Вы находитесь на странице: 1из 185

Welding consumables Part 5 - MIG/MAG and

cored carbon steel wires


Job Knowledge

Part
Part
Part
Part

1
2
3
4

To ensure that there is a consistency in composition and properties between wires from a
variety of manufacturers, specifications have been produced that enable a wire to be easily
and uniquely identified by assigning the consumable a 'classification', a unique
identification that is universally recognised.
The two schemes that are dealt with in this article are the EN/ISO method and the AWS
scheme. There are such a large number of specifications covering the whole range of
ferrous and non-ferrous filler metals, both solid wire and cored, that it will not be possible
to describe all of these here. This article therefore reviews just the carbon steel
specifications.
The identification of the solid wires is relatively simple, as the chemical composition is the
major variable although both the EN/ISO and the AWS specifications detail the strength
that may be expected from an all-weld deposit carried out using parameters given in the
specification. It should be remembered, however, that most welds will contain some parent
metal and that the welding parameters to be used in production may be different from
those used in the test. The result is that the mechanical properties of a weld can be
significantly different from those quoted by the wire supplier, hence the need to always
perform a procedure qualification test when strength is important. In addition, the
mechanical properties specified in the full designation include the yield strength. (In the
EN/ISO specifications, the classification may indicate either yield or ultimate tensile
strength).
When selecting a wire remember that the yield and ultimate tensile strengths are very
close together in weld metal but can be widely separated in parent metal. A filler metal
that is selected because its yield strength matches that of the parent metal may not,
therefore, match the parent metal on ultimate tensile strength. This may cause the cross
joint tensile specimens to fail during procedure qualification testing or perhaps in service.
The EN/ISO specification for non-alloyed steel solid wires is BS EN ISO 14341. This
specification classifies wire electrodes in the as-welded condition and in the post weld
heat-treated condition, based on classification system, strength, Charpy-V impact
strength, shielding gas and composition. The classification utilises two systems based
either on the yield strength (System A) or the tensile strength (System B):

System A - based on the yield strength and average impact energy of 47J of all-weld
metal.
System B - based on the tensile strength and the average impact energy of 27J of allweld metal.
In most cases, a given commercial product can be classified to both systems. Then either
or both classification designations can be used for the product.
The symbolisation for mechanical properties is summarised in Table 1A for classification
system A and Table 1B for classification system B. For classification system B, the 'X' can
be either 'A' or 'P', where 'A' indicates testing in the as-welded condition and 'P' indicates
testing in the post weld heat-treated condition. The symbol for chemical composition is
summarised in Table 3A and 3B of BS EN ISO 14341 based on each classification system.
For classification system A, the standard lists eleven compositions, too many to describe
completely here. Six of the wires are carbon steel with varying amounts of deoxidants, two
wires contain approximately 1% or 2.5% nickel and an additional two wires contain around
0.5% molybdenum. The designation of these wires is for example G3Si1, 'G' identifying it
as a solid wire, '3' as containing some 1.5% manganese and Si1 as containing around
0.8% silicon; G3Ni1 is a wire with approximately 1.5% manganese and 1% nickel.
Table 1A Symbols for mechanical properties based on classification system A

Symbol

Min Yield
Strength
N/mm2

35

355

38

380

42

420

46

460

50

500

UTS
N/mm
440 to
570
470 to
600
500 to
640
530 to
680
560 to
720

Min
Charpy-V Test 47 J at
Elongation Symbol
Temp C
%
22

No requirements

20

+20

20

20

-20

18

-30

4
5
6
7
8
9
10

-40
-50
-60
-70
-80
-90
-100

Table 1B Symbols for mechanical properties based on classification system B

Symbol

Min Yield
Strength
N/mm2

43X

330

49X

390

55x

460

57x

490

UTS
N/mm
430 to
600
490 to
670
550 to
740
570 to
770

Min
Charpy-V Test 27 J at
Elongation Symbol
Temp C
%
20

No requirements

18

+20

17

17

-20

-30

4
5
6
7
8
9
10

-40
-50
-60
-70
-80
-90
-100

A full designation could therefore be ISO 14341-A-G 46 5 M G3Si1 where the '-A'
designates the classification system A, the '-G' designates solid wire electrode/or deposits,
and the 'M' designates a mixed gas. An example of a System B designation could be ISO
14341-B-G 49A 6 M G3, where 'A' indicates testing in the as-welded condition.
The AWS specification AWS A5.18 covers both solid, composite stranded and cored wires
comprising six carbon steel filler metals for MAG, TIG and plasma welding in both US and
metric units.
The classification commences with the letters 'E' or 'ER'. 'E' designates an electrode. 'ER'
indicates that the filler metal may be used either as an electrode or a rod. The next two
digits designates the tensile strength in either 1000s of psi.(ksi) or N/mm 2 eg ER70 (70ksi
UTS) or ER48 (480N/mm2 UTS). However, note that there is only one strength level in the
specification.
The next two characters identify the composition, essentially small variations in carbon,
manganese and silicon contents, the wire type (solid wire (S) or metal cored or composite
wire (C)) and the Charpy-V impact values.
With one exception, the solid wires are tested using 100% CO2, the cored wires with
argon/CO2 or as agreed between customer and supplier, in which case there is a final
letter 'C' designating CO2 or 'M', a mixed gas.
The permutations in these identifiers are too many and too complicated to be able to
describe them all in sufficient detail but as an illustration, a typical designation would be
ER70S-3, a 70ksi filler metal, CO2 gas shielded and with minimum Charpy-V energy of 27J
at -20C. E70C-3M identifies the wire as a solid wire 70ksi UTS metal cored filler metal,
27J at -20C and tested with an argon/CO2 shielding gas.
The EN/ISO specification for non-alloy steel flux and metal cored wires is BS EN ISO
17632. This covers gas shielded as well as self-shielded wires. The standard identifies
electrode based on two systems in a similar way as BS EN ISO 14341, indicating the
tensile properties and the impact properties of the all-weld metal obtained with a given
electrode. Although the specification claims that the wires are all non-alloy, they can
contain molybdenum up to 0.6% and/or nickel up to 3.85%. The classification commences
with the letter 'T', identifying the consumable as a cored wire.
The classification uses the same symbols for mechanical properties as shown in Table
1A&B and a somewhat similar method to describe the composition as BS EN ISO 14341.
Thus MnMo contains approximately 1.7% manganese and 0.5% molybdenum; 1.5Ni
contains 1% manganese and 1.5% nickel. In addition to the symbols for properties and
composition, there are symbols for electrode core composition. Table 2 summarises the
symbols for electrode core type and welding position in accordance with classification
system A. Classification system B uses Usability Indicators as oppose to a one-letter
symbol for electrode core type, which can be found in Table 5B of BS EN ISO 17632.
Table 2 Symbols for electrode core type and position based on classification
system A

Flux Core
Symbol

Flux Core Type

Welding Position
Shielding
Gas

Symbol

Welding position

R
P

Rutile, slow freezing slag


Rutile, fast freezing slag

Required
Required

1
2

Basic

Required

M
V

Metal powder
Rutile or basic/fluoride
Basic/fluoride, slow freezing
slag
Basic/fluoride, fast freezing
slag
Other types

Required
4
Not required 5

W
Y
Z

All
All except V-down
Flat butt, flat and HV
fillet
Flat butt and fillet
V-down and (3)

Not required
Not required

In addition, there are symbols for gas type. These are 'M' for mixed gases, 'C' for 100%
CO2 and 'N' for self-shielded wires and 'H' for hydrogen controlled wires. A full designation
may therefore be ISO 17632-A -T46 3 1Ni B M 1 H5 in accordance with classification
system A. For classification system B, an example may be ISO 17632-B -T55 4 T5-1MAN2-UH5, where 'T5' is the usability designator, 'A' indicates test in the as-welded
condition, 'N2' is the chemical composition symbol, and 'U' is an optional designator.
The American Welding Society classification scheme for carbon steel flux cored wires is
detailed in the specification AWS A5.36. This also contains information from A5.18, but
does not officially supercede it. The full designation is ten characters in length beginning
'E' for an electrode then designators for strength, welding position, cored wire, usability,
shielding gas, toughness, heat input limits and diffusible hydrogen, the last four
designators being optional.
There are two strength levels - E7 (70ksi UTS) and E6 (60ksi UTS) followed by a
designator for welding position,'0' for flat and horizontal and '1' for all positions, including
vertical-up and vertical-down.
The next symbol 'T' identifies the wire as being flux cored and this is followed by either a
number between 1 and 14 or the letter 'G' that identifies the usability. This number refers
to the recommended polarity, requirements for external shielding, and whether the wire
can be used to deposit single or multi-pass welds. 'G' means that the operating
characteristics are not specified. The sixth letter identifies the shielding gas used for the
classification, 'C' being 100% CO2, 'M' for argon/CO2, no letter indicating a self-shielded
wire.
The non-compulsory part of the designation may include the letter 'J', confirming that the
all-weld metal test can give Charpy-V values of 27J at -40C; the next designator may be
either 'D' or 'Q'. These indicate that the weld metal will achieve supplementary mechanical
properties at various heat inputs and cooling rates. The final two designators identify the
hydrogen potential of the wire.
A full AWS A5.36 designation could therefore be E71T-2M-JQH5. This identifies the wire as
a cored, all positional wire to be used with argon/CO2 shielding gas on electrode positive
polarity. The weld metal should achieve 70ksi tensile strength, 27J at -40C, 58 to 80ksi
yield strength at high heat input, a maximum 90ksi at low heat input, and a diffusible
hydrogen content of less than 5ml of H2/100g of deposited weld metal.
This article was written by Gene Mathers, reviewed and modified by Runlin Zhou.

SUMMARY
This lecture commences with a discussion of the need for civil and structural
engineers to have a basic knowledge of the metallurgy of steel. Then the
crystalline nature of irons and steels is described together with the influence

of grain size and composition on properties. The ability of iron to have more
than one crystalline structure (its allotropy) and the properties of the
principal crystalline forms of alloys of iron and carbon are discussed.
The metallurgy and properties of slowly cooled steels are reviewed,
including the influence of grain size, rolling, subsequent heat treatment and
inclusion shape and distribution. Rapidly cooled steels are treated
separately; a brief description of quenching and tempering is followed by a
discussion of the influence of welding on the local thermal history.
Hardenability, weldability and control of cracking are briefly discussed.
Finally the importance of manganese as an alloying element is introduced.
1. INTRODUCTION

1.1 Why Metallurgy For Civil and Structural Engineers?


The engineering properties of steel, i.e. strength, ductility and resistance
against brittle fracture, depend on its crystalline structure, grain size and
other metallurgical characteristics.
These microstructural properties are dependent on the chemical composition
and on the temperature-deformation history of the steel. Heat treatments that
occur during welding may also have a large influence on the engineering
properties.
When selecting steel for welded structures, it is important to have at least a
basic knowledge of metallurgy. This knowledge is required especially when
large and complicated structures are being designed, such as bridges,
offshore structures, and high rise buildings.
Selecting materials, welding processes and welding consumables usually
requires consultation of "real" metallurgists and welding specialists. A basic
knowledge of metallurgy is essential for communication with these
specialists.
Finally, a basic knowledge of metallurgy also enables civil and structural
engineers to have a better understanding of the engineering properties of
steel and the performance of welded structures.
1.2 The Scope of Lectures in Group 2
Lecture 2.1 deals with the characteristics of iron-carbon alloys. Where
possible, direct links are indicated to the engineering properties and
weldability of steel. These subjects are covered inLectures 2.2 and 2.6
respectively.

Lecture 2.3 describes steelmaking and the forming of steel into plates and
sections. The various processes for controlling the chemical composition
and the different temperature-deformation treatments are discussed. Most of
the underlying principles described in Lecture 2.1 are applied.
Steels are available in various grades and qualities. The grade designates the
strength properties (yield strength and ultimate strength), while the quality is
mainly related to resistance against brittle fracture. Grades and qualities are
explained in Lecture 2.4. A system for choosing the right quality according
to Eurocode 3 (Annex C) [1] is presented. Some guidelines for the selection
of steel grade are given.
2. STRUCTURE AND COMPONENTS OF STEEL

2.1 Introduction
To get an impression of the metallurgical structure of steel, a piece of steel
bar can be cut to expose a longitudinal section, the exposed surface ground
and polished and examined under a microscope.
At modest magnifications, a few particles are seen which are extended in the
direction of rolling of the bar, see Slide 1. These particles are inclusions.
They are non-metallic substances which have become entrained within the
metal during its manufacture, mostly by accident but sometimes by design.
Their presence does not affect the strength but has an adverse effect on
ductility and toughness. Particular types of inclusion can greatly enhance the
machinability of steels and may therefore be introduced deliberately.

Slide 1 : Longitudinal stringers of inclusions in hot rolled steel. (x 500)


To reveal the true structure of the metal, the polished surface must be
chemically etched. When this is done, a wide diversity of microstructure
may be seen which reflects the composition of the steel and its processing,
see Slides 2 - 5.
The microstructure has a significant effect on the engineering properties as
described in later sections of this lecture.
2.2 The Components of Steel
Steels and cast irons are alloys of iron (Fe) with carbon (C) and various
other elements, some of them being unavoidable impurities whilst others are
added deliberately.
Carbon exerts the most significant effect on the microstructure of the
material and its properties. Steels usually contain less than 1% carbon by
weight. Structural steels contain less than 0,25% carbon: the other principal
alloying element is manganese, which is added in amounts up to about
1,5%. Further alloying elements are chromium (Cr), nickel (Ni),
molybdenum (Mo) etc. Elements such as sulphur (S), phosphorus, (P),
nitrogen (N) and hydrogen (H) usually have an adverse effect on the
engineering properties and during the steel production, measures are taken
to reduce their contents. Cast irons generally contain about 4% carbon. This
very high content of carbon makes their microstructure and mechanical
properties very different from those of steels.
Each of the microstructures shown in Slides 2, 3, 4 and 5 is an assembly of
smaller constituents. For example, the 0,2% C steel of Slide 2 is
predominantly an aggregate of small, polyhedral grains, in this case <20m
in size. Closer examination of one of these grains shows it to be a single
crystal. However, unlike crystals of quartz or silicon or copper sulphate,
crystals of iron (Fe) are soft and ductile. The internal structure of these
crystals is discussed later.

Slide 2 : Microstructure of hot rolled steel containing 0,2% carbon showing


ferrite (white) and pearlite colonies (dark). (x 200)

Slide 3 : Microstructure of hot rolled steel containing 0,36% carbon showing


increased proportions of pearlite (dark). (x 500)

Slide 4 : Microstructure of heat treated hot rolled steel containing 0,36%


carbon showing spheroidised pearlite (dark) in a ferrite matrix. (x 750)

Slide 5 : Microstructure of quenched hot rolled steel containing 0,36%


carbon showing bainite (x 200)
The steel of Slide 2 is an example of a polycrystalline substance which has
been made visible by polishing and etching.
(a) The surface is polished but not etched.

(b) The surface is polished and etched. Different reflections of the light
indicate different orientation of crystals (polycrystalline structure).
(c) Some etchants affect only the grain boundaries. These etchants are used
when it is required to investigate the grain structure, e.g. to estimate the
grain size.
(d) The appearance of etched grain boundaries of Figure 1c.
(e) The appearance of a steel with 0,15% carbon (enlargement 100x). The
dark areas are pearlite. The grain boundaries are clearly indicated. The dark
areas indicate the presence of carbon.
By adjusting the history of rolling and heating treatment experienced by the
steel during its production, the grain size can be altered. This technique is
useful because the grain size affects the properties. In particular, the yield
strength is determined by the grain size, according to the so-called Petch
equation:
y = o + kd-1/2
where y is the yield strength
o is effectively the yield strength of a very large isolated crystal: for mild
steel this is 50N/mm2
d is the grain size in mm
k is a material constant, which for mild steel is about 20N/mm-3/2
Thus, if the grain size is 0.01 mm, y 250N/mm2.
2.3 The Crystal Structure
The internal structure of the crystal grains is composed of iron atoms
arranged according to a regular three-dimensional pattern. The pattern is
illustrated in Figure 2. This pattern is the body-centred cubic crystal
structure; atoms are found at the corners of the cube and at its centre. The
unit cell is only 0,28nm along its edges. A typical grain is composed of
about 1015 repetitions of this unit.

This crystal structure of iron at ambient temperature is one of the major


factors determining the metallurgy and properties of steels.
Steels contain carbon. Some of it, a very small amount, is contained within
the crystals of iron. The carbon atoms are very small and can fit, with some
distortion, into the larger gaps between the iron atoms. This arrangement
forms what is known as an interstitial solid solution: the carbon is located in
the interstices of the iron crystal.
In the steels of Slides 2, 3 and 4, most of the remaining carbon has formed a
chemical compound with the iron, Fe3C, iron carbide or cementite. Iron
carbide is also crystalline but it is hard and brittle. With 0,1%C, there is only
a small amount of Fe3C in steel. The properties of such steel are similar to
those of pure iron) [2]. It is ductile but not particularly strong and is used for
many purposes where ability to be shaped by bending or folding is the
dominant requirement.
For a steel of higher carbon content, say 0,4%, as shown in Slide 5, a low
magnification shows it to be composed of light and dark regions - about
50:50 in this case. The light regions are iron crystals containing very little
dissolved carbon, as in the low carbon steel. The dark regions need closer
examination. Slide 6 shows one such region at higher magnification. It is
seen to be composed of alternate layers of two substances, iron and Fe3C.
The spacing of the laminae is often close to the wavelength of light and
consequently the etched structure can act as a diffraction grating, giving
optical effects which appear as a pearl-like iridescence. Consequently, this
mixture or iron and iron carbide has acquired the name 'pearlite'. The origin
of the pearlite and its effect on the properties of steel are revealed by
examining what happens during heating and cooling of steel.

Slide 6 : Polycrystalline structure of steel containing 0,4% carbon. (x 400)


3. IRON-CARBON PHASES

3.1 Influence of Temperature on Crystal Structure


The crystal structure of steel changes with increasing temperature. For pure
iron this change occurs at 910 C. The body-centred cubic (bcc) crystals of
Figure 2 change to face-centred cubic (fcc) crystals as illustrated in Figure 3.
For fcc crystals the atoms of iron are on the cube corners and at the centres
of each face of the cube. The body-centred position is empty.

A given number of atoms occupy slightly less volume when arranged as fcc
crystals than when arranged as bcc crystals. Thus the change of the crystal
structure is accompanied by a volume change. This change is illustrated in
Figure 4. When a piece of pure iron is heated, expansion occurs in the
normal way until the temperature of 910 C is reached. At this temperature
there is a step contraction of about % in volume associated with the
transformation from the bcc to fcc crystal structure. Further heating gives
further thermal expansion until, at about 1400C the fcc structure reverts to
the bcc form and there is a step expansion which restores the volume lost at
910C. Heating beyond 1400C gives thermal expansion until melting
occurs at 1540C. The curve is reversible on cooling slowly.

The property that metals may have different crystal structures, depending on
temperature, is called allotropy.
3.2 Solution of Carbon in bcc and fcc Crystals
When the atoms of two materials A and B have about the same size, crystal
structures may be formed where a number of the A atoms are replaced by B
atoms. Such a solution is called substitutional because one atom substitutes
for the other. An example is nickel in steel.
When the atoms of two materials have a different size, the smaller atom may
be able to fit between the bigger atoms. Such a solution is called interstitial.
The most familiar example is the solution of carbon in iron. In this way the
high temperature fcc crystals can contain up to 2% solid solution carbon at
1130C, while in the low temperature bcc crystals, the maximum amount of
carbon which can be held in solution is 0,02% at 723C and about 0,002% at
ambient temperature.
Thus a steel containing 0,5% carbon, for example, can dissolve all the
carbon in the higher temperature fcc crystals but on cooling cannot maintain
all the carbon in solution in the bcc crystals. The surplus of carbon reacts
with iron to form iron carbide (Fe3C), usually called cementite. Cementite is
hard and brittle compared to pure iron.
The amount of cementite and the distribution of cementite particles in the
microstructure is important for the engineering properties of steel.
The distribution of cementite is highly dependent on the cooling rate. The
distribution may be explained by considering the so-called iron-carbon
phase diagram, see Section 3.4.
3.3 Nomenclature
The following nomenclature is used by the metallurgist:
Ferrite or -Fe

The bcc form of iron in which up to 0,02%C by weight may be dissolve

Cementite

Iron carbide Fe3C (which contains about 6,67%C).

Pearlite

The laminar mixture of ferrite and cementite described earlier. The ove

Austenite or -Fe

The fcc form of iron which exists at high temperatures and which can c

Steel

Alloys containing less than 2% carbon by weight.

Cast Iron

Alloys containing more than 2% carbon by weight.

Steel used in structures such as bridges, buildings and ships, usually


contains between 0,1% and 0,25% carbon by weight.

3.4 The Iron-Carbon Phase Diagram


The iron-carbon phase diagram is essentially a map. The most important
part is shown in Figure 5. More details are given in Figure 6.

