Sunteți pe pagina 1din 53

USE OF MATERIALS COURSE

INTRODUCTION TO METALLURGY LECTURE NOTES

INDEX
Page

1. Introduction.................................................................................................. 4

1.1 Metals ........ ................................................................................. 4


1.2 Alloys ......... ................................................................................. 7
1.3 Solid Solutions ............................................................................. 7
1.4 Phase Mixtures ............................................................................ 8

2. Equilibrium Diagrams ............................................................................... 10

2.1 Iron - Carbon Diagram ............................................................... 10


2.2 Slowly Cooled Structures .......................................................... 12
2.3 Quenched Structures................................................................. 13
2.4 Critical Temperatures ................................................................ 15

3. Hot and Cold Working............................................................................... 17

3.1 Stress Relief .............................................................................. 17


3.2 Recrystallisation......................................................................... 17
3.3 Grain Growth ............................................................................. 18

4. Hardness and Strength ............................................................................. 19

4.1 Influence of Grain Size .............................................................. 20


4.2 The effects of Structural Changes ............................................. 20
4.3 Hardenability .............................................................................. 21

5. Heat Treatment.......... ................................................................................ 24

5.1 Hardening Mechanisms ............................................................. 24

5.1.1 Solid Solution Hardening ......................................... 24


5.1.2 Precipitation Hardening ........................................... 25
5.1.3 Work Hardening ....................................................... 26
5.1.4 Strain Ageing ........................................................... 28

1
5.2 Isothermal Transformations....................................................... 29

5.2.1 TTT Curves .............................................................. 29


5.2.2 Continuous Cooling Transformation Curves ........... 30
5.2.3 Effects of Alloys on Transformations ....................... 30

5.3 Normalising ................................................................................ 31


5.4 Full Annealing ............................................................................ 31
5.5 Quenching and Tempering ........................................................ 34
5.6 Importance of Grain Size ........................................................... 35
5.7 Grain Size Determination .......................................................... 37

6. Alloying....................................................................................................... 39

6.1 Alloying Elements in Steel ......................................................... 39

6.1.1 Carbide Formers ...................................................... 40


6.1.2 Crystal Growth ......................................................... 41
6.1.3 Corrosion Resistance .............................................. 42
6.1.4 Strength ................................................................... 42

6.2 Effects of Alloying on Steel ........................................................ 42

6.2.1 Carbon ..................................................................... 42


6.2.2 Manganese .............................................................. 42
6.2.3 Chromium ................................................................ 43
6.2.4 Nickel ....................................................................... 43
6.2.5 Molybdenum ............................................................ 44
6.2.6 Niobium .................................................................... 44
6.2.7 Vanadium ................................................................. 45
6.2.8 Aluminium ................................................................ 45
6.2.9 Sulphur..................................................................... 45
6.2.10 Phosphorus .............................................................. 46
6.2.11 Silicon....................................................................... 47

6.3 Alloy Addition in Steel - Summary ............................................. 47

7. Microstructures in Steel Arc Welds......................................................... 49

7.1 Weld Metal................................................................................. 49


7.2 Heat Affected Zone ................................................................... 51
7.3 Welding Heating Cycle .............................................................. 52
7.4 Carbon Equivalent ..................................................................... 53

2
8. Structure of Ingots.... ................................................................................ 54

8.1 Pipe........... ................................................................................. 54


8.2 Segregation ............................................................................... 55
8.3 Steel Types ................................................................................ 56

8.3.1 Rimming Steel.......................................................... 56


8.3.2 Semi-killed Steel ...................................................... 57
8.3.3 Killed Steel ............................................................... 57

3
1. INTRODUCTION

1.1 Metals

Metals are chemical elements or alloys, which in general exhibit the following
properties:

- Good conductivity of heat and electricity

- Relatively hard, yet possess properties of malleability and ductility

- Solid at normal temperatures - except Mercury

- High densities

- Low specific heats

- Reflect light and are generally white in colour -except Gold and Copper

- Difficult to penetrate with X-rays

- Magnetic to some degree but only Iron, Nickel and Cobalt are ferro
magnetic

However, it is the properties of strength, ductility and toughness which dictate


their principal uses.

All metals have a crystalline structure of which there are three common types:

Body-centred cubic
Face-centred cubic
Hexagonal close-packed

Body-centred cubic Face-centred cubic Hexagonal close packed


α - Iron (ferrite) γ - Iron (austenite) Magnesium
Vanadium Copper Titanium
Tungsten Silver Zinc
Chromium Gold Cadmium
Aluminium
Lead
Nickel
Platinum

The most economical packing arrangements for the metal atoms are the
hexagonal close packed and face centred cubic, followed the body centred
cubic structure, which occupies a larger volume (Fig 1).

4
Fig 1 The three Principal Types of Structure in which metallic elements
crystallise

Fig 2 Volume changes in Iron with Temperature

5
Fig 3 The Early Stages in the Growth of a Metallic Dendrite

The importance of different packing densities is illustrated by the fact that the
metal iron is allotropic, that is it adopts different crystal structures according to
its temperature. Up to a temperature of 910°C it h as a body-centred cubic
form, from 910°C to 1400°C it is face centred cubic , and above 1400°C it
reverts to body centred cubic until it melts at about 1535°C. Heating the
metal will not only produce an increase in volume by thermal expansion, but
also, volume changes will be caused by alterations in crystal structure.

The diagram (Fig 2) demonstrates the volume effects of different atomic


packing arrangements resulting from the allotropic changes from body
centred cubic structure to face centred cubic structure and back to body
centred cubic structure.

When a metal alloy solidifies, each crystal begins to form independently from
a nucleus or ‘centre of crystallisation’. The nucleus will be a simple unit of
appropriate crystal lattice, and from this the crystal will grow. It develops by
the addition of atoms according to the lattice pattern and rapidly begins to
assume visible proportions in what is called a dendrite. (Dendritic-tree-like).

A metallic crystal grows in this way because heat is dissipated more quickly
from a point, leading to the formation of a rather elongated skeleton (Fig 3).
The dendrite arms continue to grow and meet neighbouring dendrites which
will be orientated differently owing to their independent formation; that is, their
lattices will meet at odd angles. Hence the independent formation of each
crystal leads to the irregular overall shape of crystals and to the characteristic
grain structure of metals when in the ‘as cast’ condition.

6
1.2 Alloys

An alloy is essentially a mixture of two or more elements, the principal


component being a metallic element (the 'parent metal' or 'solvent'), so that
the resultant mixture exhibits metallic properties. A wide variety of
mechanical and physical properties may be obtained by alloying, so that
alloys, rather than pure metals, are of the greatest importance for
engineering.

If the constituent metallic atoms are chemically similar to one another, they
will crystallise as a single set of crystals, since all the atoms will behave as if
they belonged to the same species. A single-phase solid solution is then
said to form, and its microstructure is often indistinguishable from that of a
pure metal.

However, there may be a tendency for the elements to crystallise separately


to form distinct and different crystals joined at mutual grain boundaries. Such
a structure is an example of a phase mixture, which can usually be
distinguished from a single-phase solid by metallographic examination.

Note that this could include the formation of an intermetallic compound.


These compounds are in themselves of little practical value, since they tend
to be hard and brittle, but they can be important as constituents of alloy
systems.

1.3 Solid Solutions

Initially when a solid solution is formed the crystal structure is the same as
that of the parent metal - the atoms of the solute or alloying element are
distributed throughout each crystal, and a range of composition is possible.
The solution may be formed in two ways:

(a) In substitutional solid solutions the atoms share a single common


array of atomic sites (Fig 4a).

In some systems the parent metal will dissolve any proportion of the
solute and retain its original crystal structure. However, in many cases
there is a limited solubility and in order to accommodate a larger
proportion of the added alloying element a change in the initial crystal
structure becomes necessary to form a different solid solution, that is,
another phase. In this way two solid solutions may exist together over
a range of composition.

(b) In interstitial solid solutions the atoms of the solute element are
small enough to fit into the spaces between the parent metal atoms, as
illustrated (Fig 4b).

7
(a) substitutional

(b) interstitial

Fig 4 The formation of solid solutions

Because of the atom size limitation, interstitial solid solutions are less
common than substitutional solutions, although Carbon atoms can dissolve
in iron crystals in this way in steel. Similarly Nitrogen can dissolve in steel
and this is the basis of the Nitriding surface hardening process. The very
small atoms of Hydrogen will dissolve interstitially in ferrous alloys, usually
producing brittleness.

