C7 Engineering Materials Metals: Prof. Dr. Rer. Nat. S. Schmauder
C7 Engineering Materials Metals: Prof. Dr. Rer. Nat. S. Schmauder
C7 Engineering Materials Metals: Prof. Dr. Rer. Nat. S. Schmauder
Engineering Materials
Metals
Prof. Dr. rer. nat. S. Schmauder
Institute for Materials Testing,
Materials Science and
Strength of Materials
University of Stuttgart
17
1 Fundamentals of dislocation theory - 3 -
1.1 Lattice defects - 3 -
1.2 Types of dislocations - 5 -
1.3 Stress field and energy of dislocations - 8 -
1.4 Reactions of dislocations - 10 -
1.5 Movement of dislocations - 14 -
2 Plastic deformation of metals - 18 -
2.1 Glide process and flow stress - 18 -
2.2 Strengthening of single- and polycrystals - 22 -
2.3 Strength hypotheses - 26 -
2.3.1 Maximum normal stress theory - 27 -
2.3.2 Maximum shear stress theory (Tresca criterion) - 27 -
2.3.3 Maximum distortion energy theory (von Mises criterion) - 27 -
2.3.4 Example of application - 29 -
2.3.5 Example for calculation - 30 -
3 Possibilities of increasing the strength - 31 -
3.1 Strengthening mechanisms - 31 -
3.2 Solid solution hardening - 32 -
3.3 Dislocation hardening - 34 -
3.4 Grain refinement (fine grain and grain-boundary hardening) - 37 -
3.5 Particle strengthening (precipitation and dispersion hardening) - 40 -
3.6 Combined methods - 43 -
Contents
18
4 Influences on the material behaviour - 45 -
4.1 Specific material properties - 45 -
4.2 Stress state - 48 -
4.3 Temperature - 52 -
4.3.1 High temperatures - 52 -
4.3.2 Low temperatures - 54 -
4.4 Loading velocity - 56 -
5 Tasks - 58 -
6 Solutions - 61 -
Editor: Institute for Materials Testing, Materials Science and Strength of Materials
University of Stuttgart
Edition: 11 2013
All rights reserved
- 3 -
1 Fundamentals of dislocation theory
1.1 Lattice defects
Crystal lattices of technical materials are never completely free of defects (Ideal
crystal). The existing defects critically influence both the mechanical and physical
properties of real crystals. Technical applications of materials are, actually, only
possible due to the presence of defects. Lattice defects can be classified into groups
based on their spatial dimensions:
- 0-dimensional lattice defects or point defects
- 1-dimensional lattice defects or line defects
- 2-dimensional lattice defects or area defects
Grain boundary
Dislocation
Vacancy
Fig. 1.1: Lattice defects
1) Point defects can be either vacancies, interstitial atoms or Frenkel pairs.
Vacancies are non-occupied lattice positions. Their frequency increases with
temperature (faster vibration of the atoms, important for diffusion processes).
Interstitial atoms can be either lattice type atoms or foreign atoms. Frenkel
pairs (Frenkel defects) are composed of a vacancy and an interstitial atom.
- 4 -
These defects emerge when an atom, due to high-energy radiation, is pushed
out of its crystal lattice position into a location between lattice positions (for
example, neutron bombardment during nuclear fission).
Point defects only disturb the crystal structure within small areas. As a rule,
their impact on the mechanical properties is small. Exception: reactor materials
that are exposed to strong neutron radiation.
2) Line-defects, dislocations: The lattice is dislocated along a line, because, for
example, the atoms of half of a lattice plane are missing certain. In the vicinity
of the dislocation, the lattice is constrained because the regular structure is
disturbed. The dislocations in
3
cm 1 iron attain typically lengths of km 500 ,
which can increase to about km 10
7
after extensive deformation. Dislocations
significantly influence the mechanical properties, particularly the deformability
and strength of metals.
3) Area defects comprise primarily grain boundaries. They appear during
solidification (crystallization), because the lattices of individual grains are
arbitrarily oriented and the grain boundaries are, therefore, defects. Thus, an
interface between two grains is an area of disturbed lattices with a diameter of
several atomic radii. Impurities which are usually always present in the melt
can segregate to or in the proximity of the grain boundary, when they cannot
become incorporated into the lattice itself. Furthermore, the grain boundary is
frequently the preferential location for precipitates.
Because of the orderless structure and impurification at the grain boundary,
one would expect that the bonding strength in the boundary would be less than
that in the grain itself and, therefore, a crack would likely occur at the grain
boundary and continue inter-crystalline. This, however, is not the case. In
ductile as well as in brittle materials, a crack continues, as a rule, trans-
crystalline under a (static or cyclic) mechanical load. Exceptions hereof are
fracture of very hard materials, failure by corrosion (inter-crystralline corrosion)
and due to long-time loading.
- 5 -
1.2 Types of dislocations
Dislocations are linear lattice defects. The centre of such a disturbance the zone of
maximum lattice distortion continues along a straight or curved line, and is called
the dislocation line. Three different types of dislocations can be distinguished: edge,
screw and mixed dislocations (see fig. 1.2).
Fig. 1.2: Types of dislocations
Edge dislocations contain a step-like lattice distortion that can be thought of as
originating through an introduced lattice half-plane whose terminating edge is
located parallel to the dislocation line. In the case of a screw dislocation, the lattice
planes, which are perpendicular to the dislocation line, are deformed into helical
surfaces. A mixed dislocation is a combination between the two, where some parts
take on an edge character and other parts, a screw character. Dislocation lines
cannot end inside a perfect crystal, but at the surface or at other defects inside the
crystal, or they may also establish closed rings. The following fig. 1.3 shows a two-
dimensional model out of soap bubbles for an edge dislocation. The dislocation is
detectable in this seemingly regular bubble arrangement when the spheres are
viewed from an acute angle in the direction of their allignment.
Edge
dislocation
Mixed
dislocation
Screw
dislocation
- 6 -
Fig. 1.3: Soap bubble-model of an edge dislocation
The direction and amount of a dislocation caused by lattice distortion is represented
by the Burgers vector b
from the core of the dislocation. The stress field is only depending on r, that
means it is rotational-symmetric. A singularity ( t ) developes at the dislocation
core ( 0 r ), for which above equation cannot be applied because in this area the
relation = t G from linear elasticity theory is not valid.
- 9 -
Because of the distortion of the lattice a certain amount of elastic deformation energy
is stored in the area of a dislocation. The specific deformation energy is given as
G 2 2
1
W
2
0
t
t = = (1.6)
Since t is proportional to b (equ. 1.5) the energy of a dislocation is proportional to
the squared Burgers vector:
2
b ~ W (1.7)
Provided that the stress field is known, the energy of the dislocation can be
calculated approximately by the integration over its efficient range. For a screw
dislocation the following expression is obtained
0
1
2
r
r
ln
4
l Gb
W
t
=
O
(1.8)
where
G: shear modulus
b: Burgers vector
l: length of dislocation
0
r : radius of the core of dislocation
1
r : radius of the crystal
It is not possible to take the core of the dislocation into account because the linear
elasticity theory is no longer valid in this area.
For an edge dislocation the energy amounts approximately to:
O
=
t
= W
1
1
r
r
ln
) 1 ( 4
l Gb
W
0
1
2
(1.9)
( = Poissons Ratio)
The energy of a dislocation is distributed over a relatively large crystal volume
according to the corresponding stress field. This energy is much larger than the
energy of a vacancy.