Any point in the field of the diagram represents a steel containing a


particular carbon content at a particular temperature.
The diagram is divided into areas showing the structures that are stable at
particular compositions and temperatures.
The diagram may be used to consider what happens when a steel of 0,5%C
is cooled from 1000C (Figure 6).
At 1000C the structure is austenite, i.e. polycrystalline fcc crystals with all
the carbon dissolved in them. No change occurs on cooling until the
temperature reaches about 800C. At this temperature, a boundary is crossed
from the field labelled Austenite () to the field labelled Ferrite + Austenite
( + ), i.e. some crystals of bcc iron, containing very little carbon, begin to
form from the fcc iron. Because the ferrite contains so little carbon, the
carbon left must concentrate in the residual austenite. The carbon content of
the austenite and the relative proportions of ferrite and austenite in the
microstructure adjust themselves to maintain the original overall carbon
content.
These quantities may be worked out by considering the expanded part of the
iron-carbon diagram shown in Figure 7. Imagine that the steel has cooled to
750C. The combination of overall carbon content and temperature is
represented by point X.

All the constituents of the microstructure are at the same temperature. A line
of constant temperature may be drawn through X. It cuts the boundaries of
the austenite and ferrite field at F and A. These intercepts give the carbon
contents of ferrite and austenite respectively at the particular temperature.
If, now, the line FA is envisaged as a rigid beam which can rotate about a
fulcrum at X, the 'weight' of austenite hanging at A must balance the
'weight' of ferrite hanging at F. This is the so-called Lever Rule:
Weight of ferrite FX = Weight of austenite AX
The ratio of ferrite to austenite in the microstructure is then given by:

Thus, as the steel cools, the proportion of ferrite increases and the carbon
content of the remaining austenite increases, until cooling reaches 723C. At
this temperature the carbon content of the austenite is 0,8% and it can take
no more. Cooling to just below this temperature causes the austenite to
decompose. It decomposes into the lamellar mixture of ferrite and Fe3C
identified earlier as pearlite.
The proportions of ferrite and pearlite in the microstructure, say at 722C,
are virtually the same as the proportions of ferrite and austenite immediately
before the decomposition at 723C. Thus, referring to Figure 7 and using the
Lever Rule:
Weight of ferrite F X = Weight of pearlite F P
In this case, there should be about twice as much pearlite as ferrite.
For other steels containing less than 0,8%C, the explanation is identical
except for the proportions of pearlite in the microstructure below 723C.
This varies approximately linearly with carbon content between zero at
0,02%C and 100% at 0,8%C. A typical mild steel containing 0,2%C would
contain about 25% pearlite.
For steels containing a greater percentage of carbon than 0,8%, the structure
is fully austenitic on cooling from high temperatures. The first change to
occur is the formation of particles of Fe3C from the austenite. This change
reduces the carbon content of the residual austenite. On further cooling, the
carbon content of the austenite follows the line of the boundary between
the field and + Fe3C field. Once again, on reaching 723C the carbon

content of the austenite is 0,8%. On cooling further, it decomposes into


pearlite as before. Therefore, the final microstructure consists of a few
particles of Fe3C embedded in a mass of pearlite, see Figure 6.
4. COOLING RATE

4.1 Cooling Rate During Austenite to Ferrite Transformation and


Grain Size
During cooling of austenite, the new bcc ferrite crystals start to grow from
many points. The number of starting points determines the number of ferrite
grains and consequently the grain size. This grain size is important because
the engineering properties are dependent on it. Small grains are favourable.
By adding elements like aluminium and niobium, the number of starting
points can be increased. Another important factor is the cooling rate. When
cooling is slow, the new ferrite grains develop from only a few most
favourable sites. At high cooling rates, the number of starting points will be
much higher and the grain size smaller. Slides 7 - 9 shows steels with
various grain sizes, produced at different finish rolling temperatures.

Slide 7 : Microstructure of pearlite. (x 1000)


Another important factor is that, when a fine grained steel is heated to a
temperature in excess of about 1000C, some of the austenite grains grow
while neighbouring grains disappear. This grain growth occurs during
welding in the so-called heat affected zone (HAZ). This is a 3-5 mm wide

zone in the plate adjacent to the molten metal. Microstructural changes in


the heat affected zone usually give rise to a deterioration of the engineering
properties of the steel.
4.2 Slowly Cooled Steels
4.2.1 Influence of carbon on the microstructure
The iron-carbon phase diagram in Figures 5 and 6 shows that, for structural
steel (between 0,1% and 0,25% carbon), the formation of ferrite starts at
about 850C and ends at 723C. It will be remembered that ferrite can
contain hardly any carbon. Consequently, the austenite phase transforms to
ferrite and cementite (Fe3C).
When the cooling rate is slow, the carbon atoms have time to migrate to
separate "layers" in the microstructure and to form the structure called
pearlite, as shown before in Slides 2, 3, 4 and 5. The ferrite in this mixture is
soft and ductile. The cementite constituent is hard and brittle. The mixture
(pearlite) has properties between these two extremes.
The tensile strength properties of a steel containing both ferrite and pearlite
roughly scale according to the proportions of these constituents in the
microstructure as seen in Figure 8.

The dramatic effect of carbon content on toughness is shown in Figure 9.


Increasing pearlite content decreases the upper shelf toughness and increases
the ductile-brittle transition temperature.

Figures 8 and 9 illustrate one of the difficulties in the choice of carbon


content. Increasing the carbon content is beneficial in that it improves yield

strength and ultimate tensile strength, but is undesirable in that it reduces


ductility and toughness. A high carbon content may also cause problems
during welding, see Section 4.3.
In European Norm 10025, Table 3, [3] the chemical composition for flat and
long products is given. An extract is presented in Figure 10. The designation
S235 JR, for example, indicates that the yield strength is at least 235
N/mm2. It is emphasised that the compositional values in the table are
maximum values. Many steelmakers achieve much lower levels, resulting in
better ductility, resistance against brittle fracture, and weldability.

The lowest carbon content that can be achieved easily on a large scale is
about 0,04%. This content is characteristic of sheet or strip steels intended to
be shaped by extensive cold deformation, as in deep drawing.
Carbon contents of more than 0,25% are used in the wider range of general
engineering steels. These steels are usually put into service in the quenched
and tempered state (see below) for a great multiplicity of purposes in
mechanical engineering. High strength bolts for some structural applications
would also be steels of this type.
4.2.2 The need for control of grain size
The mechanical properties of steel are affected by grain size. Slides 8 and 9
show microstructures of two samples of the same batch of mild steel which
have been treated, by methods outlined in Section 4.2.3, to give different
grain sizes. Reduction in grain size improves yield strength but also has a
profound effect on the ductile/brittle transition temperature, see Figure 11.
Thus, there are several benefits from the same microstructural charge. This
is an unusual circumstance in metallurgy where adjustments to improve one
property often mean a worsening of another and a compromise is necessary.
An example of such compromise relates to carbon content, already
discussed above.

Slide 8 : Microstructure of typical hot rolled structural steel containing


0,15% carbon and showing white ferrite grains and pearlite colonies. (x
200)

Slide 9 : Refined microstructure of controlled rolled structural steel


containing 0,15% carbon (white ferrite grains and pearlite colonies. (x 200)

4.2.3 Grain size control by normalising


In Section 4.2.1 the transformations that can occur when steels are cooled
slowly are described. To form ferrite and pearlite from austenite, the carbon
atoms in the steel must change their positions. The diffusion processes
which transport the atoms within the solid occur at rates which depend
exponentially on temperature. The rate of cooling also affects these
transformations.
If the cooling rate is increased the transformations occur faster. In addition,
the diffusional processes cannot keep up with the falling temperature. Thus,
a steel cooled very slowly in a furnace keeps close to the requirements of the
phase diagram. But the same steel, removed from the furnace and allowed to
cool in air, may undercool before completing its sequence of
transformations. This more rapid cooling has two effects. First it tends to
increase slightly the proportion of pearlite in the microstructure. Secondly it
produces ferrite with a finer grain size and pearlite with finer lamellae. Both
of these microstructural changes give higher yield strength and better
ductility and toughness.
Furnace cooled steels are known as fully annealed steels. Air cooled steels
are known as normalised steels.
Grain size can also be affected by the temperature to which the steel is
heated in the austenite range. The grains of austenite coarsen with time, the
rate of coarsening increasing exponentially with temperature. The
coarsening is important because the transformation to ferrite and pearlite on
cooling starts at the grain boundaries in the austenite. If the new structures
start growing from points which are further apart in a coarse grained
austenite, the grain size of the resulting ferrite is itself coarser. Thus, steels
should not be overheated when austenitising before normalising.
The temperature to which the steel is heated before cooling in air is usually
referred to as the normalising temperature. The requirements of the last
paragraph mean that this temperature should be as low as possible, as long
as the structure is single phase austenite. A glance at the phase diagram of
Figure 5 shows that the normalising temperature decreases as the carbon
content increases from zero to 0,8%. It should lie in the hatched band shown
in Figure 12.

4.2.4 Microstructural changes accompanying hot rolling of steels


Structural steel sections are produced by hot rolling ingots or continuously
cast strand into the required forms. The rolling processes have important
effects on the development of the microstructure in the materials.
The early stages of rolling are carried out at temperatures well within the
austenite range, where the steel is soft and easily deformed. The deformation
suffered by the material breaks up the coarse as-cast grain structure but, at
these high temperatures, the atoms within the material can diffuse rapidly
which allows the deformed grains to recrystallise and reform the equiaxial
polycrystalline structure of the austenite.
Heavy deformation at low temperatures in the austenite range gives finer
recrystallised grains. If the rolling is finished at a temperature just above the
ferrite + austenite region of the equilibrium diagram and the section is
allowed to cool in air, an ordinary normalised microstructure having
moderately fine-grained ferrite results. Modern controlled rolling techniques
aim to do this, or even to roll at still lower temperatures to give still finer
grains.
If the temperature falls so that the rolling is finished in the ferrite + austenite
range, the mixture of ferrite and austenite grains is elongated along the
rolling direction and a layer-like structure is developed. If now, the section
is air-cooled, the residual austenite decomposes into fine-grained ferrite and
pearlite, with the later being present as long, cigar shaped, bands in the
material, as in Slide 10. Structural steels are not harmed by microstructures
of this sort.

Slide 10 : Microsection through a fillet weld on structural steel showing


three distinct regions: the coarse grained cast structure of the weld deposit,
the heat affected zone, and the unaffected microctructure of the parent steel.
(x 200)
If the finish rolling temperature drops further, to below 723C, the
equilibrium diagram shows that the structure should be a mixture of ferrite
and pearlite. Rolling in this range is usually restricted to low carbon steels
containing less than 0,15%C because the presence of pearlite makes rolling
difficult.
If the temperature is above about 650C, the ferrite grains recrystallise as
they are deformed, as was the case with austenite. The carbide laths (Fe3C)
in the pearlite become broken and give rise to strings of small carbide
particles extending in the direction of rolling, see Slide 11. The ferrite from
the pearlite becomes indistinguishable from the rest of the ferrite.

Slide 11 : Macrosection through a butt weld on hot rolled steel plate, typical
of line pipe weld.
If rolling is done at ambient temperature, the pearlite is broken up in the
same way, but the ferrite can not recrystallise. It work-hardens, i.e. the yield
and ultimate tensile strength of the steel increase, and the ductility
decreases, see Figure 13. As cold rolling continues, the force required to
continue deformation increases because of the increasing yield strength.
Furthermore, the steel becomes less ductile and may begin to split. The
amount of cold rolling that can be done is therefore very much smaller than
that which can be achieved when the steel is hot.

Of course, cold working need not be applied by rolling. Any way of


deforming the material causes work hardening. For example, high strength
steel wire is made by cold-drawing, imparting large deformations. In
another example, one type of reinforcing bar is made by twisting square
section bar into a helical form. The cold-deformation produced in this way is
not large but causes significant work hardening.
To restore the ductility and at the same time reduce the work hardened state
of the material, it is necessary to reform the isotropic, polycrystalline
structure of the ferrite. Re-heating to temperatures between about 650C and
723C allows the ferrite to recrystallise. The carbide particles are unaffected
by this treatment.
Thus, there is another technique for controlling the grain size of steel. The
greater the amount of deformation before the recrystallisation treatment and
the lower the temperature of the treatment, the finer is the final grain size.
Because this type of treatment does not involve the formation and
decomposition of austenite, it is known as sub-critical annealing. The
resulting microstructure has good ductility and deep drawing characteristics.
Sheet steels of low carbon content (< 0,1%C) are usually supplied in this
condition. Objects such as motor car body panels are formed from such
steels by cold pressing.
If the material is heated into the austenite range, subsequent cooling reforms
the normalised microstructure.
4.3 Rapidly Cooled Steels
4.3.1 Formation of martensite and bainite
Normalising causes steels to undercool below the requirements of the phase
diagram before the austenite transforms into fine ferrite and pearlite. Still
further increases in cooling rate give further undercooling and still finer
microstructures.
Very rapid cooling by quenching into cold water, causes the formation of
ferrite and pearlite to be suspended. The internal diffusion-controlled
rearrangement of atoms needed to form those products cannot occur
sufficiently rapidly. Instead, new products are formed by microstructural
shear transformations at lower temperatures. Very fast cooling gives
martensite: its microstructure is shown in Slide 12. When martensite forms,
there is no time for the formation of cementite and the austenite transforms
to a highly distorted form of ferrite which is super saturated with dissolved
carbon. The combination of the lattice distortion and the severe work
hardening resulting from the shear deformation processes necessary to

achieve the transformation cause martensite to be extremely strong but very


brittle.

Slide 12 : Longitudinal section of hot rolled structural steel showing dark


bands of pearlite in a ferrite matrix. (x 200)
Less rapid cooling can give a product called bainite,. This is similar to
tempered martensite where much of the carbon has come out of solution and
formed fine needles of cementite which reinforce the ferrite.
4.3.2 Martensite in welded structures
Civil engineering structures are not heat-treated by heating to, say, 900C
and quenching into water. However, there is one important circumstance
which can produce martensite in localised parts of the structure, and that is
welding. The weld zone is raised to the melting temperature of the steel and
the immediately adjacent solid metal is heated to temperatures well within
the austenite range. When the heat source is removed, the whole region
cools at rates determined mainly be thermal conduction into the surrounding
mass of cold metal. These rates of cooling can be very large, exceeding
1000C per second in some cases and can produce transformation structures
such as martensite and bainite. The properties of rapidly cooled steels and
the influence of carbon content on the nature of the transformation product ferrite and pearlite, or bainite, or martensite - are discussed below.

Figure 14 shows the hardness of martensite as a function of its carbon


content. Reheating martensite to temperatures up to about 600C causes
cementite to precipitate which causes the steel to soften and become much
tougher. This reheating is known as tempering. The extent of these changes
increases as the reheating temperature increases, as shown in Figure 15.
Tempering at 600C produces an extremely tough material. What is more,
its ductile-brittle transition temperature is lower than for the same steel in
the normalised condition. Bainite has properties similar to those of tempered
martensite.

4.3.3 Quenching and tempering


The process of quenching and tempering, when allied to changes of steel
composition, can produce a bewilderingly wide range of properties. Steels
heat-treated in this way are used for a multiplicity of general engineering
purposes which demand hardness, wear resistance, strength and toughness.
Once again, compromises must be struck between these desirable properties
but generally quenched and tempered steels exhibit optimum combinations
of strength and toughness. For structural purposes quenched and tempered
plate is used in large storage tanks, hoppers, earthmoving equipment, etc.
Martensite produced in a weld heat-affected zone as a result of single pass
welding would be in its hard and brittle untempered condition. Furthermore,
the formation of martensite from austenite is accompanied by a volume
expansion of approximately 0,4%. This expansion, together with the uneven
thermal contractions taking place as a result of uneven cooling, can produce
local stresses of sufficient magnitude to crack the martensite. Because this
type of cracking occurs after the HAZ has cooled, it is referred to as cold
cracking. The cracking problem can be further aggravated if the weld has
picked up hydrogen. Sources of hydrogen during welding might include
moisture from the atmosphere or damp welding electrodes. Hydrogen
dissolved in the weld metal diffuses to the hard HAZ where it initiates
cracks at sites of stress concentration. This diffusion can lead to cracking
which occurs some time, even days, after the welding is completed. Hard
HAZs of low ductility are less able to cope with this problem than are softer
and more ductile materials. This type of cracking is called delayed cracking
or hydrogen cracking.
Avoidance of cold cracking and hydrogen cracking requires that the material
should not be overhardened. As a rule of thumb, as-welded hardnesses of
less than about HV = 350 are considered to be acceptable. In modern fine
grain low carbon steels the "allowable" hardness may be increased to HV =
400 or even HV = 450.
The danger of hydrogen cracking may also be present in high strength
quenched and tempered steels, e.g. 10.9 bolts (Re 900 N/mm2 and
Rm 1000 N/mm2). When such bolts are electroplated with zinc or
cadmium, hydrogen may be picked up from the plating bath. Usually
cracking does not occur until sometime after tightening bolts when the
hydrogen has diffused to the sites of stress concentration at the thread roots.
4.3.4 Control of martensite formation
Martensite forms because ferrite and pearlite did not! If follows that
metallurgical factors which promote the formation of ferrite and pearlite

inhibit the production of martensite. The ability of a steel to form martensite


rather than ferrite and cementite is called hardenability. Note that this term
does not refer to the absolute value of hardness obtained, but to the ease of
formation of martensite.
The most convenient method of assessing hardenability is the so called
Jominy end quench. A rod-shaped sample is austenitised and then quenched
by spraying water onto one end face such that different cooling rates are
produced along the length of the bar. Thereafter, a flat is ground along its
length and the hardness measured as a function of distance from the
quenched end.
Some typical results are shown in Figure 16 for three different steels. For a
carbon steel containing 0,08%C and 0,3%Mn, cooling rates at 700C of
greater than about 50C s-1 are necessary to form martensite. On the other
hand in the 0,29%C, 1,7%Mn steel, martensite forms at much slower
cooling rates. It is mainly the increased carbon content that causes this
difference. In the alloy steel illustrated, martensite is formed even at very
slow cooling rates.

The significance of these curves depends very much on what is being


produced. If it is a thick-section gear wheel, the alloy steel would be ideal. It
could be cooled gently and still produce martensite, the gentle cooling being
an advantage because it would reduce stresses arising from differential
contraction rates, and hence reduce the possibility of quench cracking.

Thereafter, it could be tempered to achieve the desired combination of


strength and toughness. On the other hand, for a welded joint, the plain
carbon steel would be preferable in which it is difficult to form martensite
and the hardness of any martensite produced would be relatively low.
Welding presents particular problems for the metallurgist. Slide 13 shows a
micro section through a typical structural weld. The micro structures range
from the coarse grained cast structure of the weld deposit, to the heat
affected zone (HAZ) and to the unaffected microstructure of the parent
metal. Both the deposited weld metal and the HAZ must have adequate
strength and toughness after welding.

Slide 13 : Microstructure of martensite (x 500)


For welding, a steel of low hardenability is therefore required. Hardenability
is affected by steel composition, including not only carbon content but other
alloying elements as well. To take all of these factors into account, the
concept of the carbon equivalent value is used. There are a number of ways
of calculating carbon equivalents for use in different circumstances. In the
context of welding:

C.E. =
If the CE is lower than about 0,4%, the steel can be welded with little or no
trouble from martensite and HAZ hydrogen cracking. As indicated before,

the cooling rate is also an important factor, which means that during
welding, thick plates are more susceptible to hydrogen cracking than thin
plates. To reduce susceptibility to martensite formation, the cooling rate
(between 800C and 500C) can be reduced by preheating the plates before
welding.
5. INCLUSIONS

5.1 Sulphur, Phosphorus and Other Impurities


One tonne of steel, a cube with sides of about 0,5m, contains between
1012 and 1015 inclusions which can occupy up to about 1% of the volume.
The total content is largely determined by the origins of the ores, coke and
other materials used to extract the metal in the first place, and by the details
of steelmaking practice.
The principal impurities which worry steelmakers are phosphorus and
sulphur. If not at very low concentrations, these impurities form particles of
phosphide and sulphide which are harmful to the toughness of the steel.
Typically, less than 0,05% of each of these elements is demanded. Low
phosphorus contents are relatively easily attained during the refining of the
pig iron into steel, but sulphur is more difficult to remove. It is controlled by
careful choice of raw materials and, in modern steelmaking, by extra
processing steps to remove it.
Manganese is always added to steels. It has several functions but the
important one in this context is that it combines with the sulphur to form
manganese sulphide (MnS). If the manganese were not present, iron
sulphide would form which is much more harmful than MnS.
Some of the inclusions are too small to be seen with optical microscopes and
must be detected by more elaborate methods. Among this group, which are
mainly equiaxial in shape, are nitrides of aluminium and titanium which are
deliberately introduced in order to inhibit the processes which lead to
coarsening of grain size.
Other inclusions, large enough to be seen readily with the optical
microscope, include entrained particles of slag, deoxidation products and
manganese sulphide. At hot rolling temperatures, these inclusions are plastic
and are elongated in the rolling direction. The result is shown in Figure 1.
The properties of steels containing such inclusions reflect both the volume
of the inclusions and the anisotropy of their shapes, see Figures 17 and 18.