1.4 Phase Mixtures

A phase, present in an alloy as a separate entity, can be pure metal, a solid


solution or an intermetallic compound. Any mixtures of two or more of these
can occur. In binary systems, that is those of two elements, generally not
more than two phases can exist together.

8
2. EQUILIBRIUM DIAGRAMS

Thermal equilibrium (also known as Phase or Constitutional) diagrams are


of great importance in metallurgy for with their aid it is possible to determine
exactly the structure of a particular alloy at any given temperature, provided
the alloy has been allowed to reach a state of equilibrium. Thus the phases
present, their quantities and the chemical composition of each phase can be
shown with precision. The diagrams are constructed principally by thermal
analysis but also with microscopic studies, the examination of volume
changes, X-ray diffraction and other techniques.

Equilibrium can be considered as a state of balance ultimately arrived at by


the components at the temperature of the system concerned. However, in
some cases such a state would take a very long time to be reached while in
others it may never be reached at the temperature in question. For example,
if an alloy is rapidly cooled by quenching to room temperature, chemical and
physical changes may be suppressed such that they will never take place
unless the alloy is reheated to allow them to occur. Very slow cooling must
then follow.

2.1 Iron-Carbon Equilibrium Diagram (Fig 5)

Steel may be defined as an alloy of Iron and Carbon (up to about 1.7%C).
Here it may be helpful to recall the allotropic nature of iron and that up to
910°C it has a body centred cubic crystalline form known as alpha α Iron,
from 910°C-1400°C a face centred cubic structure, g amma γ Iron, reverting to
body centred cubic delta δ. Iron above that temperature. These terms are
modified in steel to Ferrite, Austenite and δ Ferrite. Other phases in the
equilibrium structure are Cementite the inter-metallic compound Fe3C, and,
Pearlite a phase mixture known as a Eutectoid consisting in this case of
alternate layers of Cementite and Ferrite. Pearlite contains about 0.83%C.

Ferrite and δ Ferrite, the body centred cubic structures dissolve only very
small amounts of carbon: less than 0.01% at room temperature. The face
centred cubic Austenite however, is capable of dissolving up to nearly 2%C
at 1150°C although this structure will change on re aching the Lower Critical
Temperature 723°C below which the Eutectoid reactio n will be complete. (Fig
6)

9
Fig 5 The Iron-Carbon phase diagram

10
Fig 6 Part of the Iron-Carbon Thermal-equilibrium diagram

Note that the Lower Critical Temperature 723°C bel ow which all Austenite
has been converted to Ferrite and Cementite is commonly known as the A1
temperature. The temperature above which the structure will be wholly
Austenite, the Upper Critical Temperature, is known as the A3 temperature.
Also the temperature above which the steel reverts to a wholly body centred
cubic δ ferrite is known as the A4 temperature.

2.2 Slowly Cooled Structures

The most important reaction in steel is the decomposition of austenite on


cooling. Consider the slow cooling of a steel of 0.83%C content (i.e. of the
eutectoid composition) (Fig 7a); at 723°C the structure will transform to an
eutectoid mixture consisting of alternate lamellae or plates of ferrite and
cementite.

A steel of higher carbon content (known as a 'hyper-eutectoid' steel) (Fig


7b), 1.20%C, will remain austenitic down to the temperature around, say,
870°C at which the solvus line is crossed, so that Fe3C will start to be
precipitated at the austenite grain boundaries.

Continued cooling and precipitation of cementite Fe3C will reduce the carbon
content of the austenite until it reaches that of the eutectoid 0.83%C. When
the temperature falls to below 723°C, this residual austenite will transform to
pearlite, and the final microstructure will be cementite plus pearlite.

11
A lower carbon steel (i.e. a 'hypoeutectoid steel) of 0.4%C (Fig 7c) will begin
to transform when the temperature falls below the solvus line by the
precipitation of ferrite at the austenite grain boundaries.

Continued cooling and precipitation of ferrite will increase the carbon content
of the austenite until it reaches that of the eutectoid 0.83%C. At 723°C this
remaining austenite will transform to pearlite resulting in a final structure of
ferrite plus pearlite.

2.3 Quenched Structures

The previous microstructures form in plain carbon steels which have been
moderately slowly cooled (e.g. by cooling in air) from temperatures within the
austenitic phase field, say from 50°C above the low er boundary line CED.
This is called a 'normalising' heat treatment, but medium and high-carbon
steels are very commonly subjected to more complex treatments in order fully
to exploit their properties. These treatments involve, first, heating the alloy
into the austenite phase field, as before, but then quenching it in water or
brine which suppresses diffusion and thus the formation of ferrite and
cementite. Under these conditions the austenite transforms by a process not
involving diffusion into a metastable distorted form of body-centred iron
known as 'Martensite' (Fig 7d). This process is extremely rapid and the
transformation may be completed in a few microseconds.

Fig 7a Structure of Eutectoid Steel (1000X) (Pearlite)

Fig 7b Structure of Hypereutectoid Steel (1000X)

12
Fig 7c Structure of Hypoeutectoid Steel (1000X)

Fig 7d Martensite (2000X)

All the carbon originally dissolved in the Austenite at high temperature


remains after quenching in interstitial solution in the Martensite crystals. This
has the effect of distorting the lattice from cubic to tetragonal symmetry. This
lattice distortion by the dissolved carbon has the effect of hardening the
structure and the resulting steels will not only be hard but brittle, for which
there is little practical application. A second heat treatment called tempering
is therefore required and this will reduce hardness and brittleness. If
Martensitic steel is reheated to the temperature range 200-600°C (below the
Austenite forming temperature) it rapidly decomposes to form body centred
cubic ferrite and particles of Cementite.

This structure is on an extremely fine scale, the size of the carbide particles
being dependent on the time and temperature of the treatment. The higher
the temperature and the longer the time, the softer and less brittle the
product.

13
2.4 Critical Temperatures

There has been reference to the A1 lower critical temperature, the A3 upper
critical temperature and the A4 temperature above which austenite reverts to
δ ferrite. Examination of the Iron-Carbon equilibrium diagram shows that the
latter change will not occur when the carbon content approaches and
exceeds 0.5%C. In such cases liquid steel will solidify directly to austenite.

Concern may be felt regarding the non-appearance of an A2 temperature and


also the absence of a beta phase in the equilibrium diagram. Originally the
designation A2 was given to the temperature 769°C, the Curie poin t, at which
iron ceases to be magnetic, a fact that was expected to be accompanied by a
phase change. This proved not to be the case and the A2 has no structural
significance.

It cannot be emphasised too strongly that the structural changes in the


diagram and the temperatures at which they occur refer to conditions of
equilibrium. In practice it is found that, on heating, it is necessary to exceed
the equilibrium temperatures to achieve the expected structural changes.
Such temperatures would normally be about 50°C abov e the A1 and A3 and
can be determined with reasonable accuracy for particular conditions of
heating. Here the lower critical point is termed the Ac1 and the upper critical
point the Ac3 temperature. The suffix C has been derived from the French
‘chauffage’ meaning heating. In a similar way it may be expected that on
cooling it is necessary to reach temperatures lower than those of the
equilibrium diagram to obtain the appropriate structural changes. Thus, for
example, in hot rolling operations, which are accompanied by continuous
cooling, temperatures of 100-150°C below equilibriu m become necessary for
the expected structural changes. In such cases AR1 and AR3 temperatures
are referred to, the suffix R again being French, derived from
‘Refroidissement’, cooling.

The Effect of Heating Rates on Lower and Upper Critical Temperatures


Steel AISI 1045 (0.45%C)

Equilibrium Temperature HEATING RATES °C/SEC FROM 700-1000°C


3 30 300 1400
Ac1 723 780 790 800 840
Ac3 770 820 830 860 935

14
3. HOT AND COLD WORKING

In a process such as rolling, the crystals making up a metal are deformed in


the direction of rolling. If the temperature of the operation is such that the
deformation is accompanied by simultaneous recrystallisation the process is
Hot Working. Here relatively minor ductility changes can be expected and
these can be attributed largely to improved crystal structures and better
dispersion of non-metallic inclusions. On the other hand if there is no
recrystallisation and the crystals become progressively more deformed on
working then the process is Cold Working. In this case major property
changes arise with a loss of ductility and marked increases in hardness and
strength.

