel tilt boundaries, one composed of positive edge dislocations
and the other of negative edge dislocations, which are arranged such that for each
positive edge dislocation in one boundary there is a negative edge dislocation on its
slip plane in the other boundary. If the distance between these tilt boundaries is very
large, the specific energy of formation of the pair of boundaries should be just twice
that needed for the formation of one of the boundaries.

17
物理冶金 Grain Boundary Energy 魏茂國

- Suppose that there is a positive and negative pair of edge dislocations on the same
slip plane, and that the pair is brought together until their separation is r0 = b/,
where  is the factor that accounts for the strain energy at the dislocation cores. At
this distance the core energies of the 2 dislocations should cancel each other.
Consequently, on separating the dislocations from this position, the work of
separation should equal the interaction energy of the pair.
- Next consider that the positive edge dislocation lies in an infinitely long tilt boundary
and that the negative edge dislocation is a single dislocation. The force (per unit
length) attracting the negative edge dislocation toward the positive boundary is xyb.
It may be deduced that the energy per unit length of a dislocation in either boundary
equals one half the total interaction energy obtained by the following equation:
1 
wbd    bdx
r0 xy
(6.7)
2
where xy is obtained fro Eq. 6.6.
b x
 xy  
21  v  d 2 [sinh 2 (x / d )]
(6.6) 18
物理冶金 Grain Boundary Energy 魏茂國

- Now if we let  = x/d and 0 = b/d, Eq. 6.7 becomes:


b 2  d
wbd  
4 1  v d sinh 
 2 0
(6.8)

- Multiplying the solution to this equation by 1/d, the number of dislocations per unit
area, yields b, the energy per unit area of the boundary.
b 2
 b  wbd / d  [ 0 coth  0  ln(2 sinh  0 )] (6.9)
4 1  v 
- If the use of Eq. 6.9 is limited to tilt boundaries with a tilt angle smaller than a few
degrees, then by Eq. 6.2, we may take  = b/d where b is the Burgers vector and d the
separation between a pair of adjacent boundary dislocations. In this case, 0 will be
small and Eq. 6.9 may be written:
b
b   [ln( / 2 )  ln   1]
4 1  v 
(6.10)

where b is the energy per unit area of the boundary,  the shear modulus, b the
Burgers vector,  the tilt angle of the boundary,  a factor accounting for the
19
dislocation core energy, and v Poisson’s ratio. This is the Shockley-Read equation.
物理冶金 Grain Boundary Energy 魏茂國

- In Fig. 6.8, the solid line represents the Shockley-Read equation, and the data points
are experimental values.
- It is interesting to compare a plot of the small-angle solution, Eq. 6.10, with that of
the large-angle solution, Eq. 6.9. This is done in Fig. 6.9, where it is again assumed
that the metal is iron, with  = 86 GPa, b = 0.248 nm, v = 0.3, and  = 4. Note that,
under these conditions, the 2 curves do not appear to deviate significantly for angles
less than 0.15 radians (8.6).

Fig. 6.8 Relative grain-boundary energy


as a function of the angle of mismatch
between the crystals bordering the
boundary. Solid-line theoretical curve; 20
dots experimental data of Dunn for silicon iron.
物理冶金 Grain Boundary Energy 魏茂國

- A significant factor in the b against  equations is the size of the lower integration
limit, r0 = b/, where  is selected to account for the core energy. That  is an
important factor is clearly shown in Fig. 6.10, where 3 plots of b against ,
corresponding to  values of 1, 2, and 4 respectively, made using the large-angle
equation (Eq. 6.9), re given.

Fig. 6.9 The surface energy of a tilt boundary, b, as a Fig. 6.10 The effect of  on the b against
function of its angle of tilt, , as obtained with the  curves.
small-angle and large-angle equations.
21
物理冶金 Low-Energy Dislocation Structures, LEDS 魏茂國

 Low-energy dislocation structures, LEDS


- A tilt boundary does not possess a long-range shear stress field.
- In Fig. 6.11, the magnitudes of the contributions to xy of the dislocations in the
boundary are plotted against n, the ordinal number of the dislocations, over the
interval from n = -50 and n = +50.
Note that large positive contributions come from dislocations with small n values,
however; as n increases either positively or negatively, the contributions become
negative.
The result is that the net shear stress
xy at point x becomes very small
when x lies at any reasonable
distance from the tilt boundary.