In recent years, a number of practices have been introduced which aim to


reduce the inclusion content in the molten steel before it is cast into ingots.
Sulphur contents of 0,01% or less are now regularly produced. These
processes produce what have become known as 'clean steels'. The
expression is relative. Clean steels still contain many inclusions, but are
significantly tougher than ordinary steels. Inclusion shape control is also
practised in better quality steels. Additions of calcium or cerium and other
rare earth elements to the refined molten steel combine with the sulphur in
preference to the manganese. Sulphides of these elements appear in the final
microstructure as equiaxial particles and are not so deleterious to the
through-thickness ductility of the material as elongated MnS inclusions.
Steels treated in these ways are used in applications where toughness is of
paramount importance and where the extra cost can be justified. Examples
include high integrity pressure vessels, oil and gas pipelines and the main
legs of offshore platforms. The introduction of continuous casting has also
improved the quality of conventional structural steels.
5.2 Manganese in Structural Steels
It has been noted earlier that the residual sulphur impurity in steel is less
harmful when formed into particles of MnS rather than iron sulphide. The
presence of small amounts of manganese in the steel confers several other
benefits. In normalised steels, it tends to increase the amount of
undercooling before the start of the formation of ferrite and pearlite. This
gives finer grained ferrite and more finely divided pearlite. Both of these
changes improve strength and reduce the ductile/brittle transition
temperature. The dissolution of the manganese atoms in the ferrite crystals
also improves the strength of the ferrite. These effects on properties are
summarised in Figures 19 - 21.

If the manganese content is increased too much, its effect ceases to be


beneficial and can become harmful because it increases hardenability, i.e.
promotes martensite formation. It is for this reason that a maximum
manganese content is specified: For S355 in Table 3 of EN 10025 this
maximum is 1,7% by weight, see Figure 16. A convention has also grown
that distinguishes between plain carbon steels, i.e. steels containing <
1%Mn, and carbon manganese steels i.e. >1%Mn.
6. CONCLUDING SUMMARY

Steels used for structural purposes generally contain up to about


0,25%C, up to 1,5%Mn and with carbon equivalents of up to
0,4%. They are mostly used in the hot-rolled, normalised or
controlled-rolled conditions, although low carbon steels might be
used in the cold-rolled and annealed condition. Production
processes aim to produce low inclusion contents and small grain
size to improve strength, ductility, toughness and reduce the
ductile/brittle transition.
The elastic modulus of steel is virtually independent of
composition and treatment.
The upper limits on the proportions of carbon and other alloying
elements are determined by the effect of carbon equivalent on
weldability, and by the effect of carbon on the ductile/brittle
transition temperature. All steels contain manganese, partly to
deal with impurities, such as sulphur, and partly because its
presence has a beneficial effect on the ductile/brittle transition
and strength.
In recent years the development of so-called micro-alloyed steels
or HSLA (high strength low alloy) steels has taken place. These
steels are normalised or controlled rolled carbon-manganese
steels which have been 'adjusted' by micro-alloying to give higher
strength and toughness, combined with ease of welding. Small
additions of aluminium, vanadium, niobium or other elements are
used to help control grain size. Sometimes, about 0,5%
molybdenum is added to refine the lamellar spacing in pearlite
and to distribute the pearlite more evenly as smaller colonies.
These steels are used where the improved properties justify the
ext

he Metallurgy Of Carbon Steel


The best way to understand the metallurgy of carbon steel is to study the Iron Carbon

Diagram. The diagram shown below is based on the transformation that occurs as a result
of slow heating. Slow cooling will reduce the transformation temperatures; for example:
the A1 point would be reduced from 723C to 690 C. However the fast heating and
cooling rates encountered in welding will have a significant influence on these
temperatures, making the accurate prediction of weld metallurgy using this diagram
difficult.

Austenite This phase is only possible in carbon steel at high temperature. It has
a Face Centre Cubic (F.C.C) atomic structure which can contain up to 2% carbon
in solution.

Ferrite This phase has a Body Centre Cubic structure (B.C.C) which can hold
very little carbon; typically 0.0001% at room temperature. It can exist as either:
alpha or delta ferrite.

Carbon A very small interstitial atom that tends to fit into clusters of iron
atoms. It strengthens steel and gives it the ability to harden by heat treatment. It
also causes major problems for welding , particularly if it exceeds 0.25% as it
creates a hard microstructure that is susceptible to hydrogen cracking. Carbon
forms compounds with other elements called carbides. Iron Carbide, Chrome

Carbide etc.

Cementite Unlike ferrite and austenite, cementite is a very hard intermetallic


compound consisting of 6.7% carbon and the remainder iron, its chemical symbol
is Fe3C. Cementite is very hard, but when mixed with soft ferrite layers its
average hardness is reduced considerably. Slow cooling gives course perlite; soft
easy to machine but poor toughness. Faster cooling gives very fine layers of
ferrite and cementite; harder and tougher

Pearlite A mixture of
alternate strips of ferrite and
cementite in a single
grain. The distance between
the plates and their thickness is
dependant on the cooling rate
of the material; fast cooling
creates thin plates that are
close together and slow
cooling creates a much coarser
structure possessing less
toughness. The name for this
structure is derived from its
mother of pearl appearance
under a microscope. A fully
pearlitic structure occurs at
0.8% Carbon. Further
increases in carbon will create
cementite at the grain
boundaries, which will start to
weaken the steel.

Cooling of a steel below 0.8% carbon When a steel solidifies it forms


austenite. When the temperature falls below the A3 point, grains of ferrite start to
form. As more grains of ferrite start to form the remaining austenite becomes
richer in carbon. At about 723C the remaining austenite, which now contains
0.8% carbon, changes to pearlite. The resulting structure is a mixture consisting of
white grains of ferrite mixed with darker grains of pearlite. Heating is basically
the same thing in reverse.

Martensite If steel is cooled rapidly from austenite, the F.C.C structure rapidly changes to
B.C.C leaving insufficient time for the carbon to form pearlite. This results in a distorted
structure that has the appearance of fine needles. There is no partial transformation
associated with martensite, it either forms or it doesnt. However, only the parts of a section
that cool fast enough will form martensite; in a thick section it will only form to a certain
depth, and if the shape is complex it may only form in small pockets. The hardness of
martensite is solely dependant on carbon content, it is normally very high, unless the carbon
content is exceptionally low.

Tempering The carbon trapped in the martensite transformation can be released by heating
the steel below the A1 transformation temperature. This release of carbon from nucleated
areas allows the structure to deform plastically and relive some of its internal stresses. This
reduces hardness and increases toughness, but it also tends to reduce tensile strength. The
degree of tempering is dependant on temperature and time; temperature having the greatest
influence.

Annealing This term is often used to define a heat treatment process that produces some
softening of the structure. True annealing involves heating the steel to austenite and holding
for some time to create a stable structure. The steel is then cooled very slowly to room
temperature. This produces a very soft structure, but also creates very large grains, which are
seldom desirable because of poor toughness.

Normalising Returns the structure back to normal. The steel is heated until it just starts to
form austenite; it is then cooled in air. This moderately rapid transformation creates
relatively fine grains with uniform pearlite.

Welding If the temperature profile for a typical weld is plotted against the carbon
equilibrium diagram, a wide variety of transformation and heat treatments will be observed.

Note, the carbon equilibrium diagram shown above is only for illustration, in reality it will be heavily
distorted because of the rapid heating and cooling rates involved in the welding process.

a) Mixture of ferrite and pearlite grains; temperature below A1, therefore microstructure
not significantly affected.
b) Pearlite transformed to Austenite, but not sufficient temperature available to exceed
the A3 line, therefore not all ferrite grains transform to Austenite. On cooling, only
the transformed grains will be normalised.
c) Temperature just exceeds A3 line, full Austenite transformation. On cooling all
grains will be normalised
d) Temperature significantly exceeds A3 line permitting grains to grow. On cooling,
ferrite will form at the grain boundaries, and a course pearlite will form inside the

grains. A course grain structure is more readily hardened than a finer one, therefore if
the cooling rate between 800C to 500C is rapid, a hard microstructure will be
formed. This is why a brittle fracture is most likely to propagate in this region.

Welds The metallurgy of a weld is very different


from the parent material. Welding filler metals
are designed to create strong and tough welds,
they contain fine oxide particles that permit the
nucleation of fine grains. When a weld solidifies,
its grains grow from the course HAZ grain
structure, further refinement takes place within
these course grains creating the typical acicular
ferrite formation shown opposite.

[ Home ] [ Table of Content ] [ Next ]


Pure iron is only slightly harder and stronger than copper. Its great ductility
and formability are conducive to making hand art, but its low strength is not
very practical for industrial engineering designs. With the addition of
carbon to iron, steel is created, providing the strength required for modern
engineering applications. Its mechanical properties rise to the occasion,
limited only by the designer's imagination.
However, it is the phenomenon of allotropy in iron that yields the almost
unlimited range of properties of steel. To our good fortune, allotropy in
iron is retained even in the presence of other alloying elements in steel,
allowing for many forms of heat treatable steel alloys to produce a variety of
properties for various applications.
Allotropy of Iron
Iron, as shown in figure 1, exists in three crystal (atomic) allotropes,
namely: alpha () iron, delta () iron, and gamma () iron. The -iron

form exists below 1625oF (885oC) while -iron is stable above 2540oF (
1395oC). Gamma iron exists at the temperatures between these two
ranges. It is the allotropy of iron that allows for these crystal structures to
change with temperature.
At room temperature, the -iron crystal structure has its atoms arranged in a
geometric pattern known as body-centered cubic or bcc (figure 2) . This
atomic arrangement of iron atoms is magnetic up to 1420oF (770oC), called
the curie temperature. This temperature was of practical importance to the
early blacksmiths who used an iron horseshoe magnet with a steel bar across
the two ends for temperature measurement. When the steel bar fell from
the magnet, the blacksmith knew the approximate temperature of the hearth
and was able to adjust the heat treat schedule accordingly. Above the curie
temperature is still bcc but is no longer magnetic. Slow heating of -iron to
1625oF (885oC) produces an allotropic change to gamma () iron, a facecentered cubic (fcc) crystal structure which is non-magnetic (figure 2). A

change from one crystal structure to another is called a transformation and


the temperature at which it occurs is called the transformation temperature.
When fcc - iron is slowly heated above 2540oF (1395oC) it transform back
to bcc iron. To distinguish the elevated temperature bcc iron from its lower
temperature counterpart, it is given its own name, delta () iron. This -iron
is non-magnetic and exists until the temperature is raised to 2800oF
(1540oC) which causes melting of the solid -iron to liquid iron. Since
atoms in the liquid iron have no distinct arrangement (each atom moving
freely within the liquid) there no longer exists a crystal structure above the
melting temperature.

For the allotropic transformations described, there is another driving force


equally important to the transformation temperature, namely, time. For
allotropic transformations to occur at the temperatures suggested in figure 1,
sufficient time is required for the atoms to reorganize themselves in the new
crystal structure. At the lower end of the temperature ranges for each

allotrope of iron, lower energy levels exist, so more time is required for
crystal structure transformation to occur. The interaction of time and
temperature to achieve the allotropic transformations shown in figure 1 is
called equilibrium.
Equilibrium allows metals to achieve their lowest energy state and to do so
requires a specific balance of time and temperature. When a metal is heated
or cooled very slowly, as in a controlled laboratory experiment, equilibrium
can be attained. Because equilibrium provides a metal its lowest energy
state, it is sometimes called the "Happy State", since the atoms are "happy"
at this energy level and require a change in energy to displace
them.
[ Go to Top ]
Note in figure 1 that the heating/cooling curve flattens at the allotropic
transformation temperatures. This pause in the heating/cooling cycle is
necessary for equilibrium allotropic change in the crystal structures to
occur. This form of graphical presentation was made popular by French
scientists, and consequently, the transformation temperatures are designated
by the letter A (from the French word arreter - meaning to stop), or r (from
refroidir - meaning to cool). For example, Ac3 is transformation
temperature of -iron to -iron upon heating and Ar4 is the transformation
temperature of -iron to -iron upon cooling.
The allotropic transformations illustrated in figure 1 are reversible, such taht
the transformations can occur upon slow heating or slow cooling. It is this
powerful flexibility of iron that provides the opportunity to heat treat steels
to many metallurgical conditions and associated mechanical and physical
properties.
It becomes apparent that a clear understanding of the behaviour of iron is
imperative in discussing steels and their many alloys. In this light, let's
continue this discussion by alloying the iron with carbon to make steel.
THE IRON-IRON CARBIDE SYSTEM - SLOW COOLING
Ferrite

Carbon is the most significant alloying element in steel. One of its most
pronounced effects is on the transformation temperatures as shown in figure

3. The addition of carbon to iron lowers the A3 temperature, while it raises


the A4 temperature and lowers the melting temperature. Expanding this
diagram to display the various allotropic crystal structure changes results in
the classic iron-iron carbide (Fe-Fe3C) phase diagram shown in figure
4. Although this diagram may seem quite involved at first glance, it is a
relatively simple but powerful tool in understanding the metallurgy of steels.
The area enclosed by QGPQ is a solid solution phase of carbon dissolved in
alpha iron known as alpha ( ferrite, more commonly called ferrite. Ferrite
has a body-centered (bcc) crystal structure that can only dissolve a
maximum of 0.025% C at 1340oF (725oC), with the solubility of carbon
dropping to 0.008% C at room temperature, i.e. almost pure iron.

The term ferrite was first used by the American metallurgist Professor
Henry M. Howe, and was almost certainly derived from the Latin
word ferrum, meaning iron. Since the ferrite phase at room temperature is
essentially pure iron, only containing 0.008% C, it has little commercial use
because of its extreme softness and low
strength.
[ Go to Top ]

Delta iron, with carbon contents of up to 0.1% C, exists at temperatures


above 2540oF (1395oC) and is called delta () ferrite. This area of the
diagram becomes of importance to welding when considering hot cracking
in carbon and alloy steels, since ferrite has relatively good solubility of
sulphur; where sulphur is the main cause of hot cracking. When the term
ferrite is used, it is understood that -ferrite is the subject material.

Likewise, when discussing the elevated temperature ferrite, one must use the
term delta ferrite.
As a rule of thumb, steels with < 0.25% C are called low carbon or mild
steels; steels with 0.25 - 0.50% C are called medium carbon steels; and
steels with > 0.50% C are called high carbon steels.
Austenite
The area in figure 4 enclosed by GJIEHG is a solid solution known
as austenite. Austenite is a non-magnetic, face-centered cubic (fcc) crystal
structure, that can dissolve carbon interstitially to a maximum of 2%
at 2100oF (1150oC) and is exhibited schematically in figure 5. Austenite
was first reported by Floris Osmond, a French steelworks engineer, and
named by him in honour of the eminent English metallurgist, Professor Sir
William C. Roberts-Austens.
When heat treat procedures involve heating steels in the region of Fe-Fe3C
phase diagram, the term used to describe the heat treatment
is austenizing. A steel is said to become austenitized when it has been
heated at a sufficient temperature, for the appropriate time, to achieve 100%
austenite through the thickness of the part.
Cementite
At 6.67% carbon and room temperature, ferrite is no longer stable. Instead,
the iron atoms combine with carbon atoms to form iron carbide (Fe3C),
calledcementite, existing within the boundary DOMD in figure 4. The
crystal structure of cementite is orthorhombic.

The
term
cemen
tite
was
first
applie
d by
Profes
sor
Howe
and
was
proba
bly
derive
d from
cemen
t
carbon, referring to carbon which was introduced into steel at that time by
the cementation process. Like all carbides, cementite is an extremely hard
constituent. When place in a soft matrix of ferrite, its distribution and size
produce the extraordinary range of mechanical properties that steel is noted
for.
Phases
Delta ferrite, austenite, ferrite and cementite are called phases since they are
physically homogeneous and distinct portions of the iro-iron carbide
system. With the ferrite phase occupying the left side and the iron carbide
phase the right side of figure 6, this diagram is given the name Iron-Iron
Carbide (Fe-Fe3C) phase diagram.
The areas between the single phase solid solutions of carbon in iron (i.e.
delta ferrite, austenite, ferrite and cementie) are mixtures of the two single
hpases. For example, with a carbon content of 0.4% C and temperature of
1400oF (760oC), simply draw a horizontal line starting at the intersection of
0.4% C and 1400oF (760oC) and extending in both directions until the
transformation temperatures of each end of the line are crossed. The
mixture of phases at this point will be the two phases at each end of the line,
i.e. ferrite and austenite.
[ Go to Top ]
Therefore, from figure 4: ferrite plus cementite exists within the boundary
QPNOQ; ferrite plus austenite exists within the boundary PGHP; delta
ferrite plus austenite exists within the boundary JKIJ; delta ferrite plus liquid

exists within the boundary KABK; austenite plus liquid exists within the
boundary EIBCE; austenite plus cementite exists within the boundary
HEMNH; and liquid plus cementite exists within the boundary CDMC.
Transformation Temperatures and Lines
The horizontal line PN extending along 1340oF (720oC) represents the lower
critical temperature, and is the first transformation line reached upon
heating steel from room temperature. It is designated as the A1 line.
The line GH defines the temperature at which complete transformation to
austenite is achieved upon heating steel with up to 0.8% C. In steel heat
treating terms, it is referred to as the upper critical temperature and is
designated as A3. The line HE represents the Acm temperature that borders
the lower limit of the austenitic region for steels with greater than 0.8%
C. It becomes the upper critical transformation temperature for these high
carbon steels.
The A4 transformation line (JI) outlines the temperature for the initial
transformation of austenite to delta ferrite. This temperature has little
significance in the industrial heat treatment of steels.
Although the A2 line is not a true phase transformation line, it does represent
the change from magnetic bcc ferrite to non-magnetic bcc ferrite at the
Curie temperature, 1420oF (770oC).
Eutectoid Steel
Point H represents a carbon content of 0.8% C and a temperature of 1340oF
(725oC) and is known as the eutectoid point. Thhis represents the
intersection of the two descending transformation lines, A3 and Acm, with the
horizontal transformation line A1. Steel with this composition (0.8% C) is
known as eutectoid steels.
Steels having a carbon content less than 0.8% C are called hypoeutectoid
steels and those with more than 0.8% C are called hypereutectoid steels. (A
simple reminder to keep track of these two terms is to remember
that hyper rhymes with the word higher and thus hypereutectoid steel has the
higher carbon content, i.e. > 0.8% C. By elimination, the other term,
hypoeutectoid steel, has less than 0.8% C.)
Pearlite
When a eutectoid steel (0.8% C) is cooled slowly from an austenitizing
temperature, say 1500oF (815oC), according to the Fe-Fe3C phase diagram,

no transformation will occur until the temperature reaches the eutectoid


temperature 1340oF (725oC). Upon further slow cooling below

this temperature, austenite will transform to ferrite and cementite.