- 10 -
1.4 Reactions of dislocations
It is possible that two (or more) dislocations react with each other because a large
number of dislocations is present in a real crystal. For such a reaction two conditions
are of importance. The kinematics of a possible reaction is determined under the
condition that the sum of the participating Burgers vectors is constant:
3 2 1
b b b
+ = (1.10)
E.g. in a fcc lattice (see fig 1.7):
| | | | | | 1 10
2
a
101
2
a
100 a + = (1.11)
x
y
z [001]
[010]
[100]
[10 ] 1
[101]
Fig. 1.7: Kinematics of a dislocation reaction
The energy balance is decisive for the direction of a possible dislocation reaction. It
always continues in the direction of the state having the lowest energy content.
Because the energy of a dislocation is proportional to the squared Burgers vector the
following condition is valid for the direction of the reaction:
Splitting:
2
3
2
2
2
1
3 2 1
b b b if , b b b + > + (1.12)
Merging:
2
3
2
2
2
1
3 2 1
b b b if , b b b + < + (1.13)
Lets observe some simple dislocation reactions:
It is possible that two dislocations can anihilate if their Burgers vectors possess the
same amount but different algebraic signs and if they are positioned in the same
glide plane (anihilation, see fig. 1.8).
- 11 -
They react together according to
0 ) b ( b
1 1
+ (1.14)
t
t
B
A
H
1
H
2
G
A
B
H
Fig. 1.8: Anihilation of two dislocations
- If two (edge) dislocations move on parallel glide planes, it results in a void
chain (dipole) because it contains less energy than both of the inividual
dislocations.
- An energy balance helps to explain the fact why in a crystal there occure no
double dislocations with a Burgers vector b 2
, as
2 2
b 2 ) b 2 ( > (1.15)
Both of the individual dislocations would repell each other due to their identical stress
fields and they would attain the largest distance possible to each other.
- However, on the basis of this principle a simple complete dislocation should
split up into several (incomplete) partial dislocations. Since the Burgers
vectors of incomplete dislocations are no complete lattice vectors the original
lattice would not be re-established in case of such a split up and the local
change occurring in the lattice would require additional energy. Depending on
the level of this additional energy and the type of lattice partial dislocations
may either occur or not.
- 12 -
This occurs less in the fcc lattice than in the bcc lattice. Certain partial dislocations
are possible in this case resulting in the fact that the normal stacking sequence of the
lattice planes would be disturbed with the development of so-called stacking faults
(see fig. 1.9). The energy of such a stacking fault is not that strong, that it would
prevent from the splitting into partial dislocations.
Fig. 1.9: Stacking sequence in a fcc lattice
- In recovery processes following plastic deformation (crystal recovery) or at
high temperature deformation, the arrangement of similar edge dislocations
one upon another can be observed (see fig. 1.10). They are separating crystal
areas which are inclined by a small angle o. Such a dislocation arrangement is
called small angle grain boundary.
b
o
d
o ~ o ~
o
tan b/d
< 15
Fig. 1.10: Small angle grain boundary ( d / b tan with 15 = o < o )
- 13 -
Small angle grain boundaries are coherent grain boundaries, possessing a regular
structure which is coherent with the lattice structure.
- A special completely coherent grain boundary is the twin boundary which is
present when the orientation of two crystals is mirror symmetrical to each
other (see fig. 1.11). The interface between the two crystals is called the mirror
plane.
t
Fig. 1.11: Twin boundary
Small angle grain boundaries occur at about up to b 10 d > . A special non coherent
grain boundary develops in case of larger angles as a transition zone of
approximately 0.5 nm thickness between two crystals of different orientation. It is also
called a large angle grain boundary.
In order that dislocations can react with each other they have to be flexible. The
driving force to move can be initiated by the own stress field (e.g. splitting up) or by
external stresses, i.e. by external forces and moments (mechanical load). The
ability of a dislocation to move depends on the special type of the dislocation and the
dislocation reaction. Furthermore, it depends on its position relative to the lattice
structure, i.e. to the offered glide possibilities (see section 2.1).
Experimentally, dislocations can be directly made visible with the help of electron
microscopy. Extremely thin metal foils which are enlarged by 10 to 100000 times are
used for this purpose. Since a dislocation is distorting the surrounding lattice the
conditions of deflection and reflection are slightly changed for the penetrating
electrons. In the radiation-image the dislocations appear as bright and dark lines and
after deformation also as dislocation networks.
An additional indirect method to prove dislocations is the fact that impurities in
technical metals accumulate preferably at dislocations. These precipitated impurities
- 14 -
show special etching characteristics with the consequence that the traces of
dislocations decorated with impurities can be found.
1.5 Movement of dislocations
Theoretical shear stresses in the range of 10 / G
th
= t are required to displace a
lattice plane against the neighboured plane by one atomic distance. The actual, i.e.
the experimentally measured stress
exp
t necessary for such a deformation is
smaller by the factor
3 2
10 10 . It is ) 10 10 /(
3 2
th exp
= t t . Meaning that the
deformation cannot take place simultaneously with all atoms of one plane but that a
different deformation mechanism is effective. Such a mechanism is likely if one
assumes that the lattice contains dislocations which start moving under the effect of
an externally applied stress. Because dislocations move, there develops a
permanent change of the local lattice shape as schematically shown in fig 1.12.
After traversing the complete crystal the dislocation dissolves at the surface with a
remaining step-like deformation.
Fig. 1.12: Deformation through movement of a dislocation
In this deformation mechanism the required stress is considerably lower because
each of the atoms of a lattice plane moves successively and not simultaneously. This
is demonstrated in the corrugated plate model (see fig. 1.13).
- 15 -
t
t
t
t
v
v=0
E
E
Fig. 1.13: Corrugated plate model
The trays correspond to the equilibrium-status of the atoms (= balls). If one moves a
complete lattice plane one should imagine the balls to be connected by a stiff cord
which has to be pulled strongly so that the balls slide simultaneously over the tops
into the next trays. On the other hand, if one assumes that not all the balls possess
the same distance (to each other) but partly are relocated against the neighbouring
position (similar to the atoms in an effecting area of a dislocation), and that they are
elastically connected by strong springs, each of the balls will be pulled one after each
other in the respective neighboured tray with requiring less force than in the first
case.
A similarly clear imagination of the mechanism of a dislocation movement is shown in
the carpet model (see fig. 1.14).
Fig. 1.14: Carpet model
- 16 -
To prompt a dislocation to move significantly, less stress is required than one needs
to move two complete lattice planes against each other. How high is the external
stress which causes the movement of the dislocation with the accompanying internal
stress field?
It is difficult to calculate this so-called Peierls-(Nabarro)-stress. This is because one
should not act on the assumption of an elastic continuum, for that the elasticity theory
is valid. Such an assumption would be contrary to the crystal structure of the lattice.
Therefore, one should try to precisely determine the interactions of forces between
the dislocations and the lattice.
Certainly, the energy of a moving dislocation will not be constant but will change
periodically according to the particular position (see fig. 1.15).
E
E
p
E
min
0
0
b/2
b/2
b
b
F
linear elastic
periodic
Fig 1.15: Energy of a moving dislocation
This energy is equivalent to the force that the lattice exerts on the dislocation. Thus
an external force which exceeds the maximum value of the energy (
p
E : Peierls-
energy) is needed to move a dislocation. Because of this, the calculation requires a
force-deformation-law that shows an analogous, periodic behaviour. The Peierls-
stress calculated this way is extremely sensitive to the form of this material law. This
is why a reliable theoretical estimation of the critical external force to move a
dislocation is rather difficult. Also, up to now the direct experimental determination
through measurements did not succeed in a satisfactory way.