A cold-worked metal is in a state of considerable mechanical stress,


resulting from elastic strains internally balanced. These elastic strains are
largely due to inhomogeneous deformation having taken place during cold-
working. If the metal is heated to a sufficiently high temperature the strains
will be removed; at the same time the tensile strength and hardness of the
metal will fall to approximately their original values and the capacity for cold-
work return. This form of heat-treatment is known as annealing, and is
employed when the metal is required for use in a soft but tough state or,
alternatively, when it is to undergo further cold deformation. Annealing takes
place in three stages as follows:

3.1 Stage 1 - Stress Relief

This occurs at relatively low temperatures at which dislocations are able to


move to equilibrium positions in the crystal lattice (see Heat treatment section
for an explanation of dislocations). Such small movements can reduce
internal mechanical stress without, however, producing any visible alteration
in the distorted shape of the cold-worked crystals. Moreover, hardness and
tensile strength will remain at the high value produced by cold-work and may
even increase.

3.2 Stage II - Recrystallisation

As mentioned previously, a low-temperature anneal to relieve internal stress


may sometimes be used, but generally annealing involves a definite and
observable alteration in the crystal structure of the metal. If the annealing
temperature is increased a point is reached when new crystals begin to grow
from nuclei produced in the deformed metal. These nuclei are formed at
points of high energy, such as crystal boundaries. The crystals so formed are
at first small, but grow gradually until they absorb the entire distorted structure
produced originally by cold-work. The new crystals are equi-axed in form,
that is, they do not show any directional elongation, as did the distorted cold-
worked crystals which they replace.

This phenomenon is known as recrystallisation, and it is a method


employed, in conjunction with cold-work, of course, to produce a fine-grained
structure in non-ferrous metals and alloys. Only in some cases - notably in

15
steels and aluminium bronze, where certain structural changes take place in
the solid state - is it possible to refine the grain size by heat-treatment alone.

3.3 Stage III - Grain Growth

The temperature at which recrystallisation will take place is called the


recrystallisation temperature and if the annealing temperature is above this,
the crystals will continue to grow until the structure is relatively coarse
grained. The amount of grain growth is governed to a large extent by the
annealing temperature, the duration of annealing and the degree of previous
cold-work. A high annealing temperature or a long annealing time will
encourage grain growth. Heavy deformation will lead to a small grain size,
light deformation will give rise to a larger grain size on annealing since there
will be fewer nuclei for crystal growth.

16
4. HARDNESS AND STRENGTH

Although hardness is a property difficult to describe precisely, tests have


been devised which are quick and easy to apply. Such tests are non-
destructive, unlike tensile and impact determinations, and they are therefore
attractive means of assessing the mechanical properties of metals. In
practice, routine hardness tests on manufactured parts will be applied over
limited property ranges and reliable correlations between hardness, ultimate
tensile strength, yield strength, elongation and reduction of area can be
established. As such they will provide excellent guides to the acceptability of
the items concerned (and possibly their chemical compositions), the suitability
of the manufacturing processes and heat treatment to which they have been
subjected.

Increases in hardness values (which results from an increase in C content)


will usually be associated with increases in ultimate tensile and yield strength,
but there will be reductions in ductility (elongation) expressed as reduction of
area and elongation (Fig 8). Note that an increase in C content will also be
accompanied by a reduction in impact (Charpy) toughness.

Fig 8 Diagram showing the Relationship between Carbon Content,


Mechanical Properties, Microstructure and Uses of Plain Carbon Steels
in the Normalised Condition

17
4.1 Influence of Grain Size on Hardness

Usually in metals a smaller grain size will lead to some increase in hardness
and tensile strength. This strengthening effect is due not only to the complex
intersecting slip process in the various grains but also to the grain boundaries
which are themselves obstacles to the movement of slip planes which
therefore cannot propagate freely from grain to grain.

In structural steels, processing for optimum properties is aimed at producing


very small sized relatively soft grains. In this way Yield Strength is enhanced
and the Impact properties markedly improved especially at low temperatures.

4.2 The Effects of Structural Changes on the Hardness of Steels

It must be emphasised that in steels the most profound changes in


mechanical properties occur as a result of changes in microstructure. These
are illustrated in the table below relating to the different structures obtained in
a plain carbon eutectoid steel (0.83%) which has been subjected to
isothermal transformations. (See later section).

Structure/Hardness
HARDNESS HARDNESS TEMPERATURE OF
STRUCTURE BHN ROCKWELL TRANSFORMATION
°
C C
Coarse 170 5 720
pearlite 293 31 660
Fine pearlite 388 41 580

Feathery 401 42 500


bainite (upper) 415 44 400
Acicular 555 56 280
bainite (lower) 578 58 230
Bainite & 601 60 175
Martensite
Martensite 682 66 RT

18
4.3 Hardenability

Hardenability is a measure of the depth a steel will harden on quenching.

The uniform rapid cooling of a heavy steel section is impossible even with
drastic quenching. Such a section will not harden completely to its core
whereas a thin section would be wholly martensitic. This difficulty can
however be overcome by the addition to the steel of alloying elements which
will in general increase the time available to begin and complete
transformation. Thus a martensitic structure becomes possible with the lower
cooling rates found in heavier sections. This is one of the most important
functions of alloying and to ensure the correct application of such steels, from
both the technical and economic aspects, some measure of hardenability
becomes necessary. To determine this, tests have been devised to estimate
the maximum diameter at which the required structures can be produced by
quenching. If this diameter is exceeded, hardening at the core will be
incomplete resulting in non-uniform properties.

The Jominy end-quench test (Fig 9) is widely used in evaluating the


hardenability of steel. Here a standard test piece is heated to its austenitic
region, dropped into a frame and quenched, at one end only, by a measured
jet of water at 25°C. Thus different rates of cool ing are obtained along the
length of the bar and the resulting hardness values can be determined. The
drawings illustrate the test and the graphs show typical hardness values from
the tests of three steels of differing compositions (Fig 10). It will be noted that
the depth of hardening increases markedly with growing alloy content even
though the proportion of carbon at 0.45% is the same in each case. Using
the Jominy test results as a basis it is possible, for a particular steel, to
calculate a maximum diameter at which uniform properties can be obtained.
Such a measure is known as the ideal diameter or ruling section. The ideal
diameter here represents the section at which a structure of 50% martensite
is achieved.

19
(A) The standard form of test piece used

(B) Diagrammatic representation of the apparatus used in the test

Fig 9 The Jominy End-Quench Test

20
Fig 10 The Depth of Hardening of Three Different Steels as indicated by
the Jominy Test

21
5. HEAT TREATMENT

5.1 Hardening Mechanisms

It has been calculated that the theoretical strengths of pure metals should be
much greater than those observed. These differences have been
convincingly explained by ‘dislocation theory’.

This theory has been likened to a situation where one carpet lying on top of
another is very difficult to move by pulling at one end. However a ruck in the
upper carpet will move across it very easily. This ruck represents the
dislocation moving between slip planes in a metal giving rise to plastic
deformation, which otherwise could not occur. Calculations confirm that the
stress required to make the dislocation lines move is in good agreement with
the measured yield stress so that such faults can account for the weakness of
metals. The dislocation lines finish only when they reach a metal surface or
grain boundary. If dislocation motion is impeded, for example by interaction
with other dislocations, there will usually be an increase in hardness and
strength. It should be emphasised that dislocation theory is much more
complex than this grossly over-simplified version (Fig 11).

Fig 11 The Movement of a Dislocation During Slip

5.1.1 Solid Solution Hardening

The most common reason for alloying is to increase the yield strength of a
metal. This requires the movement of dislocations to be impeded by making
alterations to the structure on an atomic scale. The effectiveness of foreign
atoms as barriers to the movement of dislocations depends first upon the size
difference between the solute atoms and those of the parent metal and
second upon the proportion of foreign atoms present in the crystal. Elastic
strain fields will be set up around the misfitting atoms making dislocation
motion more difficult.

Interstitial solute atoms such as carbon and nitrogen in body centred cubic
crystals provide an important example of solute hardening alloys. Thus

22
carbon or nitrogen dissolved in iron produces a local strain which causes a
very steep rise in yield strength with increasing solute. If only a very small
quantity of such a solute is present one effect of this high local strain is that
the solute element will tend to migrate to the space provided along the
dislocations present rather than being uniformly distributed in solution in each
crystal. This has the effect of ‘pinning’ (holding) the dislocations in place, a
phenomenon which is revealed by the presence of a Yield Point when plastic
flow sets in during a tensile test.

Substitutional solute addition is the commonest way of solution-hardening a


metal. Brass (copper-zinc) and bronze (copper-tin) are two familiar materials
which make use of this hardening effect.

5.1.2 Precipitation Hardening

Thermal treatment can be used to control the size and distribution of second
phase particles in any alloy which undergoes a phase transformation in the
solid state. Alloy systems which have a phase diagram showing a decreasing
solid solubility limit with decreasing temperature are particularly appropriate
for such treatments and particles of the second phase can often be made to
precipitate in a very finely dispersed form.