However, the transformation is unique since the carbon previously dissolved
in the austenite cannot be retained by the newly formed ferrite, due to the
low solubility of carbon in ferrite. Consequently, carbon is rejected by the
new ferrite and accumulates as cementite laths (or layars) adjacent to ferrite
layers as schematically represented in figure 6. The microstructure of
alternating laths of ferrite and cementite is called Pearlite. Eutectoid steels
(0.8% C), when slow cooled after austenitizing, will form 100% pearlite
(figure 17c).
Pearlite was first observed by the 19th century English geologist, Dr. Henry
C. Sorby, and was named pearlyte and later pearlite by Professor Howe. Its
name is said to be derived from the shiny microscopic appearance
resembling that of the mother-ofpearl.
[ Go to Top ]
The width of the alternating laths of ferrite and cementite govern the
mechanical properties of this microstructure. When pearlite is formed under
very slow cooling, the pearlite laths are wider than if cooled under relatively
faster rates. Pearlite containing wider laths is known as coarse pearlite and

is softer and weaker microstructure than pearlite with narrower laths, called
fine pearlite. It is important to remember that pearlite is not a phase of
steel, but rather a microstructure made up of two phases, namely ferrite and
cementite.
Austenite Decomposition
In terms of understanding the heat treatment of steels, the decomposition of
austenite is paramount. Consider austenite in a hypoeutectoid steel of 0.4%
C at 1550oF (843oC) and slow cooling (say 100oF/hr) to room
temperature. The following observations can be made:
1. Above the A3 line, austenite is stable and can easily dissolve the 0.4% C
into its fcc solid solution. Be aware that the higher the austenitizing
temperature reached above the A3 line and/or the longer the time at the
austenitizing temperature, the larger the austenite grain size will become.
This is called grain growth.
2. Upon cooling the fcc austenite from 1550oF (843oC), it begins to
transform to bcc ferrite at the A3 temperature, approximately 1475oF
(802oC). This phase transformation of austenite to ferrite continues as we
cool within PGHP (figure 4) region. Note that as the temperature is
decreased within this region, more ferrite is formed at the expense of losing
austenite. Since ferrite can dissolve no more than a maximum of 0.025% C,
the carbon content of the remaining (untransformed) austenite is increased
as proeutectoid (new) ferrite is formed. This continues until just above
the A1 line where the remaining austenite will contain essentially 0.8% C.
3. At the A1 line, 1340oF (725oC), the remaining austenite begins its
transformation to pearlite. As the A1 line is crossed, the remaining austenite
transform to pearlite and the resultant microstructure is a mixture of ferrite
and pearlite.
4. Cooling from just below the A1 line, where ferrite and pearlite are now
present, produces no further phase changes. The room temperature
microstructure will remain ferrite and pearlite.
Ferrite grain size and pearlite volume fraction are a key factor in
determining low temperature impact toughness. The smaller the final grain
size and the lower the pearlite amount, the higher the low temperature
impact toughness will become.
Consider austenitizing a hypereutectoid steel of 1.0% C at 1550oF (843oC)
followed by slow cooling (say 100oF/hr). The following observations can
be made:

1. Above the Acm line, austenite is stable and can easily dissolve the 1.0%
C. Again, the higher the temperature reached above the Acm line and/or the
longer the time at that temperature, the larger the austenite grain will
become. Also remember that to achieve an austenitizing condition,
sufficient time at the austenitizing temperature is required to produce 100%
austenite through the thickness of the steel part.
2. As the Acm temperature, about 1450oF (787oC), is met upon cooling,
austenite begins to give up (called precipitation) some of its carbon, thus
forming the new phase, cementite (Fe3C). The amount of austenite
decreases as new cementite is formed, with decreasing temperature
approaching the A1 line.
3. At the A1 line, 1340oF (72oC), sufficient carbon has been precipitated
from the austenite solid solution that it now retains only 0.8% C. This is
the eutectoid composition, and hence, the remaining austenite transforms to
pearlite upon further cooling.
4. Once the A1 line has been crossed, the resultant microstructure consists
of cementite and pearlite. Cementite is peresent within the pearlite or as a
network around the pearlite grains (see figure 17c and e). There are no
further phase changes as the steel cools to room temperature.
Eutectic and Eutectoid Reactions
It is important to distinguish between the eutectoid and eutectic reactions in
the iron - carbon system. The eutectoid reaction at 0.8 and 1340oF (725oC)
involves one solid solution phase (austenite) transforming on cooling to a
mixture of two solid solution phases (ferrite + cementite). By comparison,
the eutectic reaction at 4.3% C and 2100oF (1150oC) involves one liquid
phase transforming on cooling into a mixture of two solid solution phases
(ledeburite and cementite).
Fe-Fe3C Phase Diagram
Reactions
[ Go to Top ]
Industrial fabrication conditions restrict the application of the iron-iron
carbide phase diagram, since:
1. Commercial additions of other elements (Mn, Si, Cr, Ni, Mo, etc.)
shiftnthe position of the transformation lines, i.e. changing the
transformation temperatures, with the extent of the change depending on the
element and the amount added.

2. Faster rates of heating annd cooling, such as in welding and quenching,


greatly exceed the equilibrium rates (i.e. slow cooling and cooling), so that
the transformation reactions are shifted, delayed, or simply do not have
sufficient time to occur.
However, the diagram can be used in many industrial heat treatment
application of plain carbon steels and as a rough guide for alloy steels and
when considering welding or any other thermal process.
THE IRON-IRON CARBIDE SYSTEM - FAST COOLING
Bainite
If austenite is allowed to cool faster than the rates required to produce a
ferrite-pearlite structure, then at temperstures below about 1025oF (550oC)
another constituent, bainite starts to separate along with pearlite. At these
faster cooling rates, the potential for austenite to transform to ferrite and
pearlite is suppressed by the inability of the carbon atoms to move fast
enough to their equilibrium positions. The main reason for this occurrence
is the lack of heat-energy retained in the material with the faster cooling
rates; remembering that sufficient time and tempersature (energy) is
required for carbon atom diffusion to produce the ferrite-pearlite
transformations from austenite.
In 1934 the term bainite was initiated to honour Edgar C. Bain by his
colleagues at the Kearney Laboratory - Jose Vilella, John Zimmerman, E.S.
Davenport, E.L. Roff and Robert Aborn, In fact, Bain and associates were
not the first to produce the bainite microstructure, since Portevin had done
so in 1911, but at that time it was impossible to interpret the phase with the
existing technology. Bainite was formerly referred to by the now obsolete
terms, sorbite and troostite.
Depending on the temperature of formation, bainite varies from a fine
mixture of ferrite and cementite to lens-shaped needles of ferrite and no
visible cementite. The temperature range in which a eutectoid steel (0.8%
C) forms bainite is approximately 975-530oF (525 and 275oC). Since
bainite shows a substantial variation in microstructure from the highest to
the lowest temperatures of formation, the terms upper and lower bainite are
used to more accurately described the microstructure.

Upper bainite is rather featherly-appearing microstructure, while lower


bainite is much more acicular (figure 7), resembling its close cousin,
tempered martensite. Since bainite structures are composed of iron carbide
and ferrite, often supersaturated with carbon, the distinction between upper
and lower bainite is significant considering there can be major differences in
mechanical properties. For the most part, bainite is harder, stronger and
tougher at low temperatures than ferrite-pearlite or stright pearlitic
microstructures, in steels of equivalent carbon contents. This
microstructural interpretation becomes important when attempting to resolve
failure mechanisms involving these steels in H2S gas (sour)
environments. Unfortunately however, it can be extremely difficult to
distinguish a steel microstructure as upper or lower bainite, and even at
times with martensite.
Bainite is not referenced in the Fe-Fe3C phase disgram since its production
involves faster cooling rates than those allowed for in this phase
diagram. To predict the formation of bainite upon cooling from austenite,
other diagrams must be used, specifically, isothermal transformation (ITT)
diagrams, sometimes called time-temperature transformation (TTT)
diagrams. These diagrams involve isothermal cooling, meaning cooling at
ac

onstant (iso) temperature (thermal).


Bain and his associates created many ITT diagrams for steel, though they
have limited direct use in industrial applications since isothermal cooling
conditions are rarely used outside the laboratory. However, modified ITT
diagrams to accommodate continuous cooling conditions are useful for
commercial practice. The most functional diagrams of this type are the
modified continuous cooling transformation (CCT) diagram for engineering
steels,of which a popular series was produced under the direction of M.
Atkins of the British Steel Corporation.
[ Go to Top ]
Figure 8 shows the modified CCT diagram for a low carbon (0.18% C)
steel. The diagram is read by drawing a vertical line from the section
thickness (bar diameter) and cooling medium of interest, upwards to the top
of the diagram. Following this line downwards from the A (austenite)
region results in the room temperature microstructure produced upon
continuous cooling within the selected medium. This information provides
very useful data since microstructure prediction for industrial cooling is now
possible and thus, properly prediction for the steel.

Martensite
If austenite is very rapidly cooled, diffusion ocntrolled transformation to
ferrite, pearlite and even bainite may not be possible. Instead, the austenite
changes its crystal structure by a diffusionless shearing mechanism that

moves blocks of a
toms. The carbon originally dissolved in the solid solution of austenite,is
now trapped in a ferrite structure.
Since ferrite has an extremely low solubility of carbon, its crystal structure
becomes distorted to accommodate the presence of the trapped carbon,
resulting in a volume expansion. This new microstructure is
called martensite, named by Osmond in a tribute to Professor Adolf
Martens, a German railway engineer who in 1878 started a center for
metallographic research.
Martensite is no longer a true body-centered cubic phase, but rather a bodycentered tetragonal (bct) structure (figure 9). The extreme distortion
imposed by the carbon atoms is said to account for the substantially higher
hardness and strength of this microstructure. The temperature at which
austenite starts to transform to martensite is termed the Ms temperature a

nd the temperature at which it is finished is called the Mf temperature. The


maximum rate of cooling required to produce 100% martensite is called
the critical cooling rate.
The atomic proof of carbon's effect on distorting, and thereby hardening the
bct structure, is exhibited in figure 10, where increasing carbon content also
increases the height or C dimension of the bct structure.
One would expect that steels of higher carbon content, being more distorted,
would produce martensite of greater hardness, and this is in fact so, as figure
11 demonstrates. Consequently, not all martensitic structures are created
equal, with their hardness, tensile strength, wear resistance and other
mechanical properties controlled by the steel's carbon
content.
[ Go to Top ]
Martensite is the product of cooling austenite at a rate equal to or faster than
the critical coo

ling rate (figure 12). In order to produce martensite, one has to initially
start with austenite, making austenite the mother of martensite. Figure 13
shows that martensite formation often initiates at the prior austenitic grain
boundaries.
Martensite starts to form on rapid cooling at the Ms temperature. The
Mstemperature decreases sharply with increasing carbon content in
steels. All other alloying elements, such as Mn, Ni, Cr, Mo, lower the Ms,
except for Co which raises the Ms. A significant effect of low
Ms temperature is incomplete austenite to martensite transformation at room
temperature. Therefore, as-quenched martensitic structure may retain
austenite as partnof its room temperature microstructure. If left

untransformed, the retain austenite at room temperature becomes an


accident waiting to happen.
Although martensite can be a very hard, wear resistant, strong material, it
lacks ductility, toughness and in all but low-carbon steels it is extremely
brittle. Consequently, martensite must be heat treated to enable parts to be
used for industrial purposes. Heat treatment reduces the internal strain in
the bct structure, thereby increasing ductility and toughness, at some
expense to hardness, wear resistance and strength.
Tempered Martensite
A steel through-hardened to a martensitic structure is not a satisfactory
engineering material for most applications. Despite its potential strength, it
lacks ductility and toughness, often to the point where its full strength
cannot even be measured since failure is so easily initiated. In order to
develop ductility and toughness, the quenched steel is further treated by
tempering

Martensite is not a stable constituent, and on heating it will decompose to


its stable products, ferrite and cementite. The extent of this decomposition
will depend upon tempering temperature and time at temperature. At high
tempering temperature and/or long periods of time, decomposition of
martensite can be so complete that it approaches the mechanical properties
of ferrite (soft, ductile,low strength and hardness). At low tempering
temperature and/or short tempering times, decomposition is minimal and the
nartensite remains hard and strong with slight increases in ductility and
toughness. Thus, the appropriate choice of tempering temperature and time
at temperature is required to achieve the specified mechanical properties
necessary for the intended application.
In tempering fully quenched (martensitic) steels, it should be cautioned that
a loss in ductility may result from prolonged heating between 500 and 650 oF
(260 and 340 oC). Between these temperatures, the notch ductility of the
steel (assessed by impact tests) is reduced. This phenomenon is calledtemper
embrittlement or blue
brittleness.
[ Go to
Top ]
The effect of all alloying elements is to reduce the rate at which martensite
will temper. Thus, at a given tempering temperature, and for a given time,
the alloy steel will show a greater hardness than the unalloyed steel. The
design of steels and cooling conditions to produce required amounts of
martensite are the subject of technology referred to as hardenability.
Hardness vs. Hardenability

The measure of a steel's ability to harden to depth is its hardenability. Steels


with high hardenability are those that require slower cooling rates for
martensite formation. However, it is the carbon content of a steel that
determines the maximum hardness feasible. The effect of carbon on
hardness is demonstrated in figure 14

An important factor influencing the maximum hardness that can be achieved


is mass of the metal. In a small section, the heat is removed quickly, thus
exceeding the critical cooling rate of the steel. As section size increases, it
becomes increasingly difficult to remove the heat fast enough to exceed the
critical cooling rate and thus avoid formation of nonmartensitic products.
An example of the mass effect is shown in figure 15, which illustrates the
effect of section size on surface hardness. For small sections up to 0.5
inches (13 mm), full hardness of about 63 HRC is achieved. As the
diameter of the quenched piece is increased, cooling rates and hardness
decrease because the critical cooling rate for this specific steel was not
exceeded. Thus, figure 15 also serves as an example of a low-hardenability
steel. Plain carbon steels are characterized by their low hardenability, with
critical cooling rates exceeded only in thin sections. Hardenability of all
steels is directly related to critical cooling rates. The lower the critical
cooling rate, the higher the hardenability for a given steel, almost regardless
of carbon contant.

Alloying steel with elements such as nickel, chromium, and molybdenum


can also be used to make it more difficult for the diffusion controlled
transformation of austenite to occur. As a result, martensite can be formed
with less drastic cooling, such as oil quenching. Still greater alloying can
yield "air hardenable" alloys. Slower cooling rates to produce martensite
are beneficial since fast cooling introduces high surface residual stresses
which may cause quench cracking. Quench cracks arise when a steel is
quenched and undergoes stresses resulting both from thermal contraction
and from a volume expansion (2 to 4%) which accompanies the
transformation of austenite to martensite.
Although alloying elements can increase a steel's hardenability, they do not
increase the steel's maximum hardness possible. Hardness is determined
principally by the amount of carbon.
The factors which increase hardenability work not only to produce
martensite but also to form other microstructures. Thus, hardness graduants
in bars of various diameters, cooled at various rates, can be
estimated. Continuous cooling transformation diagrams, such as in figure
8, demonstrate the various cooling conditions and related microstructures.
Grain Boundary
Metals generally consist of regions called crystals or grains where the atoms
are arranged in regular geometric patterns such as bcc or fcc. Although the
geometric pattern of atoms is fixed for grains of a particular material, the
grains are oriented randomly with respect to the x, y, and z directions. As a

result there is a disarray of atoms where the grains meet each other, called
grain boundaries. This disarray of atoms along grain boundaries can be
exposed by etching techniques that allow grains of the metal to be examined
and measured.
Metallographic
Examination
[ Go to Top ]
Etching techniques are used on polished surfaces to reveal the metal grains
and the various phases of the metal. Microscopic observation of this type is
called metallographic examination and the metal images observed are
called microstructures. Metal samples must be specificaaly

prepared for the purpose and the science of sample preparation, examination
and photography of the microstructures is called metallography.

To examine the microstructure of a metal with an optical microscope, the


area to be examined is first polished. Polishing leaves a mirror-like metal
surface which is smooth and highly reflecting, but covered with thin film of
metal which is plastically deformed by the abrasive action of the final
polishing operation (figure 16). To reveal the true metal structure, the
deformed surface layer must be removed. The various structural
components of the underlying metal can then be revealed. This is done by
etching. Various etchants are used to best reveal the metal structure, but in
general the etchants dissolve the distorted surface layer and then attack and
dissolve the underlying metal. Metallographic etchants are very
selective. Crystals of varying orientation are attacked more rapidly than the
body of grains, and various structural components are attacked at different
rates. Thus, by developing hills and valleys, plateaus of varying levels, etch
pits of varying orientation, and similar differentiating effects, the structure
of metal can be revealed.
In an optical microscope where light is passed through the microscope tube
and reflected from the specimen to the observer's eye, the specimen appears
bright. Where the intensity of reflected light is decreased by scattering
from a roughened surface, the specimen appears less bright, and where the
light is reflected so that none passes back through the microscope tube, the
specimen appears dark (figure 16). Examples of ferrous microstructures are
shown in figure 17.
The grain growth characteristics of hypoeutectoid steels taht have been
deoxidized with silicon are said to be normal in that the austenitic grain size
increases continuously and progressively as the austenitizing temperature is
raised above the A3 temperature. Austenitic grain growth is also time
dependent, the grains continuing to grow at any one temperature.
The austenitic grain size of annealed or normalized medium carbon steels
can readily be observed because proeutectoid ferrite precipitates along the
austenitic grain boundaries during slow cooling. Thus bands of ferrite
outline pre-existing austenite grain boundaries
It is not so easy, however, to recognize the sites of the austenitic grain
boundaries in low carbon steels when a large volume fraction of ferrite is
present. Similarly, for quenched and tempered steels (martensitic), special
etching techniques are required to reveal the prior austenitic grain size. A
suggested etchant to reveal prior austenitic grains in steels fully hardened to
martensite is 1 g of acid, 5 ml. of HCl and 95 ml. of ethyl alcohol (see
ASTM Standard E112 appendix 3 for more details).
Grain coarsening of austenite is reversible. Several new grains of austenite
can be nucleated in the volume that had been occupied by one former

austenite grain, and that the size to which these new grains grows depends
primarily on the new austenitizing temperature. Thus the new austenite
grain size will generally be smaller than the former grain size if the new
austenitizing temperature is lower than the previous
one.
[Go to Top ]

However, a small austenitic grain size is usually not always recovered in a


single reaustenitizing heat treatment, depending on the initial size of the
grains. For larger grain sizes, several reaustenitizing heat treatments may
be required to obtain uniform and small final grain size. Keeping in mind
that the lower the austenitizing temperature the greater the grain refinement.
Some steels are treated during the steelmaking process with grain
refinement alloying elements, such as Al, Nb, V, Ti and Zr, which inhibits
austenitic grain growth. The austenitic grain size after heating at normal
austenitizing temperatures is then much smaller than for normal steels. The
product then is commonly called fine-grained steel.
Reducing the ferrite grain size by this or other methods results in increased
yield strength, which varies approximately with the reciprocal of the square
root of the ferrite grain diameter (d-1/2). Reducing the ferrite grain size also
increases the toughness, which is the one factor that improves both the yield
strength and toughness simultaneously. For example, many proprieary line
pipe steel specifications contain requirement on ferrite grain size to
minimize the risk of brittle fracture.
Grain Size
Grain size is commonly measured according to ASTM Standard Method
E112. Determining The Average Grain Size. This standard lists three
methods for determining granin size, namely the comparison Procedure,
Planimetric (Jeffrys') Procedure, and Intercept Procedure. Because of their
purely geometric basis they are quite independent of the metal concerned
and may also be used for the measurement of grain, crystal,or cell size of
nonmetallic materials.
In materials having two or more constituents, the grain size usually refers to
that of the matrix, except that in those materials where a second phase of
sufficient amount, size or continuity to be significant, the grain size may be
reported separately. Minor constituent phases, inclusions, and additives are
not normally considered.
It is important in using these methods to recognize that the measurement of
grain size is not precise, but an estimate. A metal grain is a threedimensional shape of varying sizes. The grain cross section produced by
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.
The comparison procedure is very 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 18. The ASTM Grain
Size Number corresponds to a certain number of grain/in2 according to
Table 1.
Table 1
ASTM No.
0
1
Grains/in2
0.1
1

10

16

32

64

128

256

512

The relationship between the Grain Size Number and the number
of Gr
ains/i
n2 is
given
by the
expre
ssion:

n=

(N-1)

wh
ere N
=
AST
M
Grain
Size
Num
ber
n=
numb
er of
Grain
s/in2 a
t the
specif
ied
magn
ificati

on.
In the planimetric (Jeffry's) procedure a known area is inscribed in the
observed field and the grains within this area are counted and multiplied by
the Jefferies' multiplier. The product will be the number of grains per
square millimeter.
The intercept method has two procedures: the lineal (Heyn) procedure and
circular procedure. Both methods involve placing a grid pattern on the field
of observation and counting the number of grains at each intercept within a
selected area.
Effect of Alloying Elements in Steel

[ Go to Top ]

With alloying elements, it is important to determine whether they are


carbide, austenite or ferrite formers and the purpose for being added to the
steel. Each individual element transfers specific properties to the steel,
according to the amount added. The presence of several elements can
enhance one another, resulting in a synergistic effect. However, there are
alloying elements that do not influence a particular property in the same
direction as others and may in fact counteract one another. Alloying
elements in steel only provides the potential for specific properties. These
properties may not actually be achieved until processing and heat treatment
have been carried out. The principal effects of the alloying elements on
steel are as follows.
Carbon (C) Melting point 6404oF (3540oC)
C is the most important and influential alloying element in steel. In
addition to carbon, however, any unalloyed steel will contain
silicon,manganese, phosphorus and sulphur, which occur unintentionally
during manufacture. The addition of further alloying elements to achieve
special effects and intentional increase in the manganese and silicon
contents result in alloy steel. With increasing C content, the strength and
hardenability of the steel increase, but its ductility, forgeability, weldability
and machinability (using cutting machine tools) are reduced. Corrosion
resistance to water, acids,and hot gases is practically unaffected by the
carbon.
Manganese (Mn) Melting point 2230oF (1221oC)
Mn is normally present in all commercial steels. It is essential to steel
production, not only in melting but in rolling and other processing
operations.