However, because the energy of a dislocation is proportional to the squared Burgers
vector, it can be concluded that the Peierls-energy is the lower, the smaller the glide
vector and the denser occupied the glide plane is. For that reason the movement of
the dislocation takes generally place in the densest occupied lattice-planes and here
- 17 -
it follows the densest occupied lattice-direction. When the dislocation movement
ends (at grain boundaries or on the surface), a small permanent or plastic
deformation of the size of the Burgers vector appears. The macroscopic plastic
deformation of the material results from the total of all dislocation movements.
- 18 -
2 Plastic deformation of metals
2.1 Glide process and flow stress
Plastic deformation is a consequence of glide processes of neighbouring lattice
planes. They are caused by dislocation movements and are triggered by shear
stresses when these reach a certain critical level. The extent of gliding processes
depends on the number of the sliding possibilities.
The sliding possibilities are resulting from the glide planes and glide directions. Glide
planes are the most densely filled lattice planes and sliding directions the most
densely filled lattice directions. The combination of a glide plane with a glide direction
in this plane is called glide system. The number of glide systems of a certain lattice
type results as a product consisting of the number of glide planes and the number of
glide directions on those planes. Fig. 2.1 shows the glide systems of the fcc, bcc and
hcp lattices.
Lattice Unit Cell
exemplarily:
Glide Syst.
Glide Plns. Glide Dirs.
fcc
(Al, Cu, Ni,
aust. St.)
x
[100]
z
[001]
y
[010]
[01 ] 1 [1 0] 1
[ 01] 1
(1
1
1
)
(111) [101]
12 3 4 =
bcc
(Fe, Mo, W,
ferr. St.)
x
z
y
(1
1
0
)
[1 ] 11
[ 1 ] 1 1
(110)
(112)
(123)
[111]
12 2 6 =
12 1 12 =
24 1 24 =
hcp
(Mg, Zn)
(0001)
Basal
Planes
Principal
Axes
3 3 1 =
Fig. 2.1: Glide systems of the fcc, bcc and hcp lattices
- 19 -
For bcc metals, it is not possible to fix a single glide plane. The following glide planes
are found in bcc metals, (110), (112) and (123) in which the diagonal [111] is
contained.
The number and orientation of the glide systems considerably influences the plastic
behaviour of materials. Thus, at room temperature, fcc metals possess a very high
deformability whereas bcc metals are less deformable. The glide planes in bcc
metals are not the densely packed planes as in fcc lattices. Hexagonal metals are
only deformable to a certain extent. With increasing temperature it is possible that
other or additional glide systems may become effective. This can be seen when
comparing high temperature strength or the tendency to creep in perlitic-ferritic and
austenitic steels at high temperatures.
The glide process and thus the plastic deformation initiate when a certain critical
shear stress
0
t has been obtained in a glide system. Which of several possible glide
systems will be activated first is on the one hand depending on the position of the
glide system relative to the external force. On the other hand it depends on the lattice
type itself. Concerning a certain glide system (with a given sliding plane and sliding
direction), the dependence of t as a function of direction can be deduced as follows
(see fig. 2.2):
g
n
1
o
1
o
0
t
_
F
F
A
0
A
1
Fig. 2.2: Determination of
t
(A1: Glide plane; g: Glide direction)
- 20 -
In a rod-shaped specimen subjected to force F the normal stress in the direction of
loading is dominating
0
0
A
F
= o (2.1)
On the sliding plane with the normal vector n
1
which is inclined by the angle _
against the normal of
0
A the following stress is acting because ) cos( A A
1 0
_ =
_ =
_
= = o cos
cos A
F
A
F
p
0
0 1
1
(2.2)
This stress is inclined to
1
A and possesses a shear component in the given (sliding)
direction g of
0 0 1
) cos( ) cos( cos p o o t = _ = = (2.3)
Therefore, t depends on the orienting factor ) cos( ) cos( _ = for a given
0
o . It can
be shown that t attains its maximum value
0 max
2
1
o t = for 5 . 0
max
= for = = _ 45 (2.4)
In a polycrystal sliding will therefore start as soon as
0
o has reached the value
0 max 0
2 2 t t o = = (2.5)
where
0
t is the critical shear stress of the respective glide system.
In case of a single crystal the level of
0
o at the beginning of yielding (yield stress
F
o ) is a strong function of the orientation of the slide system relative to
0
o .
Therefore, in terms of yield initiation, single crystals behave very anisotropic.
However, in polycrystals the flow stress is being averaged because of the arbitrary
orientation of the crystallites. This average flow stress of a polycrystal is
considerably higher than the lowest flow stress in a single crystal of the same metal
(see fig. 2.3).
- 21 -
250
200
150
100
50
0
0 50 100 150 200
Single Crystal
Polycrystal
Zn (hcp)
c / %
o
/
M
P
a
Fig 2.3: Stress-strain-diagram of a single crystal and a polycrystal of tin at room
temperature
The opinion that plastic deformation is based on the movement of dislocations has
been confirmed by comparing a single crystal and a whisker (thread-like single
crystal, hair crystal). The single crystal contains a large number of dislocations
whereas the dislocation density of the whisker prior to deformation is virtually equal
zero. The yield point of the single crystal is accordingly low compared to the whisker
which deforms elastically by about 5 %. Thus, the yield stress attains the magnitude
of the theoretical strength (see fig. 2.4). Then, dislocations spontaneously develop
causing a sharp drop in the flow stress in this case.
0 10 20 30
1
10
10
2
10
3
10
4
o
F
/
M
P
a
c
pl
/ %
0 10 20 30
o
th
1
10
10
2
10
3
10
4
o
F
/
M
P
a
c
pl
/ %
Single Crystal Whisker
Fig. 2.4: Strengthening curves of o-iron, a) single crystal, b) whisker
- 22 -
2.2 Strengthening of single- and polycrystals
Plastic deformation (microscopics the sliding, macroscopic: yielding) is initiated if
1) glide systems are available in such an orientation that sliding is possible (at best
under 45 to the tensile direction) and if
2) the shear stress in these glide systems has reached the (specific) critical value
0
t which initiates gliding.
As soon as plastic deformation has been put into action, i.e. the dislocations started
moving, one should expect the following. Sliding should continue under a constant
shear stress
0
t until the activated dislocations have traversed the complete crystal
each leaving one sliding step at the crystal surface. In a single crystal which is
oriented such that sliding takes place in one and the same sliding system (single
glide) it would mean that the total plastic deformation had to take place under an
almost constant shear stress. However, this is only partly the case, as shown in
fig. 2.5 where a typical shear stress-sliding curve is shown for an fcc single crystal
oriented for single glide (gliding a = shear ).
