Handbook No.1 - Metallography PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 125

METALLOGRAPHY

This chapter presents a selection of metallographic concepts


and methods essential to the understanding of basic processes
in powder metallurg y in par ticular iron powder
metallurgy. The somewhat shorthanded presentation of some
items may stimulate the reader to collect further information
from pertaining special literature.

1. METALLOGRAPHY

TABLE OF CONTENTS

1.1 THE NATURAL STATES OF MATTER ................................... 3


1.2 THE CRYSTALLINE STRUCTURE OF METALS ...................... 4
1.3 DIFFUSION IN METALS.................................................... 19
1.4 BINARY PHASE DIAGRAMS OF METALS ............................ 27
1.5 THE IRON-CARBON SYSTEM............................................ 40
1.6 TRANSFORMATION DIAGRAMS OF STEELS....................... 52
1.7 INFLUENCE OF THE MICROSTRUCTURE ON THE
PROPERTIES OF STEEL .................................................. 62
REFERENCES

1.1. THE NATURAL STATES OF MATTER

1.1 The natural states of matter


All matter surrounding us consists of atoms, ions and molecules, which, depending on
their mutual distances, are under the inuence of stronger or weaker reciprocal forces.
Depending on the strength of their reciprocal forces, these particles form gases, liquids or
solids.
In gases. The particles (i.e. atoms, ions or molecules) move about highly irregularly at
very high speeds (some 100 m s-1 at R.T.). Their free length of way, , (i.e. the distance a
particle can move before colliding with another one) is large compared to the particle
diameter. Their density is in the order of 1019 / cm3. This situation is called random
order. See Fig. 1.1 a.
In liquids. The particle density is in the order of 1022 / cm3. The particles move about
rather irregularly but are more crowded than in the gaseous state, each particle having
approximately ten very close neighbors. Signicant for liquids is the frequent local
formation of crystal-like agglomerates extending over several particle diameters. Because
of the high thermal energy of the particles, these agglomerates are very short-lived and
there is no correlation between more distant particles. This situation is called instable
short-range order. See Fig. 1.1 b. In contrast to gases and liquids, amorphous and
crystalline materials exhibit, within a certain range of temperatures, a considerable
rigidity they are solids.
In amorphous solids. The particle density is of the same order of magnitude as in
liquids, i.e. 1022/cm3. Similar to liquids, a short-range order exists between particles. But
in contrast to liquids, due to the heavily restricted mobility of the particles in amorphous
solids, this order is stable and, logically, it is called stable short-range order. See Fig. 1.1 c.
Amorphous materials have no sharply dened melting point. The transition from their
solid to their liquid state occurs gradually. They can, in fact, be looked at as liquids with
an extremely high viscosity at room-temperature. Classic examples are glasses. But today,
by means of special processes, also metals can be transformed into an amorphous state.
In crystalline solids. Such as metals, the particle density is in the order of 1023 /cm3 ,
i.e. ten times higher than in liquids and amorphous materials. The particles are
homogeneously distributed and form a strict geometrical structure. See Fig. 1.1 d. Their
mobility is restricted to very small vibrations about xed equilibrium sites. These
vibrations increase with temperature. This situation is called crystalline long-range order.
The crystalline state possesses the lowest internal energy of all states of matter and,
therefore, is stable below a sharply dened melting point. The above mentioned states
are not specic to any particular kind of material. In principle, but not always in
practice, any type of material can adopt any of these states.

1-3

1. METALLOGRAPHY

Figure 1.1 Distribution of atoms in gases, liquids, amorphous and crystalline solids (schematically);
a. Gas: statistical distribution, Number of atoms per cm3 = 1019, = free length of way.
b. Liquid: instable close-range order, Number of atoms per cm3 = 1022.
c. Amorphous solid: stable close-range order, Number of atoms per cm3 = 1022.
d. Crystal: crystalline order, Number of atoms per cm3 = 1023.

1.2 The crystalline structure of metals


Our further considerations are focused mainly on the crystalline in particular the
metallic state of matter. The formation of crystalline structures is controlled by certain
binding-forces between the atoms. Geometrical descriptions of crystalline structures
refer to so-called ideal crystals only and do not take into account disturbances occurring
in real crystals, such as surfaces, grain-boundaries, dislocations, vacancies, foreign atoms
and other irregularities. In the following paragraph, we will discuss;
1) various types of bonding between atoms,
2) structures of ideal crystals,
3) stacking faults and disturbances in real metal crystals, and
4) plastic deformation in metal crystals.

1-4

1.2. THE CRYSTALLINE STRUCTURE OF METALS

1.2.1 Bonding between atoms


Crystal structures owe their existence to various types of interatomic forces between
neighboring atoms. Four characteristic types are recognized: Van der Waals, covalent, ionic
and metallic.
Van der Waals force. This is the only appreciable force exerted between well separated
atoms and molecules. It is a weak attractive force that acts between all atoms and is
responsible e.g. for the condensation of noble gases and chemically saturated molecules
to liquids and solids at low temperatures. Its explanation requires quantum mechanics
because it involves uctuations of the electronic charge in the atom.
When two atoms approach fairly closely these uctuations can occur in unison so
that one atom has its electrons slightly nearer the other nucleus whenever this nucleus
happens, through the movements of its own electrons, to be more exposed in this
direction than usual.
Ionic bond. Neighboring atoms exchange electrons and the so formed positive and
negative ions are pulled together electrostatically as, for example, in NaCl-crystals.
Covalent bond. Partly empty electron shells of neighboring atoms overlap so that their
electrons belong to both atoms. This leads to strong bonds between the atoms because in
the overlapping shells their electrons are in a lower state of energy than in separate shells.
One important feature of these bonds is that they can exist between atoms of the same
type between which there can be no ionic bonding. Examples are H2 ,O2 and Cl2. One
other important feature of these bonds is that they act in preferential directions as, for
example, in the tetrahedral structures of the CH4-molecule and of diamond.
Metallic bond. This type of bond can be explained only on the basis of the following
quantum-mechanical principle. All atoms in the metal crystal form one common band
of electron shells to which each individual atom contributes one electron. These
electrons, not being bonded to individual atoms, can move freely inside the entire metal
crystal. They form a so-called electron gas.
According to an obsolete electrostatic theory, the negatively charged electron gas
presses the negatively charged metal ions closely together. But the correct explanation is
that the mentioned arrangement constitutes the lowest possible quantum state of energy
for the metal crystal and allows the atoms to be packed in the closest possible way, To
this kind of close packing, metals owe their good plastic formability. The freely moving
electrons are responsible for the high electrical and thermal conductivity of metals and
for the characteristic metallic gloss.
The overwhelming majority of all chemical elements crystallize in metal structures,
approximately one third of them each in BCC-, in FCC- and in HCP-structure.

1-5

1. METALLOGRAPHY

See Table 1.1. The geometrical characteristics of these particular structures are described
in the following paragraph.
Table 1.1. Crystal structures and lattice constants of some metals. [T.1.1]
Structure

Metal

BCC

-Fe
-Fe

1390

3,147

3,165
900

Ni

3,306
3,524

467

Cu

3,560
3,619

1000

3,651

Ir

3,839

Pt

3,923

Al

4,049

Au

4,078

Ag

4,086

Pb

4,950

-Co

2,506; 4,068

1,62

Zn

2,665; 4,947

1,86

-Ti

2,950; 4,679

1,59

Cd

2,979; 5,617

1,89

Mg

3,209; 5,210

1,62

(ideal)

1-6

2,932

Mo

-Fe

Ratio of Lattice
Constants c/a

2,86

2,884

-Co

HCP

Lattice Constants
a) 10-10 m; c) 10-10m

Cr

-Ti
FCC

Temperature C

1,633

1.2. THE CRYSTALLINE STRUCTURE OF METALS

Figure. 1.2. Unit cell in a crystal cell; lattice


constants: a, b, c.

1.2.2 Ideal crystals


A crystal structure has three general properties: periodicity, directionality and completeness.
Periodicity is the regular repetition in space of the atomic unit of the crystal (wallpaper
pattern). It is the basis of the remarkable plastic properties of crystals. Directionality of
the crystal structure appears in the fact that properties like conductivity and elasticity
vary with the direction of their measurement. Completeness is simply the lling of all
crystal sites dened by the periodic structure with the required atoms.
In crystallography, it is customary to present a crystal structure graphically as a space
lattice within which the sites of atoms are marked by points (or small circles). These
points are referred to as lattice points. The basic unit of the crystal structure which repeats
itself in all three directions of space is called unit cell. The dening edges of this cell are
referred to as lattice constants, (usually a, b, c). See Fig. 1.2.
The lattice points lie at the intersections of three families of parallel planes, called lattice
planes. The distance between neighboring lattice planes is usually designated by the
triple-indexed symbol d h k l. See Fig. 1.3.
The spatial orientation of any lattice plane is clearly dened by its so-called Miller indices
(h k l), which are determined in the following way:
Choose a coordinate system with the origin at the corner of one unit cell and with its
axes parallel to the edges a, b, c of this unit cell.
Determine the intersections of the plane with the tree axes, x y z, of this coordinate
system.
1-7

1. METALLOGRAPHY

LP
LP

LP-family

LP
LP-distance

Figure. 1.3. Lattice planes (LP) in


a crystal lattice.

LP-distance
LP

1-8

LP

LP

Express, in units of a, b, c, the distances of these intersections from the origin of the
coordinate system: ma, nb, pc.
Take the reciprocals 1/m, 1/n, 1/p and nd three natural numbers related to one
another in the same way as the reciprocals: 1/m : 1/n : 1/p = h : k : l. Put the result
between round brackets: (h k l).
Example: The plane, shown at Fig. 1.4, intersects the coordinate axes at 2a, 4b and
1c respectively. The above procedure applied to this plane yields m = 2, n = 4, p = 1.
Thus, it follows: 1/2 : 1/4 : 1 = 2 : 1 : 4, and the Miller indices of this plane are:
(h k l) = (214).

1.2. THE CRYSTALLINE STRUCTURE OF METALS

Figure. 1.4. Derivation of


Miller indices.

Lattice planes and their Miller indices play an important roll e.g. in X-ray structural
analysis. Each family of planes is responsible for specic X-ray reections from which the
respective Miller indices can be calculated and conclusions can be drawn as to the type of
crystal structure.
Some signicant lattice planes of the cubic crystal system and their respective Miller
indices are shown schematically at Fig. 1.5.

Figure. 1.5. Miller indices for some important lattice planes in cubic crystals.

Not only lattice planes but also lattice directions can be described by Miller indices. A
lattice direction is dened as the direction of a local vector r = u a + v b + w c, originating
from the corner of a unit cell, a, b, c being the edge vectors of the cell (unit vectors of the
coordinate system x, y, z).
The statement of the three coordinates of the vector in square brackets [uvw]
denitely identies a given lattice direction. In the cubic system (and only there), a

1-9

1. METALLOGRAPHY

lattice direction [uvw] is always perpendicular to the lattice plane (hkl) of same indices
(h = u, k = v, l = w). The coordinates of the local vector r, put in double square brackets
[[uvw]], specify a lattice point. Some lattice directions and lattice points are indicated in
the schematic drawing at Fig. 1.6.

Figure. 1.6. Lattice points and lattice directions in the


cubic crystal system.

For the characterization of lattice planes, directions and points in hexagonal crystal
systems, four instead of three Miller indices are required. Their detailed description can
be found in pertaining special literature and has no bearing on our further discussions.
As already mentioned, all metals crystallize either in a body-centered-cubic (BCC),
in a face-centered-cubic (FCC) or in a hexagonal close-packed (HCP) structure. The
respective unit cells of these structures are shown at Fig. 1.7.
The dice-shaped unit cells of the BCC- and of the FCC-structure are each presented
in two different ways: to the left as abstract point lattice, and to the right as
conglomerate of spheres. The sphere-model conveys a fairly realistic picture of the
packing structure of the atoms inside the crystal.

1-10

1.2. THE CRYSTALLINE STRUCTURE OF METALS

BCC

FCC

HCP

FCC

Figure. 1.7. Unit cells of some important crystal structures: body-centered cubic (BCC); face centered cubic
(FCC); and hexagonal close-packed (HCP). NA = number of atoms/elementary cell;
NN = number of nearest neighbors; PD = packing density.

From the unit cells presented at Fig. 1.7, three signicant parameters can be derived, viz.
the number of atoms per unit cell, NA, the number of nearest neighbors, NN, and the
packing density PD. The number of atoms, NA, is easily obtained in the following way:

In the case of the BCC-cell, putting together an eighth of an


atom from each corner of the cell, and adding the one whole atom from the center
yields 2 whole atoms per unit cell: NA = 2.
1-11

1. METALLOGRAPHY

In the case of the FCC-cell, putting together an eighth of an


atom from each corner of the cell, and putting together to 3 whole atoms the 6
halves of an atom from the faces of the cell, yields 4 whole atoms per cell: NA = 4.
In the case of the HCP-cell, putting together the 12 sixth of an
atom from the corners of the cell, putting together to 1 whole atom the two halves of
one atom from the hexagonal faces, and adding the three atoms from inside the cell
yields 6 atoms per cell: NA = 6.

Obviously, for the BCC-cell, the number of nearest neighbors is NN = 8. The number of
nearest neighbors for the FCC- and for the HCP-cell becomes evident when looking at
the plane of close-packed atoms (shaded), within which each atom has 6 nearest
neighbors. In addition, each atom has 3 nearest neighbors each in the two planes above
and beneath the shaded plane. All together, they add up to 12 atoms: NN = 12.
Both, in the FCC- and in the HCP-structure, atoms are packed at maximal possible
density, viz. PD 74 %. The packing density of the atoms in the BCC-structure is only
PD 68 %. The reader can easily verify these two gures by dividing the volume of
atoms belonging to one unit cell (NA/4 r3/3) with the volume of this cell (a3 ).
At Fig. 1.7, for comparison, fully close-packed lattice planes in the HCP- and in
the FCC-structure are marked by shading. The diagram at Fig. 1.8 (a) shows a fully
close-packed plane. The spheres or atoms lie in three sets of close-packed lines, physically
equivalent and symmetrically orientated to one another. Both, the HCP- and the
FCC-crystal structure are formed by stacking a number of close-packed planes on one
another in a certain stacking sequence.

1-12

1.2. THE CRYSTALLINE STRUCTURE OF METALS

FCCH

HCP

CC-Twinn

Figure. 1.8. Arrangement of atoms in a close-packed plane (a), and in successive close-packed planes (b) and
(c). Below: stacking sequences of close-packed planes in FCC and HCP crystals.

The diagram at Fig. 1.8 (b) shows an element of the stacking sequence, a layer B laid on
a similar layer A in the most closely packed way. The layers have the same orientation
and each B atom rests symmetrically in a hollow provided by three adjoining A atoms.
We notice that only one-half of the hollows are used by B atoms and that an equally
close-packed arrangement would result if we had used the other hollows, i.e. the C
positions in diagram (c), instead.
Any stacking sequence chosen from the A, B and C positions is fully close-packed
provided positions of the same type are not used by neighboring planes. For example,
ABACBABAC is close-packed but ABBAAACCBBC is not.
1-13

1. METALLOGRAPHY

The stacking sequence ABABAB forms the HCP-structure, and ACBACBACB


forms the FCC-structure. In FCC-structures, the stacking sequence occasionally forms
mirror images of itself like ABC(A)CBA(C)ABC.
This kind of stacking faults is called twinning. Crystal twins occur frequently in
copper, nickel and austenitic steels and appear on micrographs as parallel stripes inside
the crystal grains.
1.2.3 Real Metal Crystals
All metals used in technical applications (apart from single crystals and whiskers) have a
polycrystalline structure, i.e. they are composed of many randomly oriented small crystal
grains agglomerated to one another. When cooling down a molten metal below its
melting point, at rst, very small dendrites (crystal nuclei) precipitate from the melt.
These dendrites gradually develop into small crystallites which randomly oriented oat
around in the melt. Independently of one another, the oating crystallites continue to
grow, gradually depriving the melt of atoms. Eventually, they get so crowded in the
diminishing melt that they begin to agglomerate. See Fig. 1.9.
Because of their original random orientation, a mist between crystal lattices occurs
at the boundaries of adjoining crystallites (or crystal grains). In these grain boundaries,
atoms are not quite so closely packed as in the undisturbed lattice inside the crystal
grains. On a polished and etched metallographic section, grain boundaries usually
appear as thin dark lines.
Apart from grain boundaries, there are several other kinds of lattice disturbances in
real crystals. One distinguishes between zero-, one-, two-, and three-dimensional lattice
disturbances.
Figure. 1.9. Formation of a polycrystalline
structure during solidication of a liquid metal
(schematically);
left: nucleation period (dendrites oating in the
melt);
right: end of solidication ( agglomerated crystallites).