Mn deoxidizes steel. It compounds with sulphur to form Mn Sulphide


(MnS), thus reducing the undesirable effect of the iron sulphide (FeS). This
is of particular importance in free-cutting steel since it reduces the risk of
hot shortness.
Mn reduces the critical cooling rate, thus increasing hardenability. Yield
point and strength are increased by addition of Mn and, in addition,
increases hardness penetration depth. Steels with Mn contents > 12% are
austenitic because Mn is an austenite former and stabilizer. An example of
12% Mn steel is the Hadfield manganese steels that can achieve high
degrees of work hardening, where the surface is subjected to impact stress
while the core remains tough. For this reason, these high Mn steels are
used in the mining industry in jaw crushers.
Steels with Mn contents of 18% or less remain unmagnetizable even after
pronounced cold forming, as well as remaining tough at subzero
temperatures. The coefficient of thermal expansion increases as a result of
Mn, while thermal and electrical conductivity are reduced.
Silicon (Si) Melting point 2577oF (1414oC)
Si is contained in all steel in the same way as manganese, as iron ores
incorporate a quantity of it according to their composition. Si is not a
carbide former but enters into solution in the ferrite. Si is not a metal but a
metalloid, as are phosphorus and sulphur.
One of the most important applications of silicon is its use as a deoxidizer in
molten steel. Silicon is usually present in fully deoxidized alloy steels in
the amounts up to 0.35%, insuring the production of sound, dense ingots. It
promotes graphite precipitation and restricts the gamma phase significantly,
increases strength and wear resistance (Si-Mn heat treatable steels), and
significantly increases the elastic limit, thus being a useful alloying element
in spring steels.
Aluminum (Al) Melting point 1216oF (658oC)
Al is used for deoxidation and for control of inherent grain size. When
added to steel in small amounts, it produces a fine austenitic grain size. In
fact, of all the alloy elements, aluminum in prescribed amounts, is the most
effective in controlling grain growth. Titanium, zirconium, and vadanium
are also effective grain qrowth inhibitors, but have adverse effects on
hardenability because their carbide compounds are very stable and difficult
to dissolve in austenite prior to quenching. Al does not form a acrbide.
Al is used as alloying addition in amounts of 0.95% to 1.30% in the most
popular nitriding steel. The extremely high hardness of the nitrided case is

due to the formation of a hard, stable aluminum nitride compound. The


amount of Al present in nitriding steels is considerably in excess of the
amount necessary to produce a fine austenitic grain size in other steels.
Phosphorus (P) Melting point 111oF (44oC)
P is usually regarded as a tramp element in steel since it is present in iron
orc. P produces primary segregation on solidification of the steel melt and
the possibility of secondary segregation in solid state due to the noticeable
restriction of the gamma phase. It is difficult to achieve homogeneous
distribution of P in steel, such that P contents are usually limited to 0.030.05%.
[ Go to Top ]
Sulphur (S) Melting point 244 F (118oC)
S produces the most pronounced segregation of all steel accompanying
elements. Iron sulphide (FeS) leads to hot shortness, as the low melting
point sulphide eutectics surround the grains, so that only slight cohesion
between grains occur and during hot forming the grain boundaries tend to
fracture. FeS also become susceptible to hydrogen-induced cracking in
many environments, most notably where H2S gas is present, such that
0.001% S pipeline steels are common. As sulphur possesses a considerable
affinity for manganese, it is combined in the form of Mn Sulphide (MnS) as
this is the least dangerous of all inclusions, as it is distributed in point form
in the steel. S significantly reduces toughness. It is added intentionally to
steels for automatic machining up to 0.4%, as the friction on the tool cutting
edge, reduced by sulphur's lubricating action, permits increased tool life. In
addition, short chips occur when free-machining steels are machined. S
decreases weldability by promoting hot cracking.
o

Chromium (Cr) Melting point 3488oF (1920oC)


Cr is a strong carbide and ferrite former that among several advantages,
increases the edge-holding quality and wear resistance of steel cutting
tools. Cr reduces the critical rate of cooling necessary for martensite
formation, thereby increasing hardenability and allowing these steels to
become oil and air-hardened. Notch toughness is reduced, but ductility
suffers only slightly. Weldability decreases in pure chromium steels with
increasing Cr content. The tensile strength of the steel increases by 11.514.5 ksi (80-100 N/mm2) per 1% Cr addition. While increasing Cr contents
improve oxidation resistance, particularly at higher temperatures, a
minimum content of about 12% chromium is necessary for corrosion
resistance of steels, i.e. stainless steels.
Cr raises the A1 and A3 critical points, especially when large amounts of
chromium are present. The eutectoid carbon content is found to be lowered

by chromium additions, by an amount varying with quantity present. At


2.0% chromium, the eutectoid forms with 0.62% carbon. With 12.0%
chromium, eutectoid carbon drops to under 0.40%.
Nickel (Ni) Melting point 2647oF (1453oC)
Nickel as an alloying element in alloy steels is an austenite former and is
soluble in all proportions in both gamma and alpha iron. It is not a carbide
former. In combination with chromium, nickel produces alloy steels with
greater hardenability, higher impact strength, and fatigue resistance than
possible with carbon steels. Ni produces a significant increase in notch
toughness, even in the low temperature range, and is therefore alloyed for
increasing toughness in case-hardening, heat-treatable and low temperature
toughness steels.
Ni depresses the Ac and Ar critical points. It lowers the carbon content of
the eutectoid which, with a 3.5% nickel steel, is reduced to 0.70% carbon.
As a result of increasing the gamma loop, Ni in content of > 7% imparts
austenitic structure to stainless steels, down to well below room
temperature. Ni on its own only makes the steel rust resistant, even in high
percentages, but in austenitic Cr-Ni stainless steels (AISI 300 series), results
in resistance to the effect of reducing chemicals. Resistance of these steels
in oxidizing substances is achieved by means of Cr. At temperatures above
1100oF (593oC), austenitic steels have greater high temperature strength, as
their crystallization temperature is high.
Molybdenum (Mo) Melting point 4752oF (2622oC)
Mo in steel can form a solid solution with the ferrite phase and also,
depending on the Mo and carbon content, can form a complex carbide. Mo
is ususlly alloyed together with other elements, such as Cr and Ni. Mo raises
the Ac3 critical point when added in the usual amounts (0.10 to 0.60%) for
alloy steels. When Mo is in solid solution in austenite prior to quenching,
the reaction rates for the transformation of austenite become considerably
slower as compared with a carbon steel, resulting in deeper hardening
steel. A strong carbide former, the cutting properties with high speed steel
are improved by Mo.
Mo steels in the quenched condition require a higher tempering temperature
to attain the same degree of softness as comparable carbon or alloy
steels. This resistance to tempering contributes to the ability of these steels
to retain their strength at elevated temperatures. They show, because of this
effect, a considerable resistance to "creep" under sustained loads below their
elastic limit at temperatures up to 1100oF (593oC).

Mo promotes grain refinement and increases yield strength. It belongs to the


elements which increase corrosion pitting resistance and is therefore used
frequently with high alloy Cr steels and with austenitic CrNi steels. High
Mo contents reduce susceptability to pitting corrosion,as in type 317
stainless steel containing 3.0-4.0% Mo.
Vanadium (V) Melting point 3139oF (1726oC)
V is a strong carbide former and promotes grain refinement. The complex
carbides formed by V additions are quite stable, thus providing increase in
wear resistance, edge holding quality and high temperature
strength. Similarly, V offers significant improvement in retention of temper
and reduction of overheating sensitivity are achieved with its addition. It is
used primarily as additional alloying element in high speed, hot forming and
creep resistant steels.
[ Go to Top ]
It dissolves to some degree in ferrite, imparting strength and toughness. As
with other strong carbide formers, vanadium raises the critical points and
decreases the carbon content of the eutectoid. V restricts the gamma phase
and shifts the Curie point at elevated temperatures.
Titanium (Ti) Melting point 3141oF (1727oC)
On account of its very strong affinity for oxygen, nitrogen, sulphur and
carbon, Ti has a promounced deoxidizing, dentriding, sulphur bonding and
notable carbide forming action. Its carbide-forming tendency is so strong
that a 0.50% carbon steel will have practically no tendency to quench harden
when 1.5 to 2.0% Ti is addeed.
Used widely in stainless steels as carbide former for stabilization against
intercrystalline corrosion, Ti also possesses grain refining properties. Ti is
a strong ferrite former and stabilizer, thereby restricting the gamma
phase. In high concentration, it leads to precipitation processes and is
added to permanent magnet alloys on account of achieving high coercive
force. Ti increases creep rupture strength through formation of nitrides.
Niobium/Columbium (Nb/Cb) Melting point 3542oF (1950oC) Tantalum
(Ta) Melting point 5486oF (3030oC)
These elements occur almost exclusively together and are very difficult to
separate to separate from one another, so that they are usually used
together. They are very strong carbide formers, thus they are alloyed
particularly as stabilizers of stainless steels. Both elements are ferrite
formers and thus reduce the gamma phase. The A3 temperature is raised
and the A4, or upper austenite limit, is lowered. Due to the increase in high
temperature strength and creep rupture strength of Nb (Cb), it is frequently

alloyed to high-temperature austenitic boiler steels. Ta has a neutron high


absorption cross-section; only Ta/Nb (Cb) is considered for use in reactor
steels.
One of the advantages of using Nb (Cb) for grain refinement is its low
deoxidizing power does not introduce undesirable oxide inclusions into the
steel. This fine grain size and the decreased hardenability attributed to
columbium increases ductility of steels marginally and toughness
significantly.
Boron (B) Melting point 4172oF (2300oC)
B is usually added to steel to improve hardenability, that is, to increase the
depth of hardening during quenching and thus causes an increase in core
strength in case-hardening steels. B-treated steels will usually have a B
content in the range of 0.0005 to 0.003%.
Because B possesses a high cross section for neutron absorption, it is used to
alloy steels for controllers and shields of atomic energy plants. A reduction
in weldability must be expected in B alloyed steels.
Calcium (Ca) Melting point 1562oF (850oC)
Ca is used together with Si for deoxidation. It increases scaling resistance
of heating conductor materials. It is also added to pipeline steels for use in
sour (H2S) gas service to shape control (spherodize) nonmetallic inclusions,
such as MnS.
Nitrogen (N) Melting point -346oF (-210oC)
As an alloying element, N extends the gamma phase and stabilizes the
austenitic structure. In austenitic steels, such as the AISI 300 series of
stainless steels, N increases strength and above all the yield point plus
mechanical properties at elevated temperatures. As a result of nitride
formation, N permits high surface hardness to be achieved during nitriding.
Selenium (Se) Melting point 423oF (217oC)
Se is used in free-machining steels to improve machinability. much like
sulphur.
Steel Heat Treating Terms
Crystal allotropes of iron, phases and microconstituents have been
previously discussed and ithas been shown that steel can be thermally heated
to a broad range of properties. Several heat treating terms are used to
describe the thermal conditions under which different microstructures,
phases, and crystals may be present. These are described as follows:

Annealing
The annealing process is intended to optimize the steel's machinability and
formability. In manufacturing steel products, machining and forming are
often employed. Quenched and tempered steel may not machine or bend
very easily and annealing is often necessary to manufacture steel
components economically. Annealing is used after cold forming
operations, since during forming, the deformed areas of the steel may
become work-hardened and susceptible to fracture.
Full
Annealing
[ Go to Top ]
Steel is heated 50 to 100oF (10 to 38oC) above the A3 for hypoeutectoid
steels, and above the Acm for hypereutectoid steels, with slow controlled
cooling, resulting in soft andductile microstructures that have commercially
maximized machinability and formability. In full annealing, cooling must
take place very slowly so that a coarse pearlite is formed. When the term
annealing is applied to steels it is assumed that full annealing was
performed.
Normalizing
The process of normalizing consists of heating to a temperature 50 to 100oF
(10 to 38oC) above the A3 and allowing the part to cool in still room
temperature air. The actual temperature required for this depends on the
composition of the steel, but is usually around 1550 to 1650oF (840 to
900oC) for most low and medium carbon steels. Normalizing can be
described as a homogenizing or grain-refining treatment. Within any piece
of steel, the composition is usually not uniform throughout. That is, one
area may have more carbon than the area adjacent to it. These
compositional differences affect the way in which the steel will respond to
heat treatment. If the steel is austenitized, the carbon can readily diffuse
throughout, and the result is a reasonably uniform composition from one
area to the next. The steel is then more homogeneous and will respond to
the heat treatment is a more uniform way.
The grain-refining effects of normalizing become of prime importance in
designing steel microstructures for low temperature service. By refining
the grain size, more grain boundaries are formed and the energy necessary to
impact fracture is increased in order to "push" the crack across the grain
boundaries. For example, ASTM A 516 grade 70 pressure vessel plate steel

is commonly specified to have a notch toughness (Charpy) of 20 ft-lb (27 J)


minimum at -50oF (-46oC). To achieve this notch toughness requirement,
an ASTM grain size No. 7 or similar is typical and is most commonly
attained by normalizing.
Quenching or Hardening
A fully hardened steel is defined as having a 100% martensitic structure,
since it is the hardest strucutre that the steel can obtain. In order to achieve
a complete martensitic structure, austenite must first be formed throughout
the section thickness of the steel, called austenitizing. To ensure complete
austenitization through-thickness of the part, the appropriate temperature
above A3 must be reached for a sufficient amount of time. This is followed
by rapid cooling (quenching) in water, oil or air, with or without agitation,
depending on the hardenability of the particular steel.
Tempering
Tempering is generally applied to hardened or quenched steel to improve
mechanical properties, for the most part, tensile strength, ductility and
toughness. Tempering, formerly called drawing, is performed by heating a
quenched part to some point below the lower critical transformation
temperature for sufficient time depending on its size, commonly 2 hours or
more. Most steels are tempered between 400 to 1100oF (205 to
595oC). As higher temperatures or longer periods of time are employed,
toughness and ductility are increased, but at the expense of reduced hardness
and tensile strength. The microstructure of quenched and tempered steel is
referred to as tempered martensite.
Stress Relieving
When a metal is heated, expansion occurs which is proportional to the
temperature rise. Similarly, upon cooling a metal, contraction
occurs. When a steel component is heated at one point more than at
another, as in welding or during forging, internal stresses are set up. During
heating, expansion of the heated area do not occur uniformly, and the
component tends to distort. On cooling, contraction is restricted from
occurring by the unyielding cold metal surrounding the heated area, as in a
weldment. The forces attempting to contract the metal are not relieved, and
when the metal is cold again, the forces remain as internal stresses, also
called residual stresses. Stresses also result from volume changes which
accompany phase transformations, such as the transformation of austenite to
martensite.

Residual stresses are harmful because they may cause distortion of steel
parts and/or may render the part susceptible to brittle fracture and stress
corrosion cracking mechanisms. To relieve these stresses, plain carbon
steel is typically heated to between9400 to 1100oF (482 to 595oC), assuring
that the entire part is heated uniformly, then slowly cooled back to room
temperature. This procedure is called relief annealing or, more
commonly,stress relieving.
AN INTRODUCTION TO THE
WELDING METALLURGY OF STEEL
Welding is the joining of two or more pieces of metal by applying heat or
pressure, or both, with or without the addition of filler metal, to produce a
localized union through fusion or recrystallization across the interface.
[ Go to Top ]
In industrial welding practice, a steel of one composition, such as pipeline
or pressure vessel steel, is likely to be welded with a steel electrode of
different chemical composition, such as AWS A5.1 classification
E7018. The majority of filler metals classified by AWS, CSA, and other
standards are based on providing crack-free welds and mechanical
properties that at least meet the minimum requirements of the base metal.
The chemical composition match, although important, is a secondary
consideration for carbon steels. However, matching filler metal chemistry
to base metal chemistry becomes increasingly important when welding alloy
and stainless steels.
As a result of the chemically nonmatching filler metal, heat distribution, and
electrical arc characteristics, the weld joint is usually a chemically
heterogeneous composite consisting of as many as six metallurgical distinct
regions.
A typical single pass weld is shown in figure 19 and consists of:
1. Composite zone
2. Unmixed zone
3. Weld interface

4. Partially melted zone


5. Heat-affected zone
6. Unaffected Base metal

Composite Zone
The combination of melted filler metal and melted base metal creates a
liquid weld pool that becomes the composite zone upon cooling. Should
the filler metal be of a different chemical composition compared to the base
metal, then the base metal is said to become diluted by the filler

metal. However, due to the electrical sitrring action of the welding arc and

thermodynamic forces, the composite zone is mainly homogeneous.


Unmixed Zone
A very tin region, typically 0.05 - 0.01 in (1.25 - 2.5 mm) surrounding the
composite zone is called the unmixed zone. The metal in this region
solidified prior to mixing with the filler metal since the temperature reached
was just above its melting point. The stirring action and time above the
melting temperature is insufficient for mixing to have taken
place. Therefore its chemical composition is essentially the same as the
base metal. Although the unmixed zone is present in all fusion welds, it is
readily visible only in those welds using a filler metal alloy of substantially
different chemical composition than the base metal.
Weld Interface
The next zone metallurgically defined in a weldment is the weld interface,
typically called the weld line or fusion line. This interface clearly separates
the unmelted base metal on one side and the solidified weld metal on the
other side. In pure metals, the transition from base metal to weld metal is
often difficult to observe metallographically because of epiaxial qrowth,
where during liquid weld metal solidification, the new solid crystals begin to
grow from the existing base metal grains. However, in carbon and alloy
steels, the weld interface is easily revealed by standard etching
techniques. This method is often used in the "field"to delineate the weld
metal (copmosite and unmixed zones) from the heat affected zone in carbon
steel welds. Two common macroetchants for this purpose are: 5 to 10%
Nital (5 to 10% concentrated nitric acid dissolved in 90% to 95% methanol
or ethyl alcohol, by volume) and 10% Ammonium Persulphate (10 g
ammonium persulphate dissolved in 100 mL of water).

Partially Melted Zone


In the base metal immediately adjacent to the weld interface, the
temperature from welding is insufficient for complete melting, i.e. just
below the liquidus temperature but above the solidus temperature, in the
solid-liquid region. This area is called the partially melted zone. In steels,
this zone becomes susceptible to hot cracking since the liquidation (melting)
of manganese sulphide inclusions can cause weak localized regions that do
not have the hot strength to withstand the heating/cooling
(expansion/contraction) cycle of welding.
Heat-Affected Zone (HAZ)
The HAZ is the subject of continuing interest since it involves a wide range
of temperatures from the welding operation that can significantly alter the
base metal's metallurgy and associated mechanical and physical
properties. See the next section.
Unaffected Base
Metal
[ Go to Top ]
The further region from the liquid weld metal (composite zone) is the
unaffected base metal. This region is defined by welding temperature below
the lower critical transformation temperature (A3), 1340oF (725oC) for
carbon steels. As such, there are no phase transformations taht occur to the
original base metal microstructure. However, since the temperature
reached in this region are sufficient to cause precipitation hardening,
tempering or stress relieving, it is important to know the original
microstructure and the potential effects of further heating.
HEAT AFFECTED ZONE (HAZ)
Single-Pass Weldments
The heat affetced zone (HAZ) extends from weld interface to the unaffected
base metal. The resultant temperature range in the HAZ extends from just
below the liquidus down to the sub-critical temperatures slightly less than
the lower critical transformation temperature, i.e. 2970 - 8

00oF (1350 - 425oC) for carbon steel. With this large temperature gradient
comes varying microstructures in steel that will depend on the peak
temperature reached, time at temperature, and cooling rate. Consequently,
the term "heat affected zone" is really a misnormer when describing it on a
metallurgical basis, since the HAZ is really made up of several distinct
metallurgical zones.
Figure 20 shows a cross section of a single-pass weldment outlining the
weld metal and HAZ. Because of varying thermal conditions as a function
of distance from the weld interface, the HAZ is actually composed of four
distinct regions, namely, the grain-coarsened-HAZ, grain-refined-HAZ,
intercritical_HAZ, and subcritical-HAZ. Each of these regions within the
HAZ possesses microstructures and associated mechanical and physical
properties that make them unique.
To define the four regions of the HAZ in metallurgical terms, the Fe-Fe3C
phase diagram provides an ideal tool. Figure 21 illustrates a single-pass
weldment and compares the HAZ microstructures produced by the heat of
welding and related to peak temperature reached, time at temperature, and
cooling rate.
Grain-Coarsened-HAZ

