Вы находитесь на странице: 1из 12

2 Theoretical Properties of Materials

2-1
THE CALCULATION OF BOND STRENGTH

Assume interatomic stresses, F, vary with distance between atoms, x, according to a sine
function, then the force-interatomic distance is represented by Fig.2-1.

Interatomic Stress
sf

a a
1 2
x

Figure 2-1 Interatomic stresses assuming a sin force-distance relation

The stress-distance relation is


 π ( x − a1 ) 
σ = k1 sin   (2.1)
 a2 

where a1 is the equilibrium distance between atoms and a2 is the distance at which the
attractive force is assumed to be very small. Then at x = a1 and x = a2 , σ = 0
At the position where x-a1 = a2/2 then

π 
σ = k1 sin  
2

The amplitude k1 is the theoretical breaking stress FF. k1 is calculated from the slope at a1.
At position x = a1+dx
σ = dσ
dx
ε=
a1
Young's Modulus is then by definition

E = a1 (2.2)
dx
dF/dx is obtained by differentiating (2.1)

dσ  π k1  π ( x − aa )  
= cos   (2.3)
dx  a2  a2 

2-2
At x = a1, the cosine term =1

dσ k1π
= (2.4)
dx a2

Substituting (2.4) into (2.2)


k1π
E = a1
a2

Assume a1 ≈ a2 , then

E
k1 =
π

or
E
σ th =
π

The application of this result to typical engineering materials is shown in Table 2.1

Table 2.1 Strength of aluminum and steel predicted by the sin approximation of the force-
distance behavior.

Material E E/B Best FF Typical FF

Steel 30 x 106 10 x 106 0.5 x 106 0.05 x 106


Aluminum 10 x 106 3 X 106 0.1 x 106 0.03 x 106

2-3
THEORETICAL STRENGTH APPROXIMATION BASED ON ATTRACTIVE AND
REPULSIVE FORCES

In this approximation the stress due to the atomic forces is the summation of the
attractive and repulsive forces (Fig 2-2).
-n
Attractive A x
σ a1

x Separation between atoms


-m
Repulsive -Bx

Figure 2-2 Interatomic stress-distance assuming the summation of attractive and


repulsive forces.

The summation results in


σ = Ax − n − Bx − m (2.5)
Integrating equation (2.5) gives the energy expression:
Ax − n +1 Bx − m +1
U =− +
n −1 m −1
These results are shown in Fig.2-3.

U
a
1

Uo

Figure2-3 Energy-distance relation assuming summation of attractive and repulsive


forces.

At a1, the atoms are at the equilibrium position in the lattice and the dissociation U0 is
given as:
− ( n −1)
Aa1 Ba − ( m −1)
U0 = − + 1 (2.6)
n −1 m −1

2-4
This approach assumes that the heat of vaporization is related to the theoretical strength
of a material.

To express B in terms of A we notice that at x = a1, σ = 0 and Aa1− n = Ba1− m then


B = Aa1( m − n ) (2.7)

Substituting (2.7) into (2.6)


(n − m) Aa1− ( n −1)
U0 = (2.8)
(n − 1)(m − 1)

We now have a simple expression for U0 in terms of the single constant, A, and the
equilibrium atomic distance, a1. Equation (2.5) can be used to calculate the theoretical
strength if we have values for m and n, and if we can express the constants A and B in
terms of the elastic modulus, E.

At the point where the attractive force equals the repulsive force, σ = 0

E = a1
dx
Differentiating equation (2.5)
E = a1 (−nAx − n −1 + mBx − m −1 ) (2.9)

Using (2.7) in (2.9)


a1n E
A= (2.10)
m−n

Maximum stress occurs at dF/dx = 0 hence − nAx − n −1 + mBx − m −1 = 0 and therefore


mB
x m−n = (2.11)
nA
Substituting (2.7) into (2.11) and solving for x
1
x = a1 ( mn ) m − n (2.12)
Using (2.5), (2.7) and (2.10) the theoretical strength of a material based on interatomic
forces is
n
E  n  m−n
σ th =   (2.13)
mm
To calculate cleavage strength the following attraction and repulsion exponents shown in
Table 2.2 can be used:

2-5
Table 2.2 Attractive and repulsive exponents

Bond Type Repulsive, m Attractive, n


Metallic 7 4
Ionic 10 2
Covalent 12 8

Using these exponent the theoretical strength of various materials is given in Table 2.3.
These strengths are sometimes referred to as cleavage strengths since the atoms are
separated across an entire plane by tension. If the atoms are separated in shear a different
model must be used and undoubtedly a different result would be obtained.