Zero-dimensional lattice disturbances are


1) missing atoms in the regular lattice structure, (so-called vacancies),
2) foreign atoms occupying interstices between host atoms (interstitial atoms, and
3) larger or
4) smaller foreign atoms substituting atoms of the host lattice (substitutional atoms).
See Fig. 1.10.
1-14

1.2. THE CRYSTALLINE STRUCTURE OF METALS

Figure. 1.10. Zero-dimensional


lattice disturbances:
(1-1) vacancy,
(2-2) interstitial atom,
(3-3) and,
(4-4) bigger and smaller substitutional atoms.

One-dimensional lattice disturbances are so-called dislocations which occur in two


different varieties, edge dislocations and screw dislocations, forming zones of lower
packing-density which like laments stretch through the crystal structure. The schematic
drawings at Fig. 1.11 shows one edge dislocation (a) and one screw dislocation (b).

Figure. 1.11. One-dimensional lattice disturbance: (a) edge dislocation, (b) screw dislocation; in both cases,
the dislocation line is perpendicular to the drawing plane.

Two-dimensional lattice disturbances are the above mentioned grain boundaries, which
can also be imagined as arrays of edge dislocations stacked on top of one another.
See Fig. 1.12.

1-15

1. METALLOGRAPHY

Figure. 1.12. Two-dimensional lattice disturbance: a grain boundary (composed of an array


of dislocation lines stacked on top of each other)
the grain boundary face is perpendicular to the
drawing plane.

Three-dimensional lattice disturbances are non-metallic inclusions, dislocation tangles,


and small pores as occur in certain diffusion processes (Kirkendall-effect) and in
sintering processes.
1.2.4 Plastic Deformation of Metal Crystals
Dislocations play the most essential part in plastic deformation of metals. Under the
inuence of shearing-stresses, dislocations are created, and slip occurs on certain lattice
planes (mainly (111)-planes). On these slip planes, dislocation lines move like waves
along energetically favorable directions (mainly [111]-directions).
This kind of slipping is facilitated by the circumstance that bonds between adjacent
atoms on both sides of the slip plane are not broken all at the same time but successively
one after the other. See Fig. 1.13.

1-16

1.2. THE CRYSTALLINE STRUCTURE OF METALS

Figure. 1.13. Atomic bonds transiently being ruptured as dislocation


travels through crystal under the
inuence of shearing stresses.

Along a dislocation line perpendicular to the direction of slip, bonds are temporarily
suspended. As shown schematically at Fig. 1.14, the dislocation line, driven by shearingstresses, travels through the crystal achieving a small displacement of the upper half of
the crystal against its lower half.

Figure. 1.14. Dislocation line, stretching from edge dislocation (a) to screw dislocation (b), travels through a
crystal, shifting its upper half relative to its lower half by atomic distance.

In order to achieve a noticeable plastic deformation, a great many dislocation lines must
travel through the metal. The more crowded the dislocations get, the more do they
interfere with one another: They get entangled, pile up against grain boundaries (ends of
slip planes) and catch on with inclusions and other lattice disturbances. See Fig. 1.15.
The jamming and entangling dislocations build up a growing resistance against
further deformation which eventually ceases, unless shearing-stresses are increased. This
is the cause of the phenomenon of deformation hardening and a contributing cause to
dispersion hardening and other hardening processes.
1-17

1. METALLOGRAPHY

When the temperature is raised, dislocations gradually dissemble as atoms escape from
them via vacancies. This dissembling mechanism is called dislocation climbing.

Figure. 1.15. Obstacles hampering the movement of dislocations; Top left: dislocations pile up against grain
boundaries (schematically); Top right: dislocation lines traveling from left to right catch on with inclusions
(schematically); Bottom: single dislocations (black lines) and dislocation tangles (diffuse black spots) in a-iron
with 0.91 at.-% Cu, plastically deformed by 4 %, 50 000:1 replica. [1.1]

1-18

1.3. DIFFUSION IN METALS

See Fig. 1.16. Dissembling of jammed and entangled dislocations is the mechanism
behind the effect which soft-annealing has on cold-deformed metals. The fact that
metals deform easier at elevated temperatures is due the following two processes:
1) the increased number and greater mobility of vacancies in the metal facilitate the
dissembling of dislocations, and
2) the thermal vibration of the atoms in their lattice sites is intensied. Both processes
have a lowering effect on the metals resistance to deformation, i.e.on its yield point.

Figure. 1.16. Dislocation climbing:


A dislocation dissembles gradually as atoms escape
from it via vacancies.

1.3 Diffusion in Metals


All processes and reactions taking place in gases, liquids or solids are depending on the
thermally activated exchanges of atomic particles controlled by laws of thermodynamics
and statistics. In homogeneous materials, the thermally activated particles move around
in all directions without any preference. When gradients of temperature, pressure,
concentration or electrical potential occur in the material, the irregular movement of
particles is superimposed by a drift in the direction of the gradient (similar to a
mosquito swarm drifting in the wind). This drift of atomic particles is called
(directional) diffusion. Diffusion usually accomplishes a noticeable transport of material
as, for instance, in the formation or transformation of alloys and in the sintering of metal
powders.

1-19

1. METALLOGRAPHY

The movement and exchange of atoms in solid metals is feasible in various ways.
Inside the crystal lattice of the metal, small foreign atoms can move relatively easily from
interstice to interstice between the host atoms. This kind of atom movement is called
interstitial diffusion. See Fig. 1.17 (a).
The host atoms themselves and substitutional atoms can move only from one
vacancy to another. This kind of atom movement is called vacancy diffusion or,
pertaining to the host atoms, self-diffusion. See Fig. 1.17 (b). More easily than in the
close-packed lattice structure, atoms can diffuse along the less closely packed grain
boundaries and most easily along surfaces. Accordingly one speaks of volume-diffusion,
grain-boundary-diffusion and surface-diffusion.

Figure. 1.17. Interstitial diffusion (a) and vacancy diffusion (b); (schematically).

1.3.1 Laws of Diffusion


In heterogeneous materials, concentration differences tend to level, because an even
distribution of concentration constitutes a state of minimal free enthalpy. According to
Ficks rst law of diffusion, the number dn of atoms passing through an area A
perpendicular to the direction of their movement, during a short time interval dt, is
proportional to the concentration gradient -dc/dx residing at the location of area A:

1 dn
dc

= D
A dt
dx
1-20

(1.1)

1.3. DIFFUSION IN METALS

The proportionality factor D is called diffusion coefcient and is usually measured in


cm2 s-1. Equation (1.1) applies only if the concentration gradient -dc/dx does not
change during the whole diffusion process (stationary case). Such is approximately the
case, for instance, when carbon from a hot carbonaceous gas diffuses through the thin
steel wall of a furnace mufe. On its inside, the mufe is being carburized by the
carbonaceous gas, on its outside decarburized by the surrounding air. As a consequence,
a constant gradient of carbon concentration establishes itself in the wall, and carbon
diffuses through the mufe wall at constant rate.
If, in the course of diffusion, concentration changes everywhere in the material (nonstationary case), Ficks second law of diffusion applies:

c
c
=
D

t x x

(1.2)

In cases where the diffusion coefcient D can be considered being independent of the
concentration c, this equation is simplied to:

c
2c
=D
t
x2

(1.3)

Given the concentration distribution at the beginning of the diffusion process, this
partial differential equation can be solved by means of Gau error function erf ( ).
A relatively simple particular solution, applicable to many practical cases, is presented in
the following paragraph.

1.3.2 A Particular Solution of the Diffusion Equation


In many technical applications, the surfaces of two different metals are in such intimate
contact with one another that they can exchange atoms. Such is the case e.g. when iron
surfaces are protected with thin layers of zinc, copper, nickel or chromium, or in powder
metallurgy where powder particles of different metallic identity are intimately pressed
together in a compacting process.
In order not to confuse the understanding with complicated geometrical parameters,
we choose a solution of equation (1.3) applying to the simple geometrical arrangement
shown at Fig. 1.18: The end-faces of two oblong metal bars, A and B, are closely pressed
together. This simple model illustrates the gradual leveling of concentration between two

1-21

1. METALLOGRAPHY

different metals which is signicant of all diffusion processes. The concentration


distribution at the beginning of the diffusion process is extreme, viz. 100 % A-atoms in
the left bar and 100 % B-atoms in the right bar. For this model, the particular solution
of equation (1.3) is:

c (x ,t ) =

1
1 erf x / 2 D t
2

))

(1.4)

where c(x,t) is the concentration of A-atoms in B at time t and distance x from the
contact surface between A and B. A large variety of numerical values of Gau errorfunction erf ( ) can be found in relevant tables of mathematical textbooks.

1-22

1.3. DIFFUSION IN METALS

Figure. 1.18. Course of diffusion between two bars of metals A and B (schematically). Assumed are initial
concentrations of 100 % A-atoms in the left and 100 % B-atoms in the right bar. The curves are drawn according to equ. (1.4) and represent the distribution of A-atoms across the two bars after different periods of time.

Remarkable with equation (1.4) is the circumstance that the concentration c(x,t) does
not depend explicitly of the individual parameters x and t but on their combination as in
the expression x / 2 D t . This means that any given concentration of A-atoms in B
moves in x-direction (i.e. from A to B) according to the time law: x D t . This
time law is signicant of all diffusion processes and controls e.g. the speed of growth of
oxide layers on metal surfaces and the traveling speed of a carburized zone from the
surface to the center of a work piece of steel.

1-23

1. METALLOGRAPHY

1.3.3 The Diffusion Coefcient


The diffusion coefcient D is a measure of the diffusion rate. D depends on the
mechanism of atom movement inside the metal and particularly strongly of temperature
according to the following relationship:

D = D0 exp( Q / R T )

(1.5)

D0 is a characteristic property of the metal, R is the universal gas constant, T is the


absolute temperature, and Q is the activation energy.
D0 depends on the atomic weight of the diffusing atoms and on the melting temperature
of the metal in which diffusion takes place. In the case of vacancy-diffusion, the
activation energy is the sum of formation energy and migration energy of vacancies. In
the case of interstitial-diffusion, the activation energy is identical with the traveling
energy of interstitial atoms. The respective traveling energies of atoms are lower in grain
boundaries and lowest on surfaces.
Plotting the logarithm of D from equation (1.5) over 1/T , one obtains a straight line
with the slope Q/R. See Fig. 1.19. Thus, from the slope of this straight line, the
activation energy of the diffusion process can easily be calculated. Deviations from the
straight line indicate that several kinds of atomic movements with varying activation
energies are involved in the diffusion process.

1-24

1.3. DIFFUSION IN METALS

Figure. 1.19. Temperature


dependence of the coefcients
of self-diffusion of iron, for diffusion along surfaces and grain
boundaries and inside the crystal
volume (In D over1/T).

1-25

1. METALLOGRAPHY

Diffusion coefcients D, material constants D0 and activation energies Q for some


important diffusion systems are presented in Table 1.2.
Table 1.2. Physical data on some important diffusion systems
Diff.
Atom

HostCrystal

D0 (m2/s)

Activation Energy Q
KJ/mol
kcal/mol

Fe

-Fe
(BCC)

2,0 x 10 -4

241

Fe

-Fe
(FCC)

5,0 x 10 -5

-Fe

T (C)

D (m2/s)

57,5

500
900

1,1 x 10-20
3,9 x 10-15

284

67,9

900
1100

1,1 x 10-17
7,8 x 10-16

6,2 x 10 -7

80

19,2

500
900

2,3 x 10-12
1,6 x 10-10

-Fe

1,0 x 10 -5

136

32,4

900
1100

9,2 x 10-12
7,0 x 10-11

Cu

Cu

7,8 x 10 -5

211

50,4

500

4,4 x 10-19

Zn

Cu

3,4 x 10 -5

191

45,6

500

4,3 x 10-18

Al

Al

1,7 x 10 -4

142

34,0

500

4,1 x 10-14

Cu

Al

6,5 x 10 -5

135

32,3

500

4,8 x 10-14

Mg

Al

1,2 x 10 -4

131

31,2

500

1,8 x 10-13

Cu

Ni

2,7 x 10-5

255

61,0

500

1,5 x 10-22

1-26

1.4. BINARY PHASE DIAGRAMS OF METALS

1.4 Binary Phase Diagrams of Metals


1.4.1 Thermodynamics
First of all, some denitions are required:
A thermodynamic state is given through the following four parameters of state: P, V, T
and c (pressure, volume, temperature and concentration).
A heterogeneous system consists of several phases separated by interfaces and having different properties (e.g. ice-crystals or droplets of oil in Water).
A phase is in itself homogeneous and can consist of several components present in
varying concentrations (e.g. salt water consists of H-atoms. O-atoms, Na-atoms and
Cl-atoms).
Above-mentioned parameters of state given, number, type and relative amount of
different phases present in a heterogeneous system are stable, when the system is in
thermodynamic equilibrium. What is meant by this statement, will emerge from the
following consideration. A systems (e.g. a metals) internal energy E is the sum of all
individual kinetic and potential energies of its particles (metal atoms). The external
energy PV imposed on the system is the product of the systems volume V and the
external pressure P. The sum of these two energies E + PV can also be split up according
to a different aspect, namely into one part which constitutes irreversible thermal energy
TS (S = entropy) and another part G which is available to accomplish work (e.g.
transformation work). G is called Gibbs free energy. This interrelationship is expressed by
the following equation:

E + PV = TS + G

(1.6)

Thus, Gibbs free energy is given by the expression:

G = E TS + PV

(1.7)

The expression:

F = E TS

(1.7a)

is called (Helmholtz) free energy. The differential of (1.7), constitutes the change dG of
free energy:

d G = d ( E S T + PV ) = d F + d( P V )

(1.8)
1-27

1. METALLOGRAPHY

From the First and Second Law of Thermodynamics, in case T = constant and
P = constant:

d G = d( E ST + PV ) 0

(1.9)

and, in case T = constant and V = constant, it follows:

d F = d( E TS ) 0

(1.9a)

Free Energy G

Since equilibrium reactions between metal phases usually take place at constant pressure
and constant temperature but sometimes are accompanied by volume changes, equation
(1.9) applies. This means that between phases striving for equilibrium, only such
reactions can take place that do not increase the free energy, G. All reactions cease when
the phases are in equilibrium and the free energy, G has reached a minimum. The
schematic diagram at Fig. 1.20 illustrates this situation.

Solid

Liquid
Solid

Liquid
m.p

Temperature

Figure. 1.20. Change of free energy with temperature for a pure metal.

The free energies of the liquid and of the solid state of a pure metal are plotted as
functions of temperature. The two respective curves cross at the metals melting point.
Below the melting point, the free energy of the solid state is lower than the one of the
liquid state. Above the melting point, the relation between the two free energies is
reversed.

1-28

1.4. BINARY PHASE DIAGRAMS OF METALS

1.4.2 Derivation of Binary Phase Diagrams


Applying the thermodynamic rules discussed in the preceding paragraph to a
heterogeneous system involving two components A and B and two phases 1 and 2,
equilibrium, at given temperature and pressure, can be achieved only by changing the
concentrations of A and B in the two phases. The component A will diffuse from phase 1
to phase 2, or vice versa, until the changes of free energy with changing concentration are
equal with respect to both phases, viz.:

d G1 d G 2
=
dc
dc

(1.10)

The implication of this equation is that the two functions G1(c) and G2(c), plotted in a
G-c-diagram, have a common tangent when they are in equilibrium. On the basis of
these considerations, any binary diagram can be derived theoretically as long as the
respective free energies are known as functions of the concentration. These functions can
be deduced from the entropy of mixing of the involved components. A discussion of
entropy of mixing would exceed the frame of this chapter, and is not required for the
understanding of succeeding paragraphs.
Systems with Complete Miscibility in the Solid and in the Liquid State.
For a metal system as dened in the heading to this paragraph, the free energies GL(c)
and GS(c) of the liquid and of the solid phase are plotted, as functions of concentration
and temperature, in a schematic diagram as shown at Fig. 1.21. The melting point of the
components A and B are denoted TMA and TMB (> TMA ) respectively. In the diagram,
we distinguish between the following temperature ranges:
a) T > TMB
Over the entire range of concentration, the free energy curve GL(c) of the melt lies below
the free energy curve GS(c) of the solid. Over the entire range of concentration, only the
liquid phase is thermodynamically stable.
b) T = TMB
For c = 1 (100 %B), the free energy of the melt and the free energy of the solid are equal:
GL(c) = GS(c).

1-29

1. METALLOGRAPHY

c) TMA < T< TMB


In this temperature range, the two free energy curves cross, and each curve exhibits a
minimum. The equilibrium of the two phases lies on the common tangent of the two
curves in accordance with equation (1.10). The abscissas of the tangent points cL and cS
correspond to the concentration of B in the melt and in the solid phase respectively that
are in thermodynamic equilibrium.
d) T < TMA
In this temperature range, over the whole range of concentrations, the free energy curve
of the solid lies below the free energy curve of the melt, viz. in this temperature range,
only the solid phase (composed of A- and B-atoms) is thermodynamically stable.
Thus, the ranges of thermodynamic stability of the various phases in the binary T-cdiagram (phase diagram) are separated by two theoretically deduced phase borderlines
(Liquidus and Solidus).