The peak temperatures reached in the grain-coarsened-HAZ region raange


from 2000 to 2700oF (1090 to 1480oC), depending on the carbon
content. Another way to describe this temperature range in metallurgical
terms, is that it exrends from much above the upper critical transformation
temperature to just below the solidus temperature. There are two main
metallurgical conditions that occur in this region: 1) the microstructure is
austenite (for the most part) and 2) since the austenite produced is much
above the upper critical transformation temperature, grain growth will
occur. The amount of grain growth will depend on the peak temperature
and time at that temperature, i.e. the higher the peak temperature and the
longer the time at that temperature, the larger the austenite grains will grow.
Two significant metallurgical consequences result in this region: 1) since
austenite is produced, the potential for transformation to martensite upon

cooling exists, where martensite is not a desirable transformation product


due to its lack of ductility, toughness, and susceptibility to cold cracking and
2) as austenitic grain size grows, the resultant room temperature
microstructure will be similarly efr
effected, with low temperature notch (charpy) toughness being significantly
change, i.e. the larger the grain size, the lower the notch toughness.
Grain-refinementHAZ
[ Go to Top ]
This region comprises temperature from just above the lower critical
transformation temperature and up to 200oF (93oC) higher. These
temperatures are within the normalizing heat treatment range and are very
conducive to austenitic grain refinement and its associated improved low
temperature notch (charpy) toughness. On the other hand, austenite is still
produced and the likelihood of martensite must be considered.
Intercritical-HAZ
The temperatures in this region includes the intercritical ranges, between the
lower and upper critical temperatures. Some austenite is produced in this
partially transformed range, such that the potential for martensite
transformation exists. In medium and high carbon steels, this austenite can
contain large amounts of carbon which has a higher tendency to produce
martensite on cooling.
Subcritical-HAZ
The subcritical-HAZ includes the tempered area of the Fe-Fe3C phase
diagram. Should the base metal be in the tempered condition (i.e. quenched
and tempered) the heat of welding may be sufficient for further tempering,
thereby reducing the tensile strength and hardness in this region. There are
no phase transformations which take place in the tempered area since the
lower critical transformation temperature is not exceeded.
Cooling Rate
The cooling rate also varies from region to region in the HAZ. It increases
with increasing peak temperature ay constant heat input and decreases with
increasing heat input at constant peak temperature. Figure 21 displays a
typical cooling rate curve (solid line showing peak temperature) for a singlepass weldment. By drawing a tangent to this curve, it becomes apparent
that the steepest part of the curve (i.e. fastest cooling rate) is related to the

grain-coarsened-HAZ. Since this region is comprised essentially of


austenite and is linked to the fastest cooling r

ate, it therefore possesses the greatest potential for transformation to


martensite.
Toe cracking is a form of hydrogen-assisted cold cracking related to welding
and owes its name to the area of the weld where cracking initiates. The toe
of the weld is in the extreme grain-coarsened-HAZ (adjacent to the weld
interface) and is directly affiliated with the austenite to martensite
transformation in this high cooling rate region; a consequence of welding
that produces the highest potential hardness in the HAZ of a carbon steel
(see figure 22).
Multi-Pass Weldments
In multi-pass weldments, the situation is much more complex because of the
presence of reheated zones within the HAZ. The reheating of the HAZ
microsturctures by subsequent weld passes increases the inhomegeneity of
the various regions with respect to microstructure and mechanical
properties. Reaustenitization and subcritical heating can have a significant
effect on the subsequent structures and properties of the HAZ. The loss of
low temperature notch (charpy) toughness in a multi-pass HAZ is related to
small regions of limited ductility and low cleavage resistance within the
grain-coarsened-HAZ that are known as the localized brittle zones. At an
adjacent weld interface in the multi-pass HAZ, the localized brittle zones
may become aligned. These aligned brittle zones offer short and easy paths

for crack propagation. Consequently, fracture may occur along the fusion
line.
Welding Metallurgy Summary
Metals Data
The following Chapters/Tables list some of the more practical metals data
contained in today's engineering standards. Many of the various
organizations that issue metal standards and specifications throughout the
world are included. One prominent organization is the Americam Society
for Testing and Materials (ASTM). The following is an excerpt from the
1991 Annual Book of ASTM Standard Volume 00.01, describing the
identification for individual ASTM Standards:
Each ASTM Standard has a unique serial designation. It is comprised of
acapital letter indicating general classification (A, ferrous metals; B, nonferrous metals; C, cementitious, ceramic, concrete, and masonry materials;
D, miscellaneous materials; E, miscellaneous subjects; F, material for
specific applications; G, corrosion, deterioration, and degradation of
materials; ES, emergency standards; P, proposals), a serial number (one to
four digits), a dash, and the year of issue.
In each serial designation, the number following the dash indicates the year
of original adoption or, in the case of revision, the year of last
revision. Thus, standards adopted or revised during the year 1991 have as
their final number, 91. A letter following this number indicates more than
one revision during that year, that is 91a indicates the second revision in
1991, 91b the third revision, etc. Standards that have been reapproved
without change are indicated by the year of last reapproval in parentheses as
part of the designation number, for example, (1991). A superscript epsilon
indicates an editorial change since the last revision or reapproval; 1 for the
first change,2 for the second change, etc.
If a standard is written in acceptable metric units and has a comparison
standard written in inch-pound units (or other units), the metric standard is
identified by a letter M after the serial number; this standard contains "hard
metric" units.
If s standard is written in inch-pound units (or other units) and acceptable
metric units, the document is identified by a dual alphanumeric designation.
When reference is made to a standard, the complete designation should be
given. Best practice is to state the designation and title. The boldface

number(s) following the title refer to the volume(s) of the Annual Book of
ASTM Standards in which the standard appears.

SUMMARY
This lecture commences with a discussion of the need for civil and structural
engineers to have a basic knowledge of the metallurgy of steel. Then the
crystalline nature of irons and steels is described together with the influence
of grain size and composition on properties. The ability of iron to have more
than one crystalline structure (its allotropy) and the properties of the
principal crystalline forms of alloys of iron and carbon are discussed.
The metallurgy and properties of slowly cooled steels are reviewed,
including the influence of grain size, rolling, subsequent heat treatment and
inclusion shape and distribution. Rapidly cooled steels are treated
separately; a brief description of quenching and tempering is followed by a
discussion of the influence of welding on the local thermal history.
Hardenability, weldability and control of cracking are briefly discussed.
Finally the importance of manganese as an alloying element is introduced.
1. INTRODUCTION

1.1 Why Metallurgy For Civil and Structural Engineers?


The engineering properties of steel, i.e. strength, ductility and resistance
against brittle fracture, depend on its crystalline structure, grain size and
other metallurgical characteristics.
These microstructural properties are dependent on the chemical composition
and on the temperature-deformation history of the steel. Heat treatments that
occur during welding may also have a large influence on the engineering
properties.
When selecting steel for welded structures, it is important to have at least a
basic knowledge of metallurgy. This knowledge is required especially when
large and complicated structures are being designed, such as bridges,
offshore structures, and high rise buildings.
Selecting materials, welding processes and welding consumables usually
requires consultation of "real" metallurgists and welding specialists. A basic
knowledge of metallurgy is essential for communication with these
specialists.

Finally, a basic knowledge of metallurgy also enables civil and structural


engineers to have a better understanding of the engineering properties of
steel and the performance of welded structures.
1.2 The Scope of Lectures in Group 2
Lecture 2.1 deals with the characteristics of iron-carbon alloys. Where
possible, direct links are indicated to the engineering properties and
weldability of steel. These subjects are covered inLectures 2.2 and 2.6
respectively.
Lecture 2.3 describes steelmaking and the forming of steel into plates and
sections. The various processes for controlling the chemical composition
and the different temperature-deformation treatments are discussed. Most of
the underlying principles described in Lecture 2.1 are applied.
Steels are available in various grades and qualities. The grade designates the
strength properties (yield strength and ultimate strength), while the quality is
mainly related to resistance against brittle fracture. Grades and qualities are
explained in Lecture 2.4. A system for choosing the right quality according
to Eurocode 3 (Annex C) [1] is presented. Some guidelines for the selection
of steel grade are given.
2. STRUCTURE AND COMPONENTS OF STEEL

2.1 Introduction
To get an impression of the metallurgical structure of steel, a piece of steel
bar can be cut to expose a longitudinal section, the exposed surface ground
and polished and examined under a microscope.
At modest magnifications, a few particles are seen which are extended in the
direction of rolling of the bar, see Slide 1. These particles are inclusions.
They are non-metallic substances which have become entrained within the
metal during its manufacture, mostly by accident but sometimes by design.
Their presence does not affect the strength but has an adverse effect on
ductility and toughness. Particular types of inclusion can greatly enhance the
machinability of steels and may therefore be introduced deliberately.

Slide 1 : Longitudinal stringers of inclusions in hot rolled steel. (x 500)


To reveal the true structure of the metal, the polished surface must be
chemically etched. When this is done, a wide diversity of microstructure
may be seen which reflects the composition of the steel and its processing,
see Slides 2 - 5.
The microstructure has a significant effect on the engineering properties as
described in later sections of this lecture.
2.2 The Components of Steel
Steels and cast irons are alloys of iron (Fe) with carbon (C) and various
other elements, some of them being unavoidable impurities whilst others are
added deliberately.
Carbon exerts the most significant effect on the microstructure of the
material and its properties. Steels usually contain less than 1% carbon by
weight. Structural steels contain less than 0,25% carbon: the other principal
alloying element is manganese, which is added in amounts up to about
1,5%. Further alloying elements are chromium (Cr), nickel (Ni),
molybdenum (Mo) etc. Elements such as sulphur (S), phosphorus, (P),
nitrogen (N) and hydrogen (H) usually have an adverse effect on the
engineering properties and during the steel production, measures are taken
to reduce their contents. Cast irons generally contain about 4% carbon. This
very high content of carbon makes their microstructure and mechanical
properties very different from those of steels.

Each of the microstructures shown in Slides 2, 3, 4 and 5 is an assembly of


smaller constituents. For example, the 0,2% C steel of Slide 2 is
predominantly an aggregate of small, polyhedral grains, in this case <20m
in size. Closer examination of one of these grains shows it to be a single
crystal. However, unlike crystals of quartz or silicon or copper sulphate,
crystals of iron (Fe) are soft and ductile. The internal structure of these
crystals is discussed later.

Slide 2 : Microstructure of hot rolled steel containing 0,2% carbon showing


ferrite (white) and pearlite colonies (dark). (x 200)

Slide 3 : Microstructure of hot rolled steel containing 0,36% carbon showing


increased proportions of pearlite (dark). (x 500)

Slide 4 : Microstructure of heat treated hot rolled steel containing 0,36%


carbon showing spheroidised pearlite (dark) in a ferrite matrix. (x 750)

Slide 5 : Microstructure of quenched hot rolled steel containing 0,36%


carbon showing bainite (x 200)
The steel of Slide 2 is an example of a polycrystalline substance which has
been made visible by polishing and etching.
(a) The surface is polished but not etched.
(b) The surface is polished and etched. Different reflections of the light
indicate different orientation of crystals (polycrystalline structure).
(c) Some etchants affect only the grain boundaries. These etchants are used
when it is required to investigate the grain structure, e.g. to estimate the
grain size.
(d) The appearance of etched grain boundaries of Figure 1c.
(e) The appearance of a steel with 0,15% carbon (enlargement 100x). The
dark areas are pearlite. The grain boundaries are clearly indicated. The dark
areas indicate the presence of carbon.
By adjusting the history of rolling and heating treatment experienced by the
steel during its production, the grain size can be altered. This technique is
useful because the grain size affects the properties. In particular, the yield
strength is determined by the grain size, according to the so-called Petch
equation:
y = o + kd-1/2

where y is the yield strength


o is effectively the yield strength of a very large isolated crystal: for mild
steel this is 50N/mm2
d is the grain size in mm
k is a material constant, which for mild steel is about 20N/mm-3/2
Thus, if the grain size is 0.01 mm, y 250N/mm2.
2.3 The Crystal Structure
The internal structure of the crystal grains is composed of iron atoms
arranged according to a regular three-dimensional pattern. The pattern is
illustrated in Figure 2. This pattern is the body-centred cubic crystal
structure; atoms are found at the corners of the cube and at its centre. The
unit cell is only 0,28nm along its edges. A typical grain is composed of
about 1015 repetitions of this unit.

This crystal structure of iron at ambient temperature is one of the major


factors determining the metallurgy and properties of steels.
Steels contain carbon. Some of it, a very small amount, is contained within
the crystals of iron. The carbon atoms are very small and can fit, with some
distortion, into the larger gaps between the iron atoms. This arrangement
forms what is known as an interstitial solid solution: the carbon is located in
the interstices of the iron crystal.
In the steels of Slides 2, 3 and 4, most of the remaining carbon has formed a
chemical compound with the iron, Fe3C, iron carbide or cementite. Iron
carbide is also crystalline but it is hard and brittle. With 0,1%C, there is only
a small amount of Fe3C in steel. The properties of such steel are similar to
those of pure iron) [2]. It is ductile but not particularly strong and is used for
many purposes where ability to be shaped by bending or folding is the
dominant requirement.
For a steel of higher carbon content, say 0,4%, as shown in Slide 5, a low
magnification shows it to be composed of light and dark regions - about
50:50 in this case. The light regions are iron crystals containing very little
dissolved carbon, as in the low carbon steel. The dark regions need closer
examination. Slide 6 shows one such region at higher magnification. It is
seen to be composed of alternate layers of two substances, iron and Fe3C.
The spacing of the laminae is often close to the wavelength of light and
consequently the etched structure can act as a diffraction grating, giving
optical effects which appear as a pearl-like iridescence. Consequently, this
mixture or iron and iron carbide has acquired the name 'pearlite'. The origin
of the pearlite and its effect on the properties of steel are revealed by
examining what happens during heating and cooling of steel.

Slide 6 : Polycrystalline structure of steel containing 0,4% carbon. (x 400)


3. IRON-CARBON PHASES

3.1 Influence of Temperature on Crystal Structure


The crystal structure of steel changes with increasing temperature. For pure
iron this change occurs at 910 C. The body-centred cubic (bcc) crystals of
Figure 2 change to face-centred cubic (fcc) crystals as illustrated in Figure 3.
For fcc crystals the atoms of iron are on the cube corners and at the centres
of each face of the cube. The body-centred position is empty.

A given number of atoms occupy slightly less volume when arranged as fcc
crystals than when arranged as bcc crystals. Thus the change of the crystal
structure is accompanied by a volume change. This change is illustrated in
Figure 4. When a piece of pure iron is heated, expansion occurs in the
normal way until the temperature of 910 C is reached. At this temperature
there is a step contraction of about % in volume associated with the
transformation from the bcc to fcc crystal structure. Further heating gives
further thermal expansion until, at about 1400C the fcc structure reverts to
the bcc form and there is a step expansion which restores the volume lost at
910C. Heating beyond 1400C gives thermal expansion until melting
occurs at 1540C. The curve is reversible on cooling slowly.

The property that metals may have different crystal structures, depending on
temperature, is called allotropy.
3.2 Solution of Carbon in bcc and fcc Crystals
When the atoms of two materials A and B have about the same size, crystal
structures may be formed where a number of the A atoms are replaced by B
atoms. Such a solution is called substitutional because one atom substitutes
for the other. An example is nickel in steel.
When the atoms of two materials have a different size, the smaller atom may
be able to fit between the bigger atoms. Such a solution is called interstitial.
The most familiar example is the solution of carbon in iron. In this way the
high temperature fcc crystals can contain up to 2% solid solution carbon at
1130C, while in the low temperature bcc crystals, the maximum amount of
carbon which can be held in solution is 0,02% at 723C and about 0,002% at
ambient temperature.
Thus a steel containing 0,5% carbon, for example, can dissolve all the
carbon in the higher temperature fcc crystals but on cooling cannot maintain
all the carbon in solution in the bcc crystals. The surplus of carbon reacts
with iron to form iron carbide (Fe3C), usually called cementite. Cementite is
hard and brittle compared to pure iron.
The amount of cementite and the distribution of cementite particles in the
microstructure is important for the engineering properties of steel.
The distribution of cementite is highly dependent on the cooling rate. The
distribution may be explained by considering the so-called iron-carbon
phase diagram, see Section 3.4.
3.3 Nomenclature
The following nomenclature is used by the metallurgist:
Ferrite or -Fe

The bcc form of iron in which up to 0,02%C by weight may be dissolve

Cementite

Iron carbide Fe3C (which contains about 6,67%C).

Pearlite

The laminar mixture of ferrite and cementite described earlier. The ove

Austenite or -Fe

The fcc form of iron which exists at high temperatures and which can c

Steel

Alloys containing less than 2% carbon by weight.

Cast Iron

Alloys containing more than 2% carbon by weight.

Steel used in structures such as bridges, buildings and ships, usually


contains between 0,1% and 0,25% carbon by weight.

3.4 The Iron-Carbon Phase Diagram


The iron-carbon phase diagram is essentially a map. The most important
part is shown in Figure 5. More details are given in Figure 6.

Any point in the field of the diagram represents a steel containing a


particular carbon content at a particular temperature.
The diagram is divided into areas showing the structures that are stable at
particular compositions and temperatures.
The diagram may be used to consider what happens when a steel of 0,5%C
is cooled from 1000C (Figure 6).
At 1000C the structure is austenite, i.e. polycrystalline fcc crystals with all
the carbon dissolved in them. No change occurs on cooling until the
temperature reaches about 800C. At this temperature, a boundary is crossed
from the field labelled Austenite () to the field labelled Ferrite + Austenite
( + ), i.e. some crystals of bcc iron, containing very little carbon, begin to
form from the fcc iron. Because the ferrite contains so little carbon, the
carbon left must concentrate in the residual austenite. The carbon content of
the austenite and the relative proportions of ferrite and austenite in the
microstructure adjust themselves to maintain the original overall carbon
content.
These quantities may be worked out by considering the expanded part of the
iron-carbon diagram shown in Figure 7. Imagine that the steel has cooled to
750C. The combination of overall carbon content and temperature is
represented by point X.

All the constituents of the microstructure are at the same temperature. A line
of constant temperature may be drawn through X. It cuts the boundaries of
the austenite and ferrite field at F and A. These intercepts give the carbon
contents of ferrite and austenite respectively at the particular temperature.
If, now, the line FA is envisaged as a rigid beam which can rotate about a
fulcrum at X, the 'weight' of austenite hanging at A must balance the
'weight' of ferrite hanging at F. This is the so-called Lever Rule:
Weight of ferrite FX = Weight of austenite AX
The ratio of ferrite to austenite in the microstructure is then given by:

Thus, as the steel cools, the proportion of ferrite increases and the carbon
content of the remaining austenite increases, until cooling reaches 723C. At
this temperature the carbon content of the austenite is 0,8% and it can take
no more. Cooling to just below this temperature causes the austenite to
decompose. It decomposes into the lamellar mixture of ferrite and Fe3C
identified earlier as pearlite.
The proportions of ferrite and pearlite in the microstructure, say at 722C,
are virtually the same as the proportions of ferrite and austenite immediately
before the decomposition at 723C. Thus, referring to Figure 7 and using the
Lever Rule:
Weight of ferrite F X = Weight of pearlite F P
In this case, there should be about twice as much pearlite as ferrite.
For other steels containing less than 0,8%C, the explanation is identical
except for the proportions of pearlite in the microstructure below 723C.
This varies approximately linearly with carbon content between zero at
0,02%C and 100% at 0,8%C. A typical mild steel containing 0,2%C would
contain about 25% pearlite.
For steels containing a greater percentage of carbon than 0,8%, the structure
is fully austenitic on cooling from high temperatures. The first change to
occur is the formation of particles of Fe3C from the austenite. This change
reduces the carbon content of the residual austenite. On further cooling, the
carbon content of the austenite follows the line of the boundary between
the field and + Fe3C field. Once again, on reaching 723C the carbon

content of the austenite is 0,8%. On cooling further, it decomposes into


pearlite as before. Therefore, the final microstructure consists of a few
particles of Fe3C embedded in a mass of pearlite, see Figure 6.
4. COOLING RATE

4.1 Cooling Rate During Austenite to Ferrite Transformation and


Grain Size
During cooling of austenite, the new bcc ferrite crystals start to grow from
many points. The number of starting points determines the number of ferrite
grains and consequently the grain size. This grain size is important because
the engineering properties are dependent on it. Small grains are favourable.
By adding elements like aluminium and niobium, the number of starting
points can be increased. Another important factor is the cooling rate. When
cooling is slow, the new ferrite grains develop from only a few most
favourable sites. At high cooling rates, the number of starting points will be
much higher and the grain size smaller. Slides 7 - 9 shows steels with
various grain sizes, produced at different finish rolling temperatures.

Slide 7 : Microstructure of pearlite. (x 1000)


Another important factor is that, when a fine grained steel is heated to a
temperature in excess of about 1000C, some of the austenite grains grow
while neighbouring grains disappear. This grain growth occurs during
welding in the so-called heat affected zone (HAZ). This is a 3-5 mm wide

zone in the plate adjacent to the molten metal. Microstructural changes in


the heat affected zone usually give rise to a deterioration of the engineering
properties of the steel.
4.2 Slowly Cooled Steels
4.2.1 Influence of carbon on the microstructure
The iron-carbon phase diagram in Figures 5 and 6 shows that, for structural
steel (between 0,1% and 0,25% carbon), the formation of ferrite starts at
about 850C and ends at 723C. It will be remembered that ferrite can
contain hardly any carbon. Consequently, the austenite phase transforms to
ferrite and cementite (Fe3C).
When the cooling rate is slow, the carbon atoms have time to migrate to
separate "layers" in the microstructure and to form the structure called
pearlite, as shown before in Slides 2, 3, 4 and 5. The ferrite in this mixture is
soft and ductile. The cementite constituent is hard and brittle. The mixture
(pearlite) has properties between these two extremes.
The tensile strength properties of a steel containing both ferrite and pearlite
roughly scale according to the proportions of these constituents in the
microstructure as seen in Figure 8.