In Figure 12, an alloy C exists as a single phase solid solution (α) at high
temperatures but on slow cooling it becomes supersaturated with respect to
the second β phase, which therefore separates out.

The distribution of the β phase may be controlled as follows.

The alloy is first solution heat-treated at the high temperature and then rapidly
cooled by quenching into water or other cooling fluid. Solid-state diffusion is
suppressed in this way, so that the β phase cannot separate and the alloy
exists at the low temperature in an unstable supersaturated state. If the
temperature is now increased, so that diffusion can take place at a
measurable rate, the second phase will nucleate and grow.

In alloys of relatively low melting-point, there will be an appreciable diffusion


rate of solute atoms at room temperature, so that over a sufficient length of
time, the second phase will precipitate out. This effect is known as 'ageing',
but in most alloys the temperature has to be raised in order to cause
precipitation to occur and the material is said to be 'artificially aged'. The
rate of growth of the precipitate is controlled by the rate of atomic diffusion, so
that the precipitation increases with increasing ageing temperature. The size
of the precipitate becomes coarser as the ageing temperature is increased,
as shown schematically in Fig 12.

If, at any temperature, the time of heat treatment is very prolonged,


coagulation or coarsening of the particles occurs; the small ones tend to
redissolve and the large ones to grow at their expense.

23
Precipitation hardening is the term now used to describe ageing. Just as
heating will accelerate precipitation hardening, refrigeration will impede the
process.

Fig 12 Variation of precipitate size with ageing temperature

24
5.1.3 Work Hardening

Owing to the interaction of dislocations within the grains, plastic deformation


will progressively harden a metal. The dislocations will multiply with strain
and their density will continually rise with increasing deformation and thus the
number of interactions per unit volume will rise as the strain continues. When
a polycrystalline metal is stressed, each grain is deformed into a shape that is
dictated by the deformation of its neighbours which requires the operation of
several slip systems. The following graphs given in Figs 13, 14 and 15 show
the effects of cold work on mechanical properties.

Ultimate tensile strength can be increased by up to 1040 N/mm2 by reducing


the cross-section by cold working by 90%. A reduction in cross-section by
10% will reduce percentage elongation from about 30% to 4%. Thereafter
there will be little decrease to about 2%.

Fig 13 Stress-strain curves for metal subsequent to varying degrees of


prior cold-working. It will be observed that cold working brings about an
increase in tensile and yield strengths and a decrease in elongation
prior to fracture.

25
Fig 14 Effect of cold deformation on the strength and hardness of metal. It will
be seen that the spread between the yield strength and tensile strength curves
becomes less with greater amounts of deformation and a consequent
reduction in ductility.

Fig 15 Stress-strain curves for metal subsequent to varying degrees of prior


cold-working. It will be observed that cold working brings about an increase in
tensile and yield strengths and a decrease in elongation prior to fracture.

26
Gain in UTS (N/mm2)

% Reduction in Cross-Section 0.20%C 0.80%C


30 140 200
50 280 350
80 450 700

5.1.4 Strain Ageing

Strain ageing is observed when low carbon steel has been subjected to a
small amount of deformation and then allowed to age for a period. This
results in an increase in hardness and strength with reduced ductility and a
drastic reduction of toughness as measured by impact tests. Even less than
1% cold reduction can produce such ageing which can be attributed largely to
the movement of nitrogen atoms in the steel to sites which allow dislocations
to be pinned in place. The maximum effect is with about 15% reduction.
Such ageing proceeds very slowly at room temperatures and may take
several months to reach a peak. Temperature increases will however rapidly
increase the rate of hardening and at 300°C this wi ll reach the maximum
value in a few seconds.

5.2 Isothermal Transformations

So far consideration has been given to very slow cooling, when conditions will
approach equilibrium, or very fast cooling to ambient temperatures.
Variations in cooling rates between these extremes are possible and the
results can be shown in isothermal transformation curves, also known as S
curves, or time-temperature-transformation (TTT) curves. These are
constructed by taking a number of specimens of the steel in question, heating
them into the austenitic range and then quenching them in baths of different
temperatures. At predetermined time intervals individual specimens are
taken from their baths and quenched in water. They are then examined
microscopically to evaluate the extent of the transformation that has occurred.

5.2.1 TTT Curves for Steel

Figure 16 shows TTT Curves for a eutectoid plain carbon steel i.e. 0.83%C.
First to be noted is the horizontal broken line at 723°C which is the
temperature above which a eutectoid steel will be wholly stable austenite.
Below this line the austenite is unstable and the two C shaped curves indicate
the times necessary for the transformation of this austenite to ferrite plus
cementite to begin and go to completion following quenching to a particular
temperature.

The horizontal lines at the foot of the TTT diagram, strictly speaking, are not
part of the TTT curves but represent the temperatures at which the formation
of martensite will begin MS, and end MF, during the cooling of unstable
austenite through this range. Unlike the formation of pearlite and bainite the

27
transformation of unstable austenite to martensite is not dependent on time
as it occurs almost instantly. The degree to which this change takes place is
determined simply by the temperature to which the steel is cooled. Once a
transformation has gone to completion a steel will usually be water quenched
with no further change in its constitution.

Fig 16 Time - Temperature - Transformation (TTT) Curves for a Plain


Carbon Steel of Eutectoid Composition

5.2.2 Continuous Cooling Transformation Curves

The TTT curves indicate structures which are produced by transformations


which take place at a fixed temperature and specify a given ‘incubation’
period before the transformation begins. It is important to realise that there
can be no direct connection between such isothermal transformations and
those which take place under continuous cooling at a constant rate from
723°C to room temperature. Therefore it is not pos sible to super-impose
curves which represent continuous cooling onto a TTT diagram. However,
modified TTT curves which are related to continuous cooling can be
produced. They are of a similar shape to the TTT curves but are displaced to
the right.

28
5.2.3 The Effects of Alloy Additions on Isothermal Transformations

With the exception of the metal cobalt, all alloy additions to steel will increase
both the incubation period and the time necessary for transformations at
specific temperatures. In effect this will increase hardenability. These
changes are illustrated in the attached series of TTT diagrams (Fig 17). It
should be noted that alterations in composition will be accompanied by
variations in the temperatures A1, A3 and A4 and also those of MS and MF.
Attention is drawn to the fact that the diagrams are based on a logarithmic
scale so that even apparently minor movements of the TTT curves to the right
will result in major increases in incubation and transformation periods. These
will be reflected in the continuous cooling transformation diagrams
appropriate to these alloys.

5.3 Normalising

The process refers only to steel and consists of heating the metal to a
temperature about 50°C above its Upper Critical Poi nt (Ac3 temperature) or
austenitising temperature and having given adequate time for full conversion
to Austenite allowing it to cool in ‘still air’. On heating, the change from body
centred to face centred cubic structure will result in small austenite crystals
whose sizes will determine those of the body centred cubic crystals formed on
cooling. The process is therefore important in producing a refined grain size.
If too high a temperature is employed, grain growth of the austenite will occur
which will be reflected in the final cooled structure.

5.4 Full Annealing

This process is also applicable only to steel and consists of heating the
material to the appropriate normalising temperature and then cooling it very
slowly (Fig 18), usually in the furnace. The resulting structure, as with
normalised steel, will consist of ferrite or cementite with grains of pearlite but
full annealed steel will be expected to have a somewhat larger grain size.
More important is the effect of slow cooling on the structure of pearlite and
this will become more coarsely lamellar with decreasing cooling rate which
may even lead to very coarse lamellae balling up into coalesced particles of
cementite in a groundmass of ferrite. Steel subjected to full annealing
treatment will have lower tensile strength, impact strength and yield point but
higher elongation values than those of normalised steel.

29
Fig 17 TTT-curves for alloy steels (after US Steel Corp)

30
Fig 18

(A) Normalised
(B) Fully Annealed
(C) Water-quenched
(D) Water-quenched and tempered

31
5.5 Quenching and Tempering

If a plain carbon steel is sufficiently rapidly cooled from above its upper critical
temperature, i.e. within the wholly austenitic area, martensite will be formed.
This phase becomes harder and more brittle as the carbon content increases
up to the eutectoid value, 0.83%C. The hardness so attained will vary
somewhat with the rate of cooling. The upper set of values shown in Fig 19
are the result of rapid cooling while those below are of average rates.