1-30

1.4. BINARY PHASE DIAGRAMS OF METALS

T>TMB

GS
GL
Free Energy of Phases Gj (c)

a)
T=TMB
GS
GL
b)
TMB>T>TMA
GS

GL

CL

c)

CS
TMA>T

GL
GS

Temperature T

d)

TMB

Liquidus

TMA

A
e)

Solidus

Concentration c

Figure. 1.21. Deriving a binary phase diagram


from the free energy curves for the liquid and
solid phases at different temperatures.
A and B are the components of the system
which are completely miscible both in the
liquid and solid state. GL and GS are the free
energies of the liquid and of the solid state
resp. TMA and TMB are the melting points of
A and B resp. cL and cS are the concentrations
of B in the liquid and in the solid state resp.
[1.2]

1-31

1. METALLOGRAPHY

Experimentally, binary diagrams can be established by means of thermal analysis. To this


effect, from the two components of the system to be investigated, a number of alloys of
different composition are smelted and subsequently cooled. By means of a
thermocouple, the temperature changes during cooling are registered. On the
so-obtained curves, phase transformations show up in the form of kinks.
These kinks are caused by thermal energy being released during phase transformation
(lowering of free energy) and slowing down the rate of cooling. The principle of this
procedure is illustrated at Fig. 1.22. Typical representatives of this simple binary system
are the following pairs of metals: Cu-Ni, Ag-Au and Ag-Pd.

B 1

Temperature T

Temperature T

Li

A
Time t
a)

So

c3

c2

c1

Concentration c
b)

Figure. 1.22. Experimental derivation of a phase diagram by means of thermal analysis, Left: cooling curves
for the two pure metals and for three of their alloys. Right: phase diagram derived from cooling curves. [1.3]

Eutectic Systems.
More complicated phase diagrams can be derived theoretically and established
experimentally by methods analogous to the ones described above. As an example, we
will study the binary T-c-diagram of a so-called eutectic system as presented at Fig. 1.23.
In this system, two components (metals) A and B are completely miscible in the
liquid state but only to a limited degree in the solid state.

1-32

1.4. BINARY PHASE DIAGRAMS OF METALS

L
Figure. 1.23. Experimental derivation of the phase diagram for a eutectic system of two components A and B
with limited miscibility in the solid state.

A remarkable feature of eutectic systems is the circumstance that at a certain temperature


TE (eutectic temperature) not two but three phases are in equilibrium with one another,
viz. a solid A-rich -phase, a solid B-rich -phase, and a liquid phase L (the melt). The
equilibrium relation between the three phases is represented in the diagram by a
horizontal tie-line (also called eutectic line) at the level of TE. Above this line, there are
two so-called 2-phase areas (+L) and (+L) as well as one 1-phase area (L). All three
areas join in one point E (the eutectic point) on the eutectic line.
Below the eutectic line, there is one 2-phase area (+). To their left, the 2-phase
areas border a single-phase area () and to their right, they border a single-phase area ().
In the (+L)-area, the melt is in equilibrium with the solid -phase and in the (+L)area with the solid -phase. In the (+)-area, the two solid phases and are in

1-33

1. METALLOGRAPHY

equilibrium with one another. No additional solid phases exist. To the right of the binary
diagram, the cooling-curves for three different alloys 1, 2, 3 are shown as obtained by
thermal analysis. On the cooling-curves for alloys 1 and 2, the eutectic reaction shows up
in the form of sharp kinks. Below the binary diagram, microstructures of the solidied
alloys are shown (schematically).
Alloy 1 of concentration cE (eutectic alloy) remains liquid until it passes the eutectic line.
Then, it spontaneously precipitates two different solid phases and with
concentration c and c respectively. The separation of the melt takes place according to
the eutectic reaction L + . The ratio of the relative amounts of and in the
microstructure of the solidied alloy is exactly equal to the ratio of distance E - M2 and
distance E - M1 on the eutectic line.
This relationship follows from a law called lever-rule of phases which will be explained
in paragraph 1.4.3.
Alloy 2 begins to solidify when passing the borderline between the melt (L) and the
(L+)-area. As temperature falls further, -crystallites of gradually increasing Aconcentration precipitate from the melt, depriving it of B-atoms, according to the
reaction L. As the alloy passes the eutectic line, the residual melt spontaneously
solidies, precipitating -crystallites of concentration c and -crystallites of
concentration c, according to the eutectic reaction L+. Also in this case, the abovementioned lever-rule of phases applies.
Alloy 3 begins to solidify in a similar fashion as alloy 2, gradually reducing the amount
of the remaining melt. But on arrival at the borderline between the (L+)-area and the
()-area, the residual melt has disappeared and the alloy is completely solidied as
-phase. On further cooling, the alloy passes the borderline between the ()-area and the
( + )-area and is now tending to establish an equilibrium between -phase and
-phase according to the reaction + . Consequently, -crystallites precipitate
along the grain boundaries of the solid -structure.
Binary eutectic systems are formed e.g. by the following pairs of metals: Pb-Sn
(soldering alloys), and Cu-Ag (brazing alloys). In both cases, the eutectic allows to make
alloys that melt at considerably lower temperatures than any of the pure metals they are
composed of. Fig. 1.24 shows the binary system Pb-Sn and two typical microstructures.
In the microstructure at left, the pre-eutectic Pb-rich -phase stands out as small
potato-shaped dark areas embedded in the subsequently formed eutectic. The
microstructure at right exhibits a plain eutectic with its typical arrangement of
alternating lamellae of - and -phase.

1-34

1.4. BINARY PHASE DIAGRAMS OF METALS

Figure. 1.24. Phase diagram of the system Pb - Sn. Microstructures: -phase embedded in eutectic (left) and
pure eutectic (right).

1-35

1. METALLOGRAPHY

Peritectic Systems.
An equilibrium between three phases occurs also in a family of binary diagrams called
peritectic. In these systems, the melt reacts with an already precipitated solid phase, say ,
in such a way that a different solid phase, say , is formed according to the peritectic
reaction L+ .
Fig. 1.25 shows the binary diagram of a peritectic system and the cooling curves by
means of which it is established.

Figure. 1.25. Experimental derivation of the phase diagram of a peritectic system A-B.

Eutectoid Systems.
In some binary systems, as for instance in the iron-carbon-system, in addition to the
eutectic reaction between the melt and two solid phases L + , an analogous
reaction occurs between one solid phase and two other solid phases: + which is
called eutectoid reaction.
The iron-carbon-system is of central interest not only to the metallurgy of conventional
steel but also to the powder metallurgy of iron. But before concentrating on this
important system, we want to give a more detailed explanation of the above-mentioned
lever-rule of phases and, in this context, describe the phenomenon of crystal segregation
which frequently occurs in the solidication process of alloys.

1-36

1.4. BINARY PHASE DIAGRAMS OF METALS

1.4.3 The Lever-Rule of Phases


Guided by the simple binary diagram for two components A and B as shown at Fig.
1.26, this rule is readily explained.
TB

m0
Liquid

mS

mL
x

y
Figure. 1.26. Lever-rule of phases;
mL = amount of residual melt with
concentration cL.
mS = amount of solid phase with concentration cS precipitated from the
melt.
x = (c0 - cL) and y = (cS - c0),
c0 = initial concentration of the melt
= average concentration of the solid
alloy.
r-rule: mLy = mSx.

TA

Solid Solution

CL C0

CS

Concentration of B in A

We assume that a given amount m0 of a molten alloy of concentration c0 (concentration


of B-atoms) is rapidly cooled down to a point P0 inside the 2-phase area (liquid +
solid). Through P0 we draw a horizontal line (the so-called tie-line) which cuts the two
phase-borderlines at the points PL and PS respectively. In this undercooled state, the melt
of concentration c0 has become thermodynamically instable.
According to the binary diagram, a melt of concentration cLshould now be in
equilibrium with a solid phase of concentration cS. In order to establish this new
equilibrium, crystallites of concentration cS > c0 must precipitate depriving the melt of
B-atoms. In this reaction, both, the total amount m0 = (mL + mS) of alloy and the total
amount c0m0 of B-atoms must remain unchanged.

1-37

1. METALLOGRAPHY

Thus, the following two laws apply:


(a) Conservation of total mass:

m0 = mL + mS

(b) Conservation of mass of B-atoms:

c0 m 0 = c L m L + cS m S

Inserting (a) into (b):

m L (c 0 c L ) = m S (c S c 0 )

Evident from the diagram:

x = (c 0 c L )

Lever-rule of phases:

m L x = mS y

and

y = (c S c 0 )

m0 = mass of melt before solidication = total mass of alloy


mL = mass of residual melt after established equilibrium
mS = mass of solidied phase after established equilibrium
c0m0 = mass of B-atoms before solidication
cLmL = mass of B-atoms in residual melt
cSmS = mass of B-atoms in solidied phase
The lever-rule of phases expressed in words: (amount of residual melt) x (tie-line from
balance point to liquidus line) = (amount of precipitated solid phase) x (tie-line from
balance point to solidus line).
1.4.4 Crystal Segregation
Frequently, during solidication of molten alloys, a characteristic phenomenon called
crystal segregation occurs, i.e. a heterogeneous distribution of concentration inside
individual crystal grains of the solidied alloy. Crystal segregation comes about as
follows:
As can be seen from the binary diagram shown at Fig. 1.27, solidication starts with
the precipitation of small B-rich dendrites of concentration c1 > c0. As solidication
continues, thin layers of decreasing concentration of B-atoms are successively being
deposited on the dendrites, wrapping them in onion skins so to speak. In the
outermost layer, the concentration of B-atoms is cn < c0. The varying amount of B-atoms
in the individual layers is a direct consequence of the above-mentioned lever-rule of
1-38

1.4. BINARY PHASE DIAGRAMS OF METALS

phases. Although the concentration of B-atoms in the core of each crystal grain is higher,
and in its outermost zone lower, than the original concentration of the melt, the average
concentration of the alloy has not changed during solidication.
Concentration of A in B
Liquid

Temperature

T1

TB

C0

Tn
TA
Solid Solution

Concentration of B in A

Cn

C0

C1

Figure. 1.27. Crystal segregation;


c0= initial concentration of
the melt;
c1 = concentration of dendritic crystal nuclei rst
precipitated from the melt;
cn = concentration of residual melt solidifying last.

When a molten alloy is cooled sufciently slowly, due to diffusion processes,


concentration differences caused by crystal segregation level to a certain degree already
during continued cooling. In some cases, in order to improve the properties of the
solidied alloy, remaining concentration differences have to be eliminated by means of a
subsequent heat treatment called diffusion annealing or homogenizing.
1-39

1. METALLOGRAPHY

1.5 The Iron-Carbon System


1.5.1 The Equilibrium Diagram: Iron-Carbon
This binary diagram applies to the equilibrium states of carbon-steel and cast-iron. The
term carbon-steel refers to iron-carbon alloys containing less than approximately 2 %C.
The term cast-iron refers to alloys containing between 2 and 4 %C where the Fe-Ceutectic occurs. Of interest for technical applications is mainly the iron-rich side of the
diagram stretching to the concentration of the inter-metallic phase Fe3C (6.67 %C).
This part of the diagram is shown at Fig. 1.28 and can be looked at as being
composed of three sub-diagrams: one peritectic, one eutectic and one eutectoid diagram.
In connection with the heat treatment of steel, only the eutectoid sub-diagram is of
importance. The range up to temperatures of approx. 1150 C and up to concentrations
of approx. 1.5 %C is of particular interest to the powder metallurgy of iron.
Pure iron melts at 1536 C. Below this temperature, it solidies in BCC crystal
structure, called -iron, which can dissolve max. 0.10 % carbon interstitially (at 1493
C). On further cooling, at 1392 C, -iron transforms to FCC crystal structure, called
-iron, which can dissolve max. 2.06 % carbon interstially (at 1147 C). Carboncontaining -iron is called austenite. At 911 C, pure -iron transforms again to a BCC
structure, which is called -iron or ferrite, and which is ferro-magnetic below 769 C
(Curie-point). Ferrit can dissolve max. 0,02 % carbon (at 723 C).
Iron together with carbon forms the inter-metallic phase Fe3C, a carbide called cementite,
which has an ortho-rombic structure and a carbon content of 6.67 %. Cementite is a socalled meta-stable phase because, when annealed for long times at high temperatures, it
dissembles into iron and graphite.
Depending on whether carbon occurs bound in cementite or as free graphite, one
speaks of the meta-stable or of the stable system. Transformations in steel and in white
cast-iron proceed according to the meta-stable system (fully drawn lines).
Transformations in graphite-containing gray cast-iron proceed according to the stable
system CD1.
An austenite containing less than 0.8 %C transforms partly to ferrite when it passes
the phase-borderline GS (also called A3), and below the eutectoid line PS (also called
A1), its residue separates into ferrite and cementite, forming a lamellar structure called
pearlite. An austenite containing exactly 0.8 %C transforms directly and entirely to
pearlite when it passes through the eutectoid point S.
An austenite containing more than 0.8 %C transforms partly to cementite
(secondary cementite) when it passes the phase-borderline SE (also called Acm), and
below the eutectoid line SK, its residue transforms directly to pearlite.
1-40

1.5. THE IRON-CARBON SYSTEM

Figure 1.28. Equilibrium phase diagram of the system Iron-carbon: fully drawn lines = metastable system
(Fe-Fe3C), broken lines = stable system (Fe-graphite).

With reference to these three model cases, one distinguishes between hypoeutectoid,
eutectoid (pearlitic) and hypereutectoid iron-carbon-alloys or steels. The respective
microstructures are presented schematically below the diagram. Real microstructures of
austenite, ferrite, pearlite and cementite are shown at Fig. 1.29.

1-41

1. METALLOGRAPHY

Figure 1.29. Microstructure of steel; a. Austenite (stainless steel), b. Ferrite (< 0,02 % C), c. Ferrite + Pearlite
(0,30 % C), d. Pearlite + Ferrite (0,60 % C), e. Pearlite (0,85 % C), f. Pearlite + grain boundary cementite.
[1.4]

1-42

1.5. THE IRON-CARBON SYSTEM

1.5.2 The Non-Equilibrium Diagram: Iron-Carbon


The iron-carbon diagram as shown at Fig. 1.28 like the binary diagrams treated further
above pertain to states of equilibrium and can, in principle at least, be deduced
theoretically from classical thermodynamics. With cooling rates as usually practiced in
the steel-producing and steel-working industry, states of equilibrium are not always easily
achievable and, in many cases not even desirable.
On the contrary, it is just the aimed provocation of certain meta-stable states that
offers outstanding technical advantages. With their corresponding microstructures, the
value scale of properties like tensile strength, elongation, hardness and toughness can be
largely extended. By varying the additional parameters time and cooling-rate, ironcarbon-based materials can be optimally adapted to a great variety of applications.
At the present state of development of thermodynamic theories, it is still not possible to
quantitatively describe the states of non-equilibrium as occurring during forced cooling
of steels. Therefore, empirically-experimentally established transformation diagrams are
still the only basis on which heat treatments of steel can be intelligently planned.
Such transformation diagrams, known as TTT-diagrams (for Time, Temperature,
Transformation), exist today for all common types of steel and can be found in
brochures made available by the steel producers. In paragraph 1.6, these diagrams will be
discussed in detail.
Transformations in the solid state require rearrangement of atoms by diffusion and in
some cases also certain mechanisms of shearing the crystal lattice. Since diffusion is a
time- and temperature-dependent process, the attainment of equilibrium is the more
retarded the more the steel is undercooled. At sufcient degrees of undercooling, metastable and instable crystal structures (and corresponding microstructures) occur which
considerably deviate from equilibrium structures.
With increasing cooling rate, the phase-borderlines in the equilibrium diagram: ironcarbon (Fig. 1.28), are moving towards lower temperatures, particularly the borderlines
GSE (A3 and Acm) and PSK (A1) of the austenite area. The diagram at Fig. 1.30 shows
the displacement of the A3-point and the of the A1-point with increasing cooling rate for
a steel with 0.45 %C. As can be seen from this diagram, the A3-point moves faster
towards lower temperatures than the A1-point.
As a consequence, with increasing cooling rate, the formation of pre-eutectoid ferrite
is more and more reduced and, eventually, entirely suppressed as A3 and A1 join in a
common A-point. At still higher cooling rates, Bainite (AB) and Martensite (MS) are
formed.