The dramatic effect of carbon content on toughness is shown in Figure 9.


Increasing pearlite content decreases the upper shelf toughness and increases
the ductile-brittle transition temperature.

Figures 8 and 9 illustrate one of the difficulties in the choice of carbon


content. Increasing the carbon content is beneficial in that it improves yield

strength and ultimate tensile strength, but is undesirable in that it reduces


ductility and toughness. A high carbon content may also cause problems
during welding, see Section 4.3.
In European Norm 10025, Table 3, [3] the chemical composition for flat and
long products is given. An extract is presented in Figure 10. The designation
S235 JR, for example, indicates that the yield strength is at least 235
N/mm2. It is emphasised that the compositional values in the table are
maximum values. Many steelmakers achieve much lower levels, resulting in
better ductility, resistance against brittle fracture, and weldability.

The lowest carbon content that can be achieved easily on a large scale is
about 0,04%. This content is characteristic of sheet or strip steels intended to
be shaped by extensive cold deformation, as in deep drawing.
Carbon contents of more than 0,25% are used in the wider range of general
engineering steels. These steels are usually put into service in the quenched
and tempered state (see below) for a great multiplicity of purposes in
mechanical engineering. High strength bolts for some structural applications
would also be steels of this type.
4.2.2 The need for control of grain size
The mechanical properties of steel are affected by grain size. Slides 8 and 9
show microstructures of two samples of the same batch of mild steel which
have been treated, by methods outlined in Section 4.2.3, to give different
grain sizes. Reduction in grain size improves yield strength but also has a
profound effect on the ductile/brittle transition temperature, see Figure 11.
Thus, there are several benefits from the same microstructural charge. This
is an unusual circumstance in metallurgy where adjustments to improve one
property often mean a worsening of another and a compromise is necessary.
An example of such compromise relates to carbon content, already
discussed above.

Slide 8 : Microstructure of typical hot rolled structural steel containing


0,15% carbon and showing white ferrite grains and pearlite colonies. (x
200)

Slide 9 : Refined microstructure of controlled rolled structural steel


containing 0,15% carbon (white ferrite grains and pearlite colonies. (x 200)

4.2.3 Grain size control by normalising


In Section 4.2.1 the transformations that can occur when steels are cooled
slowly are described. To form ferrite and pearlite from austenite, the carbon
atoms in the steel must change their positions. The diffusion processes
which transport the atoms within the solid occur at rates which depend
exponentially on temperature. The rate of cooling also affects these
transformations.
If the cooling rate is increased the transformations occur faster. In addition,
the diffusional processes cannot keep up with the falling temperature. Thus,
a steel cooled very slowly in a furnace keeps close to the requirements of the
phase diagram. But the same steel, removed from the furnace and allowed to
cool in air, may undercool before completing its sequence of
transformations. This more rapid cooling has two effects. First it tends to
increase slightly the proportion of pearlite in the microstructure. Secondly it
produces ferrite with a finer grain size and pearlite with finer lamellae. Both
of these microstructural changes give higher yield strength and better
ductility and toughness.
Furnace cooled steels are known as fully annealed steels. Air cooled steels
are known as normalised steels.
Grain size can also be affected by the temperature to which the steel is
heated in the austenite range. The grains of austenite coarsen with time, the
rate of coarsening increasing exponentially with temperature. The
coarsening is important because the transformation to ferrite and pearlite on
cooling starts at the grain boundaries in the austenite. If the new structures
start growing from points which are further apart in a coarse grained
austenite, the grain size of the resulting ferrite is itself coarser. Thus, steels
should not be overheated when austenitising before normalising.
The temperature to which the steel is heated before cooling in air is usually
referred to as the normalising temperature. The requirements of the last
paragraph mean that this temperature should be as low as possible, as long
as the structure is single phase austenite. A glance at the phase diagram of
Figure 5 shows that the normalising temperature decreases as the carbon
content increases from zero to 0,8%. It should lie in the hatched band shown
in Figure 12.

4.2.4 Microstructural changes accompanying hot rolling of steels


Structural steel sections are produced by hot rolling ingots or continuously
cast strand into the required forms. The rolling processes have important
effects on the development of the microstructure in the materials.
The early stages of rolling are carried out at temperatures well within the
austenite range, where the steel is soft and easily deformed. The deformation
suffered by the material breaks up the coarse as-cast grain structure but, at
these high temperatures, the atoms within the material can diffuse rapidly
which allows the deformed grains to recrystallise and reform the equiaxial
polycrystalline structure of the austenite.
Heavy deformation at low temperatures in the austenite range gives finer
recrystallised grains. If the rolling is finished at a temperature just above the
ferrite + austenite region of the equilibrium diagram and the section is
allowed to cool in air, an ordinary normalised microstructure having
moderately fine-grained ferrite results. Modern controlled rolling techniques
aim to do this, or even to roll at still lower temperatures to give still finer
grains.
If the temperature falls so that the rolling is finished in the ferrite + austenite
range, the mixture of ferrite and austenite grains is elongated along the
rolling direction and a layer-like structure is developed. If now, the section
is air-cooled, the residual austenite decomposes into fine-grained ferrite and
pearlite, with the later being present as long, cigar shaped, bands in the
material, as in Slide 10. Structural steels are not harmed by microstructures
of this sort.

Slide 10 : Microsection through a fillet weld on structural steel showing


three distinct regions: the coarse grained cast structure of the weld deposit,
the heat affected zone, and the unaffected microctructure of the parent steel.
(x 200)
If the finish rolling temperature drops further, to below 723C, the
equilibrium diagram shows that the structure should be a mixture of ferrite
and pearlite. Rolling in this range is usually restricted to low carbon steels
containing less than 0,15%C because the presence of pearlite makes rolling
difficult.
If the temperature is above about 650C, the ferrite grains recrystallise as
they are deformed, as was the case with austenite. The carbide laths (Fe3C)
in the pearlite become broken and give rise to strings of small carbide
particles extending in the direction of rolling, see Slide 11. The ferrite from
the pearlite becomes indistinguishable from the rest of the ferrite.

Slide 11 : Macrosection through a butt weld on hot rolled steel plate, typical
of line pipe weld.
If rolling is done at ambient temperature, the pearlite is broken up in the
same way, but the ferrite can not recrystallise. It work-hardens, i.e. the yield
and ultimate tensile strength of the steel increase, and the ductility
decreases, see Figure 13. As cold rolling continues, the force required to
continue deformation increases because of the increasing yield strength.
Furthermore, the steel becomes less ductile and may begin to split. The
amount of cold rolling that can be done is therefore very much smaller than
that which can be achieved when the steel is hot.

Of course, cold working need not be applied by rolling. Any way of


deforming the material causes work hardening. For example, high strength
steel wire is made by cold-drawing, imparting large deformations. In
another example, one type of reinforcing bar is made by twisting square
section bar into a helical form. The cold-deformation produced in this way is
not large but causes significant work hardening.
To restore the ductility and at the same time reduce the work hardened state
of the material, it is necessary to reform the isotropic, polycrystalline
structure of the ferrite. Re-heating to temperatures between about 650C and
723C allows the ferrite to recrystallise. The carbide particles are unaffected
by this treatment.
Thus, there is another technique for controlling the grain size of steel. The
greater the amount of deformation before the recrystallisation treatment and
the lower the temperature of the treatment, the finer is the final grain size.
Because this type of treatment does not involve the formation and
decomposition of austenite, it is known as sub-critical annealing. The
resulting microstructure has good ductility and deep drawing characteristics.
Sheet steels of low carbon content (< 0,1%C) are usually supplied in this
condition. Objects such as motor car body panels are formed from such
steels by cold pressing.
If the material is heated into the austenite range, subsequent cooling reforms
the normalised microstructure.
4.3 Rapidly Cooled Steels
4.3.1 Formation of martensite and bainite
Normalising causes steels to undercool below the requirements of the phase
diagram before the austenite transforms into fine ferrite and pearlite. Still
further increases in cooling rate give further undercooling and still finer
microstructures.
Very rapid cooling by quenching into cold water, causes the formation of
ferrite and pearlite to be suspended. The internal diffusion-controlled
rearrangement of atoms needed to form those products cannot occur
sufficiently rapidly. Instead, new products are formed by microstructural
shear transformations at lower temperatures. Very fast cooling gives
martensite: its microstructure is shown in Slide 12. When martensite forms,
there is no time for the formation of cementite and the austenite transforms
to a highly distorted form of ferrite which is super saturated with dissolved
carbon. The combination of the lattice distortion and the severe work
hardening resulting from the shear deformation processes necessary to

achieve the transformation cause martensite to be extremely strong but very


brittle.

Slide 12 : Longitudinal section of hot rolled structural steel showing dark


bands of pearlite in a ferrite matrix. (x 200)
Less rapid cooling can give a product called bainite,. This is similar to
tempered martensite where much of the carbon has come out of solution and
formed fine needles of cementite which reinforce the ferrite.
4.3.2 Martensite in welded structures
Civil engineering structures are not heat-treated by heating to, say, 900C
and quenching into water. However, there is one important circumstance
which can produce martensite in localised parts of the structure, and that is
welding. The weld zone is raised to the melting temperature of the steel and
the immediately adjacent solid metal is heated to temperatures well within
the austenite range. When the heat source is removed, the whole region
cools at rates determined mainly be thermal conduction into the surrounding
mass of cold metal. These rates of cooling can be very large, exceeding
1000C per second in some cases and can produce transformation structures
such as martensite and bainite. The properties of rapidly cooled steels and
the influence of carbon content on the nature of the transformation product ferrite and pearlite, or bainite, or martensite - are discussed below.

Figure 14 shows the hardness of martensite as a function of its carbon


content. Reheating martensite to temperatures up to about 600C causes
cementite to precipitate which causes the steel to soften and become much
tougher. This reheating is known as tempering. The extent of these changes
increases as the reheating temperature increases, as shown in Figure 15.
Tempering at 600C produces an extremely tough material. What is more,
its ductile-brittle transition temperature is lower than for the same steel in
the normalised condition. Bainite has properties similar to those of tempered
martensite.

4.3.3 Quenching and tempering


The process of quenching and tempering, when allied to changes of steel
composition, can produce a bewilderingly wide range of properties. Steels
heat-treated in this way are used for a multiplicity of general engineering
purposes which demand hardness, wear resistance, strength and toughness.
Once again, compromises must be struck between these desirable properties
but generally quenched and tempered steels exhibit optimum combinations
of strength and toughness. For structural purposes quenched and tempered
plate is used in large storage tanks, hoppers, earthmoving equipment, etc.
Martensite produced in a weld heat-affected zone as a result of single pass
welding would be in its hard and brittle untempered condition. Furthermore,
the formation of martensite from austenite is accompanied by a volume
expansion of approximately 0,4%. This expansion, together with the uneven
thermal contractions taking place as a result of uneven cooling, can produce
local stresses of sufficient magnitude to crack the martensite. Because this
type of cracking occurs after the HAZ has cooled, it is referred to as cold
cracking. The cracking problem can be further aggravated if the weld has
picked up hydrogen. Sources of hydrogen during welding might include
moisture from the atmosphere or damp welding electrodes. Hydrogen
dissolved in the weld metal diffuses to the hard HAZ where it initiates
cracks at sites of stress concentration. This diffusion can lead to cracking
which occurs some time, even days, after the welding is completed. Hard
HAZs of low ductility are less able to cope with this problem than are softer
and more ductile materials. This type of cracking is called delayed cracking
or hydrogen cracking.
Avoidance of cold cracking and hydrogen cracking requires that the material
should not be overhardened. As a rule of thumb, as-welded hardnesses of
less than about HV = 350 are considered to be acceptable. In modern fine
grain low carbon steels the "allowable" hardness may be increased to HV =
400 or even HV = 450.
The danger of hydrogen cracking may also be present in high strength
quenched and tempered steels, e.g. 10.9 bolts (Re 900 N/mm2 and
Rm 1000 N/mm2). When such bolts are electroplated with zinc or
cadmium, hydrogen may be picked up from the plating bath. Usually
cracking does not occur until sometime after tightening bolts when the
hydrogen has diffused to the sites of stress concentration at the thread roots.
4.3.4 Control of martensite formation
Martensite forms because ferrite and pearlite did not! If follows that
metallurgical factors which promote the formation of ferrite and pearlite

inhibit the production of martensite. The ability of a steel to form martensite


rather than ferrite and cementite is called hardenability. Note that this term
does not refer to the absolute value of hardness obtained, but to the ease of
formation of martensite.
The most convenient method of assessing hardenability is the so called
Jominy end quench. A rod-shaped sample is austenitised and then quenched
by spraying water onto one end face such that different cooling rates are
produced along the length of the bar. Thereafter, a flat is ground along its
length and the hardness measured as a function of distance from the
quenched end.
Some typical results are shown in Figure 16 for three different steels. For a
carbon steel containing 0,08%C and 0,3%Mn, cooling rates at 700C of
greater than about 50C s-1 are necessary to form martensite. On the other
hand in the 0,29%C, 1,7%Mn steel, martensite forms at much slower
cooling rates. It is mainly the increased carbon content that causes this
difference. In the alloy steel illustrated, martensite is formed even at very
slow cooling rates.

The significance of these curves depends very much on what is being


produced. If it is a thick-section gear wheel, the alloy steel would be ideal. It
could be cooled gently and still produce martensite, the gentle cooling being
an advantage because it would reduce stresses arising from differential
contraction rates, and hence reduce the possibility of quench cracking.

Thereafter, it could be tempered to achieve the desired combination of


strength and toughness. On the other hand, for a welded joint, the plain
carbon steel would be preferable in which it is difficult to form martensite
and the hardness of any martensite produced would be relatively low.
Welding presents particular problems for the metallurgist. Slide 13 shows a
micro section through a typical structural weld. The micro structures range
from the coarse grained cast structure of the weld deposit, to the heat
affected zone (HAZ) and to the unaffected microstructure of the parent
metal. Both the deposited weld metal and the HAZ must have adequate
strength and toughness after welding.

Slide 13 : Microstructure of martensite (x 500)


For welding, a steel of low hardenability is therefore required. Hardenability
is affected by steel composition, including not only carbon content but other
alloying elements as well. To take all of these factors into account, the
concept of the carbon equivalent value is used. There are a number of ways
of calculating carbon equivalents for use in different circumstances. In the
context of welding:

C.E. =
If the CE is lower than about 0,4%, the steel can be welded with little or no
trouble from martensite and HAZ hydrogen cracking. As indicated before,

the cooling rate is also an important factor, which means that during
welding, thick plates are more susceptible to hydrogen cracking than thin
plates. To reduce susceptibility to martensite formation, the cooling rate
(between 800C and 500C) can be reduced by preheating the plates before
welding.
5. INCLUSIONS

5.1 Sulphur, Phosphorus and Other Impurities


One tonne of steel, a cube with sides of about 0,5m, contains between
1012 and 1015 inclusions which can occupy up to about 1% of the volume.
The total content is largely determined by the origins of the ores, coke and
other materials used to extract the metal in the first place, and by the details
of steelmaking practice.
The principal impurities which worry steelmakers are phosphorus and
sulphur. If not at very low concentrations, these impurities form particles of
phosphide and sulphide which are harmful to the toughness of the steel.
Typically, less than 0,05% of each of these elements is demanded. Low
phosphorus contents are relatively easily attained during the refining of the
pig iron into steel, but sulphur is more difficult to remove. It is controlled by
careful choice of raw materials and, in modern steelmaking, by extra
processing steps to remove it.
Manganese is always added to steels. It has several functions but the
important one in this context is that it combines with the sulphur to form
manganese sulphide (MnS). If the manganese were not present, iron
sulphide would form which is much more harmful than MnS.
Some of the inclusions are too small to be seen with optical microscopes and
must be detected by more elaborate methods. Among this group, which are
mainly equiaxial in shape, are nitrides of aluminium and titanium which are
deliberately introduced in order to inhibit the processes which lead to
coarsening of grain size.
Other inclusions, large enough to be seen readily with the optical
microscope, include entrained particles of slag, deoxidation products and
manganese sulphide. At hot rolling temperatures, these inclusions are plastic
and are elongated in the rolling direction. The result is shown in Figure 1.
The properties of steels containing such inclusions reflect both the volume
of the inclusions and the anisotropy of their shapes, see Figures 17 and 18.

In recent years, a number of practices have been introduced which aim to


reduce the inclusion content in the molten steel before it is cast into ingots.
Sulphur contents of 0,01% or less are now regularly produced. These
processes produce what have become known as 'clean steels'. The
expression is relative. Clean steels still contain many inclusions, but are
significantly tougher than ordinary steels. Inclusion shape control is also
practised in better quality steels. Additions of calcium or cerium and other
rare earth elements to the refined molten steel combine with the sulphur in
preference to the manganese. Sulphides of these elements appear in the final
microstructure as equiaxial particles and are not so deleterious to the
through-thickness ductility of the material as elongated MnS inclusions.
Steels treated in these ways are used in applications where toughness is of
paramount importance and where the extra cost can be justified. Examples
include high integrity pressure vessels, oil and gas pipelines and the main
legs of offshore platforms. The introduction of continuous casting has also
improved the quality of conventional structural steels.
5.2 Manganese in Structural Steels
It has been noted earlier that the residual sulphur impurity in steel is less
harmful when formed into particles of MnS rather than iron sulphide. The
presence of small amounts of manganese in the steel confers several other
benefits. In normalised steels, it tends to increase the amount of
undercooling before the start of the formation of ferrite and pearlite. This
gives finer grained ferrite and more finely divided pearlite. Both of these
changes improve strength and reduce the ductile/brittle transition
temperature. The dissolution of the manganese atoms in the ferrite crystals
also improves the strength of the ferrite. These effects on properties are
summarised in Figures 19 - 21.

If the manganese content is increased too much, its effect ceases to be


beneficial and can become harmful because it increases hardenability, i.e.
promotes martensite formation. It is for this reason that a maximum
manganese content is specified: For S355 in Table 3 of EN 10025 this
maximum is 1,7% by weight, see Figure 16. A convention has also grown
that distinguishes between plain carbon steels, i.e. steels containing <
1%Mn, and carbon manganese steels i.e. >1%Mn.
6. CONCLUDING SUMMARY

Steels used for structural purposes generally contain up to about


0,25%C, up to 1,5%Mn and with carbon equivalents of up to
0,4%. They are mostly used in the hot-rolled, normalised or
controlled-rolled conditions, although low carbon steels might be
used in the cold-rolled and annealed condition. Production
processes aim to produce low inclusion contents and small grain
size to improve strength, ductility, toughness and reduce the
ductile/brittle transition.
The elastic modulus of steel is virtually independent of
composition and treatment.
The upper limits on the proportions of carbon and other alloying
elements are determined by the effect of carbon equivalent on
weldability, and by the effect of carbon on the ductile/brittle
transition temperature. All steels contain manganese, partly to
deal with impurities, such as sulphur, and partly because its
presence has a beneficial effect on the ductile/brittle transition
and strength.
In recent years the development of so-called micro-alloyed steels
or HSLA (high strength low alloy) steels has taken place. These
steels are normalised or controlled rolled carbon-manganese
steels which have been 'adjusted' by micro-alloying to give higher
strength and toughness, combined with ease of welding. Small
additions of aluminium, vanadium, niobium or other elements are
used to help control grain size. Sometimes, about 0,5%
molybdenum is added to refine the lamellar spacing in pearlite
and to distribute the pearlite more evenly as smaller colonies.
These steels are used where the improved properties justify the
extra cost.

Iron is abundant in the universe, found in the sun and many types of stars in
considerable quantity. The core of the earth is thought to be made up of nickel
and iron, and to be hotter than the Sun's surface. This intense heat from the
inner core causes material in the outer core and mantle to move around
(convection currents).

(Note: Funny how we don't really know eh? - We know it gets hotter as you dig
deeper, but we can only guesstimate how hot it is at the center of the earth. Even
at 12km underground the calculations of scientists were out by more than
100%.)

Carbon Steel
Steel is an alloy of iron (Fe) and carbon (C), with 0.2 to 2.04%
carbon by weight. Carbon is the most cost-effective alloying
material for iron, but various other alloying elements are used such
as manganese, chromium, vanadium, and tungsten.
Carbon Steel

ANSI
def'n

General
Def'n

Low carbon
steel

0.05
0.15%

<0.1%

Mild Steel

0.160.29%

Medium
carbon steel

0.30
0.59%

High carbon
steel

0.6
0.99%

Ultra-high
carbon steel
Cast Iron

0.1-0.25%

Applications and properties


Soft, ductile. Easy to form.