I II III
(fcc)
G/300
G=d /d t
t
Fig. 2.5: Shear stress-sliding curve of an fcc single crystal (schematic)
- 23 -
One can differentiate between three regions: the relatively flat regime I of easy glide,
the strongly increasing linear regime II and finally the regime III with decreasing
slope. These areas can be interpreted as follows:
The dislocations in the undeformed crystal of regime I (with a density of
2 8 6
cm 10 10
= estimated from measurements at small areas in the electron
microscope) are activated by the stress
0 I
t t = . They move mainly without
resistance, i.e. without interaction with other dislocations, completely through the
crystal. When this possibility of deformation is exhausted, additional dislocations
which are less capable of sliding as well as unfavourably oriented sliding systems
have to be activated or new dislocations have to be generated requiring a higher
stress. This is the transition to regime II, the hardening. New dislocations may
develop if a dislocation line has been arrested in its movement. In single crystals,
knots with other dislocations develop on crossings of different glide planes. In
polycrystals knots develop also on grain boundaries or with other dislocations so that
they cannot slide anymore (stationary dislocations). If the shear stress increases
further, closed dislocation rings may develop detaching themselves and spreading
into the crystal (see fig. 2.6).
t
Fig. 2.6: Frank-Read-source
The blocked dislocation line is thus a source of permanent new dislocations (Frank-
Read-Source) and the density of dislocations can be increased from
2 8 6
cm 10 10
= for the undeformed state up to
2 12
cm 10
= by plastic
deformation. However, this also causes interactions between the individual
dislocations, e.g. the superposition of the residual stress fields of several
dislocations. Consequently, the shear stress for the movement of the dislocation
increases appropriately. This is especially pronounced when groups of dislocations
develop, e.g. as in the case of the pile-up of dislocations in front of an obstacle (see
fig. 2.7).
- 24 -
Such stress concentrations may arise to such a magnitude that a (micro)crack may
develop triggering failure.
t
Fig. 2.7: Dislocation pile-up in front of an obstacle
In regime II a correlation obviously prevails between the critical shear stress
v
t in
the hardened state, the critical shear stress
0
t of the undeformed crystal and the
density of dislocations to be expressed as
+ = t t b G
0 v
(2.6)
The slope of the ( ) t, curve in regime II (the hardening coefficient) is given for all
fcc metals as
300
G
d
d
~
t
(2.7)
The reduced slope in regime III can be explained by the fact that apart from the
previous deformation mechanism another process has been activated facilitating the
bypassing of obstacles. The dislocations may leave their original glide plane from
stress
III
t t = and jump over to other glide planes (climbing of edge dislocations
and cross gliding of screw dislocations).
If the crystal is oriented such that gliding does not only take place in one (optimum
located) glide system (single glide) but that several (not optimum located) glide
systems contribute to the deformation (multiple glide) one can observe the following.
Regime I nearly disappears and the slope of the hardening curve in regime II as well
as
III
t increase, consequently, the shape change of the crystal becomes more
difficult.
- 25 -
This is even more pronounced if one changes from single crystals to polycrystals
because the random distribution of orientations forces to multiple slip in the
crystallites. Likewise, the critical shear stress
I
t at the beginning of sliding is higher
than for the single crystal because of the statistically distributed grain orientations.
If one converts the critical shear stress into a critical normal stress (using the Taylor
factor of 3.06) the result will be the yield point of the polycrystal:
0 0
06 . 3 t o = (2.8)
assuming the same frequency of all orientations relative to the external stress o.
However, this value does not yet consider the finite grain size. Strictly speaking it
only applies to a polycrystal of infinite large grain size, resp. for a medium
orientation of the sliding systems of a single crystal. However, the grain size itself is
also of influence on the yield stress.
A fundamental difference to the single crystal lies also in the fact that in the case of
the polycrystal the deformation of each grain has to be compatible with that of the
neighbouring grains. In contrary to single crystals, where the dislocations which exit
at the free surface of the crystal may dissolve themselves easily in glide steps, the
deformation of a single grain has to act in such a manner that no holes or cracks
develop at the grain boundaries. Thus, the traversing dislocations pile up at grain
boundaries, forming far reaching stress fields. Therefore, the grain boundaries
interact with the dislocation movement and exert a stress which is the more effective
the shorter the running path, i.e. the smaller the grain diameter. This is expressed in
the experimentally well confirmed (Hall-Petch-) relation:
d
k
R
0 e
+ = o (2.9)
e
R is the measured yield point,
0
o the average single crystal yield point, k a
material constant and d the average grain diameter.
Consequently, the yield point of pure metals (i.e. of a homogenous phase) results
from the stress
0
o , necessary for the movement of dislocations, a portion
1
o A which
expresses the impedement of the movement by other dislocations, and a part
2
o A
which indicates the impedement by grain boundaries.
- 26 -
d
k
b G R
0 2 1 0 e
+ o + = A + A + = o o o o (2.10)
(, k := constants)
Additional external influences which may modify the yield point of pure metals are
mainly the temperature and the loading velocity. Thus,
e
R always increases with
decreasing temperature and increasing deformation velocity to different degrees,
depending on the various types of lattices. For example, the yield point of bcc metals
is stronger temperature dependent as compared to fcc metals.
2.3 Strength hypotheses
When a component is subjected to multiaxial loading conditions, the point of failure of
the material has to be determined. The material parameters of uniaxial loading can
be applied to the component, when the multiaxial stress state can be transferred into
a single effective stress that is equivalent to the multiaxial stress state:
Uniaxial strength condition: all
o o s
= s
o
o o
D
F
m
Y
e
all
/S
/S R
/S R
W
+ +
uniaxial uniaxial
(Tension, pressure, bending) (Tension test, fatigue test)
Multiaxial state:
3
2
1
failure
of theory
v
o (effective stress)
+ +
triaxial virtually uniaxial
Here, o
1
, o
2
and o
3
are the principal stresses ordered as o
1
> o
2
> o
3
. Several
theories of failure have been developed to calculate the effective stress o
v
. The
theory used as a basis for the calculation depends on the mode of failure (brittle or
ductile). The inner cause of failure sliding, separation or disruption yield results in
R
e
= Yield stress
R
m
= Tensile strength
o
W
= Cyclic strength
S
Y
, S
F
, S
D
= Safety factors:
S
Y
= 1,2 2 (1,5), yielding
S
F
= 2 3 (2), ductile material failure
S
F
= 4 9 (4), brittle material failure
S
D
= 1,5 2,5 (2), cyclic loading
- 27 -
macroscopically different modes of failure: Plastic flow, brittle fracture or fatigue
failure. If the mode of failure is unknown, the effective stress has to be calculated
with several possible theories of failure.
Multiaxial strength condition: all
v
o o s
(2.11)
The primarily used theories of failure are the maximum normal stress theory,
maximum shear stress theory and the maximum distortion energy theory.
2.3.1 Maximum normal stress theory
The decisive stress used in this theory is the maximum normal stress that appears in
the material. This theory assumes that failure appears by means of brittle fracture,
normal to the direction of the maximum normal stress o
max
. Hence the theory is
primarily used for inherently brittle materials.
Maximum normal stress theory:
1 max v
o o o = = (2.12)
2.3.2 Maximum shear stress theory (Tresca criterion)
This theory assumes the highest load of the material to be caused by the highest
shear stress. The toughness limit for multiaxial loading is reached, when the
maximum shear stress in the material is equal to the maximum shear stress under
uniaxial loading.
The maximum shear stress theory assumes that failure occurs as shear failure; it is
applied for ductile materials that fail by shear fracture. Hence the theory is primarily
used for ductile materials.
Maximum shear stress theory:
max 3 1 min max v
2t o o o o o = = = (2.13)
2.3.3 Maximum distortion energy theory (von Mises criterion)
The maximum distortion energy theory uses the distortion energy stored in the
material as the decisive criterion for failure. This criterion is independent of the stress
condition and relevant as soon as the distortion energy obtains a material-dependent
value, failure occurs as plastic deformation. Hence the theory is primarily used for
ductile materials.