1-43

1. METALLOGRAPHY

Figure 1.30. Inuence of cooling rate on the position of the transition points A3 and A1 for a steel with
0.45 % C. [1.5]

The inuence of cooling rate on the displacement of A3 and the A1, as just described for
a steel with 0.45 %C, applies analogously to the entire eutectoid part of the iron-carbon
diagram, i.e. to all plain carbon steels.
The diagram at Fig. 1.31 illustrates the situation. Pertaining cooling rates are stated
along the right ordinate of the diagram. With reference to cooling rates, one
distinguishes between the following ve steps of undercooling:
Undercooling step 0 (equilibrium).
Equilibrium transformation at very low cooling rate.
Undercooling step I.
Already at a cooling rate of 1 Ks-1, a reduced formation of pre-eutectoid ferrite in undereutectoid steels and of pre-eutectoid cementite in over-eutectoid steels becomes
noticeable. At further increased cooling rate, the formation of pre-eutectoid ferrite is
gradually hampered to such a degree that it occurs only in steels of very low carbon
contents.
The austenite transforms to pearlite of increasingly ner lamellar structure which,
when just about resolvable by light-microscopic methods, is called sorbite. In
hypoeutectoid steels, the formation of pre-eutectoid cementite is completely suppressed.
Undercooling step II.
An extremely ne-lamellar pearlite, called trostite, is formed. Its lamellar structure is
resolvable by means of electron-microscopic methods only. The united undercooling
steps I and II are also called pearlite step.
1-44

1.5. THE IRON-CARBON SYSTEM

Undercooling step III.


In this so-called bainite step (below approx. 450 C), the diffusion of iron atoms has
practically ceased, and a formation of pearlite is no longer possible. Only the interstitial
carbon atoms can still diffuse in this temperature range. Already before the BCC
austenite lattice shears into the FCC ferrite lattice, cementite begins to precipitate. The
resulting microstructure consists of nely dispersed cementite particles in a matrix of
acicular ferrite. It is a kind of dispersion-hardened ferrite, so to speak. The orientation of
the ferrite needles is related to the initial austenite lattice.

Figure 1.31. Inuence of cooling rate on the position of phase boundary lines in the eutectoid part of the
Fe- Fe3C - diagram. [1.6]

1-45

1. METALLOGRAPHY

Undercooling step IV.


In this so-called martensite step, the undercooling of the austenite is so extreme that the
diffusion of carbon atoms has ceased and, hence, the formation of cementite has become
impossible. The FCC lattice of the extremely undercooled austenite shears diffusionless
into a tetragonally distorted (BCT) ferrite lattice. Since the so formed ferrite is extremely
supersaturated with carbon, high internal stresses occur in its lattice. The tetragonal
distortion of the lattice and the internal stresses increase with increasing carbon content.
See Fig. 1.32. This is the cause of the high hardness of martensite.

Figure 1.32. Inuence of carbon content on the axis lengths of the tetragonal martensite lattice. [1.7]

1-46

1.5. THE IRON-CARBON SYSTEM

1.5.3 Mechanisms of Austenite Transformation


To further the understanding of austenite transformations, it will be helpful to discuss in
more detail the various mechanisms of diffusion and structural changes involved in the
formation of ferrite, pearlite, bainite and martensite.
In between the close-packed iron atoms in the FCC lattice of austenite, there are two
types of interstices: the smaller tetrahedral and the larger octahedral interstices as shown
in the schematic diagrams at Fig. 1.33. Carbon atoms dissolved in austenite occupy
octahedral interstices only, where they cause the least possible lattice distortions and
internal stresses. To each unit cell of the austenite lattice belong four octahedral
interstices (one at its center and a quarter of one on each of its twelve edges). Thus, in
each unit cell, there are four positions available to be lled by carbon atoms. Since also
four iron atoms belong to each austenite cell, the theoretically possible concentration of
carbon atoms in austenite would be 50 at.-% compared to their maximal concentration
of 9.51 at.-% in reality. This means that, even at maximal carbon concentration, most
octahedral interstices in the austenite lattice remain unlled.

Figure. 1.33. Tetrahedral (a)


and octahedral (b) interstices
in the FCC lattice of austenite.
= possible sites of carbon
atoms.

Ferrite Formation.
Pure -iron does not contain any carbon atoms at all. In hypoeutectoid steels
(< 0.8 wt.-% = 3.7 at.-%), at best, only one out of seven unit cells of austenite contains
one carbon atom. Between 911C and 723C, because of the sparse distribution of
carbon atoms, parts of the FCC austenite lattice can easily shear into the BCC ferrite
lattice, while carbon atoms, not dissolvable in ferrite, get enriched in the residual parts of
the austenite. This shearing mechanism requires a minimum of rearrangement of iron
atoms in the respective crystal lattices, as becomes evident from the schematic diagram at
Fig. 1.34. In the drawing representing two adjacent FCC cells, certain iron-atoms

1-47

1. METALLOGRAPHY

(lled circles) are connected by tie-lines in such a way that they form a body-centeredtetragonal (BCT) cell. It does not take much imagination to comprehend that a slight
stretching of its horizontal and/or slight upsetting of its vertical edges turns this BCT cell
into an ordinary BCC cell. The theoretically possible positions of carbon atoms are
indicated by small black spots on the edges of the cell. Since the described shearing
mechanism occurs almost instantly, one could say that, on cooling, the FCC lattice of
austenite snaps into the BCC lattice of ferrite (or vice versa on heating).

Figure. 1.34. Geometrical interrelation between the FCC lattice of austenite and the BCC lattice of ferrite.
= possible sites of carbon atoms.

Cementite Formation.
The maximal solubility of carbon in austenite is 2.06 wt.-% = 9.51 at.-% and occurs at
1147C (eutectic temperature). This means that scarcely twenty percent of all
theoretically possible positions are lled with carbon atoms (i.e. less than one carbon
atom per austenite cells). A denser lling with carbon atoms (supersaturation) would
increase internal stresses in the austenite lattice to such extent, that the corresponding
increase in elastic energy would exceed the formation energy of cementite.
Consequently in case of supersaturation with carbon, the austenite tends to lower its
free energy by transforming to cementite. The tendency increases with increasing degree
of undercooling. The concentration of carbon in cementite is 25 at.-%, i.e. at least close
to three times higher than in austenite. Thus, in order to form cementite, considerable
amounts of carbon atoms must be adequately displaced by diffusion.
Not enough with carbon diffusion, iron atoms too must be rearranged by
diffusion in order to enable the austenite to transform partly into cementite. Thus, the
formation of cementite depends not only on the austenites opportunity to lower its free
energy but also on the actual mobility of the carbon atoms (which decreases with

1-48

1.5. THE IRON-CARBON SYSTEM

increasing degree of undercooling). At temperatures where the mobility of carbon atoms


in the austenite lattice has ceased, cementite cannot be formed, and martensite occurs
instead. See further down.
Pearlite Formation.
Below the eutectoid temperature (723C), austenite separates into cementite and ferrite,
thereby lowering the free energy of the alloy. Nucleated at the austenite grainboundaries, lamellas of cementite and ferrite grow side by side gradually towards the
center of the austenite grains as shown in the schematic diagrams at Fig. 1.35. The
lamellar structure occurs because the precipitating cementite completely deprives its
austenitic neighborhood of carbon triggering its transformation to ferrite. With
increasing degrees of undercooling, i.e. decreasing temperature, the diffusion rate of
carbon decreases and, consequently, the lamellar structure of the pearlite becomes ner
(and harder).

Figure 1.35. Nucleation of pearlite colonies at austenite grain boundaries (left) and growth of pearlite lamellas in austenite grains (right). [1.8]

Martensite Formation.
Martensite formation is based on the same mechanism as ferrite formation. The
difference lies in the cooling rate. At low cooling rates, carbon atoms have sufcient time
to diffuse and form cementite before the lattice of the residual austenite shears into the
ferrite lattice. At very high cooling rates, as in quenching procedures, carbon atoms
cannot move fast enough to avoid getting trapped inside the spontaneously forming
ferrite. Since ferrite, under equilibrium conditions, cannot dissolve more than
0.02 wt.-% = 0.09 at.-% carbon, it is now extremely supersaturated with carbon, and its

1-49

1. METALLOGRAPHY

lattice correspondingly (tetragonally) distorted (ref. Fig 1.32).


The high internal stresses arising from this distortion are the cause of the high hardness
of martensite. In microstructures, the tetragonally distorted ferrite, i.e. martensite,
appears in the form of relatively coarse needles oriented at certain angles inherited from
the initial austenite lattice.
The small areas in between the needles are residual austenite. See micrograph at
Fig. 1.36 a. When annealing martensite at temperatures above MS , nely dispersed
cementite precipitates from the supersaturated ferrite. See micrograph at Fig. 1.36 b.
Thus, the high level of internal stresses in the tetragonal ferrite lattice is reduced, and the
brittleness of the martensite correspondingly lowered.

Figure. 1.36. Acicular martensite (a) and tempered martensite with nely dispersed cementite
(b). [1.9]

Bainite formation.
Bainite is of great technical interest because it offers a very favorable combination of
hardness and toughness not easily attainable with martensite. According to differences in
cooling conditions, one distinguishes between two kinds of bainite: upper bainite and
lower bainite.
1-50

1.5. THE IRON-CARBON SYSTEM

Upper bainite occurs at temperatures around 500 C where diffusion rates of carbon are
relatively higher. Lower bainite occurs at temperatures around 300 C (but above MS)
where diffusion rates of carbon are relatively lower.
In the case of upper bainite, small parallel ferrite platelets, nucleated by shearing
processes, grow gradually inside the austenite. Moving ahead of the growing ferrite
platelets, carbon atoms accumulate in the residual austenite which gets increasingly more
restricted. Eventually, the residual austenite becomes so restricted and so much enriched
in carbon that it transforms to cementite jammed between the ferrite platelets. See
micrograph at Fig. 1.37a.
In the case of lower bainite, ferrite platelets form in much the same way as in upper
bainite but faster. At the same time, the diffusion rate of carbon has dropped so much
that carbon atoms cannot move fast enough to avoid getting trapped inside the fast
growing ferrite platelets. In this respect, the mechanism of lower bainite formation is
quite similar to the mechanism of martensite formation.
But in the case of lower bainite, temperatures are high enough to initiate
precipitation of nely dispersed cementite particles inside the ferrite platelets
immediately after they have been formed. See microstructure at Fig. 1.37 b. One could
say that lower bainite is a kind of self-tempered martensite. Indeed, comparing Fig. 1.36
b with Fig. 1.37 b, it can be seen that the microstructures of lower bainite and tempered
martensite have a certain resemblance.

Figure 1.37. Lower (a) and upper (b) bainite in a bainitic steel (0,87 % C; 0,44 % Mn; 0,17 % Si; 0,21 %
Cr; 0,39 % Ni). [1.10]

1-51

1. METALLOGRAPHY

1.6 Transformation Diagrams of Steels


As briey mentioned in paragraph 1.5.2, the goal of commercial heat treatment of steel
is attaining optimal combinations of properties adaptable to a large variety of
applications. This can only be achieved by steering the formation of adequate
microstructures via strictly controlled cooling rates and holding times.
Adequate heat-treating procedures cannot be designed on the basis of equilibrium
diagrams. To this end, a particular kind of empirically established transformation
diagrams are required, usually called Time-Temperature-Transformation diagrams, or
TTT-diagrams for short. According to the procedure by which they are established, one
distinguishes between Isothermal Transformation diagrams (ITT-diagrams) and
Continuous Cooling Transformation diagrams (CCT-diagrams).
We concentrate our interest mainly on ITT-diagrams, since these are usually better
suited for the planning of reliable heat-treating procedures than CCT-diagrams (apart
from the fact that the latter are of limited availability). ITT-diagrams are today available
for practically all common types of steel and can be found e.g. in brochures edited by
steel-makers.
Before going into details, a general feature of phase transformations must be briey
explained. Phase transformations in metals (like any kind of phase transformations) pass
through two characteristic periods: a period of nucleation and a period of growth.
Nucleation is a process in which small nuclei of a new phase precipitate at random from
the mother-phase. Growth means that nuclei, which incidentally exceed a critical
minimum size, begin to grow on account of the mother-phase, eventually replacing it
entirely.
The growth rate of the new phase is proportional to the residue of untransformed
mother-phase, i.e. at the beginning, the new phase grows fast, but as the amount of
mother-phase shrinks, the growth rate of the new phase slows down accordingly. This
mechanism yields the characteristic s-shaped transformation curve as shown
schematically at Fig. 1.38.

1-52

1.6. TRANSFORMATION DIAGRAMS OF STEELS

Degree of Transformation

Nucleation
Growth

Time
Figure 1.38. Typical s-shaped course of isothermal transformation passing from a period of nucleation to a
period of growth.

1.6.1 ITT-Diagrams
In order to experimentally derive an ITT-diagram, a series of small specimens of the steel
to be investigated are austenitized at a temperature above A3. From austenitizing
temperature, the specimens are individually quenched to certain pre-selected
temperatures, where they are held for various lengths of time. Subsequently they are
quenched to R.T. On the so-treated specimens, the amount of residual austenite is
determined by means of quantitative metallographic methods. Plotting the so-obtained
amounts of residual austenite against respective holding times, characteristic s-curves (of
type as discussed above) emerge one for each holding temperature. On each of these
curves, three points of time are marked:
1) the point of beginning transformation (100 % austenite),
2) the point where the amount of residual austenite has decreased to 50 % (50 %
transformation) and
3) the point where the amount of residual austenite has dropped to 10 % or less
(90 - 100 % transformation). Transferring these three points from each individual
S-curve to a temperature-time-diagram, as illustrated schematically at Fig. 1.39, one
obtains an ITT-diagram with three characteristic nose-shaped curves, representing
begin, half-completion and end of austenite transformation respectively.
1-53

1. METALLOGRAPHY

Figure 1.39. Empirically derived ITT-diagram for a hypoeutectoid carbon steel (right) and its relation to the
equilibrium phase diagram Fe-Fe3C (left). Austenitizing above A3. A = austenite, F = ferrite, P = pearlite,
B = bainite, [1.11]

The following details are noted:


At low degrees of undercooling, i.e. at temperatures close below A1, rather long periods of time elapse before transformation begins (hours or days). Transformation product: pearlite (P) and residual austenite (A).
With increasing degrees of undercooling, transformation begins correspondingly
sooner. Transformation product: P + A.
At medium degrees of undercooling, i.e. at temperatures between approx. 550 and
500 C, transformation begins almost instantly (after less than 1 second). Transformation product: P + A.
With further increasing degrees of undercooling, the begin of transformation is
delayed again ( up to approx. 20 seconds). Transformation product: A + B (bainite).

1-54

1.6. TRANSFORMATION DIAGRAMS OF STEELS

At quenching to temperatures below approx. 280 C, large parts of austenite transform instantly (without preceding period of nucleation) into martensite (M). The
lower the quenching temperature, the smaller is the amount of residual austenite.
Transformation product: x% A + (100 - x)% M.

The nose of the transformation curves between 500 and 600 C is called pearlite-nose
because above its tip, austenite transforms to pearlite. The nose-shape of the
transformation curves has a simple explanation. Close below A1, the diffusion rate of
carbon is high, but, due to the low degree of undercooling, the probability of cementite
nucleation is very low. With increasing degrees of undercooling, the probability of
cementite nucleation increases accordingly, but the diffusion rate of carbon decreases. As
a result of these two counteracting processes, transformation times adopt a minimum
between low and high degrees of undercooling.
Carbon steels with less than 0.25 %C are not well suited for heat-treating procedures
aiming at the formation of martensite or bainite. Their pearlite-nose lies too far left in
the ITT-diagram and cannot be bypassed at technically practicable cooling rates.
A cooling curve just touching the pearlite-nose is called critical cooling curve. When
the carbon content is increased or when alloying elements like Cr, Ni, Mn, Si, or W are
added, the pearlite-nose is shifted toward the right in the diagram, the critical cooling
rate is correspondingly decreased, and the formation of martensite and bainite
facilitated.
Certain combinations of Cr-, Ni-, Mn-, and Mo-additions entail transformation curves
with two noses. Underneath the pearlite-nose a so-called bainite-nose occurs as in the
ITT-diagram shown at Fig. 1.40.

1-55

1. METALLOGRAPHY

32

Figure 1.40. ITT-diagram with pronounced bainite-nose for a steel with 0,42 % C; 0,68 % Mn; and
0,93 % Cr. [1.12]

When cooling a big work piece of steel from austenitizing temperature, center and
surface are cooling at different rates. Then, it may happen that the cooling curve
pertaining to the surface bypasses the pearlite-nose, while the curve pertaining to the
center passes through it. In such a case, a martensitic microstructure occurs at the surface
and a pearlitic-bainitic microstructure at the center of the piece.
See schematic diagram at Fig. 1.41. An analogous problem arises, when a heat of
smaller pieces is cooled, because pieces located in the outer parts of the heat are cooling
faster than those buried deeper inside. This kind of problem is avoided if the heat is
quenched to a temperature somewhat above martensite formation (MS) and held there
until transformation is complete.