Low tensile strength, but it is cheap and m


surface hardness can be increased through
carburizing.

Balances ductility and strength and has go


0.25-0.45% resistance; used for large parts, forging an
automotive components.
0.45-1.0%

Very strong, used for springs and high-stre


wires.

Very hard - knives, punches. Usually anyth


1.0-1.50% 1.2% would require other alloys to prevent
1.02.0%
(>1.5% rare) brittleness. Very high carbon content can b
using powder metallurgy.

2.5-4.0%

Lower melting point, easy casting, lower to


and strength than steel.

Carbon percentages in various steel applications;

Varying the amount of alloying elements and the way they


incorporated into the steel (solute elements, precipitated phase)
influences such properties as hardness, ductility and tensile
strength of the resulting steel. With increased carbon content steel
becomes harder and stronger than iron, but also more brittle. The
maximum solubility of carbon in iron (in austenite region) is 2.14%
by weight, occurring at 1149 C; higher concentrations of carbon or
lower temperatures will produce cementite (very brittle). Add any
more carbon and you get cast iron, which has a lower melting
point and is easier to cast.
Wrought iron containing only a very small amount of other
elements, but contains 13% by weight of slag in the form of
particles elongated in one direction, giving the iron a characteristic
grain. It is more rust-resistant than steel and welds more easily. It
is common today to talk about 'the iron and steel industry' as if it
were a single entity, but historically they were separate products.
Steel has been produced for thousands of years, but it became common after
more efficient production methods were devised in the 17th century. The
Bessemer process in the mid 1800's made steel relatively inexpensive for massproduced goods. Further refinements in the process, such as basic oxygen
steelmaking, further lowered the cost of production while increasing the quality of
the metal. Today, steel is one of the most common materials in the world and is a
major component in buildings, tools, automobiles, and appliances.

Get pdf: XLER_International_Compare.pdf

VIDEO: Properties and Grain Structure. BBC 1973


Don't laugh at the date - this video beats all those pathetic modern ones that give you a fancy intro but
nothing more than a talking head. They never venture out of the studio. This old video is fabulous for a
clear introduction to steel grain structure.

Part 1: What is a grain? (Video 11MB)


1. The patches seen on a galvanised object
are crystals or grains of zinc.
2. All metals are made up of grains, but they are usually
invisible (too small to see or same shine/colour).
3. Etching Process: Mirror finish, powerful acid, washed and
sealed.
4. In a pure metal, the grains are different colours because of
the way they reflect the light.
5. Tiny crystals grow outward until they meet. Each fully grown
crystal is called a grain.
Part 2: Recrystallisation (Video 13MB)
1. Before cold working the grains are similar size and shape
2. Cold working elongates the grains, increases hardness and
strength increases, reduces ductility.
3. At 350C, new grains form in the Al to replace old grains.
Called recrystallisation
4. Recrystallisation softens, lowers the strength, ductility
increased
5. Excessive recrystallisation temp gives poor mechanical
properties
Part 3: Heat Treatment of Steel (Video 23MB)
1. Steel grains are too small to be visible - need a microscope
approx 250 times magnification.
2. Ferrite: Light coloured. Made of iron. Ductility to the steel
3. Pearlite: darker coloured. Layers of Iron + Iron Carbide.
Hardess and strength to the steel

4. 100% Pearlite is about 0.8%C. Pearlite, recrystallisation


temperature 720C.
5. Normalising - cooled in air, reduced grain size and more
uniform shape, toughness increased
6. Quenching - increases hardness. Not enough time for pearlite
to form, so a needlelike strcture forms - martensite. Very
hard and brittle.
7. Tempering - (after quenching) restores toughness. Modifies
the martensite needles with small flakes of carbon. This gives
hardness AND toughness.
8. 0.1%C steel (Mild Steel). Recrystalisation 900C. Not enough
carbon to produce martensite.

Iron-Carbon Equilibrium Diagram


Excellent link (Cambridge
University): http://www.msm.cam.ac.uk/phasetrans/2008/Steel_Microstructure/SM.html
An Equilibrium diagram is a graph of the different structural
arrangments that occur within a range of an alloying element.

This diagram shows how iron and carbon combines IF it is cooled


slowly (in equilibrium). Under 2% is steel, over 2% is heading into
the cast iron range where carbon tends to coagulate (clump
together). Cementite Fe3C has 6.67%C and it is basically a
ceramic. The eutectoid (pearlite) at E has 0.83% C, less carbon is
a hypoeutectoid steel (A), and greater is hypereutectoid (B). Alpha
iron (ferrite), gamma iron (austenite, which only exists at high
temperature), and delta iron (another high temp structure).
Two very important phase changes take place at 0.83%C and at
4.3% C. At 0.83%C and 723C, the transformation is eutectoid,
called pearlite. These 2 phases separate out in layers. From
gamma (austenite) --> alpha + Fe3C (cementite)
At 4.3% C and 1130C, the transformation is eutectic, called
ledeburite. L(liquid) --> gamma (austenite) + Fe3C (cementite).

This is cast iron.


BTW. Since carbon is much lighter than Fe, the actual atomic
% Carbon (by counting atoms) is actually about 4 times higher than
the %C by weight. So it's not quite so amazing now is it? I mean
like how 0.5% Carbon can totally transform soft iron... it's really
about 2% if you count atoms - not mass.

Summary of Fe-C structures (grains)

Austenite (-iron; hard). Only exists above 723C, which is


when the FCC -iron structure occurs. Can dissolve up to
2.1%C by mass. Non-magnetic, soft (hance we have hotworking)

Cementite (Iron carbide Fe3C, 6.67%C by mass. There are


twelve iron atoms and four carbon atoms per unit cell, so
33% carbon atoms). Very hard and brittle.

Ledeburite (The Ferrite-Cementite eutectic, 4.3% carbon.)

Ferrite (-iron, -iron; soft). No carbon, BCC. Soft and ductile.

Pearlite (88% ferrite, 12% cementite, which is 0.83%C)

Martensite. Occurs when cooling is too rapid to form pearlite,


so it locks carbon into the grain. Quench-hardening of steel.

Micrographs (photos from a microscope).


(A) = 0.1%C ferrite/pearlite, (B) = 0.25%C more pearlite, (C) =
0.83%C all pearlite, (D) = 1.4%C pearlite/cementite

Close-up view of pearlite showing layers of ferrite (white) and


cementite (dark).

Large FC Equilibrium Diagram

Large Print Version 2000x2658px

Slip
When a piece of metal is deformed, it is the grains that are deformed. A grain is a
crystal, an orderly arrangement of atoms in a grid. If atoms are stretched apart
this is elastic deformation because the atoms are held together by electron
attractions - which acts like a spring. However, permanent (or plastic)
deformation means the atoms actually slip past each other in layers or planes.

Real crystals do not slip in a whole plane at once. This would take a very high
force. Instead, the imperfections in the crystal allow slip to travel one atom at a
time. The wider the range of effected atoms, the more ductile (easily slipping) the
grain is. Here is an example of an imperfection called a dislocation which can
travel easily through the crystal.

Here is an actual example of slip. (we are not just making this up!)

A scanning electron micrograph of a single crystal of cadmium deforming by


dislocation slip on 100 planes, forming steps
on the surface.
The following animation shows a lattice of atoms (such as in a metal). There are
only 2 ways to distort the atoms - axial (tension and compression) and shear
(sideways).
This animation shows only the elastic portion of the stress/strain curve, where no
atomic slip occurs.
Further information here:
http://www3.nd.edu/~manufact/MPEM_pdf_files/Ch03.pdf

Hardening of Steel
There are 2 ways to make the grain stronger. Get rid of
all imperfections (pretty impossible task, although this is why very
fine fibres can give crazy strengths), or stop the slip from travelling
all the way through the grains. Carbon and other elements act as a
hardening agent, preventing dislocations in the iron atom crystal
lattice from sliding past one another. Carbon is tiny compared to Fe
- no wonder it messes up the slip so much.
Quenching of steel is not shown on the Fe-C equilibrium diagram, because
quenching is not in equilibrium! (i.e. cooling is too fast for the carbon to get itself
in the right structure - typically pearlite)
After quenching, tempering is used to add toughness to the steel. Tempering
must be at a temperature below recrystallisation. An oven is best for tempering,
but it can be done by flame by judging the colour of the steel. Temper colours
can be used as a guide to temperature. Alloys such as stainless steel form thinner
films than do carbon steels for a given temperature and hence produce a colour
lower in the series. For example, pale straw corresponds to 300C for SS, instead
of 230C for CS. The colours colder than the reds (below 500C) are actually
discolouration of oxides, not the actual radiation glow of the temperature itself.
(Which would be infra-red, and invisible)

Radiation Colour

Celcius

Farenht

Tempering Applicatio

Yellow-White

1539C

2800F

Highest melting point (0%C pure

Bright Yellow

1130C

2066F

Lowest melting point (4%C cast

Yellow

1093C

2000F

Copper melts at 1084C, Gold 10

Dark yellow

1038C

1900F

Orange yellow

982C

1800F

Orange

927C

1700F

Orange red

871C

1600F

Bright red

816C

1500F

Red

760C

1400F

Medium red

704C

1300F

Dull red

649C

1200F

Aluminium melts 600-660C

Slight red

593C

1100F

Toughening for constructional st

Very slight red, mostly grey

538C

1000F

Toughening for constructional st

Dark grey

427C

0800F

Toughening for constructional st


change 410

Oxidation Colour

Celcius

Farenht

Blue

302C

0575F

Saws for wood, springs

Dark Purple

282C

0540F

Cold chisels, setts for steel

Purple

271C

0520F

Press tools, axes

Brass melts 930C

Steel recrystalisation temperatur

Tempering Applic

Brown/Purple

260C

0500F

Punches, cups, snaps, twist drills

Brown

249C

0480F

Taps, shear blades for metals

Dark Straw

241C

0465F

Milling cutters, drills

Light Straw

229C

0445F

Planing and slotting tools

Faint Straw

199C

0390F

Quenching:
Speed of quenching: Higher carbon steels can be quenched more slowly, but a
lower C steel will need to be rapidly quenched to have any hardening effect.
Quenching Rate: (FASTEST) Salt water > water > oil > air. (SLOWEST)
For a complex and expensive job, it is better to have a slow quenching alloy
because it is less sensitive to variations of temperature. This is why most tool
steels for things like injection moulding tools are OIL quenched. Water quenching
is fine for simple shapes that can be controlled more easily.
Induction Hardening. Induction Hardening where electric induction (rapid
magnetic changes) heat up the steel which is quickly followed by quenching in
water jet. Alternative to flame or oven.

Induction Hardening. http://www.thermobondflame.com/Services.page?i=4

How to harden the surface: CASE Hardening.


Heat treat = quench > Martensite (arrests slip).
Heat outside surface > quench. Local flame or induction quenched with water
(gears).
Carbon penetrating the outside surface > quench. Carburizing (Heat up in carbon
packing). Heated solutions.

Flame Hardening of steel roll.: http://www.thermobondflame.com/Services.page?i=2

Alloy Steels
Effects of alloying elements on tool steel properties:

Carbon: Raising carbon content increases hardness slightly


and wear resistance considerably. Dramatic increases to
hardness & strength when heat treated.

Manganese: Small amounts of of Manganense reduce


brittleness and improve forgeability. Larger amounts of
manganese improve hardenability, permit oil quenching (less
severe queching required - which reduces quenching
deformation).

Silicon: Improves strength, toughness, and shock resistance.

Tungsten: Improves "hot hardness" - used in high-speed tool


steel. Very dense (heavy)

Vanadium: Refines carbide structure and improves


forgeability, also improving hardness and wear resistance.

Molybdenum: Improves deep hardening, toughness, and in


larger amounts, "hot hardness". Used in high speed tool steel
because it's cheaper than tungsten.

Chromium: Improves hardenability, wear resistance and


toughness.

Nickel: Improves toughness and wear resistance to a lesser


degree.

Including these elements in varying combinations can act synergistically,


increasing the effects of using them alone. (For example, cetain alloying elements
can allow more carbon, where so much carbon would be unworkable in a plain
carbon steel). Another example is the interesting way stainless steel (Chrome and
Nickel added to Iron) is quite corrosion resistant.

Steels Identification Codes


AISI-SAE Coding system (American Iraon and Steel Institute - Society of
Automotive Engineers). A 4 digit code, the first 2 digits give the general steel
type, and the last 2 digits are the % Carbon x 100. For example, 1010 is plain
carbon srteel with 0.10% C, 5120 is a chromium steel with 0.20% C. More
detail here

American Steel codes: From Higgins: Materials for Engineers aand Technicians 5th Ed. 2010. p21

The BSA (British Standards Association) uses a 6 digit code. The digits separated
into 3 groups as shown below. For example, a steel with the code 070M20 would
be 070 = carbon or carbon-manganese steel, M = mechanical property
specification, 20 = carbon content 0.20%.

British Steel codes: From Higgins: Materials for Engineers aand Technicians 5th Ed. 2010. p20

The UNS number (short for "Unified Numbering System for Metals and Alloys") is
a systematic scheme in which each metal is designated by a letter followed by
five numbers. It is a composition-based system of commercial materials and does
not guarantee any performance specifications or exact composition with impurity
limits. Other nomenclature systems have been incorporated into the UNS
numbering system to minimize confusion. For example, Aluminum 6061 (AA6061)
becomes UNS A96061. Below is an overview of the UNS system, with special
emphasis on common commercial alloys. As with any system, there are
ambiguities such as the distinction between a nickel-based superalloy and a highnickel stainless steel.
Axxxxx - Aluminum Alloys
Cxxxxx - Copper Alloys, including Brass and Bronze
Fxxxxx - Iron, including Ductile Irons and Cast Irons
Gxxxxx - Carbon and Alloy Steels
Hxxxxx - Steels - AISI H Steels
Jxxxxx - Steels - Cast
Kxxxxx - Steels, including Maraging, Stainless Steel, HSLA, Iron-Base
Superalloys
L5xxxx - Lead Alloys, including Babbit Alloys and Solder Alloys
M1xxxx - Magnesium Alloys
Nxxxxx - Nickel Alloys

Rxxxxx - Refractory Alloys R03xxx- Molybdenum Alloys R04xxx- Niobium


(Columbium) Alloys R05xxx- Tantalum Alloys R3xxxx- Cobalt Alloys R5xxxxTitanium Alloys R6xxxx- Zirconium Alloys
Sxxxxx - Stainless Steels, including Precipitation Hardening Stainless Steel and
Iron-Based Superalloys
Txxxxx - Tool Steels
Zxxxxx - Zinc Alloys

Tool steels
Tool steels are covered in Australian Standard AS1239 and is
virtually the same as the American AISI tool steel
classification.(Similarly with British Standard 4659)
For example: AS 1239 grade H13 hot work tool steel containing
0.35% carbon, 5.0% chromium, 1.5% molybdenum and 1%
vanadium would be written as X40CrMoV51 in DIN (German).
High Speed Steels, for example: AS 1239 grade M2 Containing
0.85% carbon, 4.0% chromium, 5.0% molybdenum, 6.0%
tungsten, 2.0% vanadium would be written as S 6-5-2 in DIN.

Steels Selector

Large Size (400kB): steel_types_large.jpg


Printable Size (1.7MB): steel_types_fullsize.jpg
Common steel grades in Australia (Edcon)

Cast Iron
When too much carbon is added to steel, the carbon cannot dissolve into the
solution and creates a totally different structure. From the Fe-C diagram we saw
earlier, Cast Iron forms in the range of 2% to 7% carbon (by weight).

There are many types of cast iron, but Grey Cast Iron is the most familiar, often
used for machine tool bases. It is useful and popular for several reasons.
Firstly, the melting temperature is lower, which makes it easier to cast. This is
because the eutectic is at 4.3% C, giving a melting point of only 1147oC. This
eutectic produces a new grain called ledeburite, which is a mixture of austenite
and cementite. (Remember Pearlite? It was a eutectoid and made of layers of
ferrite and cementite). But since a eutectoid is a low point in the liquid-solid
transition, it is the melting point.
Secondly, Grey Cast Iron is great for machine bases. Normally, so much carbon
would be a nightmare of brittleness due to extreme martensite. But it turns out
that with the right cooling, excess carbon forms flakes of graphite. This is
completely different to all these Fe-C grains we've been talking about - like
Ferrite and Cementite and Pearlite and Ledeburite. Instead, graphite is like an
inclusion in the metal, and it gives Grey Cast Iron the damping properties suitable
for machine bases. It is a material with low tensile strength however, so GCI is
usually used where it is in compression. GCI is prone to hardening due to excess
heat however, so it is not easy to weld. More often it would be brazed.

Photo micrograph of Grey Cast Iron showing graphite flakes in a ferrite


matrix. Source

Comparative qualities of cast irons (Wikipedia)

Nominal
composi
Name tion [%
by
weight]

Yield
stren
Tensil Elongat Hardn
Form
gth
e
ion [%
ess
and
[ksi
stren
(in
[Brinel
conditi (0.2
gth 2 inche
l
on
%
[ksi]
s)]
scale]
offset
)]

Uses

Cast

50

0.5

260

Engine
cylinder
blocks,
flywheel
s, gears,
machine
-tool
bases

White C 3.4,
cast Si 0.7,
iron Mn 0.6

Cast
(as
cast)

25

450

Bearing
surfaces

Mallea C 2.5,

Cast

33

52

12

130

Axle

Grey
cast
C 3.4,
iron Si 1.8,
(ASTM
Mn 0.5
A48)

ble Si 1.0,
iron Mn 0.55
(ASTM
A47)

C 3.4,
Ductil
P 0.1,
e or
Mn 0.4,
nodula
Ni 1.0,
r iron
Mg 0.06
C 2.7,
Ni- Si 0.6,
hard Mn 0.5,
type 2 Ni 4.5,
Cr 2.0

(anneal
ed)

Cast

Sandcast

bearings
, track
wheels,
automot
ive
cranksh
afts

53

70

55

18

170

Gears,
camshaf
ts,
cranksh
afts

550

High
strength
applicati
ons

Glossary
1.

Alloy: A metallic substance that is composed of two or more


elements.

2. Austenite: Face-centered cubic iron or an iron alloy based on


this structure.
3. Bainite: The product of the final transformation of austenite
decomposition.
4. Body-centered: A structure in which every atom is surrounded
by eight adjacent atoms, whether the atom is located at a
corner or at the center of a unit cell.
5. Cementite: The second phase formed when carbon is in
excess of the solubility limit.
6. Critical point: Point where the densities of liquid and vapor
become equal and the interface between the two vanishes.
Above this point, only one phase can exist.
7. Delta iron: The body-centered cubic phase which results when
austenite is no longer the most stable form of iron. Exists

between 2802 and 2552 degrees F, has BCC lattice structure


and is magnetic.
8. Eutectic: A eutectic system occurs when a liquid phase
tramsforms directly to a two-phase solid.
9. Eutectoid: A eutectoid system occurs when a single-phase
solid transforms directly to a two-phase solid.
10.
Face-centered: A structure in which there is an atom at
the corner of each unit cell and one in the center of each face,
but no atom in the center of the cube.
11.
Ferrite: Body-centered cubic iron or an iron alloy based
on this structure.
12.
Fine pearlite:Results from thin lamellae when cooling
rates are accelerated and diffusion is limited to shorter
distances.
13.
Hypereutectoid: Hypereutectoid systems exist below the
eutectoid temperature.
14.
Hypoeutectoid: Hypoeutectoid systems exist above the
eutectoid temperature.
15.

Ledeburite: Eutectic of cast iron. It exists when the

carbon content is greater than 2 percent. It contains 4.3


percent carbon in combination with iron.
16.
Liquidus Line: On a binary phase diagram, that line or
boundary separating liquid and liquid + solid phase regions.
For an alloy, the liquidus temperature is that temperature at
which a solid phase first forms under conditions of equilibrium
cooling.
17.
Martensite: An unstable polymorphic phase of iron
which forms at temperatures below the eutectoid because the
face-centered cubic structure of austenite becomes unstable.
It changes spontaneously to a body-centered structure by
shearing action, not diffusion.
18.
Microstructure: Structure of the phases in a material.
Can only be seen with an optical or electron mircoscope.
19.
Pearlite: A lamellar mixture of ferrite and carbide
formed by decomposing austenite of eutectoid composition.
20.
Phase: A homogeneous portion of a system that has
uniform physical and chemical characteristics.

21.

Phase diagram: A graphical representation of the

relationships between environmental constraints, composition,


and regions of phase stability, ordinarily under conditions of
equilibrium.
22.
Polymorphic: The ability of a solid material to exist in
more than one form or crystal structure.
23.

Quench: To rapidly cool

24.
Solidus Line: On a phase diagram, the locus of points at
which solidification is complete upon equilibrium cooling, or at
which melting begins upon equilibrium heating.
25.
Solubility: The amount of substance that will dissolve in
a given amount of another substance.

DVDs:
Assignment:
Heat Treat

Вам также может понравиться