Maximum distortion energy theory:
( ) ( ) ( )
2
1 3
2
3 2
2
2 1 v
2
1
o o o o o o o + + = (2.14)
- 28 -
Equivalent stresses according to the maximum normal stress theory, the maximum
shear stress theory and the maximum distortion energy theory for 2D and 3D stress
states are summarized in Table 2.1
Equivalent stresses
Stress state
Theory of
failure
triaxial (3D) biaxial (2D)
3 2 1
; ; o o o
)
3 2 1
0 ; ;
+
= o o o
xy y x
; ; t o o
t b
;t o
Maximum normal
stress theory
1
(Tension)
3
(Compression)
1
o
( )
2
xy
2
y x
y x
4
2
1
2
t o o
o o
+ +
+
2
t
2
b
b
4
2
1
2
t + +
Maximum shear
stress theory
(Tresca criterion)
3 1
o o
1
o
( )
)
2
xy
2
y x
4
++
t o o +
2
t
4
2
b
t + o
Maximum distortion
energy theory
(von Mises criterion)
( ) ( ) ( )
2
1 3
2
3 2
2
2 1
2
1
o o + o o + o o
2 1
2
2
2
1
o o o + o
2
xy
3
y x
2
y
2
x
t + o o o + o
2
t
3
2
b
t + o
( ) ( ) stress normal Maximum Tresca if ; if valid only
(3D)) (triaxial 1 column to according 0 ; 0 ; 0 and 0 ; 0 ; 0 cases the For
V V
2
xy y x
2
xy y x
)
V 3 2 1 3 2 1
)
o o t o o t o o
o o o o o o o
= > s
< < = < = >
+ +
+
Table. 2.1: Theories of failure
- 29 -
2.3.4 Example of application
Thin-walled hollow cylinder under internal pressure.
Tangential stress:
Axial stress:
Radial stress:
r
o =
2
p
Normal stresses in the planes of symmetry are principal stresses:
1 t
o o =
2 a
o o =
(2.17)
3 r
o o =
Assumption:
Stress is independent of
wall thickness s.
Tolerable for: 2 , 1 d / d
i a
s
d
i
= inner diameter
d
a
= outer diameter
2sb b pd
t i
= o
2s
pd
i
t
= o (2.15)
2
i
d
4
p
t
= ( )
i m m
d d s d
l
~ t o
l
o =
s 4
pd
i
(2.16)
o
t
o
t
o
a
o
a
o
a
- 30 -
2.3.5 Example for calculation
Dimensions: mm 500 d
i
= , mm 25 s = ,
2
mm / N 10 bar 100 p = =
2
N/mm 100
25 2
500 10
1 t
=
= = o o
2
2 l
mm / N 50
25 4
500 10
=
= = o o
2
3 r
mm / N 5
2
10
= = = o o
Effective stress:
Normal stress criterion:
2
1 v
mm / N 100 = = o o
Shear stress criterion:
2
3 1 v
mm / N 105 ) 5 ( 100 = = = o o o
Distortion energy criterion:
2
1 3
2
3 2
2
2 1 v
) ( ) ( ) (
2
1
o o o o o o o + + =
2 2 2 2
v
mm / N 91 ) 105 ( 55 50
2
1
= + + = o
DIN 2413: Calculation of steel pipes (shear stress criterion)
3 1
o o o =
v r t
o o =
|
.
|
\
|
+ =
|
.
|
\
|
=
1
s
d
2
p
2
p
s 2
d p
i
i
s
s d
2
p
i
+
=
s 2
d p
m
C
0
400
500
600
700
1
2
Fig. 3.5: Precipitation hardening of an AlCuMg-alloy
In the case of slow cooling from the field (1) Cu diffuses after transgressing the
solubility of the -solid solution crystal and forms the intermetallic phase Al
2
Cu
(equilibrium condition (2) at RT). In case of fast cooling the diffusion will be
- 41 -
suppressed and an Al-lattice develops which is over-saturated with Cu. This
increases already strongly the toughness. However, the real hardness increase takes
place during temperature ageing (cold hardening) or - even better - accelerated at
120 175C (warm hardening) by the subsequent precipitation of Al
2
Cu in
submicroscopically fine distribution (one-phase decomposition). The hardened
condition is a deadlocked interim state in direction of the thermodynamic
equilibrium.
The extent of the strength increase is a function of magnitude, shape, distribution and
properties of the precipitated particles. Of further influence is the type of the
interaction between dislocations and precipitated particles. There are two principally
different possibilities:
1. The particles can be cut (sheared) by the dislocation (cutting mechanism, see
fig. 3.6a). They are included in the plastic deformation.
2. The particles have to be obviated so that dislocation rings develop around the
particles similar to the Frank-Read-Source. These dislocation rings are not
participating in the plastic deformation (Orowan mechanism, bypassing-
mechanism, see fig. 3.6).
t
t
a)
b)
Fig. 3.6: Schematic of the cutting mechanism (a) and of the bypassing mechanism
(b) according to Orowan
The increase in strength is much higher in the bypassing mechanism than in cutting
because new dislocations are developing causing additional hardening of the glide
- 42 -
planes. New dislocations are not nucleated during cutting and since the cross section
of the particles reduces, it is even possible that the obstacle effect may be reduced.
The fact whether the particles precipitate coherent (oriented to the lattice) or
incoherent is also of influence. Incoherent particles prevent the movement of
dislocations much more and increase the yield limit more than coherent ones which
are closer to the properties of homogeneous solid solutions. The bypassing
mechanism prevails mainly at large distances and high hardness of the precipitated
particles while in case of small distances and soft particles the cutting mechanism is
dominating.
The increase in the yield limit caused by precipitated particles is given in a first
approach by the relation :
1
P
3 / 1 1
P
d f b G
~
D b G
~
o ~ o = Ao (3.5)
Where
~
is a constant, G the shear modulus, b the Burgers vector, D the distance of
the particles in the original lattice and f and
P
d are the volume fraction and the
diameter of the particles, resp. (
3
P
f D / d ~ ).
The precipitation hardening is of great importance for non-ferrous metals which are
not subjected to transformation hardening (e.g. Al-alloys in aircraft construction, Ni-
alloys for turbine blades). But the precipitation processes are also important to steel
in a positive or negative sense. Precipitates show a negative effect if they cause
embrittlement. This can be the case during ageing where so called blue-brittleness
is caused by nitrogen-precipitates or during tempering following hardening,
quenching and annealing (temper embrittlement, stress relief cracking). From a
positive point of view precipitates are employed, for example, in heat-resistant steels
and in newly developed iron alloys such as the micro alloyed building steels and the
so-called martensite-hardening steels.
The micro-alloyed steels contain little or no perlite and they contain alloying elements
such as Ti, V, Zr, Nb in magnitudes of approx. 0.1 %. They develop carbides and
carbon nitrides which act as precipitation particles as well as crystallisation nuclei.
This combination of precipitation hardening and grain refinement is done most
effectively under thermo-mechanical treatment (combined rolling and heat
treatment under well defined conditions).
- 43 -
The martensitically aged steels are nearly carbon-free with high content on Ni (may
be replaced by Co), Al, Mo, Ti and Nb. Example: X2NiCoMo18 8 5 (0.02% C,
18% Ni, 8% Co, 5% Mo+Ti). Following quenching from austenitising a not very hard,
martensite structure with bcc structure remains (not C-martensite as in steel,
martensite development generally = diffusionless phase transformation in a solid
state, mostly together with a change in volume). Tempering at 450 500 C causes
in the martensitic base lattice inter-metallic phases such as Ni
3
Ti, Ni
3
Mo and FeAl
effectuating the actual increase in hardness. Martensite aged steels (Maraging
steels) possess very high tensile strengths ( MPa 2400 1150 R
m
= ) together with
high yield strengths ( 94 . 0 R / R
m e
~ ), good toughness and weldability.