Fig. 6.11 The magnitude of the shear-stress


contribution, at point x on the slip plane, of a
dislocation in the tilt boundary as a function of its 22
location in the boundary.
物理冶金 Low-Energy Dislocation Structures, LEDS 魏茂國

- There remains to be considered the 2 normal stress components of the boundary


stress at point x, namely xx and yy. Because of the anti-symmetrical nature of the
equations for xx and yy (see Eq. 4.9), the net normal stress contributions from all the
dislocations in a tilt boundary sum up to zero for both of the boundary stresses.
- Fig. 6.12 shows that as d decreases, the energy per dislocation also decreases,
indicating that there is a driving force that attracts the edge dislocations to a tilt
boundary, where we is the energy per unit length of a random dislocation.
- The tilt boundary considered here is
only one of a very large number of
dislocation arrays that possess the
property of having a low strain
energy. Kuhlmann-Wilsdorf has
proposed that these be known as low-
energy dislocation structures, or LEDS.

Fig. 6.12 The variation of wgb/we with the spacing 23


between the dislocations in a tilt boundary.
物理冶金 Low-Energy Dislocation Structures, LEDS 魏茂國

- Grain and subgrain boundaries normally fall into the LEDS category.
- One of the earliest forms of LEDS was proposed by G. I. Taylor in 1934. This
consists of an equilibrium array of alternating rows of positive and negative edge
dislocations arranged so that the 4 nearest neighbors of a given dislocation have the
sign opposite to that of the given dislocation (Fig. 6.13).
- Several other LEDS are shown in Fig. 6.14. These include a simple dipolar mat in
Fig. 6.14A that may be formed by the locking of dislocations of opposite sign moving
on adjacent slip planes because of the screening interaction of the stress fields of
dislocation of opposite sign.
- Fig. 6.14B shows another simple form
of a LEDS. This is a dipolar wall formed
by kinking on one of the crystallographic
slip planes of a crystal.

24
Fig. 6.13 The LEDS known as the Taylor lattice.
物理冶金 Low-Energy Dislocation Structures, LEDS 魏茂國

- During plastic deformation there is a


tendency to form cells with a low internal (A)
dislocation density and boundaries between (B)

the cells composed of dislocation tangles


(Fig. 6.15).
A typical example of this type of
microstructure is shown in the electron
micrograph of Fig. 6.15B. The driving
Fig. 6.14 (A) A dipolar mat that can form as a result
force for the formation of this structure is of the interaction between dislocations of opposite
the strain energy decrease associated with sign moving on a pair of adjacent and parallel slip
planes. (B) When a kink band forms in a crystal, a
the formation of the tangles. dipolar array of edge dislocations of a different type
- There exists a significant empirical is created.

relationship between the cell size and the dislocation density,



 (6.11)

where  is the average cell diameter,  a constant, and  the dislocation density.25
物理冶金 Low-Energy Dislocation Structures, LEDS 魏茂國

(A) 9% strain (B) 26% strain


Fig. 6.15 These transmission electron micrographs illustrate dynamic recovery in nickel deformed at
77 K. Note that even at this low temperature there is a definite tendency to form a cell structure that
increases at high strains.

26
物理冶金 Dynamic Recovery 魏茂國

 Recovery (回復)
- During recovery, some of the stored internal strain energy is relieved by virtue of
dislocation motion, as a result of enhanced atomic diffusion at the elevated
temperature.
- There is some reduction in the number of dislocations, and dislocation
configuration are produced having low strain energies. Physical properties such as
electrical and thermal conductivities and the like are recovered to their precold-
worked states.
- However, even after recovery is complete, the grains are still in a relatively high
strain energy state.

27
物理冶金 Dynamic Recovery 魏茂國

 Dynamic recovery (動態回復)


- As may be seen on annealing, the basic effect of high temperature recovery is the
movement of the dislocations resulting from plastic deformation into subgrain or
cell boundaries. This process can actually start during plastic deformation. When
this happens, the metal is said to undergo dynamic recovery.
- In Fig. 6.15, note that for small strains (9%) the dislocation arrangement tends to be
roughly uniform. However, when the strain becomes large (26%) a definite trend
toward a cell structure becomes clearly evident.