Table 2.3 Theoretical Cleavage Strength of Various Materials

Bond Material n m F/E E, msi Fth, ksi Factual,


ksi

Fe 4 7 0.0677 30 2,032 50
Cu 4 7 0.0677 16 1,084 70
Al 4 7 0.0677 10 667 20
Metallic Ti 4 7 0.0677 16 1,084 80
Be 4 7 0.0677 44 2,980 80
KCl 2 10 0.0669 .19 13 0.145
Al2O3 2 10 0.0669 60 4,012 35
Ionic Glass 2 10 0.0669 10 669 2
Covalent SiC 8 12 0.0370 50 1,852 55

What is most striking about these results is that the theoretical strengths are at least two
orders of magnitude less than commonly measured strengths. This observation applies to
all types of bonding: metallic, ionic and covalent.

THEORETICAL SHEAR STRENGTH

One explanation of why the strength of metals is below the theoretical strength is that
even when loaded in tension the deformation occurs by shear. If we start with the
assumption that the metal is a perfect crystal and deformation occurs by half the crystal
sliding past the other half across the entire plane as illustrated in Fig. 2-4, a simple
mathematical model can be created.

2-6
shear stress

b
X

Figure 2-4. Shear taking place on a crystalline plane.

Let us envision this process occurring in a metal with a close packed crystal structure.
The stress is minimum at the equilibrium positions (before any sliding occurs) and
increases in a sinusoidal manner to a maximum when two atoms on adjacent planes,
separated by a distance h, have been displaced by the distance b/4 from each other.
Further displacement reduces the stress needed for the plane to slide past each other
reaching zero at the next position of equilibrium. The shear stress at various sliding
distances, x is
 2π x 
τ = k2 sin   (2.14)
 b 
where k2 is a constant that describes the amplitude of the sine wave.
The shear modulus G near the equilibrium position is given by

G=h (2.15)
dx
Differentiating (2.14) and combining with (2.15) gives
G 2k2π  2π x 
= cos   (2.16)
h b  b 
At x = 0
Gb
k2 = (2.17)
2π h
Substituting (2.17) into (2.14) and evaluating at x = b/4 gives
Gb
τ th = (2.18)
2π h

We can calculate a value for τ th by using known values of G, b and h. The latter two
quantities depend on crystal structure. For a face-centered-cubic metal b = a0 6 and
h = a0 3 . Substituting these values into (2.18) gives

G
τ th =
9

2-7
By converting shear strength to tensile strength and shear modulus to tensile modulus, we
arrive at
E
σ th =
12

This value for theoretical strength is near the value calculated for theoretical tensile
strenght, E/15, and is an order of magnitude larger than that measured for any metal.
Shear actually takes place by the glide of dislocations on a slip plane that requires
considerably less force than the model pictured above.

DISLOCATION MODEL OF SLIP

In the dislocation model an extra plane of atoms is inserted into the lattice creating a
linear fault terminating on the slip plane (Figure 2-5). Instead of requiring all the atoms
on the entire slip plane to move simultaneously the dislocation model requires only the
atoms around the linear fault to move.

slip plane

b
Figure 2-5. A dislocation in a crystal.

The theoretical shear strength for this model is then the frictional force of the dislocation
as it moves through the lattice. This frictional shear is called the Peierls stress and given
as:
2π W
E −
τ= e b

(1 −ν )
2

where b is called the Burgers vector and W is the width of the dislocation, and is
approximately 3b. The theoretical shear strength is then
E
τ th = 6
10

2-8
The shear stress calculated using Peierls stresses under estimates the shear strength of
most metals. Hence the simple models of shear either grossly over estimates or under
estimates the shear strength of real materials.

CLEAVAGE STRENGTH FOR REAL MATERIALS

Only perfect unflawed materials are expected to have strengths approaching the
theoretical strength. Real materials have numerous flaws and the fracture strength of
brittle materials depends on the size of flaws contained in the structure. The cleavage
fracture strength can be expressed as:
EG
σf =
acπ (1 −ν 2 )
where G is referred to as the strain energy release rate and ac is the critical flaw size
required for it to propagate catastrophically through the structure.

This model implies that the elastic strain energy stored in a cracked structure under load
provides all the energy required to propagate a crack completely though it. The "rate"
(with respect to crack length) of decrease of elastic strain energy is a measure of the
fracture toughness of the material. This model predicts fracture stresses very close to
measured strengths and significantly lower than theoretical cleavage strengths. These
observations leads to the conclusion that the strengths of common engineering materials
are controlled by size of flaws they contain.