The particle hardening is also used in austenitic steels, e.g. by precipitating Ni
3
Al in
Fe-Ni-Al-alloys. Finally, the nitration of steels should be mentioned. The nitrogen,
which is diffusing into the steel causes the formation of nitrides together with Al and
Cr dissolved in the nitrogen steels. These nitrides precipitate in submicroscopically
fine distribution, causing an extremely high hardness of the nitride layer.
3.6 Combined methods
It can be concluded from the basic mechanisms of strength increase that, a high yield
point can be expected if one succeeds in
- increasing the concentration on dissolved atoms,
- increasing the dislocation density,
- reducing the grain size and
- maintaining small distances between the particle (see fig. 3.7).
1/2
0
R
e
0
0
P
Mo
Cr
0
R
e
0
0
C
0
d
-1/2
R
e
0
0
f =
0
,1
d
T
-1
R
e
0
0
Fig. 3.7: Hardening of o-Fe through different basic mechanisms (schematic)
1/2
- 44 -
There were attempts to combine the different methods to obtain a specially intensive
increase in strength. The most prevalent example for such a combination is the
hardening of Fe-C-alloys by martensitic transformation (steel hardening). Solid
solution hardening acts by interstitial C-atoms together with a large number of
dislocations (
2 10
cm 10
~ ), caused during the martensite formation. The yield limit
results in
SS D 0 e
R o o o A + A + = (3.6)
In the alloyed (e.g., heat-resistant) steels one can take advantage of the precipitation
hardening
P SS D 0 e
R o o o o A + A + A + = (3.7)
In fine-grained steels the grain boundary influence has to be added to the influence
of martensitic transformation (if they are alloyed, possibly also the precipitation
hardening). The yield point results then in
P GB SS D 0 e
R o o o o o A + A + A + A + = (3.8)
In alloys where no lattice transformations occur combinations of mechanical
strengthening by dislocations is tried to combine with precipitation hardening (thermal
mechanical treatment). The yield limit is then given as
P D 0 e
R o o o A + A + = (3.9)
However, it is only partly possible to obtain an effective combination of various
hardening mechanisms because of mutual interactions of the different mechanisms.
Therefore, it is only possible to a certain extent to predict quantitatively an increase in
the yield limit as a simple addition of the single contributions.
- 45 -
4 Influences on the material behaviour
The term Behaviour of Materials circumscribes the reaction of the material to an
external, i.e. externally applied, load. The reaction will be deformation and finally
fracture. Therefore, the term material behaviour comprises the deformation and the
fracture as function of load.
If subjected to mechanical loading the material behaviour is mainly a function of
- Specific material properties
- Stress state
- Temperature
- Loading velocity.
These influences shall be examined more closely in the following.
Thereby the metal physical aspects which are related to the atomic composition of
the material are not of priority for these factors of influence but the phenomenological
ones, i.e. the obvious phenomena which characterise the material behaviour.
4.1 Specific material properties
The specific properties of a certain material can be obtained in the tensile test in the
form of a stress-strain-diagram. This is shown in fig. 4.1 where diagrams of different
materials are given representative for various groups of materials. The influence of
the material particularities on the stress-strain-curves is reflected in these diagrams.
First of all one can differentiate between tough and brittle materials. Tough materials
possess an extreme plastic deformability whereas the brittle materials (fig. 4.1, cast
iron) deform nearly purely elastic until fracture. For the stress it is found that with
increasing strength, i.e. with increasing deformation resistance the toughness resp.
the deformability decreases. Austenitic steels represesent a peculiarity in that they
possess a relatively high tensile strength and a very strong hardening behaviour and
a high deformability.
- 46 -
1400
1000
800
600
400
200
S
t
r
e
s
s
/
M
P
a
o
Strain / % c
spring steel
high alloy
steel
fine grain
building steel
cast
iron
austenit. steel
0 10 0 10 20 20 0 10 20 30 40
general
building steel
Fig. 4.1: Stress-strain-diagrams of various materials
The stress-strain diagram characterises the strength and deformation behaviour of a
certain material. One can also obtain completely different stress-strain-curves for the
same material, depending on the state of this material. This can be seen in fig. 4.2,
where the stress-strain curves of material E335 are shown for an annealed,
quenched and tempered, and a hardened condition. It can be seen that the
deformability is reduced the more an appropriate heat treatment increases the
strength.
1400
1000
800
600
400
200
0 5 10 20 15
Strain / % c
S
t
r
e
s
s
/
M
P
a
o
tempered
annealed
E335
hardened
Fig. 4.2: Stress-strain-diagrams for various material states of boiler steel E335
- 47 -
The material properties can be changed by additional alloying elements within wide
ranges and thus be adjusted to certain purposes. The most important accompanying
element in iron is carbon. It is essential to obtain sufficient strength (and, therefore,
C is not really considered as an alloying element). This can be seen in fig. 4.3 where
the dependence of the strength and deformation characteristics of unalloyed steels
on carbon is shown.
R , R
e p0.2
Z
A
10
800
600
400
200
0
R
,
R
/
M
P
a
m
p
0
.
2
A
,
Z
/
%
1
0
0
20
40
60
80
0,1 0,2 0,3 0,4 0,5 0,6 0,7
C- Content / %
Fig. 4.3: Influence of carbon content on material properties (schematic)
Deformation hardening is also able to change considerably the material properties,
fig. 4.4. The deformability is then reduced due to the anticipated plastic deformation
and the strength increases accordingly.
80%
40%
20%
0%
cross section reduction
0
10 20 30
0
200
400
600
800
1000
Strain / % c
S
t
r
e
s
s
/
M
P
a
o
Fig. 4.4: Influence of pre-deformation on the stress-strain-diagram (steel with
0,07 % C)
- 48 -
4.2 Stress state
The stress state developing in a component as a reaction to the external load is
determined explicitly by the three principal stresses. Depending on their presence,
one differentiates between uniaxial, biaxial or triaxial stress states. The stress state is
called triaxial or spatial if all three principal stresses are different from zero. The
stress state is biaxial or plane if one, and uniaxial, if two of the three principal
stresses are equal to zero. Uniaxial stress states occur generally in rod-like
components. Biaxial stress states dominate mainly in plane components and three-
axial stress states may develop in expanded (e.g. thick-walled) components.
A very clear graphics of the stress state is possible with the aid of Mohrs stress
circles (Christian Otto Mohr 1835 1918; in Stuttgart 1867 1873). As an example,
a triaxial stress state is shown in fig. 4.5. Each of the three circles represents the
stress state in the planes given by the two appropriate principal stresses.
t
12
t
13
t
23
t
o
1
o
o
2
o
3
o
2
Fig. 4.5: Mohrs stress circle for a triaxial stress state
Each point of a circle characterises the normal stress
t of
a certain plane defined by the angle .
- 49 -
R
e
o
1 o
3
=0
t
t
t
o
o
o
o
T
t
F
t
max
o
F
o
1
o
3
o
T
t
F
t
max
o
1
o
3
t
F
o
T
Fig. 4.6: Influence of stress state on material behaviour (schematic)
The influence of the stress state on the material behaviour can be very well explained
by Mohrs stress circles, as shown in fig. 4.6. One can draw into the
F
t t = (start of
yield) and
T
o o = (fracture) into the t o diagram of Mohr.