1-56

1.6. TRANSFORMATION DIAGRAMS OF STEELS

Figure 1.41. Cooling curves for surface and center of a small (fully drawn lines) and of a big (dotted lines)
work piece (schematically). [1.13]

1.6.2 Heat Treatment of Steel based on ITT-Diagrams


The properties of steel can be improved considerably through special heat treatments, i.e.
certain sequences of heating, quenching and tempering. Properties like strength,
hardness, toughness and ductility, depend essentially, if not entirely, on the
microstructure of the steel. Heat-treating procedures must, therefore, be especially
designed to attain the required microstructures.

1-57

1. METALLOGRAPHY

For the design of reliable heat treatments, ITT-diagrams are indispensable.


Depending on the kind of application, one or the other of the following four
standard heat treatments are frequently applied:
a) Quenching and tempering.
b) Martempering.
c) Austempering (Bainitizing).
d) Isothermal soft-annealing.
The schematic ITT-diagrams shown at Fig. 1.42 illustrate these procedures which, in
detail, can be described as follows:
a) Quenching and tempering:
On quenching to temperatures well below the martensite temperature (MS), the austenite transforms to a structure consisting of martensite and smaller amounts of residual austenite. In this condition, the steel is extremely hard but also extremely brittle.
In order to reduce its brittleness and increase its toughness, the steel is subsequently
tempered at temperatures above MS. Due to temperature differences between surface
and core or between thinner and thicker sections of the steel piece, cracks may arise
on quenching. Hence, this type of heat treatment is less suited for big work pieces or
pieces of complicate shape.
b) Martempering:
In this heat treating process, the risk of crack formation in the steel is avoided. The
steel is quenched in a bath at a temperature slightly above MS and held in the bath
until the core of the piece reaches bath temperature. During this period, the bath
temperature falls gradually below MS, and the austenite transforms uniformly into
martensite, whereupon the steel is removed from the bath and allowed to cool in air.
Subsequently the steel is tempered to desired hardness.
c) Austempering:
In this heat treatment, the steel is quenched to a temperature below the bainite-nose
but above MS, at which temperature the austenite is transformed isothermally to
bainite. Subsequent tempering is not required, since, even without tempering, bainite offers a favorable combination of hardness and toughness.
d) Isothermal annealing:
In many alloyed steels, there is a pronounced minimum in the ending line of the
ITT-diagram at relatively high temperatures. Assuming that a soft ferritic-pearlitic

1-58

1.6. TRANSFORMATION DIAGRAMS OF STEELS

Temperature

Temperature

Temperature

Temperature

structure is to be obtained, advantage may be taken of the time-temperature coordinates of this minimum to design a short annealing cycle. This is accomplished by
cooling the steel initially in the austenitic state as rapidly as convenient to the temperature of the minimum and holding it approximately at this temperature for the time
required to transform the austenite completely. Subsequently, the steel may be cooled
as convenient.

Figure 1.42. Four standard procedures for the heat treatment of steels: a. Quenching and Tempering; b.
Martempering; c. Austempering; d. Isothermal Annealing. [1.14]

1-59

1. METALLOGRAPHY

These examples have shown that ITT-diagrams are helpful guides in designing adequate
heat treating procedures. In industrial practice, heat treatments rarely proceed under
strictly isothermal conditions. In many cases, continuous cooling procedures are even
more practicable. Thus, one might think that CCT-diagrams would be more adequate.
But the derivation of CCT-diagrams is a rather tedious task, and would, even if feasible,
rarely be warranted since a particular CCT-diagram exactly presents but one sample;
samples from other heats, or even from other locations of the same heat, are likely to
have slightly different CCT-diagrams.
Thus, when used with discrimination and with its limitations in mind, the ITTdiagram is most useful in interpreting and correlating observed phenomena on a rational
basis, even though austenite transforms during continuous cooling rather than at
constant temperature.
Fig. 1.43 shows how the CCT-diagram for a typical structural steel is correlated with
the corresponding ITT-diagram and with end-quench hardenability. The essential
difference between the two transformation diagrams is that, in the CCT-diagram,
pearlite-nose and bainite-nose are shifted to lower temperatures and longer
transformation times.

1-60

1.6. TRANSFORMATION DIAGRAMS OF STEELS

.
Figure 1.43. CCT-diagram (thick lines) as related to corresponding ITT- diagram (thin lines) and to EQhardenability curves (top) for a steel containing 0,37 % C; 0,77 % Mn; 0,98 % Cr; and 0,21 % Mo.
Transformation products: (A) martensite, (B), (C), (D) martensite + ferrite + bainite. [1.15]

1-61

1. METALLOGRAPHY

1.7 Inuence of the Microstructure on the Properties of Steel


1.7.1 Pearlite and Spheroidite
In pearlitic structures, lamellas of soft ferrite alternate with lamellas of hard cementite.
Under the inuence of shearing stresses, plastic deformation occurs essentially only in
the soft ferrite lamellas where dislocations can move relatively easily. The thinner the
ferrite lamellas are, the more restricted, by the rigid cementite lamellas, is the mobility of
the dislocations. This explains why the ductility of pearlite increases and its hardness
correspondingly decreases with the thickness of its ferrite lamellas as shown in the
diagram at Fig. 1.44.
By means of elongated soft-annealing, the cementite lamellas can be spheroidized. In
the so transformed pearlite, also called spheroidite, small spherical cementite particles are
embedded in a coherent ferrite matrix. See microstructure at Fig. 1.45.
Spheroidite is the most ductile variety of pearlite because, in its coherent ferrite
matrix, dislocations can move more freely, restricted only to a lesser degree by the
spherical cementite inclusions.

Figure. 1.44. Hardness of pearlite as a function of


the distance between its lamellas.

1-62

Figure. 1.45. Spheroidized cementite (spheroidite) in a carbon steel with 1,30 % C.

1.7. INFLUENCE OF THE MICROSTRUCTURE ON THE PROPERTIES OF STEEL

See diagrams at Fig. 1.46. In mixed structures of pearlite and ferrite, strength and
hardness increase with increasing proportion of cementite, while toughness and ductility
decrease. See diagram at Fig. 1.47.

Figure. 1.46. Inuence of carbon content on hardness (a) and ductility (b) for a plain carbon steel
having ne and coarse pearlitic as well as spheroiditic microstructures. [1.16]

Figure. 1.47. Inuence of carbon content on the


mechanical properties of a plain carbon steel
having a ne pearlitic microstructure. [1.17]

1-63

1. METALLOGRAPHY

1.7.2 Martensite and Bainite


Martensite is the hardest and most brittle of all phases occurring in microstructures of
steel. It has virtually no ductility at all. Its hardness increases with increasing carbon
content - rst faster (up to approx. 0.4 %C) then slower, approaching a maximum at
approx. 1 %C. The diagram at Fig. 1.48 shows the inuence of carbon content on the
hardness of martensite and pearlite.
Martensite owes its high hardness not to the presence of cementite, as is the case
with pearlite, but to the high internal stresses caused by its supersaturation with carbon.
In its heavily distorted crystal lattice, the number of sliding systems available for plastic
deformation is extremely low.

Figure. 1.48. Inuence of carbon


content on hardness for plain carbon
martensitic and ne pearlitic steels.
[1.18]

Since austenite has a higher density than martensite, a noticeable volume-increase occurs
on quenching. Due to the accompanying shearing stresses, big or intricately shaped
work pieces may crack when quenched. This constitutes a serious problem in the heat
treatment of steels containing more than approx. 0.5 % C.
In the quenched state, martensite is too hard, too brittle and too crack-sensitive for
most applications. Its brittleness can be reduced and its toughness considerably increased
through tempering at temperatures between 250 and 650 C. Internal stresses in the
martensite lattice dissolve already at approx. 200 C. During tempering, carbon atoms
1-64

1.7. INFLUENCE OF THE MICROSTRUCTURE ON THE PROPERTIES OF STEEL

have sufcient time to diffuse and form cementite which precipitates in the form of
nely dispersed particles inside the martensite needles (ref. Fig. 1.36 b). Tempered
martensite has a certain resemblance to lower bainite (ref. Fig. 1.37 b). The size of the
precipitated cementite particles in martensite increases with tempering time and
temperature.
By means of varying these two parameters, the properties of hardened steel can be
optimally adapted to many different applications. The diagrams at Fig. 1.49 and
Fig. 1.50 illustrate these possibilities.

1-65

1. METALLOGRAPHY

Figure 1.49. Mechanical properties as functions of tempering temperature for a steel, quenched from 850 C
in oil, and containing 0,30 % C; 0,25 % Si; 0,60 % Mn; 0,30 % Cr; 3,30 % Ni and 0,25 % Mo. [1.19]

1-66

1.7. INFLUENCE OF THE MICROSTRUCTURE ON THE PROPERTIES OF STEEL

Figure. 1.50. Hardness as


a function of tempering
time and temperature for
a water-quenched eutectoid plain carbon steel.
[1.20]

Since the microstructure of bainite - similar to that of tempered martensite - is composed


of nely dispersed cementite particles in a ferritic matrix, bainitic steels offer favorable
combinations of hardness, strength and toughness.
The diagram at Fig. 1.51 shows the inuence of transformation temperature on
microstructure and tensile strength of a low-carbon bainitic steel.

Figure. 1.51. Relationship between temperature of maximum transformation rate and


strength of low carbon bainitic steels. [1.21]

1-67

1. METALLOGRAPHY

1.7.3 Temper Brittleness


With some types of steel, tempering may cause a marked decrease of impact strength.
This phenomenon, which is called temper brittleness, occurs either when the steel is
tempered above 575 C and subsequently slow-cooled to R.T., or when tempered
between 375 and 575 C. Steels containing Mn, Cr or Ni and, in addition, small
amounts of impurities, like Sb, P, As or Sn, are especially sensitive to temper brittleness.
When the mentioned alloying elements and impurities are present, the border
between tough and brittle tempered martensite is shifted noticeably to higher
temperatures. Crack propagation in temper-brittle steel proceeds along former austenite
grain boundaries where the impurities preferentially have segregated during solidication
of the melt. Temper brittleness can be avoided by reducing the amount of impurities or
by tempering above 575 C or below 375 C with subsequent quenching to R.T.

1-68

. REFERENCES

References
[1.1]

E. Hornbogen, Symposium: Steel-Strengthening Mechanisms, Zurich, May


5th and 6th 1969.
[1.2]
G.E.R. Schulze: Metallphysik, Akademie-Verlag, Berlin 1974.
[1.3]
W. Schatt: Einfhrung in die Werkstoffkunde, VEB Deutscher Verlag fr
Grundstofndustrie, Leipzig 1981.
[1.4]
Jrnets och stlets metallogra, Sandvikens Jernverks Aktibolag 1964.
[1.5]
W. Schatt: Einfhrung in die Werkstoffkunde, VEB Deutscher Verlag fr
Grundstofndustrie, Leipzig 1981.
[1.6]
F. Wever and A. Rose.
[1.7]
K. Honda and Z. Nihiyama.
[1.8]
A.H.Cottrell: An introduction to Metallurgy, Edward Arnold (Publishers),
London 1968.
[1.9]
K.J. Irvine, Symposium: Steel-Strengthening Mechanisms, Zurich, May 5th
and 6th 1969.
[1.10] I.J.Habraken and M.Economopoulos, Symposium: Transformation and
Hardenability in Steels, Ann Arbor, Michigan, USA 1958.
[1.11] Sandvikens Jernverks Aktiebolag: Stl - struktur och vrmebehandling,
1958.
[1.12] Atlas of Isothermal Transformation Diagrams, US Steel, Pittsburgh, 1951.
[1.13] Sandvikens Handbok, Del 7, Vol II, Sandvikens Jernverks Aktiebolag, 1964.
[1.14] Atlas of Isothermal Transformation Diagrams, US Steel, Pittsburgh, 1951.
[1.15] Atlas of Isothermal Transformation Diagrams, US Steel, Pittsburgh, 1951.
[1.16] Metals Handbook, Vol.4, 9.Edition, ASM, 1981.
[1.17] Metals Handbook, Vol.4, 9.Edition, ASM, 1981.
[1.18] E.C.Bain: Functions of the Alloying Elements in Steel, AMS, 1939.
[1.19] Sandvikens Handbok, Del 7, Vol II, Sandvikens Jernverks Aktiebolag, 1964.
[1.20] E.C.Bain: Functions of the Alloying Elements in Steel, AMS, 1939.
[1.21] K.J. Irvine, Symposium: Steel-Strengthening Mechanisms, Zurich, May 5th
and 6th 1969.
[T.1.1] W. Schatt: Einfhrung in die Werkstoffwissenschaft. VEB Deutscher Verlag fr
Grundstofndustrie, Leipzig 1981, S.55.

1-69

1. METALLOGRAPHY

1-70

PRODUCTION OF IRON
AND STEEL POWDERS
This chapter presents a short history of iron powder and a brief
account of how world production and consumption of iron
powder have developed since 1965. It describes in detail two
production methods which together account for more than 90%
of todays world production of iron and steel powders.

2. PRODUCTION OF IRON AND STEEL POWDERS

TABLE OF CONTENTS

2.1 INTRODUCTION ................................................................ 3


2.2 SHORT HISTORY OF IRON POWDER ................................... 5
2.3 THE HGANS SPONGE IRON PROCESS ........................... 7
2.4 THE HGANS WATER-ATOMIZING PROCESS ................... 10
2.5 ALLOYING METHODS...................................................... 13
2.6 DISTALOY AND STARMIX ............................................. 14

2.1. INTRODUCTION

2.1 Introduction
Iron and steel powders for the manufacturing of sintered structural components
(including sintered porous bushings) are produced in many parts of the world. The
worldwide consumption of such powders has been growing increasingly fast over the last
three decades and reached 770 000 metric tons by the end of 2002.
Over the last thirty years, the quality of iron and steel powders has been continuously
improved and the spectrum of available grades has been widened. During the same
period, compacting- and sintering-techniques have become more and more
sophisticated. This development has lead to a substantially widened range of applications
for sintered iron and steel parts.
Table 2.1 Worldwide Usage of Iron Powder for Sintered Parts
Unalloyed
Iron Powders

Year

Low-alloyed
Iron Powders

All
Iron Powders

compact densities

compact densities

compact densities

compact densities

5.5 to 7.0 g/cm3

> 7.0 g/cm3

> 6.7 g/cm3

> 5.5 g/cm3

tons

tons

tons

tons

1965

35 000

100.0

0.0

0.0

35 000

100.0

1975

120 000

92.3

10 000

7.7

0.0

130 000

100.0

1985

180 000

70.6

50 000

19.6

25 000

9.8

255 000

100.0

1995

300 000

53.6

200 000

35.7

60 000

10.7

560 000

100.0

From Table 2.1, it can be seen that in 1965 iron powders were used almost exclusively for
low- and medium-density applications, i.e. for parts having pressed densities from 5.5 to
7.0g/cm3 . First after about 1970, increasing quantities of iron powders were used for
high-density applications, i.e. for parts having pressed densities higher than 7.0g/cm3.
Between 1975 and 1985, low-alloyed iron powders appeared on the marked and have
since been used in growing quantities for medium- and high-density applications, i.e. for
parts having densities higher than 6.7g/cm3. This development of high density products
has continued after 1995.

2-3

2. PRODUCTION OF IRON AND STEEL POWDERS

It must be mentioned that, at present (2003), the worldwide production capacity of


iron and steel powders is considerably larger than the consumption. Thus, there is no
risk of shortage for many years to come.
At present there are two basically different production methods which together
account for more than 90% of the world production of iron and steel powders, viz. the
Hgans sponge-iron process and the water-atomizing process. The former process is based
on reduction of iron ore, yielding a highly porous sponge-iron which subsequently is
comminuted to powder. The latter process is based on atomization of a stream of liquid
iron (or steel) by means of a jet of pressurized water. Both processes will be described in
detail further below.
In the manufacturing of sintered parts, iron powders are always used admixed with a
small amount of lubricant in powder form in order to minimize the friction in the
compacting tool. In many cases, they are also blended with alloying elements in powder
form, like graphite, copper, nickel, molybdenum and others (in order to achieve
increased strength properties).
Since powder blends tend to segregate when transported and handled, Hgans AB
has developed special blending processes in which the alloying additives are safely bound
to the iron powder particles. Powdermixes produced according to these processes are
known as the trade names Distaloy and Starmix. These two processes are treated in
detail further below.