The fracture toughness approach can be used to determine the diameter of a fiber that
would be smaller than the critical crack length by equating the theoretical cleavage
strength to the cleavage strength for real materials:
2
E
  =
EG
 15  π ac
(
1 −ν 2 )
Assuming ν = 0.3 and solving for ac gives
G
ac = 80
E
Typical values of G and E for alumina result in a critical particle size of 4nm.

SHEAR STRENGTH FOR REAL MATERIALS

Plastic flow occurs when dislocations sweep though a material. The resistance to
dislocation motion is therefore the source of the shear strength. This resistance includes
not only the Peierls stress (lattice friction) but also the force necessary to generate new
dislocations. There are numerous mechanisms for generating dislocations one such
mechanism is called the "Frank-Reed" source (Fig 2-6). In this process a dislocation,
which is pinned by two obstacles, bows out under an applied stress. The bowing will
proceed until it encircles the obstacles and joins up behind them leaving two distinct and

2-9
separated dislocations. One of these forms a ring around the original pinning points while
the other becomes a new segment pinned between the obstacles that replaces the original
one that has expanded to form the ring. By this mechanism the Frank-Reed source can be
considered a mill that can continue to generate dislocations and hence slip as long as a
stress is applied. The shear strength according to this model is then the frictional stress
required to move the dislocation through the crystal lattice.
.

Figure 2-6. Frank-Reed mechanism for dislocation generation

The shear strength is then


2Gb
τ=
L
where L is the distance between obstacles (pinning points).

The fracture strength by continuous yielding until one plane slides completely off an
adjacent plane (in the manner of two playing cards sliding past one another) is then
4Gb
σ=
L
Minimum pinning point distance to achieve theoretical strength for shear failure is found
by equating the theoretical strength by shear to the dislocation strength model. This is
given by
60Gb
L=
E
For common engineering alloys L ≈ 6 nm .Hence, whether by cleavage or shear the
critical size to achieve theoretical strength is very small.

2-10
CONLUSIONS REGARDING FIBER STRENGTH ANALYSIS

Material is stronger in fiber form than in bulk form because of the reduced size and
probability of flaws in the very limited cross section of fibers. Stiffness and density are
not significantly affected by cross sectional area.

These observations lead to the following conclusions:

• Elastic properties of materials are a direct consequence of the atomic bond


strength. These conclusions are illustrated by the various elastic moduli possible
in the iron-carbon alloy system. The atomic bond strength for graphite normal to
the layered form is very low. Changing the type or fraction of the graphitic form
of carbon in iron can have a significant effect on elastic properties because the
stiffness of the graphite phase strongly influences the material stiffness. The
fraction of layered graphite in a 2%C-Fe alloy (cast iron) may be varied by heat
treatment, and minor alloying constituents resulting in significant changes in
Young's modulus. Some possible variations in carbon form and the resulting
Young’s modulus are shown in Table 2.4. These materials are listed in the order
of decreasing amounts of layered form of carbon with corresponding increases in
Young’s modulus.

Table 2.4 The effect of carbon form on Young’s modulus in cast iron

Young's
Cast Iron Form Modulus Form of the Undissolved Carbon
106 psi
Grey Iron 11 Graphite flakes
Compacted Graphite Iron 16 Graphite flakes and nodules
Ductile Iron 23 Graphite nodules
White Iron 28 Carbides

• Strength of materials is only indirectly connected with atomic bond strength.


Carbon dissolved in iron does not affect elastic modulus but has a profound effect
on strength. All steel compositions regardless of heat treatments have an elastic
modulus of 30,000,000 psi even though strength can vary from 10,000 psi to over
500,000 psi.

• Material strength is greatly affected by size and defect content. Fig. 2-7 shows a
large variation in strength for silicon carbide crystals depending on domain
particle size and shape.

2-11
FLEXURAL STRENGTH, MPa
WHISKERS
4000
CRYSTALS
1500

400

150

40

15 PLATELETS

1 10 100 1000
SIZE, MICROMETERS

Figure 2-7. The effect of particle size on the strength of silicon carbide

SELECTING FIBER MATERIALS USING THE PERIODIC TABLE

If light weight and strength are the most important considerations for selecting a
reinforcement material in a composite, then we can use the following guidelines in the
search for the most promising materials: The lightest elements are found in first two rows
of the periodic table of elements. If high strength is the selection criteria then select
elements or compounds that have covalent or electrovalent bonding. Some elements fall
in both categories such as boron and carbon and are indeed preferred reinforcement
materials. Compounds that meet these selection criteria are: Al2O3, BeO, SiC, B4C,
B13O2, B6Si, AlB3.

2-12

Вам также может понравиться