(A) 9% strain (B) 26% strain 28


Fig. 6.15
物理冶金 Dynamic Recovery 魏茂國

- At more elevated temperatures, the effects of dynamic recovery naturally become


stronger because the mobility of the dislocations increases with increasing
temperature. As a result, the cells tend to form at smaller strains, the cell walls
(LEDS) become thinner and much more sharply defined, and the cell size becomes
larger. Dynamic recovery is therefore often a strong factor in the deformation of
metals under hot working conditions.
- Dynamic recovery thus tends to lower the effective rate of work hardening.
- Dynamic recovery occurs most strongly in metals of high stacking-fault energies and
is not readily observed in metals of very low stacking-fault energy.
- The correspondence between the ability of a metal to undergo dynamic recovery and
the magnitude of its stacking-fault energy strongly suggests that the primary
mechanism involve in dynamic recovery is thermally activated cross-slip.

29
物理冶金 Surface Tension of the Grain Boundary 魏茂國

 Surface tension of the grain boundary


- The surface energy (G) of a grain boundary has the units of erg/cm2 or J/m2, where
1 erg/cm2 = 10-3 J/m2.
J
G  2 (6.12)
m
N m N
G  2  (6.13)
m m
But the units N/m are those of a surface tension.
- The experimental data of Fig. 6.8
show that the grain-boundary
surface tension is an increasing
function of the angle of mismatch
between grains to an angle of ~20,
and then it is essentially constant
for all larger angles.

30
Fig. 6.8
物理冶金 Surface Tension of the Grain Boundary 魏茂國

- Measurements have been made on copper, silver, and gold which show that their
surface tensions of free surfaces (external surfaces) are of the order of 1.2 to 1.8 J/m.
It has been determined that the surface tensions of large-angle grain boundaries equal
approximately one-third of the free surface tensions, and are therefore of the order of
0.3 to 0.5 J/m.
- Let the 3 lines of Fig. 6.17 represent grain boundaries that lie perpendicular to the
plane of the paper and meet in a line the projection of which is o. The 3 vectors a, b,
and c originating at point o represent, by their directions and magnitudes, the surface
tensions of the 3 boundaries. If these 3 force vectors are in static equilibrium, then the
following relationship must be true:
a b c
  (6.14)
sin a sin b sin c
where a, b, and c are the dihedral angles
(由二個平面所構成的夾角) between boundaries.

31
Fig. 6.17 The grain-boundary surface tensions at a junction of 3 crystals.
物理冶金 The structure of Metals 魏茂國

- Since crystal boundaries are regions of misfit or disorder between crystals, it is to be


expected that atom movements across and along boundaries should occur quite easily.
- Grain-boundary movements occur only if the energy of the metal as a whole is
lowered by a greater movement of atoms in one direction than in the other.
One way that the energy of a specimen can be lowered by the motion of a grain
boundary occurs when it moves into a deformed crystal, leaving behind a strain-free
crystal.
- A metal can approach a more stable state by reducing its grain-boundary area.
There are 2 ways in which this may be achieved. First, boundaries may move s as to
straighten out sharply curved regions; and, second, they may move in such a way that
some crystals are caused to disappear, while others grow in size. The latter
phenomenon, which results in a decrease in the total number of grains, is called grain
growth.
- If a metal is heated at a sufficiently high temperature for a long enough time, the
equilibrium relationship between the surface tensions and the dihedral angles can
32
actually be observed.
Materials Science and Engineering Grain Growth 魏茂國

 Grain growth (晶粒成長)


- After recrystallization is complete, the strain-free grains will continue to grow if
the metal specimen is left at the elevated temperature (Fig. 7.21d-f); this
phenomenon is called grain growth.
- An energy associated with grain boundaries is the driving force for grain growth.
As grains increase in size, the total boundary area decreases, yielding an attendant
reduction in the total energy.
- Grain growth occurs by the migration of grain
boundaries. The average grain size increases with
time. Boundary motion is just the short-range
diffusion of atoms from one side of the boundary
to the other (Fig. 7.24).

Figure 7.22 Schematic representation of grain growth 33


via atomic diffusion.
Materials Science and Engineering Grain Growth 魏茂國

- For many polycrystalline materials, the grain diameter d varies with time t according
to the relationship
d n  d 0n  Kt
d0: initial grain diameter at t = 0, K: time-independent constant,
n: time-independent constant.
The value of n is generally equal to or greater than 2.
- The dependence of grain size on time and temperature is demonstrated in Fig. 7.25.
At lower temperatures the curves are linear. Furthermore, grain growth proceeds
more rapidly as temperature increases.
This is explained by the enhancement
of diffusion rate with rising
temperature.