2-4

2.2. SHORT HISTORY OF IRON POWDER

2.2 Short history of iron powder


Industrial production of iron powder started in 1937 on the incentive of General
Motors Corporation in the USA. Hgans was active since 1922 in producing a highquality sponge iron to be used by the Swedish steel industry as high-purity melting stock
for the production of special steels like tool-steel and stainless steel.
Initial tests in 1937 showed that, owing to its high porosity, the sponge iron produced
by Hgans could readily be comminuted to iron powder. Production of iron powder
started on a small scale at Hgans in Sweden between 1937 and 1939, and regular shipments were made to the USA.
These early grades of sponge iron powder were unannealed, containing up to 2% of
reducible oxygen (H2-loss) and up to 0.15% of carbon. Consequently, their compressibility was poor, and the reduction of residual oxides had to take place in the sintering process under consumption of carbon present in the compacts. Under such conditions, it
was impossible to use carbon as an alloying element.. The bulk density (apparent density)
of these powder grades was relatively low, i.e. 2.2 to 2.4 g/cm3.
After 1940, Hgans introduced an annealing procedure in which the crude iron
powder was heated under dissociated ammonia (75% H2 + 25% N2) resulting in
considerably lower contents of residual oxygen and carbon and yielding a higher
compressibility. This substantially improved powder grade was traded under the name
MH100.24 which was the forerunner of todays NC100.24.
During World War II, iron powder metallurgy went through a dramatic development
in Germany, where, due to shortage of copper, artillery shell driving bands, compacted
and sintered from plain iron powder, were produced in very large quantities which in
1944 amounted to 30,000 metric tons per annum.
The iron powder for this purpose was mainly made by grinding wire cuttings and
sheet clippings in hammer mills of type Hametag. This process yielded a powder well
suited for the application, and such powder was actually continued to be produced in
Czechoslovakia until 1987-89. The production costs, however, were so high that the
process was very short-lived in Western Europe after the war.
In conjunction with the war effort, also other iron powder processes were developed in
Germany. The most important one was the so-called RZ-process (Roheisen-ZunderVerfahren) developed by the Mannesmann AG. Desulphurized cast-iron was atomized
by air jets. The resulting powder consisted of small agglomerates of spherical particles
having a heavily oxidized surface and a highly carbon-containing core. This crude
powder was heated in closed iron boxes, and the oxide on the particle surface reacted
with the carbon in the core, thus forming a reducing atmosphere of carbon monoxide
inside the boxes. After this treatment, the powder consisted of particles having a
2-5

2. PRODUCTION OF IRON AND STEEL POWDERS

somewhat spongy surface and a compact core. Owing to the spongy surface of its
particles, the powder had good green strength, but due its relative high amounts of
residual oxides, it had poor compressibility.
Production and use of RZ-type iron powders have virtually ceased today. A modied
version of the old RZ-process, however, is used by the Quebec Metal Powders
Corporation of Canada for the production of their Atomet powder grades. Also here, the
basis of the process is a liquid cast-iron which, differing from the old RZ-process, is
atomized by a water jet instead of an air jet. The resulting crude powder is ground and
mixed with iron oxide. In a subsequent annealing process under reducing conditions,
carbon and oxygen are eliminated from the mix. Although this type of powder is still in
production, it is gradually giving way to powder types obtained by water atomization of
liquid low carbon iron.
Immediately after the war, Husquarna Vapenfabriks AB of Sweden developed an iron
powder process on the basis of electrolytic iron. This electrolytic iron powder, being
extremely pure and having excellent compressibility, was ideally suited for the
production of high-density components. However, because of its high manufacturing
cost, it found only a very limited market, and its production was discontinued in the
early seventies.
Since 1950, Hgans sponge iron powders have also been produced by Hoeganaes
Corporation in Riverton, N.J., USA.
In 1958, Esamann AB of Sweden, a manufacturer of welding electrodes, started to
produce iron powder by water-atomization of liquid low-carbon iron. This powder had a
high bulk density and was used as a component in the coating of the companys welding
electrodes.
In 1962, the A.O.Smith Corporation in the USA was rst to use water-atomization of
liquid low-carbon iron and low-carbon steel, followed by an annealing process, for the
manufacturing of powders for sintered structural parts. These powders were an
immediate success because, due to their high compressibility, they allowed the
production of high-density parts of excellent quality.
In 1968, Hgans AB started the production of water-atomized iron and steel powders
in Sweden (taking over the Esamann-plant) and in the USA (building a new large
atomizing plant at Riverton, N.J.). Rening the water-atomizing process, Hgans soon
succeeded in making water-atomized iron powder grades of superior quality, some of
them matching the compressibility of electrolytic iron powder and thus replacing the
latter on the market.
Today (2003), Hgans AB has production facilities for iron and steel powders in
Sweden, United Kingdom, Japan, China, India, Brazil and USA as well as for Stainless
Steel Powders in Belgium and USA.
2-6

2.3. THE HGANS SPONGE IRON PROCESS

2.3 The Hgans Sponge Iron Process


The Hgans sponge iron process, is essentially a chemical process in which nely
divided iron ore is being reduced with coke breeze yielding a spongy mass of solid
metallic iron, which can readily be comminuted to iron powder. The iron ore used at
Hgans is a magnetite (powder Fe3O4) obtained from selected sources. This magnetite,
which by nature contains only very small amounts of gang and has extremely low
contents of sulfur and phosphorus, is being dressed and concentrated while still at the
mining location and then delivered to Hgans in a highly pure state.
As shown in the ow-sheet at Fig. 2.1, the process of transforming the magnetite to iron
powder proceeds as described on the next page.

2-7

2. PRODUCTION OF IRON AND STEEL POWDERS

1. Reduction Mix of Coke Breeze and


Limestone
2. Iron Ore
3. Drying
4. Crushing
5. Screening
6. Magnetic Separation
7. Charging in Ceramic Tubes
8. Reduction in Tunnel Kilns,
Approximately 1200C
9. Discharging
Figure. 2.1. Flow-sheet for the Hgans Sponge Iron Process.

2-8

10.
11.
12.
13.
14.
15.
16.
17.
18.
19.

Coarse Crushing
Storage in Silos
Chrusing
Magnetic Separation
Grinding and Screening
Annealing in Belt Furnace,
Approximately 800-900 C
Equalising
Automatic Packing
Iron Ore
Reduction Mix

2.3. THE HGANS SPONGE IRON PROCESS

The process starts with two raw materials: a reduction mix consisting of coke breeze
blended with ground limestone (1), and a pre-processed magnetite slick (2).
The magnetite slick and the reduction mix are being dried separately in two rotary ovens
(3). The slightly agglomerated dried reduction mix is crushed (4) and screened (5), and
the dried magnetite slick is passed through a magnetic separator (6). Then, both
materials are charged by means of an automatic charging device into tube-like ceramic
retorts as illustrated (7), (18), (19). These retorts have an ID of 40 cm, are 2 m long, and
consist of four tube segments of silicon carbide being stacked on top of each other.
These retorts are standing, 25 each, on rail-bounded cars which are clad with a thick
layer of refractory bricks. These cars are traveling slowly through a tunnel kiln of approx.
260 m length (8) within which the retorts are gradually heated to a maximum
temperature of approx. 1200C. As the temperature inside the retorts increases, the coke
breeze begins to burn forming CO which, in turn, begins to reduce the magnetite to
metallic iron while itself oxidizing to CO2.
The so generated CO2 reacts with the remaining coke breeze forming new CO,
which again reduces more magnetite to metallic iron. This reaction cycle continues until
all magnetite has been reduced to metallic iron and the major part of coke breeze is
burned up. Parallel to the reduction cycle, the limestone in the reduction mix binds the
sulfur arising from the burning coke breeze.
After completed reduction, the retorts are slowly cooled down again to approx. 250C
before leaving the kiln. Inside each retort, there is now a tube-like sponge iron cake with
a porosity of about 75%, a residue of unburned coke breeze, and a sulfur-rich ash. At an
automatic discharging station (9), the sponge iron tubes are pulled out off and the
remaining coke breeze and ash are exhausted from the retorts. Thereafter, the retorts are
ready to be charged again and go on a new trip through the tunnel kiln.
The sponge iron tubes (after having been cleaned from adhering coke breeze and ash) are
in several steps crushed and comminuted to a particle size below 3 mm (10). The so
obtained crude powder is intermediately ensilaged before further processing. From the
intermediate silo (11), the crude powder is passed through a specially designed chain of
magnetic separators (12), mills (13) and screens (14), in order to be rened to a particle
size below 150 m (<100 Tyler-mesh) and a well dened bulk density (apparent density).
Subsequently, the powder is passing a belt furnace (15) where it is soft-annealed at 800 1000C in hydrogen, and its remaining contents of carbon and oxygen are reduced to a
very low level. During annealing, the powder agglomerates to a very crumbly cake which
is gently comminuted again in a special mill (not shown on the ow-sheet). The so
treated powder has good compressibility and high green strength.
2-9

2. PRODUCTION OF IRON AND STEEL POWDERS

Many belt furnaces are, of course, available to take care of the huge volume of iron
powder to be treated. These modern electrically heated belt furnaces are especially
designed for the purpose, the belt width being 1500 mm.
The powder from several belt furnaces is collected in a special silo (16), where it is
homogenized in lots of 60 or 120 tons. Each lot is carefully checked with respect to
specied properties and packed and stored, ready for shipment (17).

2.4 The Hgans Water-Atomizing Process


A modern atomizing plant of Hgans AB is located at Halmstad, a small coastal town
80 km north of Hgans. The water-atomizing process, as practiced there, can be
followed on the ow-sheet shown at Fig. 2.2, and is described in detail on the next page.

2-10

2.4. THE HGANS WATER-ATOMIZING PROCESS

Figure. 2.2. Flow-sheet for the Hgans Water-Atomizing Process.

2-11

2. PRODUCTION OF IRON AND STEEL POWDERS

The raw material for this process is a carefully selected iron scrap and sponge iron from
the process described in the preceding paragraph. This raw material (1) is melted down
in an electric arc furnace of 50 tons capacity (2) where, if desired, alloying elements can
be added.
The melt is teemed slag-free through a bottom hole into a ladle (3) where it is rened
with an oxygen lance (4). The ladle is then transferred to the atomizing station (5), and
the liquid iron (or steel) is teemed slag-free through a bottom hole in the ladle into a
specially designed tundish (A).
From there, the liquid iron (or steel) ows in a well controlled stream (B) through
the center of a ring-shaped nozzle (D) where it is hit by jets of highly pressurized water
(C). The stream of liquid iron (or steel) explodes into ne droplets (E). Some of these
droplets freeze immediately to small spheres, others unite in small irregularly shaped
agglomerates while freezing.
Air, swept along by the water jet, and water vapor arising in the atomizing process, cause
supercial oxidation of the small droplets. The solidied droplets and the atomizing
water are collected in a huge container, where they are settling as a mud. This powder
mud is de-watered (6) and dried (7).
The dry powder is magnetically separated from slag particles (8), screened (9) and
homogenized (10), and eventually transported in special containers (11) to the works at
Hgans for further processing.
In the state as leaving the atomizing plant, the atomized powder particles are not only
supercially oxidized but also very hard because, due to the extremely high cooling rates
residing in the atomizing process, they have solidied in the martensitic state despite
their low carbon content.
The powder is, therefore, soft-annealed, and its surface oxides and residual carbon are
reduced in belt furnaces of the same type as described in the preceding paragraph.
Routines for homogenizing, quality checking, packaging and storing are the same as for
sponge iron powders.

2-12

2.5. ALLOYING METHODS

2.5 Alloying Methods


In order to achieve hardenable sintered ferrous materials, carbon and other suitable
alloying elements, like e.g. copper, nickel, and molybdenum, have to be introduced.
While carbon is normally admixed to the iron powder in the form of graphite, metallic
alloying elements are commonly introduced by either of the following two methods:
Method 1

Water atomization of the liquid iron alloy, resulting in a homogeneously


alloyed powder.
Method 2 Mechanically blending plain iron powder with the respective alloying
elements in powder form, and letting the actual alloying process take place
during sintering of the parts compacted from the powder mix.
Both methods have their advantages and disadvantages:
Homogeneously alloyed powders.
Advantages:
Alloying elements do not segregate when the powder is handled .
Yield fully homogeneously alloyed sintered parts.
Disadvantages:
Have low compressibility, because their particles are solution-hardened.
(See Figs. 2.3 and 2.5).
In order to change or correct the composition of a fully alloyed powder, if ever so
little, a new melt (usually 50 tons at time) will have to be atomized.
Powder mixes.
Advantages:
Have higher compressibility. (See Fig. 2.5).
No additional mixing operation is required as the powder has to be admixed with a
lubricant anyway.
The composition of a powder mix can very easily be changed or corrected by re-mixing it with additional amounts of either iron powder or alloying elements.
Disadvantages:
Yield less homogeneously alloyed sintered parts, because the admixed alloying elements (except carbon) diffuse very slowly in solid iron.
(ref. Chapter 6, Figs. 6.9 and 6.10).
Alloying elements tend to segregate when the powder mix is transported and handled. (However, powder mixes can be made segregation-proof by means of special treatments as described in the following paragraph).

2-13

2. PRODUCTION OF IRON AND STEEL POWDERS

The diagram at Fig. 2.3 shows the inuence of alloying elements on the hardness of iron.

Figure. 2.3. Inuence of some alloying


elements on the hardness of iron.

2.6 Distaloy and Starmix


In order to eliminate the segregation problem with powder mixes, Hgans AB has
developed two special processes for the production of segregation-proof iron powder
mixes. A large variety of standard and tailor-made powder mixes produced according to
these processes are offered under the trade-names Distaloy and Starmix.
The Distaloy process can be described as follows:
Alloying elements used in the Distaloy process are mainly copper, nickel and
molybdenum (but not graphite!) in the form of very nely grained powders. The process
starts with weighing-in a production lot of 30 tons of iron powder and alloying powders
in exactly controlled proportions. This lot is mixed in a double-cone mixer. Special
precautions are taken to prevent segregation of the mix when discharged from the mixer.
The so produced powder mix is heat-treated in a continuous furnace under a
reducing atmosphere at a temperature somewhat below the melting point of the lowestmelting alloying element. During this heat-treatment, the ne particles of the added
alloying elements are safely bonded to the surfaces of the coarser iron powder particles.

2-14

2.6. DISTALOY AND STARMIX

See SEM-photographs at Fig. 2.4. Due to inter-diffusion between the diffusion bonded
alloying particles and the iron particles, the latter become, to a certain extend, locally
pre-alloyed.

Figure. 2.4. SEM-photographs showing ne particles of copper, nickel and molybdenum diffusion bonded
in the Distaloy process, to the surface of an iron powder particle.

2-15

2. PRODUCTION OF IRON AND STEEL POWDERS

The so treated powder mix contains the alloying additives as nely and evenly dispersed
as possible . The fact that the iron powder particles are locally pre-alloyed has practically
no negative effect on the compressibility of the mix.
See Fig. 2.5, where the compressibility curve for a Distaloy powder is shown in
comparison with those for an ordinary powder mix, and for an atomized homogeneously
alloyed powder of identical compositions.

Figure. 2.5. Compressibility curves for


three ferrous powders produced by
different methods but having the same
chemical composition: 1.75% Ni, 1.5%
Cu, 0.5% Mo, remainder Fe.

Graphite and lubricants have to be excluded from the Distaloy process because, during
heat-treatment of the powder mix, graphite would carbonize the iron particles and spoil
the compressibility of the powder mix, and lubricants would burn-off.

2-16

2.6. DISTALOY AND STARMIX

The Starmix process uses special types of organic binders to glue graphite and lubricants
to the iron powder particles during the mixing procedure.
See SEM-photograph at Fig. 2.6.

Figure. 2.6. SEM-photograph showing


ne graphite particles glued, in the
Starmix process, to the surface of an iron
powder particle (NC100.24).

2-17

2. PRODUCTION OF IRON AND STEEL POWDERS

The Starmix process can, of course, be applied to ordinary iron powder mixes as well as
to Distaloy powders; in both cases, it yields segregation-proof press-ready powder
mixes. The gain in product consistency achieved through the Starmix process is
illustrated in the diagrams at Figs. 2.7 to 2.10.

Figure. 2.7. Graphite loss in air


de-dusting test for a
Starmix-treated powder mix and
for a conventional powder mix.

Figure. 2.8. Relative frequency of dimensional


changes during sintering
for a Starmix-treated
powder mix and for a
conventional powder mix.

2-18

2.6. DISTALOY AND STARMIX

Figure. 2.9. Variation of


carbon content in sintered
parts during mass production,
comparing Starmix-treated
and conventional mixes of
NC100.24 + 1.2%
graphite.

Figure. 2.10. Comparison of


powder properties and sintered
properties for Starmix-treated and
conventional powder mixes of
NC100.24 + 0.8% graphite
+ 0.8% Zn-stearate.

2-19

2. PRODUCTION OF IRON AND STEEL POWDERS

2-20

CHARACTERISTICS OF
IRON AND STEEL POWDERS
The most important prerequisite of a successful mass
production of high-quality sintered parts is a high and
consistent quality of the powder from which the parts are to
be made. The characteristics of the powder decide about its
compacting and sintering behavior and, eventually, about the
properties of the nished product.