T
o is the critical normal
stress which causes a local material separation which arises as a crack or as
fracture. If the material is brittle
T
o coincides with the tensile strength
m
R . However,
T
o is considerably higher (tensile test) for tough materials. The straight line
F
t t = corresponds to the shear stress hypothesis (SH). In accordance with this
hypothesis plastic deformations take place if the highest shear stress reaches a
critical value, the shear yield stress
F
t . The straight line
T
o o = is in agreement with
the normal stress hypotheses, which predicts fracture as soon as the highest normal
stress
1
o attains a critical value
T
o . Consequently, plastic yielding and fracture may
occur if the largest Mohrs stress circle touches one of these two limit lines.
- 50 -
A deformable material flows in a uniaxial or biaxial stress state, fig. 4.6a, if
F max
t t =
resp.
e 1
R = o . The same material starts yielding under a triaxial tensile stress state,
fig. 4.6b according to the same criterion
F max
t t = just at a higher flow stress. If the
stress state is finally strongly triaxial, with
F max
t t < , fracture occurs without plastic
deformation for
T 1
o o = , see fig. 4.6c.
The observation of these processes (which have only been schematically recorded in
fig. 4.6) results in the so-called Mohrs enveloping parabola shown in fig. 4.7.
brittle fracture
t
F
t
max
o
T
o
1
o
3
t
o
o o
1 3
+
2
yielding before fracture
0
fra
c
tu
re
yield
Fig. 4.7: Mohrs enveloping parabola
This is a limit curve comprising all Mohrs stress circles which are about leading to
yield and/or fracture. The limit line for leading is as before given by the straight line
F
t t = .
If a stress circle tangents firstly this straight line yielding occurs prior to fracture. If it
touches the curvature first, fracture takes place without previous plastic deformation.
The intersection between the straight yielding line and enveloping parabola
separates the deformation fractures from the brittle fractures. Concerning brittle
fractures, the lattice planes are separated under the effect of normal stresses. The
deformation fractures are generated by sliding of lattice planes against each other for
which shear stresses are responsible. The shape of the enveloping parabola which is
open towards the compression side is based on the idea that the critical shear
- 51 -
stresses causing shear fracture (shear strength of the material) are affected by the
normal stresses which simultaneously effect the sliding plane. For a certain stress
circle the highest shear stress is given as
2 / ) (
3 1 max
o o t = (4.1)
Perpendicular to the planes, which are stressed by shear stress
F
t the normal stress
is given as
2 / ) (
3 1
o o o + = (4.2)
This normal stress has the effect of an internal friction: if tensile stresses act
perpendicular to the highest loaded sliding planes the sliding strength of the material
will be reduced while it will be increased in case of compressive stresses.
The shape of the enveloping parabola is depending on the type of material (tough /
brittle). The parabola is a graphic aid to explain the material behaviour. It can not be
used as an appropriate quantitative fracture criterion because its exact determination
encouters difficulties.
The theoretical dependence of the material behaviour on the stress state as it is
shown in fig. 4.6 and fig. 4.7 is expressed in the stress-strain-diagrams of the
material concerned. An example is given in fig. 4.8 showing the stress-strain-
diagrams of a smooth and three notched tensile bars with different shape numbers
k
o from material C45. A three-dimensional inhomogeneous stress state dominates
in the notched bars and
k
o characterises the degree o f inhomogeneity. This applies
also approximately to the multiaxiality because the inhomogeneity and mulitaxiality of
the stress states are connected with each other in the notched bars. It can be seen
that the average yield limit and the achieved ultimate load which can be expressed
as notch tensile strength
n max mn
A / F R = , is the higher and the fracture strain is
lower the more sharp the notch is, i.e., the stronger the inhomogeneity and the
multiaxiality of the stress state. The triaxial tensile stress state inside the notched
bars obstructs the yielding, reduces the plastic deformability and simultaneously
increases the load carrying capacity.
- 52 -
1000
900
800
700
600
500
400
300
200
100
0 5 10 15 20
o
k
= 2,91
1,57
1,29
1,0
8
t =25
16
Strain = l/l c A
0
/ %
C45 (annealed)
S
t
r
e
s
s
=
F
/
A
/
M
P
a
o
K
Fig. 4.8: Stress-strain-diagram for different notch strengths = = o o o
n max k
/ ( shape
factor with elastic stresses; = o
max
maximum stress at the notch
ground,
n n
A / F = o )
4.3 Temperature
The mechanical properties in metallic materials are subject to different changes if
one exchanges from the normal environmental temperature (ambient temperature,
approx. 20 C) to higher and lower temperatures, respectively.
4.3.1 High temperatures
With increasing temperature the strength decreases. The values characterising yield
strength,
2 . 0 p e
R and R , respectively, and strength R
m
as well as the deformation
characteristics A
5
and Z of three materials are shown as a function of the
temperature in fig 4.9. An exception is the temperature range around 250 C in steels
with low carbon content. The tensile strength can once more pass through a
maximum whereas the deformations take their lowest values in this regime. This
phenomenon is known as blue-brittleness and caused by precipitating nitrogen in
the microstructure of the material.
- 53 -
X6 CrNiMo 17 13
21 CrMoV 5 11
St 35.8
R
m
R
m
R
p0.2
R
p0.2
R
,
m
R
/
M
P
a
p
0
.
2
Z
A
5
Z
,
A
/
%
5
1000
600
400
200
0
100
80
60
40
20
0 100 200 300 400 500 700 600
Temperature / C
800
Fig. 4.9: Temperature dependency of material properties
Crystal recovery and recrystallization occur in a deformed microstructure above
certain temperatures which are characteristic for each metal. Crystal recovery can be
explained as recovery from hardening caused by cold deformation. Recrystallization
means the new formation of the microstructure from the deformed microstructure.
Therefore, the material cannot harden anymore during deformation in the
temperature range of crystal recovery and recrystallization. Once initiated yielding
continues more or less as long as the load acts at the same level: the material
creeps.
Consequently, the crystal recovery temperature is an upper limit for the applicability
of strength values obtained in short term experiments at room temperature or at
enhanced temperatures - however, below the crystal recovery temperature. Yield
stress at temperature 0,
0 / 2 . 0 p
R and the strength at temperature 0,
0 / m
R are
important material properties up to these temperatures (with increasing temperature
- 54 -
a pronounced yield limit disappears). They are determined as in a normal tensile test.
The tensile specimen is heated up to the required temperature 0 in an oven and
then fractured. For loading above the crystal recovery or recrystallization temperature
new characteristic material properties have to be determined in long term tests. They
shall take the influence of the loading time on the strength behaviour into account.
4.3.2 Low temperatures
The deformation resistance of metallic materials and consequently also the yield limit
and the 0.2% strain limit, respectively, as well as tensile strength increase more or
less with decreasing temperature. Therefore, components which have been
sufficiently dimensioned for loading at room temperature should accordingly tolerate
the same loading at low temperature (although there are no special strength material
properties available for low temperatures). Although this is theoretically correct it only
applies for the circumstances of technical standards. This is due to the deformation
behaviour of the materials: Whereas, for one part of the metals the deformation
characteristics (such as fracture strain) remain almost constant or even increase
slightly the deformation ability at low temperatures drops suddenly for another part.