Figure 7.25 The logarithm of grain


diameter versus the logarithm of time for 34
grain growth in brass at several temperatures.
Materials Science and Engineering Grain Growth 魏茂國

- The mechanical properties at room temperature of a fine-grained metal are usually


superior (i.e., higher strength and toughness) to those of coarse-grained ones. If the
grain structure of a single-phase alloy is coarser than that desired, refinement may be
accomplished by plastically deforming the material, then subjecting it into a
recrystallization heat treatment.

35
物理冶金 Boundaries between Crystals of Different Phases 魏茂國

 Boundaries between crystals of different phases


- A phase is defined as a homogeneous body of matter that is physically distinct. The
3 states of matter (liquid, solid, and gas) all correspond to separate phases.
- A number of metals are allotropic (polymorphic); that is, they are able to exist in
different crystal structures, each stable in a different temperature range.
- A solid solution (a crystal containing 2 or more types of atoms in the same lattice)
also satisfies the definition of a phase.
- In alloys of 2 phases, 2 types of boundaries are possible: boundaries separating
crystals of the same phase, and boundaries separating crystals of the 2 phases.

36
物理冶金 Boundaries between Crystals of Different Phases 魏茂國

- In Fig. 6.18, if the surface tensions in the boundaries are in static equilibrium, then
 
 11  2 12 cos  (6.15)
2
where 11 is the surface tension in the single-phase boundary, 12 the surface tension
in the two-phase boundary, and  the dihedral angle between the 2 boundaries that
separate phase 2 from phase 1.
 12 1

 11   (6.16)
2 cos 
2

12cos(/2)

12sin(/2)

Fig. 6.18 The grain-boundary surface tensions at a junction 12sin(/2)


between 2 crystals of the same phase with a crystal of a
different phase. 12cos(/2)
37
物理冶金 Boundaries between Crystals of Different Phases 魏茂國

- The ratio 12/11 is plotted as a function of the dihedral angle  in Fig. 6.19.
Notice that as the surface tension of the boundary between 2 phases approaches half
of that of the single phase, the dihedral angle falls rapidly to zero.
- The significance of small dihedral angles is readily apparent in Fig. 6.20, where the
shape of the intersection is shown for angles of 10 and 1. As the angle approaches
zero, the second phase moves to form a thin
(A)
film between crystals of the first phase.

(B)

Fig. 6.19 The dependence of the two-phase dihedral Fig. 6.20 When the dihedral angle is small, the
angle on the ratio of the two-phase surface tension second phase tends to separate crystals of the 38
first
to the single-phase surface tension. phase. (A) dihedral angle 1, (B) dihedral angle 10.
物理冶金 Boundaries between Crystals of Different Phases 魏茂國

- When the dihedral angle becomes zero, the second phase penetrates the single-phase
boundaries and isolates the crystals of the first phase. Furthermore, the extent of
penetration also depends on the grain-boundary energy.
- Ex: bismuth in copper
The surface tension of a bismuth-copper interface is so low that the dihedral angle is
zero and a minute quantity of bismuth is capable of forming a thin film between the
copper crystals.
Whereas copper is ordinarily a metal of high ductility and capable of extensive plastic
deformation, bismuth is not. In fact, bismuth is a very brittle metal, and when it forms
a continuous film around copper crystals, the copper loses its ductility, even though
the total amount of the bismuth impurity is less than 0.05%. This loss in ductility is
observed at all temperatures at which copper is worked.

39
物理冶金 Boundaries between Crystals of Different Phases 魏茂國

- Second-phase impurities may remain in the liquid state until a temperature is reached
well below the freezing point of the major phase. The amount of harm that these
impurities (in small percentages) can do to the plastic properties of metals is a
function of the surface tension between liquid and solid.
If this interfacial energy is high, the liquid tends to form discrete globules with little
effect on the hot working properties of metals.
On the other hand, low interfacial energies lead to liquid grain-boundary films. These
are very harmful to the plastic properties of metals.

