Вы находитесь на странице: 1из 5

MATE1000 Lecture 11 Mechanical Properties of Metals

(Callister Chapter 6, p 112-134; Chapter 8, p 209-232)

Metals deform by slip the sliding of one block of material over another. This manifests
itself in the formation of slip steps on the surface of the material. Slip occurs in both single
crystals and in polycrystals (an assembly of a number of grains or crystals) and results in
plastic or permanent deformation of the material. Slip is the result of the movement of
line defects called dislocations in the crystal. It is much easier to cause slip via dislocation
movement than it is to slip (slide) one block of material over another in one hit.

The Stress-Strain Curve


One of the most important mechanical properties of a material is its strength the ability of
the material to withstand stress without failure. Strength can be measured in a variety of
loading situations tension, compression, torsion, etc. When materials are deformed in
simple tension, a typical stress-strain curve results:

The important features of this stress-strain curve are:- as the stress (load) is increased, the
specimen initially deforms elastically, then plastically (permanently) in a uniform fashion,
with work hardening (stress increasing with the amount of plastic strain), until it reaches a
point (the UTS) when the deformation becomes non-uniform (necking commences), the
engineering stress falls and finally the specimen fractures.

Definitions
Elastic region where stress is proportional to strain Hookes Law is obeyed and the slope
of the stress-strain curve (/) is equal to Youngs modulus. No permanent deformation.

1
Yielding onset of plastic (permanent) deformation, where the strain does NOT recover
on unloading. First point of departure from truly elastic behaviour.
Yield strength stress at which yielding or plastic deformation begins. Used for stress-
strain curves that show a sharp transition from elastic to plastic behaviour.
Proof strength stress after a measurable amount of plastic deformation eg 0.1% proof
stress. Used for materials that show a gradual transition from elastic to plastic behaviour.
Work hardening region of plastic deformation where the material gets harder (ie.
stronger) with increasing plastic strain.
Tensile strength Also called ultimate tensile strength (UTS). Maximum engineering
stress (maximum applied load) achieved during the test. Coincides with the transition from
uniform to non-uniform plastic deformation.
Total elongation to failure nominal strain to failure, after correcting for any elastic
component of the strain. Also known as the tensile ductility.
Necking the non-uniform deformation behaviour that begins at the UTS and leads to the
formation of a neck or waist in the specimen. Since the stress is greater in this region of
reduced cross-sectional area, the deformation is concentrated in the neck, the deformation
process becomes unstable (plastic instability) and rapidly leads to failure. During necking
the engineering stress (load/original area) falls, but the true stress at the neck (load/actual
area) is still increasing. So, the apparent decrease in load after necking commences is not
an indication that the material is getting weaker, but that the cross-sectional area available to
support the applied load is decreasing.
Fracture stress the load at fracture divided by the cross-sectional area of the fracture.

For a tensile specimen of original gauge length L0 and cross-sectional area A0 subjected to a
load F;-
Engineering stress: = F/A0 Engineering strain: = dL/L0

At any point in the test where the length of the specimen is L and the cross-sectional area is
A, then:
True stress: T = F/A True strain: T = ln(L/L0)

If the final length of the broken specimen is LF and the cross-sectional area of the fracture is
AF, then:
Total elongation to failure: F = (LF L0)/L0
Total reduction in area = (A0 - AF)/A0
Fracture stress = true stress (T) at fracture = (load at fracture)/AF

Note the distinction between stress and strength. These two terms are sometimes
interchanged incorrectly! Stress is a result of an applied load ie. it is something done to
the material. Strength is a property of the material itself and is an indication of how well
the material can withstand stress.

2
Ductility and Toughness
The two materials properties ductility and toughness are related, but strictly speaking
they refer to different things. The confusion arises because the same term brittle - is used
to describe the opposite of both ductile and tough.

Ductility is a measure of how much


the material deforms before fracture
and is given by the total elongation
to failure (F) or the total
reduction in area. In the example
on the right material A is ductile,
material B is not very ductile ie. it
is brittle.
Ductile materials often have low
strengths, but there are examples of
low strength materials that are also
brittle eg. polystyrene foam.