Fig 19 The effect of carbon content on hardness of martensite

Occasionally such high martensitic hardnesses are useful for applications


requiring good wear resistance but more often they are the essential starting
points for tempered structures. Thus a steel may be hardened by heating it to
30-50 degrees C above its upper critical temperature and quenching it in
some medium to produce the desired rate of cooling. In this condition, a tool
steel for example, will be hard and brittle and some treatment to relieve
internal stresses and to reduce brittleness is needed. Tempering will cause
martensite to transform and the higher the temperature the nearer will the
structure revert to the stable pearlitic type appropriate to that temperature.
Tempering is always carried out below the lower critical level and when the
desired temperature is reached may be followed by an immediate water
quench, if the alloy is susceptible to temper embrittlement. Rate of cooling
from the tempering temperature has little other effect.

Up to 200°C tempering will provide only some stre ss relief but from 230-
400°C the martensite will change to form a new cons tituent consisting of a
very finely dispersed granular mixture of ferrite and cementite. This phase is
much tougher than martensite and somewhat softer and less brittle.
Toughness may be defined as the resistance to fracture by impact.
Increasing tempering temperatures causes coalescence of the cementite

32
particles until at about 400°C they become visible at magnifications of about
500 times. Subsequent increases of tempering temperature to above 550°C
will cause strength to fall away with no rise in ductility.

The importance of the fact that steels can be heat treated to give marked
changes in properties cannot be over-emphasised. Thus it is possible by
correct treatment to produce a relatively soft and ductile material that can be
cold worked and machined to the desired shape and size and then to change
its properties, drastically increasing its hardness and strength, making it far
more suitable for its eventual service life. Compared with the pearlitic and
bainitic structures of the same hardness, tempered martensites will have
higher yield, tensile and impact strengths compatible with a high level of
ductility. For most purposes therefore they are the most desirable structures.

Tempering and Water Quench

0.40%C Steel As Quenched 200°C 350°C 450°C 600°C


Hardness HV 670 620 440 370 220
Tensile N/mm2 1850 1700 1420 1200 880
Elongation % 5 12 14 16 24
Charpy Impact (J) 11 19 14 30 75

5.6 The importance of grain size

As steel is heated through the critical range it is recrystallised with the


formation of fine grains (Fig 20).

Fig 20 Schematic representation of the effect of temperature and grain-


refining elements on grain size

33
These grains tend to grow as the temperature is raised, but the growth is
more or less abrupt at a characteristic temperature - with the formation of
first some coarse and some fine grains and finally all coarse grains, as
indicated above. Prolonged heating also tends to increase the grain size.

Each steel has a characteristic coarsening-temperature range and if that


temperature is below the maximum commonly used for hardening, the steel is
classed as coarse grained. If the steel is treated with effective amounts of
aluminium, niobium, vanadium or titanium, or a combination of these
elements, it does not coarsen at the usual hardening temperatures and is
classed as fine grained.

Coarse-grained steels (ASTM grain size 1-5) have some advantages in easier
machining, higher creep strength at elevated temperatures, greater
hardenability, and possibly less notch sensitivity in fatigue, but they have
lower notched-bar impact strength and the impact resistance falls off rapidly
at subnormal temperatures. Fine-grained steels (ASTM grain size 5-8 plus)
have much better toughness, and virtually all heat-treating alloy steels used in
machine construction are treated during melting so that they will be fine
grained. Steels that develop a mixture of coarse and fine grains tend to give
erratic properties, and are unlikely to be used.

The grain-coarsening temperature of fine-grained steels varies somewhat but


can be expected to be in the neighbourhood of 980°C for aluminium-treated
steel. This temperature may, however, vary widely with the composition and
prior treatment. The coarsening temperature of high-carbon steels tends to
be relatively low but grain growth may be restrained by undissolved carbides.

5.7 Grain Size Determination

Grain size is commonly measured according to ASTM Standard Method E


112, Determining The Average Grain Size. In materials having two or more
constituents, the grain size usually refers to that of the matrix. Minor
constituent phases, inclusions, and additives are not normally considered.

It is important in using these methods to recognise that the measurement of


grain size is not precise, but an estimate. A metal grain is a three-
dimensional shape of varying sizes. The grain cross section produced by a
random plane (surface of observation) is dependent upon where the plane
cuts each individual grain. Thus, no two fields of observation can be exactly
equal.

34
Of the three methods listed in the standard, the comparison procedure is most
popular since it takes the least time to carry out. This method involves viewing
grains in a microscope and comparing them at the same magnification, 75X or
100X, to charts defined in ASTM E112, with two examples shown in Figure 21.
The ASTM Grain Size Number corresponds to a certain number of grains/in2
according to Table below:

ASTM No
0 1 2 3 4 5 6 7 8 9 10
2
Grains/in
0.5 1 2 4 8 16 32 64 128 256 512

The relationship between the Grain Size Number and the number of grains/in2
is given by the expression:

Where:

N = ASTM Grain Size Number n = 2 (N – 1)


n = number of grains/in2 at the specified magnification

Fig.21 Examples of ASTME E 112 Comparison procedure grain size


charts for No. 5 and 8

Thus they may be added to improve mechanical properties. They may be


used to enhance resistance to corrosion or high temperature oxidation.
Further, they may be present to develop special characteristics such as those
of an electrical or magnetic nature, strength at high temperatures or for the
steel to remain austenitic at room temperatures.

35
6. ALLOYING

6.1 Alloying Elements in Steel

The principal function of alloying elements in steel is to improve its properties.


Thus they may be added to improve mechanical properties. They may be
used to enhance resistance to corrosion or high temperature oxidation.
Further, they may be present to develop special characteristics such as those
of an electrical or magnetic nature, strength at high temperatures or for the
steel to remain austenitic at room temperatures.

The alloying elements added may either simply dissolve in the ferrite or they
may combine with some of the carbon, forming carbides.

The principal effects which these alloying elements have on the


microstructure and properties of a steel include those on the allotropic
transformation temperatures.

Some elements, notably nickel, manganese, cobalt and copper, raise the A4
temperature and lower the A3 temperature, as shown in Figure 22 (A). In this
way these elements, when added to a carbon steel, tend to stabilise austenite
and increase the range of temperature over which austenite can exist as a
stable phase. Other elements, the most important of which include
chromium, tungsten, vanadium, molybdenum, aluminium and silicon, have
the reverse effect, in that they tend to stabilise ferrite α by raising the A3
temperature and lowering the A4, as indicated in Figure 22 (B). Such
elements restrict the field over which austenite may exist, and thus form what
is often called a ‘γγ loop’.

The elements of the γ-stabilising group generally have a face-centred cubic


lattice. Since this is the same as austenite, these elements will retard the
transformation of austenite to ferrite. At the same time these elements retard
the precipitation of carbides, and again this has the effect of stabilising
austenite. The a-stabilising elements are usually those with a body-centred
cubic lattice. These will dissolve more readily in ferrite than in austenite, and
at the same time diminish the solubility of carbon in austenite. In this way
they stabilise ferrite. As shown in Figure B, progressive increase in one or
more of the α-stabilising elements will cause a point to be reached beyond
the confines of the γ-loop, where the austenite cannot exist at any
temperature.

36
Fig 22 Relative effects of the addition of an alloying element on the allotropic
transformation temperature at A3 and A4
(A) Tending to stabilise γ (B) Tending to stabilise α

6.1.1 Carbide or Graphite Formers (Fig 23)

Some alloying elements form very stable carbides when added to a plain
carbon steel. This generally has as hardening effect especially when the
carbides formed are harder than iron carbide. Such elements include
chromium, tungsten, molybdenum, titanium and manganese. When one or
more of these elements is present, a structure containing complex carbides is
often formed.

37
Fig 23 The condition in which alloying elements are present in steel

Other elements have a graphitising effect on the iron carbide; that is, they
tend to make it unstable so that it breaks up, releasing free graphitic carbon.
This effect is more evident if no carbide stabilisers are present. Elements
which tend to cause graphitisation include silicon, nickel and aluminium.
Therefore, if it is necessary to add appreciable amounts of them to a steel, it
can be done only when the carbon content is extremely low. Alternatively, if
the carbon content needs to be high, one or more of the elements of the first
group, namely the carbide stabilisers, must be added in order to counteract
the effects of the graphitising element.

6.1.2. Crystal Growth

The rate of crystal growth is accelerated, particularly at high temperatures, by


the presence of some alloy additions, notably chromium. Care must therefore
be taken that steels containing elements in this category are not overheated
or, indeed, kept for too long at an elevated temperature, or brittleness, which
is usually associated with a coarse grain size, will result.