40
物理冶金 Boundaries between Crystals of Different Phases 魏茂國

- Ex: sulfur in iron


Iron sulfide is liquid at temperatures well below the freezing point of iron. This range
includes the temperature range normally used for hot rolling iron and steel products.
The surface energy of the iron sulfide to iron boundary is very close to one-half that
of the boundary between iron crystals, and the liquid sulfide forms a grain-boundary
film that almost completely separates the iron crystals.
Since a liquid possesses no real strength, iron or steel in this condition is brittle and
cannot be hot-worked without disintegrating. In such a case as this, when a metal
becomes brittle at high temperatures, the metal is said to be “hot short”, that is, its
properties are deficient at hot-working temperatures.
 Manganese has a very strong ability to combine with the sulfur in a steel to form
globules, which are solid at the rolling temperatures of steel. The MnS particles have
a much less deleterious effect on the hot rolling properties.

41
物理冶金 Boundaries between Crystals of Different Phases 魏茂國

- For dihedral angles greater than zero but less than 60, the equilibrium configuration
is one where small quantities of the second phase run as a continuous network along
the grain edges of the first phase; that is, along the lines where 3 grains of the first
phase would meet in the absence of the second phase.
According to Fig. 6.19, a dihedral angle of 60 corresponds to a ratio of interphase to
single-phase surface tensions of 0.582.
Many liquid-solid interfaces possess surface tensions less than 0.600 of the solid-
boundary surface tension. Therefore, when liquids and solid coexist, there is a strong
possibility that the liquid will form a continuous network throughout the metal.
Since diffusion in liquids is more rapid than in solids, this network forms a convenient
path for rapid diffusion of elements (both good and bad) into the center of the metal.

42
物理冶金 Boundaries between Crystals of Different Phases 魏茂國

- When the dihedral angle between the interphase boundaries is greater than 60, the
second phase no longer forms a continuous network unless it is present as the major
phase. The second phase, if present in small quantities, now forms discrete globular
particles, usually along the boundaries of the first-phase crystals (Fig. 6.21).

Fig. 6.21 When the dihedral angle is larger (~60), the second phase (when present in small quantities)
tends to form small discrete particles, usually in the boundaries of the first phase. 43
相變態 Equilibrium in Polycrystalline Materials 魏茂國

 No torque at grain boundary


If the boundary energy is independent of orientation, the
torque term is zero and the grain boundary behaves like
a soap film. The boundary tensions 12, 23 and 13 must
balance.
 13 sin180   3    12 sin180   2 
 
  13 sin 3   12 sin 2  13  12 13sin(180-3) 13
sin 2 sin 3
3-90
  23   12 cos180   2    13 cos180   3 
  23   12 cos  2   13 cos  3 90 180-3  cos(180- )
23 13 3
180-2 12cos(180-2)
 sin 2  90
  23   12 cos  2    12  cos  3
 sin 3  2-90
12
   12sin(180-2)
  23   12   cos  2 sin 3  sin 2 cos  3 
 sin 3
Fig. 3.17 The balance of grain

boundary tensions for a grain
     
  23   12   sin 2   3    12   sin360   1  boundary intersection in
 sin 3   sin 3  metastable equilibrium.
      23  
  23   12   sin 1  23  12   13  12
 sin 3  sin 1 sin 3 (3.13)
sin 1 sin 2 sin 3 44
相變態 Equilibrium in Polycrystalline Materials 魏茂國

 Measuring grain-boundary energy


- One method of measuring grain-boundary energy is to anneal a specimen at a high
temperature and then measure the angle at the intersection of the surface with the boundary,
Fig. 3.18.

- In this case, the presence of any torque terms has been neglected, though an approximation
may introduce large errors.
If the solid-vapor energy (SV) is the same for both grains, balancing the interfacial tensions
gives
 
2 SV cos    b (3.14)
2

Therefore if SV is known, b can be calculated.

Fig. 3.18 The balance of surface and grain boundary tensions at the intersection of a grain
boundary with a free surface.
45
物理冶金 Grain Size 魏茂國
(A)

(B)

(C)

Fig. 6.22 Grains removed from a beta brass specimen that appeared to be of a nearly uniform grain size on a
46
metallographic section. Three of some 15 size classifications are shown in these photographs. They represent
the smallest, the largest, and a size from the middle of the range.
物理冶金 Grain Size 魏茂國