Toughness is a measure of the work


done in fracturing the specimen it
is the area under the stress strain
curve. Toughness depends on both
the elongation to failure and the
strength of the material. In the top
example on the right material A is
tougher than material B, as well as
being more ductile. In the bottom
example on the right material A is
more ductile than B, but the
toughness of the two materials is the
same.

Polycrystalline Materials
Polycrystalline materials are stronger than single crystals because the different orientations
of adjoining grains make it difficult to transmit slip (deformation) from one grain to the next.
To maintain coherency between grains in a polycrystalline material (no sliding along the
grain boundaries, no voids or holes at the grain boundaries), each grain must deform to a
shape that is dictated not only by the applied stress, but also by the changes in shape of the
neighbouring grains. This requires each grain to have at least 5 independent slip systems
the Von Mises criterion.
Cubic metals (Al, Cu, Fe) normally satisfy this criterion and deform relatively easily (low
yield stress, ductile). Hexagonal metals (Zn, Mg, Ti) often have less than 5 independent slip
systems that can operate easily (at low stresses), so they tend to be brittle and difficult to
deform.

3
Fatigue
Fatigue is the failure of a material under conditions of fluctuating load. The majority of
metal component failures are due to fatigue. Fatigue failures start with the formation of a
crack at the surface (at a slip step or a region of stress concentration) and the crack grows
incrementally with successive load cycles. Eventually the crack becomes so large that the
remaining cross section of intact material is unable to withstand the applied stress and failure
occurs.
Fatigue failure is more likely at high
stress amplitude (S) and a large number of
cycles (N) of the fluctuating load. Hence
the representation of fatigue behaviour
using S/N curves see right.
Steels (ferrous metals and alloys) tend to
show a fatigue limit a stress (~0.5UTS)
below which fatigue failure will not
occur. Non-ferrous metals (such as Cu
and Al) and their alloys do not show a
fatigue limit and can fail at quite low
stresses, albeit after a very large number
of cycles eg Al fails at ~0.25UTS after
108 cycles.
Fatigue failure can be prevented, or its
consequences reduced by:
a) Lowering the applied stress
b) Eliminating stress concentrations, including polishing the surface
c) Hardening the surface carburising, nitriding, shot peening, etc.
d) Changing from a non-ferrous alloy to a ferrous alloy
e) Rigorous quality control via non-destructive testing for flaws
f) Providing component redundancy.

Creep
Creep is the time-dependent plastic strain of a material under load at high temperatures
(relative to the melting temperature of the material ie above 0.4Tm).
A material that can easily withstand an applied stress () at low temperature might deform
slowly under the same stress at temperatures in excess of 0.4Tm. The strain () now depends,
not just on the stress (), but also on the temperature (T), and the time (t).
Unlike low temperature deformation, creep deformation of polycrystalline materials involves
the sliding of grains past each other and the formation of voids (holes) at grain
boundaries. After the instantaneous deformation, the creep curve of a material typically
shows three stages see diagram on next page. In the first stage (primary creep) the
deformation slows down with time, until it reaches a constant rate steady state creep in
the secondary stage. Towards the end of secondary creep, grain boundary sliding and void
formation commence. This leads to the accelerating rate of deformation in the tertiary
stage and eventually to failure (creep rupture).

4
Note that the secondary creep rate
increases with both temperature
and applied stress.
A creep-resistant material requires:
a) A high melting point
b) Stability at high temperatures
c) High strength at temperature
d) No grain boundaries at least
none at a large angle (~90) to
the applied stress
A typical example of the demand for creep-resistant materials is in the turbine blades of jet
engines. The starting point is the nickel-based superalloys, which have adequate toughness,
good corrosion/erosion resistance and high temperature strength. The turbine blade
performance can be improved by keeping the blades cool ie. by incorporating cooling
channels in the blade itself. The next step is to use directional solidification to ensure that
any grain boundaries run along the length of the blade (parallel to the stress). This can be
carried further by making single crystal turbine blades and avoiding grain boundaries
altogether.

Вам также может понравиться