Fortunately, grain growth is retarded by other elements, notably niobium and


vanadium, whose presence thus produce a steel which is less sensitive to the
temperature conditions of heat-treatment.

38
6.1.3 Corrosion Resistance

The corrosion-resistance of steels is substantially improved by the addition of


aluminium, silicon and chromium. These metals form thin but dense and
adherent oxide films which protect the surface of the steel from further attack.
Of the elements mentioned chromium is the most useful when mechanical
properties have to be considered. When nickel also is added in sufficient
quantities, the austenitic structure is maintained at room temperature.

6.1.4 Strength

One of the main reasons for alloying is to effect improvements in the


mechanical properties of steel. These improvements are generally the result
of physical changes already referred to. For example, hardness is increased
by stabilising the carbides; strength is increased when alloying elements
dissolve in the ferrite; and toughness is improved by refinement of the grain
size.

6.2 The Effects of Alloying on Steel Strength and Hardenability

6.2.1 Carbon

It has been shown that increasing carbon content in plain carbon steels will
lead to increases in hardness and, up to the eutectoid composition (0.83%C)
increases the tensile and yield strengths. Corresponding to these changes
will be decreases in ductility, malleability and impact strength. Increasing the
carbon content of martensitic steels may be expected to show increases in
hardness up to the eutectoid point. Tempered martensitic structures will show
increases in hardness and brittleness up to about 1.4%C. Hardenability in
plain carbon steels is at a maximum at the eutectoid composition. Increasing
or decreasing the carbon content from the eutectoid composition will lead to
lower hardenability, i.e. it becomes increasingly more difficult to obtain
martensitic structures, and increasingly rapid quenching is required.

6.2.2 Manganese

Some manganese is present in nearly all steels. It is usually below 1% and it


is only when this amount is exceeded that it is regarded as a deliberately
added alloying element. Like nickel, manganese stabilises austenite but
unlike nickel it also has the effect of stabilising the carbides, by itself forming
Mn3C. It has a considerable strengthening effect on ferrite and increases
hardenability.

39
6.2.3 Chromium

Very small amounts of chromium when added to carbon steel will cause a
considerable increase in hardness. At the same time strength is raised with
some loss of ductility which is not apparent with less than 1.0%Cr. The
increase in hardness is mainly due to the fact that chromium is a carbide
stabiliser forming hard carbides itself or double carbides with Fe3C.
Chromium is a ferrite stabiliser and austenite may be eliminated entirely when
more than 11%Cr is added to pure iron.

The main disadvantages in the use of chromium as an alloying element is its


tendency to promote grain growth, with its attendant brittleness. Care must
therefore be taken to avoid overheating or holding for too long at the normal
heat-treatment temperature.

Steels containing small amounts of chromium and up to 0.45% carbon are


used for axle shafts, connecting-rods and gears; whilst those containing more
than 1.0% carbon are extremely hard and are useful for the manufacture of
ball-bearings, drawing dies and parts for grinding machines.

Chromium is also added in larger amounts - up to 21% - and has a


pronounced effect in improving corrosion-resistance, due to the protective
layer of oxide formed. This oxide layer is extremely thin, and these steels
take a very high polish. They contain little or no carbon and are therefore
completely ferritic and non-hardening (except by cold-work).

Stainless steels which have only chromium as the main alloying element and
have C levels that exceed 0.1% are of the martensitic type, the structure
being obtained by rapid cooling. If these steels are allowed to cool slowly,
carbides will be precipitated, with consequent loss in corrosion-resistance
(sensitisation).

6.2.4 Nickel

The addition of nickel to a plain carbon steel tends to stabilise the austenite
phase.

For example, the addition of 25% nickel to pure iron renders it austenitic, and
so non-magnetic, even after slow cooling to room temperature.

Nickel makes the carbides unstable and tends to cause them to decompose
to graphite. It is therefore inadvisable to add nickel by itself to high carbon
steels and most nickel steels are of the low carbon type. Generally the 3.5%
nickel steels are the most widely employed, those with about 0.12%C being
used mainly for case hardening, while the 0.30%C types are used for
structural purposes and in engines for shafting and axles etc. Nickel also has
a grain refining effect enabling the alloys to be employed in case hardening
as grain growth will be limited during prolonged treatment at about 900°C.

40
Nickel was one of the first elements added to plain carbon steels, its main
advantage being to increase tensile strengths without adversely affecting
ductility. Its effect of lowering critical temperatures for heat treatment is also
an advantage in cost reduction.

Nickel in moderate amounts also increases hardenability allowing slower


quench rates than in plain carbon steels.

6.2.5 Molybdenum

Molybdenum is a strong carbide stabiliser and relatively small amounts will


markedly enhance the properties of plain carbon steels. Usually it is added in
quantities ranging from 0.12-0.65% depending on the intended application of
the steel. More often it is used in combination with manganese, nickel and/or
chromium to give a noticeable improvement in properties. Hardenability is
increased especially when used in conjunction with chromium. One of the
metal's advantages is that it will reduce the tendency to temper brittleness in
low nickel low chromium steels when heat treated at 250-400°C. In general
Ni-Cr-Mo alloy steels possess the best all-round combination of properties,
especially when high tensile strength and good ductility are required in large
components. Such steels are relatively free from the mass effects of heat
treatment, the transformation rates of the Ni-Cr steels being even further
reduced by the presence of molybdenum which contributes considerably to
hardenability. Its presence will also raise the high temperature strength and
creep resistance of high temperature alloys and it is added to stainless steels
in proportions of up to 3.0% to improve corrosion resistance especially in
chlorides and acids.

The disadvantages of molybdenum steels are that they require higher


tempering temperatures to obtain properties comparable to plain carbon
steels and they need a longer holding period at quenching and normalising
temperature to ensure complete solid solution of the molybdenum iron
carbide.

6.2.6 Niobium (US Columbium)

Plain low carbon steels for structural purposes benefit greatly from small
additions of niobium and quantities as low as 0.02-0.04% will increase Yield
Point by 62-110N/mm2 and Ultimate Tensile Strength 54-85N/mm2. This will
give a higher Yield to Ultimate Tensile Strength ratio than in plain carbon
steels. Above 0.04%Nb this strengthening effect falls away rapidly.

Niobium is a strong carbide former but its effect on low carbon steels is
thought to be mainly by precipitation hardening. It has a marked grain
refining function sometimes reinforced by aluminium and/or vanadium
additions. Such refined grain structures give high impact values.

Niobium as a very strong carbide stabiliser, is also added to some stainless


corrosion resisting steels (type 347) of the 18/8 chromium-nickel variety to
prevent weld decay sensitivity.

41
6.2.7 Vanadium

Plain vanadium steels are manufactured to a very limited extent, but


chromium-vanadium steels containing up to 0.2% vanadium are widely used
for small and medium sections. The mechanical properties resemble those of
nickel-chromium steels, but usually show an advantage in respect of the limit
of proportionality and percentage reduction in area. Chromium-vanadium
steels are also easier to forge, stamp and machine, but are more susceptible
to mass effects of heat-treatment than the corresponding nickel-chromium
steels.

Vanadium has a strong carbide-forming tendency. It also stabilises


martensite and low temperature tempered martensite on heat-treatment and
increases hardenability. Like nickel, it restrains grain growth of the austenite.
One of the most important effects of vanadium is that it induces resistance to
softening at high temperatures provided that the steel is first heat-treated to
absorb some of the vanadium carbide into solid solution. Consequently
vanadium steels are used for hot-forging dies, extrusion dies, die-casting dies
and other tools operating at elevated temperatures.

6.2.8 Aluminium

The presence of aluminium in plain carbon and low alloy steels will tend to
stabilise ferrite but as it will normally be in very low concentrations (<0.050%)
its direct effect on hardenability will be negligible. It can however act as a
strong grain refiner (levels >0.015%) which may reduce hardenability
somewhat although improving tensile and impact strengths. It is widely used
as a deoxidant in low carbon mild steels where ductility is the principal quality.
In such steels the presence of small amounts of nitrogen introduced during
steel making will cause strain ageing to the detriment of their properties. A
small amount, up to about 0.08%, of residual aluminium by forming aluminium
nitride will suppress this phenomenon in cold rolled annealed tempered
grades. In offshore structures a Al to N ratio of at least 2:1 is sometimes
specified to overcome strain ageing during welding.

6.2.9 Sulphur

The element sulphur will tend to decrease hardenability somewhat, but as,
with a few exceptions, considerable efforts are made to eliminate it from plain
carbon and alloy steels, it is usually present only in small amounts and its
effect in this respect may be neglected.