 Grain size
- The size of the average crystal is a very important structural parameter in a
polycrystalline aggregate of a pure metal or single-phase metal. Unfortunately, this
is a difficult variable to define precisely.
- The most useful parameter for indicating the relative size of the grains in a
microstructure is l determined by the linear intercept method. This quantity is called
the mean grain intercept and is the average distance between grain boundaries.
- In making this measurement, one can lay a straight edge of perhaps 10 cm in length
down on the photograph and then count the number of boundaries that the edge of
the instrument crosses. This measurement should be repeated several times, placing
the straight edge down on the photograph in a random fashion.
- The total number of intersections, when divided by the magnification in the
photograph, yields the quantity Nl, the average number of grain boundaries
intercepted per centimeter. The reciprocal of this quantity is l, which is often used
as a parameter for indicating the approximate grain size.
1
l (6.17)
47
Nl
物理冶金 Grain Size 魏茂國

- Quantitative metallography has shown that the reciprocal of l, that is Nl, is directly
related to the amount of grain-boundary surface area in a unit volume.

Sv  2 N l (6.18)

where Sv is the surface area of the grain boundaries per unit volume.

48
物理冶金 Effect of Grain Boundaries on Mechanical Properties 魏茂國

 Effect of grain boundaries on mechanical properties: Hall-Petch relation


- The smaller the grain size the greater the hardness or flow-stress, where the flow-
stress is the stress in a tension test corresponding to some fixed value of strain.
- The hardness of a crystal may vary with the orientation of the crystal.
- In Fig. 6.23, it is possible to write an empirical relationship of the form

H  H 0  k H d 1/ 2 (6.19)

where H is the hardness, d is the


average grain diameter, kH is the
slope of the straight line drawn
through the data, and H0 is the
intercept of the line with the
ordinate axis and corresponds to the
hardness expected at a hypothetical
infinite grain size.
Fig. 6.23 The hardness of titanium as a function of 49
the reciprocal of the square root of the grain size.
物理冶金 Effect of Grain Boundaries on Mechanical Properties 魏茂國

- In Fig. 6.24, tensile test data is plotted for titanium specimens tested at room
temperature. Here the flow-stress corresponding to 3 different strains (2%, 4%, and
8%) are plotted against d-1/2, and straight lines are drawn through the data
corresponding to linear relationships of the form

   0  kd 1/ 2 (6.20)
where  is the flow-stress and 0 and k are constants.
- Grain boundaries act as obstacles to slip dislocations, causing dislocations to pile up
on their slip planes behind the
boundaries.
- In the fine-grained materials a much
larger applied stress is needed to
cause slip to pass through the
boundary than is the case with
coarse-grained materials.
Fig. 6.24 The flow-stress of titanium as a function 50
of the reciprocal of the square root of the grain size.
物理冶金 Grain Size Effects in Nanocrystalline Materials 魏茂國

 Grain size effects in nanocrystalline materials


- Nanocrystalline is a terminology that has been used in recent years to describe
materials whose grain size does not exceed 10 nm.
- Sanders et al. have investigated tensile behavior of high-purity Cu and Pd nano-
crystalline metals with grain sizes in the range of 10 to 110 nm.
The hardness of the samples was found to follow the coarse-grained Hall-Petch
relation down to grain size around 16 nm. Below this size, the hardness values were
found to deviate from the relation and level off (Fig. 6.25).

Fig. 6.25 Plot of yield strength and Vickers


microhardness (divided by 3) as a function of the
inverse square root of the grain size. Also shown
is an extrapolation to small grain sizes of the
coarse-grained Hall-Petch relation for Cu and the 51
change in the bulk density.
物理冶金 Grain Size Effects in Nanocrystalline Materials 魏茂國

The yield strength dropped off dramatically as grain size fell below 110 nm. The
failure of the finer-grained samples also occurred with little plastic deformation with
a fracture surface perpendicular to the stress axis.
Dislocation-assisted deformation mechanisms that operate in coarse-grained materials
may not operate effectively in nanocrystalline materials.
- In Fig. 6.25, the bulk density of the finer grained materials was appreciably lower
than that of the coarse-grained ones, indicating the presence of appreciable porosity in
the samples.

52
Fig. 6.25
物理冶金 Effect of Grain Boundaries on Mechanical Properties 魏茂國

- Fig. 6.26 shows yield stress data as a function of grain size for several metallic
systems. The line with slope of one in the figure represents the Hall-Petch relation
given by Eq. 6.2.
For grain sizes smaller than ~30 nm, the yield strength actually decreases with
decreasing grain size.