This very often sudden drop in deformation properties occurs preferably in metals
with body centred cubic lattices ( o-iron, unalloyed and low-alloyed steels) and with
hexagonal lattices (tin, magnesium), but not in such ones with face centred cubic
lattices ( -iron, austenitic steels, aluminum, copper, nickel). Fig. 4.10 demonstrates
this for the example of three materials of which their strength and deformation
characteristics is shown as a function of temperature.
The sudden loss in the ability of plastic deformation of certain materials at low
temperatures is of importance to the strength behaviour of technical components
because of the danger of a brittle fracture. The highest stress arises mostly as a
more or less marked peak of which the magnitude is so bound during the strength
calculation of components of tough material that this value remains below the yield
point of the material with a certain additional safety.
- 55 -
R
m
R
e
A
5
Z
Z
R
p0.2
R
m
A
5
1600
1200
800
400
0
100
80
60
40
20
0
-200 -160 -120 -80 -40 0 40
Temperature / C
D
e
f
o
r
m
a
t
i
o
n
/
%
S
t
r
e
s
s
/
M
P
a
A
5
Ti Al 6 V 4
X 2 NiCr18 16
S235
R
p0.2
Fig. 4.10: Strength and deformation properties of different materials at low
temperatures
Nonetheless it can be the case that a stress peak reaches the yield point or exceeds
it by short term over-loading (shocks) or as a consequence of residual stresses (as
they appear in weldments) which, however, can hardly be taken into account. Such
stress increases above the theoretical maximum allowable stress value cause
generally no damage as long as plastic deformation of the material is sufficiently
possible. As soon as a stress peak reaches the yield limit of the material it cannot get
larger but effects a (mostly locally limited) plastic deformation of the material causing
the original peak of the stress distribution to flatten. The deformation ability of a tough
material is normally sufficient at room temperature to carry such unexpected stress
peaks by local plastic yielding. However, if the ability of plastic deformation has been
lost, as it may be possible in materials with body centred cubic and hexagonal
lattices at low temperatures the same load may cause a brittle fracture.
- 56 -
The critical temperature at which the embrittlement initiates is not a constant for a
component of a certain material but depends on various factors. With increasing
notch sharpness, increasing deformation velocity, by previous cold deformation
and/or unfavourable operational temperatures (ageing) the critical temperature is
shifted to higher temperatures up to room temperature or above. The temperature
which causes the change from ductile fracture to a brittle one is called transition
temperature and is determined in the notch impact bending test: A notched specimen
is fractured by a pendulum hammer. The required impact work
v
A may be taken as a
measure for the deformation ability under difficult conditions. The notch impact
energy possesses principally the same temperature dependence as the other
deformation parameters. However, the strong drop appears already at higher
temperatures because of the unfavourable test conditions.
The loss in toughness of certain materials at low temperatures is not directly taken
into account by the strength calculation because the allowable stress is determined
from the strength properties and not from the deformation properties. Therefore, if
there is a risk of temperature embrittlement in a component according to the
operation conditions this has to be taken into consideration in the selection and
treatment of the material.
4.4 Loading velocity
The resistance of deformation is influenced by the loading velocity because the
plastic deformation requires a certain time after exceeding the yield limit, much more
than the pure elastic one. As shown in fig. 4.11 the strength parameters
e
R and
m
R
increase with increasing loading velocitiy, the yield limit more than the tensile
strength. Cross section reduction Z and fracture strain
5
A are less influenced by the
loading velocitiy, they remain constant or decrease only slightly. Therefore, the
fracture work increases for low and medium loading velocities. When yield limit
e
R
and tensile strength approach the same values there will result a strong drop in the
deformation values
5
A and Z and consequently the material fails in a brittle fracture.
Generally this occurs only at low temperatures. In fig. 4.12 it is demonstrated that in
impact tensile tests with Armco iron the transition temperature for fracture increases
with increasing impact velocity. Mainly unstabilized steels are susceptible to impact
like loads (ageing embrittlement).
- 57 -
R
m
R
m
R
e
R
e
Z
Z
A
5
A
5
Loading velocity v
Fig. 4.11: Influence of deformation velocity on material properties (schematic, for
unalloyed building steels, temperature around transition temperature)
ARMCO-iron
annealed
25 m/s 36 58
70
60
50
40
30
20
10
0
-80 -60 -40 -20 0 20
Temperature / C 0
C
r
o
s
s
s
e
c
t
i
o
n
r
e
d
u
c
t
i
o
n
Z
/
%
0
trans
= -80 C
Influence of loading velocity on transition temperature of cross section
reduction
- 58 -
5 Tasks
Task 1: Lattice Defects
a) What kind of lattice defects do exist?
b) Explain the impact of defects on the mechanical behaviour of metals.
Task 2: Energy of Dislocations
a) Derive the relationship
|
|
.
|
\
|
t
=
0
1
2
r
r
ln
4
l b
G W for the energy of a screw dislocation.
b) How large, approximately, is the energy of a screw dislocation of 0.1 mm length in
body centered Fe? (G = 83100 N/mm
2
, a = 0.287 nm)
Task 3: Dislocations
a) What type of dislocations do exist?
b) Explain the difference between a complete dislocation and an incomplete (partial)
dislocation.
c) What determines the process of dislocation reactions?
d) Explain the impact of dislocations on the shear stress required for plastic
deformations.
- 59 -
Task 4: Burgers Vector
a) Explain the meaning of a Burgers vector b.
b) What kind of dislocation is it if the angle between Burgers vector and dislocation
line amounts to 0, 45, 90 degrees, respectively?
c) Where is the energy of a dislocation coming from and how is it related to b?
Task 5: Glide Systems, Hardening
a) Explain the expression glide system.
b) Explain the reasons for hardening of metallic materials (in area II of the (t, ) -
curve)
Task 6: Dislocation Hardening, Bauschinger Effect
a) Explain the effect of dislocation hardening.
b) What should be considered when cold deformation is applied for strengthening of
metallic materials?
c) Explain the Bauschinger Effect (BE).
d) What is the reason for the BE?
e) How can the BE be avoided or reduced?
- 60 -
Task 7: Precipitation Hardening
a) Which prerequisites have to be fulfilled with respect to precipitation hardening?
b) What provides the hardening in precipitation hardening materials? What types of
interactions do appear?
c) Give a few examples for unexpected precipitation hardening.
d) What are the mechanisms related with transformation hardening of steel?
Task 8: Behaviour of metallic materials
a) What are the relevant influencing factors with respect to materials behaviour under
mechanical loading?
b) Use Mohrs circle to explain the influence of the stress state on plastic
deformability and fracture of a ductile component.
Task 9: Properties of metallic materials
a) Depict the qualitative behaviour of R
e
, R
m
, A
5
and Z as a function of T (<RT) for an
unalloyed (bcc) steel.
b) Why is the transition temperature of the notch impact energy for unalloyed steels
much higher as compared to the transition temperature of the strain (A
5
) and area
reduction (Z) at fracture?
c) Explain the expression blue brittleness. What is the reason for it?
- 61 -
6 Solutions
Task 1: Lattice Defects
- 62 -
Task 2: Energy of Dislocations
- 63 -
- 64 -
Task 3: Dislocations
- 65 -
Task 4: Burgers Vector
- 66 -
Task 5: Glide Systems, Hardening
- 67 -
Task 6: Dislocation Hardening, Bauschinger Effect
- 68 -
Task 7: Precipitation Hardening
- 69 -
Task 8: Behaviour of metallic materials
- 70 -
- 71 -
Task 9: Properties of metallic materials