It is the most deleterious impurity commonly present in steel. If precautions


are not taken to render it harmless it will form the brittle sulphide, FeS. This is
soluble in molten steel, but when solidification takes place the solid solubility
falls to an equivalent of 0.03% sulphur. If the effects of extensive coring are
also taken into account amounts as low as 0.01% sulphur may cause
precipitation of the sulphide at the grain boundaries. In this way the austenite
grains will become virtually coated with brittle films of ferrous sulphide. Since
this sulphide has a fairly low melting point, the steel will tend to crumble

42
during hot-working. Being brittle at ordinary temperatures, ferrous sulphide
will also render steel unsuitable for cold-working processes, or, indeed, for
subsequent service of any type.

The sulphur is rendered harmless by the presence of manganese with which


it combines preferentially to form manganese sulphide MnS. This MnS is
insoluble in molten steel and some will be lost in the slag while the rest will be
present as globules distributed throughout the steel and not associated with
the structure when solidification takes place. These globules are plastic at hot
working temperatures and the tendency of the steel to crumble is thus
removed while the globules will be rolled out as threads in subsequent
operations.

With modern steel making practices levels of <0.01%S and 0.02%P are
commonly achieved. The demand for high grade weldable steels for stringent
applications such as off-shore pipelines has led to the availability of grades
with less than 0.01%S and 0.01%P. These have become possible largely by
the selection of high-grade raw materials which themselves have low
proportions of these elements and the introduction of new steelmaking
techniques the most notable being vacuum degassing. With such steels MnS
will also be low which will reduce the extent of non-metallic inclusions with
consequent benefits to weldable steels for critical structural applications.

Free cutting steels are made with a deliberately high level of sulphur, from
0.15-0.25%S in plain carbon steels, with about five times the manganese
content to ensure its presence a globular MnS. This sulphide is widely
distributed so that on machining the cutting tool forms small chips which are
easily handled.

6.2.10 Phosphorus

In the same way as sulphur, phosphorus is almost always considered an


undesirable element in steel and its proportion is usually kept to a minimum.
It will increase hardenability but as it is present in such a small amount its
effect and that of sulphur are generally thought to cancel each other out.

Phosphorus forms the brittle phosphide Fe3P, which is soluble in ferrite. In


solution, it has a considerable hardening effect but it must be rigidly controlled
to amounts in the region of 0.05% or less because of the brittleness also
introduced, particularly if Fe3P should appear as a separate constituent in the
microstructure. Nowadays 0.02% is commonly produced and high grade
materials with 0.01%P are available.

43
6.2.11 Silicon

Silicon is widely used as a deoxidant in steelmaking and it is often present in


quantities of 0.2-0.5%. Like nickel it encourages graphite formation and it
must therefore be kept low in high carbon steels. It dissolves in ferrite where
it has a substantial strengthening effect. Hardenability is only moderately
affected by the element. Silicon is included in some heat resisting steels in
amounts up to 1.5% as it aids high temperature resistance to oxidation.

6.3 Alloying Additions in Steel - Summary

The effects of individual alloying elements are shown in tabular form below:

Alloying Effect on Effect in Effect on Effect on


Element Strengthening Forming Transformation Hardenability
Ferrite Carbides Temperatures
Manganese Strong Weak Lowers Strong
Silicon Strong None Raises Moderate
Phosphorus Strong None Lowers Moderate
Nickel Moderate None Lowers Moderate
Chromium Weak Moderate * Strong
Copper Moderate None Lowers Weak
Molybdenum Strong Strong Raises Strong
Vanadium Weak Strong Raises Mild
Tungsten Moderate Strong Raises Moderate
* Raises or lowers depending on carbon content

1. Elements which tend to form carbides - Cr, W, Ti, V, Nb, Mo, Mn.

2. Elements which tend to graphitise the carbide - Si, Co, Al, Ni. Elements
from category 1. should be present to avoid graphitisation when small
additions of the above are exceeded.

3. Elements which tend to stabilise the Austenite - Mn, Ni, Co, Cu.

4. Elements which tend to stabilise the Ferrite - Cr, W, Mo, V, Si.

The Austenite phase may disappear when a closed γ loop is formed. If no


carbon is present in the alloy the γ loop will disappear when the following
elements are present in these approximate percentages.

Cr 12.8%, S 2.0%, W 6.0%, P 0.5%, Al 1.1%, Ti 0.75%, Mo 4.0%

44
7. MICROSTRUCTURES IN STEEL ARC WELDS

7.1 Weld Metal

The large number of variables inherent in welding such as the welding process
itself, the final composition of the melt, welding speed and thermal cycle make it
unrealistic even to attempt to predict with any precision the microstructures that
can arise. Nevertheless, some general points can be made. In the welding
operation itself a pool of molten metal is retained within the work piece. This
may be likened to molten metal within an ingot mould, and on freezing, large
columnar crystals will form and since the weld pool is small, and in the case of
a single arc weld, solidification rapid, they will usually occupy the complete
cross-section of weld metal. This metal has a characteristic cellular-dendritic
structure of cast metal consisting of the rather coarse columnar austenitic
grains which curve into the weld centre line and a fine cellular network within
the grains. The transformation products resulting from the decomposition of
these grains is dependant mainly on the rate of cooling of the weld metal.

Although it has been pointed out that accurate predictions of weld metal
structure are not possible a schematic CCT (Continuous Cooling Time) (Fig 24)
in which general changes arising with differing cooling rates and the effects of
heat input, alloying elements and slag inclusions are related in a qualitative
way. Thus C, N, Mn etc tend to move the diagram to the right to give more time
for transformations. If cooling curves are superimposed on such a diagram it
will be clear that rapid cooling rate curves will approach the vertical whereas
slow cooling rate curves will approach the horizontal. All phases intersected by
such curves will be present in the final structure.

It may be expected that the strength and hardness of welds would be generally
high as rapid cooling rates tend to promote lower temperatures of
transformation and there are usually large quantities of impurities and alloying
elements. On the other hand the impact properties and toughness of such
steel welds can be impaired especially at the weld centre line which is the area
of maximum segregation where the columnar crystals meet. To this must be
added the deleterious effects of the coarse columnar structure with mixtures of
transformation products which may result in the precipitation of carbides which
on impact tend to rupture to produce brittle fractures of the pro-eutectoid ferrite.

45
Fig 24 Continuous Cooling Time. Schematic CCT diagram for steel weld
metal, summarising the possible effect of microstructure and alloying on
the transformation products for a given weld cooling time

In a plain carbon steel of 0.15-0.20%C with slow cooling and small under-
cooling below A3, a structure of ‘blocky’ ferrite with pearlite will be formed. A
medium-slow cooling rate with larger undercooling below A3 will result in a form
of ferrite known as ‘Pro-eutectoid’ ferrite and on further cooling the production
of ferrite needles known as ‘Widmanstatten’ side plates and the remaining
austenite at even lower temperatures transforming into pearlite and cementite.
Another possible phase in such weld metals at undercooling near A1 is
‘Acicular Ferrite’ the transformation to which is enhanced by the presence of
strong carbide formers such as Mo or Cr. The presence of acicular ferrite gives
improved toughness to such welds.

The acicular ferrite is formed intragranularly resulting in randomly orientated


short ferrite needles with a basket weave feature. This interlocking nature
together with the fine grain size provides maximum resistance to crack
propagation by cleavage. Notch toughness increases with increasing volume
fraction of acicular ferrite in the weld metal. The formation of either grain
boundary ferrite, ferrite side plates or upper bainite is detrimental to weld metal
toughness since these microstructures provide easy crack propagation paths.

Medium high cooling rates result in the transformation occurring below A1 with a
fast initial ferrite growth resulting in rapid carbide concentration at the austenite-
ferrite interface causing the precipitation of cementite which is again followed by
very rapid ferrite growth with the same consequences. This is known as the
‘Periodic Pearlite’ reaction. A fast cooling rate results in the formation of the
phase ‘Upper Bainite’, while a very fast cooling rate may produce ‘Lower
Bainite’ or ‘Lath martensite’.

46
7.2. Heat Affected Zone (HAZ)

To be satisfactory, dilution of a weld metal must occur by melting part of the


base metal which in the region of the melt will be exposed to similar high
temperatures. Subsequently, conditions of rapid cooling under conditions of
severe restraint will be imposed and as a result of this thermal cycle the original
microstructure and properties of the metal close to the weld will be changed.
This region is normally referred to as the HAZ.

The HAZ can be conveniently divided into a series of sub-zones and illustrated
in fig 25 are those of a 0.15%C transformable steel.