53
Fig. 6.26 Compilation of yield stress data for several metallic systems.
物理冶金 Coincidence Site Boundaries 魏茂國

 Coincidence site boundaries


- Kronberg and Wilson demonstrated that the rotation could produce surfaces,
separating the new crystal from the old, that contained a number of positions where
the atoms in both crystals were in coincidence. These were called coincidence sites.
- In Fig. 6.27, net A corresponds to the atomic arrangement on the (111) plane of the
parent crystal where the atom centers are indicated by open circles.
Net B corresponds to the (111) plane in a secondarily recrystallized crystal. Here
the atom centers are designated by filled-in circles.
The 2 nets are rotated relative to each other by 22 about the [111] pole.
The points where atoms on the 2 nets coincide are shown as larger open circles.
Note that the coincidence sites also define a similar but larger hexagonal net than
the nets A and B.
A boundary of this type, with a large fraction of coincidence sites, is considered to
have valuable properties, such as a higher mobility or the ability to support a rapid
rate of grain growth.
54
物理冶金 Effect of Grain Boundaries on Mechanical Properties 魏茂國

Fig. 6.27 A coincident site boundary obtained by a 22 rotation across a (111) plane.
The reciprocal density, , of the coincident sites, known as large open circles, is 7.
55
物理冶金 Density of Coincidence Sites 魏茂國

 Density of coincidence sites


- By a direct count, Kronberg and Wilson were able to show that the number of
coincidence sites in the boundary of Fig. 6.27 was equal to 1/7 of the atoms in either
net A or B.
- The fraction of atoms in coincidence, at a boundary of this type, is generally known
as the density of coincidence sites. The reciprocal of the density is more commonly
used as a parameter to describe a coincidence site boundary.
This is usually designated by the Greek symbol .
Thus, for the boundary in Fig. 6.27,  = 7.

56
Fig. 6.27
物理冶金 Ranganathan Relations 魏茂國

 Ranganathan relations
- The theory of coincidence site boundaries received a larger advance from a paper
published by Ranganathan in 1966.
- There are 4 basic factors involved in a coincidence site lattice.
The first is the axis [hkl] about which the rotation occurs; the second is the rotation
angle, , about this axis; the third is the coordinates of a coincidence site in the
coincidence site net on (hkl), the plane normal to the axis of rotation; and the fourth
is , the reciprocal of the density of coincidence sites in the (hkl) net.
- These 4 factors are not all independent. Ranganathan was able to formulate
equations giving both  and  in terms of Miller indices and the coordinates (x,y) of
a coincidence site in the plane (hkl).
- If  is even it should be divided by multiples of 2 until an odd number is attained.
  2 N 1 / 2 tan 1 ( y / x ) (6.21)

  x2  y2 N (6.22)
57
N  h2  k 2  l 2 (6.23)
物理冶金 Examples Involving Twist Boundaries 魏茂國

 Examples involving twist boundaries


- First consider a simple cubic lattice rotated about one of its cube axes, [100]. This
case is treated in Fig. 6.28 in which drawing shows the (100) plane. In this
illustration the x and y axes are the [010] and [001] directions, respectively.
Rotation of the lattice through an arbitrary angle  about [100] will cause the x and
y axes to assume the rotated positions x’ and y’.
Since a rotation about [100] is
being considered, h = 1 while k
and l are 0, so that N = 1. A coincidence site
lattice can be formed if one takes x = 2 and
y = 1. In this case  = 212/1tan-1(1/2) = 53.1
while  = 22 + 121 = 5. Notice that the
coincidence sites define a simple lattice
with a cell whose sides in the (100) plane
equal 51/2a, where a is the lattice constant
58
of the primitive lattice. Fig. 6.28
物理冶金 Examples Involving Twist Boundaries 魏茂國

Simple cubic structure


Plane: (100)
Rotation: [100]
x: [010]
y: [001]
N = h2 + k2 + l2
= 12 + 02 + 02 = 1
(x, y) = (2, 1)
 = x2 + y2N
= 22 + 121 = 5
y
  2 N tan 1
x
1
   2 1 tan 1  53.1
2
Lattice constant
(2a ) 2  (a ) 2  5a
Fig. 6.28 A rotation of 53.1 about <100> axis of a simple cubic crystal gives this coincident site
boundary with  = 5. The coincident sites also form a lattice with a unit cell whose sides are
equal to 5a in the boundary. The cell size of the reciprocal lattice is thus five times larger than
that of the primitive lattice. 59
物理冶金 Examples Involving Twist Boundaries 魏茂國