3. CHARACTERISTICS OF IRON AND STEEL POWDERS

TABLE OF CONTENTS

3.1 GENERAL ASPECTS.......................................................... 3


3.2 PROPERTIES OF HGANS IRON POWDERS ....................... 6

3.1. GENERAL ASPECTS

3.1 General aspects


Iron and steel powders - as well as other metal powders - used in the production of
sintered parts can be characterized by three categories of properties:
1. Metallurgical properties
chemical composition and impurities
microstructure
microhardness
2. Geometrical properties
particle size distribution
external particle shape
internal particle structure (particle porosity)
3. Mechanical properties
flow rate
bulk density
compressibility, green strength and spring-back
All these powder properties are inherited from and specic to the process by which the
powder was produced. Some of them are interrelated with each other. For instance:

microstructure and microhardness are depending on chemical composition;


compressibility decreases with increasing microhardness, increasing particle porosity
and decreasing particle size;
coarser powders and powders of regular particle shape ow better than ne powders
and powders of irregular particle shape;
powders of irregular particle shape have better green strength after compacting than
powders of regular particle shape.

In the following, we present a brief denition of the aforementioned powder properties


and their relevance to processing steps in the production of sintered parts.
Metallurgical properties
are determined by chemical analysis and metallographic procedures. The chemical
composition of a ferrous powder has a great inuence upon the nal strength properties
of the sintered parts. Non-metallic impurities may have an adverse effect upon
compressibility and upon the life of compacting tools.

3-3

3. CHARACTERISTICS OF IRON AND STEEL POWDERS

Geometrical properties
viz. particle size distribution, particle shape and particle porosity, determine the powders
specic surface which is the driving force of the sintering process (Chapter 6).
Particle size distribution
is determined by sieve analysis if particle sizes are above 45 m (minimum screen
aperture). Finer powders are suspended in water and analyzed by means of laser
diffraction methods.
External particle shape
is analyzed by means of scanning electron microscopy (SEM). (See Fig. 3.1 left).
Internal particle structure
is analyzed by means of metallographic techniques. (See Fig. 3.1 right).

Figure 3.1. External particle shape (SEM) and internal particle structure (cross-section) of sponge iron
powder (NC100.24) and water-atomized iron powder (ASC100.29).

3-4

3.1. GENERAL ASPECTS

Flow rate
or rather its reciprocal value, is the time in seconds which an amount of 50 g dry powder
needs to pass the aperture of a standardized funnel. Flow rate is inuenced by type and
amount of lubricant admixed to the powder. The ow rate decides about how fast a
compacting tool can be lled with powder, and thus is a limiting factor in the
compacting cycle of the powder press.
Bulk density (Apparent density)
is determined by lling the powder through a standardized funnel into a small cup,
leveling-off the surplus powder on top of the cup, and dividing the weight of powder
contained in the cup by the cup volume (25 cm3). Apparent density is inuenced by
type and amount of lubricant admixed to the powder. The apparent density of the
powder decides about the required lling depth of the compacting tool.
Compressibility
is the name of a curve obtained by plotting the compact densities of a series of small
cylindrical powder compacts ( 25 mm), over the applied pressures. Compact density is
the weight of a powder compact divided by its bulk volume. Compact density is
inuenced by type and amount of lubricant admixed to the powder. Green density is the
compact density of a small cylindrical powder compact ( 25 mm) pressed with a
standardized pressure (either 4,2 tons/cm2 or 600 N/mm2).
The compressibility of the powder decides about how high a compacting pressure is
needed to achieve a desired compact density.
Green strength
is the bending strength of a green (i.e. compacted but not sintered) rectangular test bar.
Green strength increases with increasing compact density and is inuenced by type and
amount of lubricant admixed to the powder. Sufcient green strength is required to
prevent compacts from cracking during ejection from the compacting tool and prevent
them from getting damaged during handling and transport between press and sintering
furnace. The more complex and delicate the shape of a compact, the higher its green
strength should be. If the green strength of compacts is high enough, they may even be
machined prior to sintering (e.g. undercuts, traverse slots and holes).
Spring-back
is the elastic expansion of a cylindrical powder compact ( 25 mm) after ejection from
the compacting die. Its value is expressed as the difference between the OD of the
compact and the ID of the (empty) die divided by the ID of the die. Spring-back
increases with increasing compacting pressure and is inuenced by type and amount of
lubricant admixed to the powder, and by the elasticity coefcient of the die material in
which the powder is compacted.
The spring-back value is important for calculating the exact dimensions of the
compacting tool in relation to the required dimensions of the compact.
3-5

3. CHARACTERISTICS OF IRON AND STEEL POWDERS

3.2 Properties of Hgans Iron Powders


As has been mentioned in the preceding chapter, Hgans AB produces two different
kinds of ferrous powders:
sponge-iron powders, and
water-atomized (unalloyed and low-alloyed) iron powders.
Typical structural differences between these two kinds of powders appear on SEMphotographs and cross-sections of representative powder particles shown at Fig. 3.1. The
external shapes of both particles are irregular and fairly similar to one another. But the
sponge iron particle has, as its name suggests, a spongy internal structure, while the
water-atomized particle is internally compact.
Both kinds of powder are specially treated to yield various standard grades having
different properties to t different applications. These standard grades are also used in
various press-ready powder mixes, i.e. blended with lubricants and alloying additives like
graphite, copper, nickel, molybdenum, phosphorous and others.
Properties of three sponge-iron grades NC100.24, SC100.26, MH80.23, and of
two water-atomized iron powder grades ASC100.29, ABC100.30, are presented in the
diagrams at Figs. 3.2 and 3.3 and in Table 3.1.
Table 3.1 Properties of some Hgans Iron Powders
Powder
grade

Approx.
particle size
range m

Apparent
density
g/cm3

NC100.24

20 180

2.45

SC100.26

20 180

2.65

MH80.23

40 -200

ASC100.29

20 -- 180

ABC100.30

30 -200

C
%

Green
density

Green
strength

1)

1)

g/cm3

N/mm2

<0,01

7,02

54

29

0,10

<0,01

7,12

48

2,30

34

0,3

0,08

6,27 (2)

24(3)

2,98

24

0,08

0,003

7,20

40

3,02

24

0,05

0,002

7,27

40

1)

compacted at 600 N/mm2 in a lubricated die.


compacted at 4.2 t/cm2 in a lubricated die.

3)

measured at a green density of 6.0 g/cm2

31

H2-loss
%

0,20

2)

3-6

Flow
s/50g

3-7

3.2. PROPERTIES OF HGANS IRON POWDERS

Figure.3.2. Particle Size Distribution, Flow and Apparent Density of ve different iron powder grades.

3. CHARACTERISTICS OF IRON AND STEEL POWDERS

3-8
Figure.3.3. Compressibility, Green Strength and Spring-back of ve different iron powder grades.

3.2. PROPERTIES OF HGANS IRON POWDERS

NC100.24
is one of the most widely used grades in the manufacturing of sintered parts. Due to the
irregular surface and the spongy structure of its particles, it has high green strength; and
due to its low contents of oxygen and carbon, it has good compressibility.
SC100.26
has the best compressibility of all Hgans sponge-iron powder grades. Its green strength
is slightly lower and its apparent density slightly higher than for NC100.24. It is very
useful in cases where parts with high density are to be achieved in one single pressing
operation.
MH80.23
is especially designed to match the requirements for self-lubricating bearings. Its particle
size range is chosen to give a pore structure optimal for this application. MH80.23 can
also be added to powder mixes in small quantities to dramatically improve green
strength.
ASC100.29
is a water-atomized iron powder which due to its high purity and its compact particle
structure has very high compressibility. ASC100.29 can be compacted in one single
pressing operation at moderate pressures to densities of up to 7.2 g/cm3. It is also
suitable as basic material for soft-magnetic applications.
ABC100.30
is a water-atomized iron powder of outstanding compressibility and chemical purity. It is
especially well suited for the production of high-density structural components.
Densities of up to 7.4 g/cm3 are achievable in one single pressing operation. ABC100.30
is also used in applications where very good soft-magnetic properties are required.
An interesting feature of practical importance must be mentioned here: The apparent
density of iron powders increases with increasing mixing time. This phenomenon can be
utilized in some cases to match a slightly under-dimensioned lling depth in the
compacting tool. See diagrams at Fig. 3.4.

3-9

3. CHARACTERISTICS OF IRON AND STEEL POWDERS

a sponge grade
i.e NC100.24

b atomized grade
i.e ASC100.29

Figure.3.4 Inuence of Mixing


Time upon Apparent Density of
two different iron powder grades
admixed with three different lubricants.

Apart from unalloyed iron powder grades, Hgans AB also produces water-atomized
low-alloyed steel powders trade-name Astaloy , and a variety of diffusion alloyed steel
powder trade-name Distaloy. (The Distaloy process has been described in some
detail in the preceeding chapter).

3-10

3.2. PROPERTIES OF HGANS IRON POWDERS

Properties of some Astaloy and Distaloy powders are presented in Table 3.2.
Table 3.2 Properties of some Astaloy and Distaloy powders
Powder
grade

Approx.
particle size
range m

Apparent
density
g/cm3

Green
density

Green
strength

1)

1)

g/cm3

N/mm2

Astaloy A

20 180

3.30

23

0.10

0.01

7.01

17

Astaloy Mo

20 180

3.28

23

0.07

0.01

7.14

22

Distaloy SA

20 150

2.83

27

0.13

<0.01

7.09 2)

41

Distaloy SE

20 150

2.82

28

0.12

<0.01

7.11

39

Distaloy DC

20 180

3.28 2)

24 2)

0.10

<0.01

7.16

22

Distaloy DH

20 180

3.41 2)

22 2)

0.10

0.01

7.13

29

Distaloy HP

20 180

3.37 2)

22 2)

0.10

0.01

7.07

25

1)

compacted at 600 N/mm2 in a lubricated die.

2)

rust inhibitor admixed.

Flow
s/50g

H2-loss
%

C
%

Astaloy A
is a water-atomized steel powder alloyed with 1.90% Ni, 0.55% Mo and 0.20% Mn.
Astaloy A is primarily intended for powder forging, but is also used for pressed and
sintered structural components. It has very good hardenability and can be used for casehardened as well as through-harden parts with non-critical tolerances.
Astaloy Mo
is a water-atomized steel powder alloyed with 1.50% Mo. Astaloy Mo has high
compressibility, exhibits an homogeneous microstructure after sintering, and has optimal
hardenability. These properties make it an excellent choice for surface-hardened
components requiring high surface hardness in combination with good core toughness.
Distaloy SA
is based on the sponge-iron grade SC100.26, to which 1.75% Ni, 1.5% Cu, and 0.5%
Mo in form of very nely dispersed powders have been diffusion-bonded. Distaloy SA is
recommended for densities up to 6.9 g/cm3 after single pressing. With additions of
graphite, a tensile strength of 600 N/mm2 can be achieved in one pressing and sintering
operation, and so produced parts respond well to heat-treatment.

3-11

3. CHARACTERISTICS OF IRON AND STEEL POWDERS

Distaloy SE
is also based on the sponge-iron grade SC100.26, but contains 4% Ni, 1.5%Cu, and
0.5% Mo. Due to its higher nickel content, it has better hardenability than Distaloy SA
and is especially well suited for large sintered parts requiring heat-treatment.
Distaloy DC (DC = Dimensional Control)
is based on Astaloy Mo to which 2% nely dispersed nickel powder has been diffusionbonded. It thus contains 2% Ni + 1.47% Mo. Parts compacted from Distaloy DC
admixed with graphite exhibit high strength and very small scattering of dimensions
after sintering, independent of compact density; and the sintered parts respond very well
to subsequent heat-treatment.
Distaloy DH (DH = Direct Hardening)
is based on Astaloy Mo to which 2% nely dispersed copper powder has been diffusionbonded. It thus contains 2% Cu + 1.47% Mo. Parts compacted from Distaloy DH
admixed with graphite can be hardened directly from sintering heat (in belt furnaces
equipped with a rapid cooling system) and exhibits small scattering of dimensions after
sintering and hardening.
Distaloy HP (HP = High Performance)
is based on Astaloy Mo to which 4% nely dispersed nickel powder and 2% nely
dispersed copper powder have been diffusion-bonded. It thus contains 4% Ni + 2%Cu
+ 1.41% Mo. Parts compacted from Distaloy HP admixed with graphite exhibit
dimensional changes close to zero during sintering and adopt very high strength values.
More detailed information about all powder grades described above, as well as about
many other interesting powder grades and powder mixes can be obtained from Hgans
AB in form of special brochures and electronic data.

3-12

The index comprises all chapters from the three handbooks.


Each index word is followed by the chapter number and
relevant page number.

HGANS HANDBOOK FOR SINTERED COMPONENTS

A
A1-point 1-43
A3-point 1-43
ABC100.30 3-6, 3-9
activated sintering 6-17
activation energy 1-24, 1-26
adhesive friction 4-25
alloying methods 2-13
alloying systems 9-8
Astaloy 85 Mo 9-27
Astaloy Mo 9-27
Distaloy AB 9-27
Distaloy AE 9-27
Distaloy DC 9-27
Distaloy DH 9-28
Distaloy HP 9-28
Distaloy SA 9-27
Distaloy SE 9-27
influence of carbon content 9-14
iron-carbon 9-12
iron-copper, iron-copper-carbon 9-15
iron-copper-nickel-carbon 9-24
iron-copper-nickel-molybdenum-carbon 9-27
iron-phosphorus-carbon 9-20
plain iron 9-8
amorphous solids 1-3
stable short-range order 1-3
apparent density 3-5
ASC100.29 3-4, 3-6, 3-9
Astaloy A 3-11
Astaloy Mo 3-11
atoms per unit cell 1-11
austempering 1-58, 1-59
austenite
I-2

INDEX

area 1-43
grain 1-49, 1-68
residual 1-50, 1-58
transformation 1-47
axial density distribution 4-22
axial pressure 4-14
B
bainite
formation 1-50
lower 1-50
nose 1-55, 1-56
step 1-45
upper 1-50
bainitic steel 1-51
bainitizing 1-58
ball-sizing 7-22, 7-23, 7-24
BCC 1-6, 1-11
BCT 1-46, 1-48
belt furnace 2-9
binary phase diagrams
eutectic system 1-32, 1-33
blistered sintered iron parts 6-44
bonding between atoms 1-5
covalent bond 1-5
electron gas 1-5
ionic bond 1-5
metallic bond 1-5
Van der Waals force 1-5
Boudouard reaction 6-36, 6-45
bridging phenomena 5-5
bulk density 3-3, 3-5
burn-off zone 6-25

I-3

HGANS HANDBOOK FOR SINTERED COMPONENTS

C
carbon
potential 6-42
precipitation 6-47, 6-48
precipitation from gas mixtures 6-46
precipitation inside pores 6-45
restoring zone 6-25
carbon-steel 1-40
cast-iron 1-40
CCT-diagram 1-52, 1-60, 1-61
Astaloy Mo + 0,60%C 9-40
Distaloy AE + 0,50%C 9-34
Distaloy DH + 0,40%C 9-36
Distaloy HP + 0,50%C 9-38
Distaloy SA + 0,45%C 9-32
cementite
formation 1-48
cementite reaction 6-36
ceramic retorts 2-9
chamfers, fillets and tapers
chamfers 8-6
chamfers and burrs 8-7
corners and edges facing the core rod 8-9
corners and edges facing the die 8-9
fillets 8-7
rounded-off edges 8-8
spherical end 8-10
tapered sides formed by the die 8-10
tapered sides formed by upper punches 8-11
chemical composition and impurities 3-3
clearance between sliding tool members 5-28
close-packed 1-13
close-packed lines 1-12
close-packed planes 1-12
I-4

INDEX

coke breeze 2-7


compact density 3-5, 5-21
compacting cycle 5-4, 5-10
compacting cycle for a cylindrical bushing 5-9
compacting cycle for a two-level part 5-11
compacting in a cylindrical die 4-4
compacting of metal powders 4-1
compacting pressure 4-3
compacting punches 4-3
complete miscibility 1-29
component with flange and blind hole 5-16
composition of the powder mix 6-3
compressibility 3-3, 3-5
constants 1-7
contact areas 4-7
continuous cooling transformation diagrams 1-52
control of sintering atmospheres 6-43
cooling curves 1-32, 1-34
critical 1-55
cooling rate 1-44, 1-45
cooling zone 6-25
corrosion protection 10-34
phosphatizing 10-37
steam treatment 10-34
covalent bond 1-5
crack formation 5-12, 5-13, 5-14
crack propagation 1-68
cracked ammonia 6-25, 6-40
cracking of sintered iron powder parts 6-47, 6-48
crude powder 2-9
crystal
grains 1-14
lattice 1-8