Fig 25 Various regions of the HAZ in a single pass weld

Changes taking place in the Solid-liquid transition zone depend largely on the
composition of the metal and whether the austenite melts directly or is
transformed to delta ferrite. In the latter case boundary segregation will result
in reduced grain growth and a layer of small crystals, but in the former there will
be no such changes. In the grain-coarsened HAZ pro-eutectoid ferrite
networks feature prominently for lower carbon equivalent x 2 grades but with
higher steels this will be reduced in favour of lower temperature transformation
products such as Widmanstatten side plates with possibly some martensite
mainly in the mid-grains. In the grain refined HAZ the structure tends to be fine
grained ferrite-pearlite. The intercritical HAZ can yield a wide range of probable
structures, depending on the cooling rate, and may be pearlite, upper bainite,
autotempered martensite or high carbon martensite. The subcritical HAZ
produces its most notable change in the degradation of pearlite to spheroidal
particles of cementite. The zone of unaffected base material at lower
temperatures may result in embrittlement of the structure caused by dynamic
strain ageing.

The production of high strength structural materials is mainly based on


developing a product with as small a grain size as possible but the severity of
the weld thermal cycle is such that the structure is completely modified near the
weld. In high energy processing such as submerged arc it is not uncommon for
the grain size to be increased tenfold or more in the zone closest to the fusion
line. In steels of high Carbon equivalent (above 0.40) it is likely that the grain
growth zone will contain martensite. Thus the microstructure of the grain

47
growth zone more than other zones in the HAZ will determine the properties of
the weld.

The thermal cycle that is to say the heating time, the time at temperatures or
dwell time, and the cooling time are important and must be considered together
with the presence of precipitates and their solubility at high temperatures.
Moreover the previous thermal and mechanical history is also important. Thus
for example the original steel may have been in the cold rolled condition, it may
have been annealed or normalised or subjected to other treatments. With
these aspects in mind the sub-zones may be considered separately since each
one will have a different type of microstructure with appropriately different
properties. Notwithstanding the complexities involved such structure can be
predicted with greater certainty then those arising in the weld metal.

7.3 Welding Heating Cycle

The importance of the welding heating cycle is not always fully recognised. The
temperature increases can be 200-300°C/sec and thus α to γ transformations
will be above that of equilibrium and the phase will be substantially superheated
before the change occurs. The degree of superheating will be affected by the
welding process and the thermal cycle is such that in most metals some grain
growth will occur in the HAZ which not only affects strength and toughness but
also influences grain size in the weld metal. In most steels the presence of
carbide or nitride particles tends to hinder grain growth and it has been shown
that such growth occurs predominantly at temperatures above the equilibrium-
solubility limits of such particles. Also it has been observed that most grain
growth occurs during the heating part of the thermal cycle and in some steels
experiments indicate that only about 20% of growth occurs in the cooling cycle.

In general some degree of grain growth control can be exerted either by limiting
the duration of the weld thermal cycle or by precipitate pinning, that is, mainly
by Carbides and Nitrides in steels. The best way to control grain size is to use
large quantities of very small precipitates but this is not always easy to achieve.

The employment of multi-run welds will result in a refinement of microstructure


and improvements in toughness with reductions in residual stress when
compared with a single run of the same cross-section.

48
7.4 Carbon Equivalent (Ceq)

Weldable structural steels are normally capable of being welded without the
need for any pre-heat because the fectors which cause high hardness and the
likely presence of martensite would be absent in such welds. These
characteristics are usually expressed by reference to the Carbon Equivalent
value. This is assessed in terms of how the alloying elements affect the
transformation characteristics including the martensite start temperature Ms of
the steel. An empirical formula has been devised for Carbon Equivalent (CE)
and adopted by the International Institute of Welding as follows:

A weldable structural steel will normally have a CE value not greater than
0.40%. Other formulae such as PCM used in Japan are also available. It should
be realised that such formulae and the values calculated from them are
principally guides and circumstances may arise when departures from them are
justified.

Essentially a Carbon Equivalent is a measure of the hardenability of a steel.

49
8. STRUCTURE OF INGOTS

When molten metal is poured into a metal ingot mould, solidification will usually
begin with the formation of small chill crystals at the mould-metal interface (Fig
26). These are followed by long columnar crystals growing at right angles to
the mould surface. The size of these crystals will depend on the rate of
solidification and they will be small if this is rapid, increasing in size with slower
rates. With very slow solidification, as with a sand mould for example, the
columnar and chill crystals will not generally form but dendrites will develop into
irregular equi-axed grains. Where columnar crystals have formed, planes of
weakness can be expected at sharp corners and in extreme cases even at the
ingot centre.

Fig 26 Ingot Solidification

8.1 Pipe

As a layer of cooled metal forms around the ingot walls, contraction occurs
resulting in a fall of the level of liquid. Successive layers of solid form, each
accompanied by a fall in the liquid level, the fall increasing as the volume of
liquid decreases. In this manner a central cavity is formed, known as Primary
Pipe. With the ingot mould designed ‘narrow end up’ (Fig 27) a conical
volume of metal still remains liquid after the top portion of the ingot is solid.
Solidification of this metal will give rise to further cavities known as Secondary
Pipe.

50
Fig 27 Effects of Ingot Taper and Hot Top/Feeder Head

This can be prevented by using a mould ‘wide end up’ when shrinkage will be
counteracted by molten metal feeding down from the top. The primary pipe can
also be reduced by using a brick-lined top on the mould - called a hot top or
feeder head, in which the metal will remain molten for a prolonged period acting
as a reservoir of molten metal, which will consequently feed the ingot.

8.2 Segregation (fig.28)

Molten steel contains soluble impurities - sulphur and phosphorus - and soluble
alloying elements together with insoluble impurities or slag particles in
suspension. The first crystals to separate contain less impurity than the average
composition and those elements which lower the freezing point, such as
sulphur, phosphorus, carbon, silicon and manganese collect in the last portions
to solidify. This phenomenon is known as segregation and in the case of steel
ingots means that there is a concentration of sulphur, phosphorus and carbon
in the centre and upper portions of the ingot. Associated with the pipe they are
largely removed when it is discarded.

Segregation also occurs as the microscopic scale. Impurities and soluble


alloying elements are rejected by the solidifying crystal with the result that an
alloy rich area surrounds each crystal. This can be clearly seen in carbon steel
castings, the original crystal lower in alloy content tends to be ferritic whilst the
more highly alloyed crystal surround is pearlitic.

51
Fig 28 Segregation in Steel

8.3 Steel Types

Steel, manufactured from impure pig-iron or scrap, is first purified to a great


extent by a process of oxidation which inevitably leaves the molten material
with excessive oxygen which must be reduced to render it suitable for casting to
an ingot. The steel type is defined by the degree of deoxidation.

8.3.1 Rimming Steel (Fig 29)

The structure of this type of ingot is characterised by an outer envelope or rim


of solid, comparatively pure steel, with the inner core of less solid or pure
character. Such ingots are made from steel of an effervescing nature, i.e. steel
in which deoxidation either in the ladle or in the mould has been intentionally
limited with the result that a free evolution of gas progressively takes place
during freezing. The evolution of gas towards the centre of the ingot produces
globular blowholes which counteract cooling and solidification shrinkage and
are subsequently welded together during rolling. The thickness of the rim and
the proportion of blowholes will both decrease with increasing carbon content.
This type of steel improves the yield of ingot to finished product and will provide
rolled material of good surface finish especially important in applications for low
carbon grades such as cold rolled sheets and strip. Rimming steel is not
usually available with carbon contents exceeding 0.35% and is not acceptable
for critical structural applications.

52
Fig 29 Rimming Steel

8.3.2 Balanced or Semi-killed steel

A balanced steel is a non-piping steel in which no observable gas evolution


takes place but where sufficient gas is formed during solidification to balance or
offset normal shrinkage. Numerous blow-holes are produced so the central
shrinkage cavity or pipe is prevented or considerably reduced in extent.
Balanced steels are only partially deoxidised and small additions of aluminium
are made to the moulds, if necessary, to reduce the oxygen content to that
amount which will give an ingot with a flat or slightly bulging top.

8.3.3 Killed steel

Killed steel is steel that has been completely deoxidised by additions of


manganese, silicon and sometimes aluminium, before casting, so that there is
practically no evolution of gas during solidification and sound ingots are
obtained. The shrinkage cavity or pipe is limited to upper portions of the ingot
or in the ‘feeder head’ with which such ingots are usually provided. Steel used
for continuous casting is always fully killed.

53

S-ar putea să vă placă și