- Another possibility, involving a rotation about [100], occurs when one chooses x = 3
and y = 1. This is illustrated in Fig. 6.29, where  = 211/2tan-1(1/3) = 36.9 and
 = 32 + 121 = 10. The proper value for  is 10/2 or 5. This is the same as for the
first rotation where x = 2 and y = 1. Note that the same coincidence site lattice may be
obtained by either a rotation of 53.1 or 36.9. The sum of these angles equals 90.
This result is related to the 4-fold symmetry about a <100> cubic axis.

60
物理冶金 Examples Involving Twist Boundaries 魏茂國

Simple cubic structure


Plane: (100)
Rotation [100]
N = h2 + k2 + l2
= 12 + 02 + 02 = 1
(x, y) = (3, 1)
 = x2 + y2N
= 32 + 121 = 10
’ = 10  2 = 5
y
  2 N tan 1
x
1
   2 1 tan 1  36.9
3
Lattice constant
(2a ) 2  (a ) 2  5a

Fig. 6.29 A 36.9 rotation about a <100> axis also produces a coincident site boundary with
 = 5 due to the symmetry of the simple cubic lattice. 61
物理冶金 Tilt Boundaries 魏茂國

 Tilt boundaries
- A tilt boundary can be created by rotating two sections of the same crystal through
53.1 and joining them together, as shown in Fig. 6.31. Note that the boundary
itself is drawn as a series of steps that conform, in both the A and B crystal sections,
to the Ranganathan coordinates x = 2 and y = 1. Thus, the distance between steps
and thus, between coincidence sites on the boundary, equals (x2+y2)1/2 = 51/2a. This
distance, p, is called the structural periodicity.

Fig. 6.31 The coincident site boundaries in


Fig. 6.27 to 6.29 are twist boundaries. This
figure illustrates a coincident site boundary
formed by a tilt of 53.1 about a <100> axis
of a simple cubic crystal.  also equals 5
for this boundary and its structural
periodicity, p, is equal to an edge of the
coincident site lattice.

Structure: S.C., Plane: {100}


Rotation: <100>, (x, y): (2, 1) 62
物理冶金 Tilt Boundaries 魏茂國

- A coincidence site lattice with the same -for example,5-may be obtained with a
twist rotation of 36.9 about a <100> direction. In this case, the Ranganathan
coordinates are x = 3 and y = 1. The corresponding tilt boundary is shown in Fig.
6.32. Notice that its structural periodicity, p, is equal to a diagonal of a coincidence
site lattice cell. For a 53.1 rotation, p equaled an edge of this cell. Thus, the 36.9
and 53.1 rotations yield tilt grain boundaries with different structural periodicities.

Fig. 6.32 A tilt of 53.1 about <100>


of a simple cubic lattice also yields a
coincident site boundary,
but in this case the structural
periodicity, p, is equal to the diagonal
of the coincident site unit cell.  again
is equal to 5.
Structure: S.C., Plane: {100}
Rotation: <100>, (x, y): (3, 1)

63
物理冶金 Tilt Boundaries 魏茂國

- If one considers the atoms as touching spheres, the illustrated boundaries would
contain a pair of overlapping atoms just to the right of each coincidence site on these
boundaries, as is demonstrated in Fig. 6.33 for the 36.9 boundary of Fig. 6.32.
As pointed out by Aust, the problems of this overlap could be relieved by removing
an atom from each of these pairs or by a relative translation of the lattices above and
below the grain boundary as suggested in Fig. 6.34.
This results in an alternation of the ledges between 2 crystal halves and the resulting
boundary is known as a relaxed coincidence boundary.

64
物理冶金 Tilt Boundaries 魏茂國

Fig. 6.33 If the atoms of the diagram in


Fig. 6.32 are drawn as hard balls instead of
points, one finds that there is an overlap
of the atoms just to the right of the
coincident sites on the boundary.

Fig. 6.34 As suggested by Aust, the


overlapping of the atoms at the
boundary could be relieved by a
relative translation of the lattices
Structure: S.C., Plane: {100} above and below the boundary.
Rotation: <100>, (x, y): (3, 1), shifted 65 65

You might also like