I-5

HGANS HANDBOOK FOR SINTERED COMPONENTS

segregation 1-38, 1-39


twins 1-14
crystalline solids 1-3
crystalline long-range order 1-3
crystallites 1-14
Curie-point 1-40
D
deburring and cleaning 10-29
abrasive blasting/shot blasting 10-29
barreling 10-29
electrolytic-alkaline cleaning 10-29
ultra-sonic cleaning 10-29
vibratory deburring 10-29
decarbonization and carbonization 6-34
decomposing lubricant 6-45
decrease of maximum shearing stress 4-10
deformation strengthening of powder particles 4-9
degree of homogenization 6-12, 6-13
dendrites 1-14
dendritic crystal nuclei 1-39
densification 4-3
densifying the powder 5-6
density 4-3
density of the powder compact 6-4
density-pressure curves 4-4
depth of fill 5-21
design features 8-6
designing a compacting tool 5-19
dew point 6-42
die being withdrawn 5-7
die cavity 4-3
die compacting 4-3
die lubrication 7-10
I-6

INDEX

dies and core rods 5-30


diffusion
annealing 1-39
coefficient 1-21, 1-24
equation 1-21
Ficks first law 1-20
Ficks second law 1-21
grain boundary 1-20, 1-25
in metals 1-19
interstitial 1-20
laws 1-20
rate 1-24
self 1-20
surface 1-20, 1-25
systems 1-26
vacancy 1-20
volume 1-20, 1-25
diffusion coefficient 6-12, 6-13
dimensional accuracy 8-4
dislocation 1-15
climbing 1-18, 1-19
edge 1-15, 1-17
line 1-15, 1-16, 1-17, 1-18
pile up 1-18
screw 1-15, 1-17
dispersion hardening 1-17
dissociation pressure 6-30
dissociation temperature 6-30
Distaloy 2-4, 2-14, 2-16, 2-18
process 2-14, 2-15
Distaloy DC 3-11, 3-12
Distaloy DH 3-11, 3-12
Distaloy HP 3-11, 3-12
Distaloy SA 3-11
I-7

HGANS HANDBOOK FOR SINTERED COMPONENTS

Distaloy SE 3-11, 3-12


double-sided densification 5-7
E
ejecting force 4-25, 4-26, 4-27
ejection principle 5-8
ejection procedure 5-13, 5-14
elastic expansion 4-27
elastic expansion of two lower punches 5-12
elastic loading 4-15, 4-18
elastic releasing 4-16, 4-18
electric arc furnace 2-12
electrolytic iron powder 2-6
Ellingham-Richardson diagram 6-28, 6-29, 6-32, 6-34
endogas 6-25, 6-40, 6-41, 6-42
entropy of mixing 1-29
EQ-hardenability 1-61
equilibrium
dissociation pressure 6-30, 6-31
temperatures 6-33
equilibrium diagram
Fe - Fe3C - C - CH4 6-37
Fe - FeO - Fe3O4 - CO - CO2 6-36
Fe - FeO - Fe3O4 - H2 - H2O. 6-35
equilibrium diagram iron-carbon 1-40
equilibrium states of carbon-steel and cast-iron 1-40
error function 1-21
eutectic 1-32
alloy 1-34
line 1-33
point 1-33
reaction 1-34
temperature 1-33
evaporation/condensation 6-5
I-8

INDEX

exogas 6-41
external 3-3
external particle shape 3-3, 3-4
F
face-centered-cubic 1-10
FCC 1-6, 1-11, 1-13
Fe- Fe3C - diagram 1-45
ferrite 1-40, 1-41, 1-43, 1-44
formation 1-47
filling density 5-21
filling the die 5-4, 5-5
first and second law of thermodynamics 1-28
flanged
bushing 7-31
connections 7-31
floating die 5-7
floating-die principle 5-8
flow rate 3-3, 3-5
formation of bridges 5-5
free energy 1-27, 1-28, 1-29, 1-30, 1-31, 1-32, 1-48, 1-49
interfaces 6-15
oxidation 6-28
free surface energy 6-5
frictional coefficient at the die wall 4-25
frictional force 4-17
functional sketch of the tool 5-20
further design considerations
shape and function 8-26
sintering behavior 8-26
tooling economy 8-25

I-9

HGANS HANDBOOK FOR SINTERED COMPONENTS

G
gases 1-3
free length of way 1-3
random order 1-3
geometrical properties 3-3, 3-4
geometrical structure of the powder particles 6-3
grain boundaries 1-14
cementite 1-42
diffusion 1-19, 1-25
grain-boundary diffusion 6-5
grain-size distribution 6-10
green density 3-5
green strength 3-3, 3-5
growth 1-53
during sintering 6-17
neck 6-6, 6-7
H
Hametag 2-5
hardenability of Astaloy and Distaloy materials 9-31
hardness of pearlite I-62
HCP I-6, I-11, I-13
heat treatment of steel 1-57
heat-treatment
through hardening 10-3
heat-treatments 10-5
austenitizing 10-5
carbonitriding 10-14
carbonizing 10-10
case hardening 10-7
controlling case depth 10-8
defining case depth 10-9
induction hardening 10-21
measuring case hardness 10-8
nitriding 10-18
I-10

INDEX

nitrocarbonizing 10-20
plasma-nitriding 10-19
precipitation hardening 10-6
quenching 10-5
tempering 10-5
through-hardening 10-5
height of compact 5-21
Helmholtz free energy 1-27
heterogeneous system 1-27
history of iron powder 2-5
Hgans 2-6
sponge iron powder 2-6
sponge iron process 2-7, 2-8
water-atomizing process 2-10
holes and wall thickness
alphanumeric characters 8-24
assemblies 8-23
blind holes 8-20
feather edges 8-21
grooves and undercuts 8-21
holes 8-17
knurls 8-22
narrow holes 8-18
special shapes 8-23
taper holes (wider end down) 8-20
taper holes (wider end up) 8-19
threads 8-22
wall thickness 8-18
hollow sphere 4-10
homogeneously alloyed powders 2-13
homogenization time 6-12
homogenizing 1-39
horizontal cracks 4-27
horizontal shearing stress 4-27
I-11

HGANS HANDBOOK FOR SINTERED COMPONENTS

hot zone 6-25


hydrogen 6-25, 6-40
hydrostatic pressure 4-11
hypoeutectoid steels 1-47
hysteresis 4-16
hysteresis curve 4-17
hysteresis of the radial pressure 4-15
I
ideal crystals 1-4, 1-7
industrial sintering atmospheres 6-39
infiltration and impregnation 10-22
impregnation with polymers 10-22
infiltration with metals 10-22
oil impregnation 10-23
influence of carbon content on hardness 1-63
influence of profiles 5-32
influence of the microstructure on the properties of steel 1-62
influence of the yield point 4-20, 4-21
inter-metallic phase Fe3C 1-40
internal 3-3
internal energy 1-27
internal particle structure 3-3, 3-4
interrelation between the FCC lattice 1-48
interstices
octahedral 1-47
tetrahedral 1-47
interstitial atom 1-14
interstitial elements 6-12
ionic bond 1-5
isostatic compacting 4-3
isostatic powder compacting 4-5
isothermal
annealing 1-58, 1-59
I-12

INDEX

soft-annealing 1-58
transformation 1-53
transformation diagrams 1-52
ITT-diagram for a hypoeutectoid carbon steel 1-54
ITT-diagrams 1-52, 1-53
J
jets of highly pressurized water 2-12
joining 10-29
adhesive techniques 10-33
brazing 10-30
riveting techniques 10-33
shrink-fitting 10-32
Sinter Braze 90 10-30
welding 10-32
K
Kuczynskis model 6-6
L
lamellas of cementite and ferrite 1-49
lattice
direction 1-9, 1-10
disturbances 1-14, 1-15, 1-16, 1-17
planes 1-7, 1-9, 1-10, 1-12, 1-16
points 1-7, 1-10
sites 1-19
structure 1-14, 1-20
lever-rule of phases 1-34, 1-37, 1-38
limited miscibility 1-33
limits to densification 4-8
liquids 1-3
instable short-range order 1-3
liquidus 1-30
I-13

HGANS HANDBOOK FOR SINTERED COMPONENTS

load distribution on punches 5-32


loading-releasing cycle 4-16
Longs model 4-16, 4-20
low melting eutectic 6-18
lubrication for sizing and coining 7-9
M
machining 10-24
drill-life test 10-25
machinability-enhancing additives 10-25
machining parameters 10-28
magnetic separator 2-9
magnetite 2-7
martensite 1-43, 1-47
and bainite 1-64
formation 1-49
martempering 1-58, 1-59
step 1-46
maximum shearing-stress 4-15
mechanical properties 3-3
mechanisms of sintering 6-5
meta stable
phase 1-40
system 1-40
metallic bond 1-5
metallurgical properties 3-3
MH80.23 3-6, 3-9
microhardness 3-3
microstructure 3-3
ABC100.30 9-11
ASC100.29 + 2%Cu 9-16
ASC100.29 + 4%Cu 9-16
Astaloy Mo + 0,60%C 9-41
Distaloy AE+ 0,50%C 9-35
I-14

INDEX

Distaloy DH + 0,40%C 9-37


Distaloy HP +0,50%C 9-39
Distaloy SA + 0,45%C 9-33
hardened Distaloy AE + 0,5%C 9-43
NC100.24 9-9
NC100.24 + 0,45%P 9-22
NC100.24 + 0,45%P+ 0,5 %C 9-22
NC100.24 +2,5%Cu + 2,5%Ni + 0,6 %C 9-26
SC100.26 + 2%Cu + 0,2%C 9-19
SC100.26 + 2,5%Cu + 2,5%Ni 9-26
SC100.26 + 2%Cu + 0,6%C 9-19
variation with distance from surface 9-42
microstructures
ASC100.29 + 0,2%C 9-13
ASC100.29 + 0,5%C 9-13
migration of vacancies 6-5
Miller indices 1-7
mixed systems 6-38
mobility of carbon atoms 1-49
modulus of elasticity 4-15
Mohrs circle 4-10
multi-platen adapter 5-17, 5-18
multiple level parts
flanges and studs 8-16
gear hub 8-16
multiple punches 8-12
profiled faces 8-15
shelf die 8-13
slot made by a punch 8-15
step core rod 8-13
step in the punch face 8-14
multiple platen systems 5-15
multiple-function presses 5-8

I-15

HGANS HANDBOOK FOR SINTERED COMPONENTS

N
natural states of matter 1-3
NC100.24 3-4, 3-6, 3-9
neck formation 6-6
neck growth 6-8
neutral type 6-4
neutral zone 5-23
nitrogen 6-25, 6-43
non-equilibrium diagram iron-carbon 1-43
non-metallic inclusions 1-16
nucleation 1-53
of pearlite 1-49
number of nearest neighbor 1-11
O
oxidation and reduction 6-27
P
P/M - parts of different complexity 8-28
packing density 1-11
parameters of influence 9-3
alloying elements 9-4, 9-5
density 9-3, 9-7
dimensional stability 9-6
heat-treating conditions 9-6
sintering conditions 9-4, 9-7
parameters of state 1-27
particle 3-3
particle porosit 3-3
particle rearrangement 4-5
particle size distribution 3-3, 3-4
parts with external flanges 7-31
parts with internal flanges 7-35
pearlite 1-42, 1-62
I-16

INDEX

formation 1-49
lamellas 1-49
nose 1-55
step 1-44
peening and plating 10-33
electroplating 10-33
peen-plating 10-33
shot peening 10-33
phase diagram 1-27
piston 7-35
automobile shock absorbers 7-36
complete sizing 7-35
double-action press 7-37
faces and bore 7-36
internal profiles 7-38
lower core rod 7-40
using upper core rod 7-38
plain bushings 7-13
advanced concept 7-15
chamfers 7-13
core rod with a bulge 7-19
density 7-13
eccentricity 7-14
fitting of bushings 7-24
proportions 7-14
serrated core rod 7-18
simple concepts 7-14
sizing by balls 7-21
spherical bushings 7-25
surface finish 7-13
tolerances 7-13
plain parts without holes 7-12
plastic deformation 4-5
plastic deformation of metal crystal 1-16
I-17

HGANS HANDBOOK FOR SINTERED COMPONENTS

plastic flow in a hollow sphere 4-11


plastic loading 4-16, 4-18
plastic releasing 4-16, 4-18
Poisson factor 4-15
polycrystalline structure 1-14
pore-free density 4-11
pore-free zones 6-10
pore-size distribution 6-10
porosity 4-3
powder mixes 2-13
powder transfer without densification 5-16
production of iron and steel powders 2-1
profiled components with holes 7-29
profiled parts with holes 7-28
properties
AB + 0,6 % C 9-30
AE + 0,5 % C 9-30
DC + 0.5% C 9-31
DH + 0,5 % C 9-31
HP + 0.5% C 9-31
influence of alloying elements 9-5
influence of carbon content 9-18
influence of copper and carbon additions 9-29
influence of nickel and copper additions 9-24
influence of phosphorus additions 9-23
influence of phosphorus and carbon additions 9-21
influence of sintered density 9-3
SA + 0,5 % C 9-30
SE + 0,5 % C 9-30
properties of Hgans iron powders 3-6
properties of tool steels 5-30
protective atmosphere in the sintering furnace 6-4
punches 5-29

I-18

INDEX

Q
quenching and tempering 1-58, 1-59
R
radial and axial pressure 4-17, 4-19, 4-20
radial pressure 4-14
radial stress 4-10
random order 1-3
rapid burn-off 6-47
RBO 6-47
real metal crystal 1-14
reducing agents 6-31
reducing-carbonizing type 6-4
reducing-decarbonizing type 6-4
reduction in area 7-6
reduction mix 2-9
re-pressing 7-4, 7-5
required filling depths 5-21
residual radial pressure 4-27
ring sizing 7-31
location plate 7-31
oil pump gears 7-31
profiled sizing ring 7-31
ring-shaped nozzle 2-12
rotary ovens 2-9
RZ-process 2-5
S
SC100.26 3-6, 3-9
scatter in density 4-28
scatter in spring-back 4-28
self diffusion 1-20
self-tempered martensite 1-51
shearing yield-stress 4-15, 4-16
I-19

HGANS HANDBOOK FOR SINTERED COMPONENTS

shock absorber piston 5-14


shrinkage
during sintering 6-17
sintering atmosphere 6-25
neutral 6-25
reducing-carbonizing 6-25
reducing-decarbonizing 6-25
sintering behavior 6-20
iron-copper 6-23
iron-copper-carbon 6-23
plain iron powders 6-20
sintering furnaces 6-25
sizing and coining 7-6
sizing bushings with thick flanges 7-34
sizing flanged bushings 7-32, 7-33
sliding friction 4-25
sliding support 5-11
slip planes 1-16
solid state sintering
heterogeneous material 6-10
homogeneous material 6-5
solidus 1-30
sorbite 1-44
space lattice 1-7
specific weight 4-3, 4-12
specific weights
metals, additives and impurities 4-12
spherical bushings 7-26, 7-27, 7-28
spheroidite 1-62
spheroidized pearlite 1-62
sponge-iron powders 3-6
spray lubrication 7-11
spring-back 3-3, 3-5, 4-25, 4-27, 4-28
s-shaped transformation curve 1-52
I-20

INDEX

stable short-range order 1-3


stable system 1-40, 1-41
stacking sequence 1-12
stage of sintering 6-5
stages in a compacting cycle
1) filling the die 5-4
2) densifying the powder 5-4
3) ejecting the compact 5-4
stages in sintering 6-16
standard dissociation temperature 6-30
Starmix 2-14, 2-17, 2-18, 2-19
Starmix process 2-17
stationary case 1-21
stationary die 5-7
stationary lower punch 5-7
stick-slip behavior 4-26
substitutional atoms 1-14, 1-15, 1-20
substitutional elements 6-12
supersaturation 1-48
with carbon 1-48
surface diffusion 1-20, 1-25, 6-5
surface lubrication by oil spray 7-9
swaging 7-6
point 7-7
radius 7-7
swelling of a compact 6-15
system
Fe - Fe3C - C - H2 - CH4 6-37
Fe - FeO - Fe3O4 - Fe3C - CO - CO2 6-35
Fe - FeO - Fe3O4 - H2 - H2O 6-35
T
tangential stress 4-10
tapering the die exit 5-13
I-21

HGANS HANDBOOK FOR SINTERED COMPONENTS

technical problems 6-25


temper brittleness 1-68
temperature and time 6-3
tempered martensite 1-50, 1-51, 1-65, 1-67, 1-68
tetragonal martensite lattice 1-46
tetragonally distorted ferrite 1-46, 1-50
theoretical density 4-3
theoretical density of iron powder mixes 4-13
theoretical density of powder mixes 4-11
thermal analysis 1-32
thermite welding 6-29
thermodynamic
equilibrium 1-27
state 1-27
thermodynamical
problems 6-27
processes 6-27
thick-walled component 7-41
tie-line 1-33
time-temperature-transformation diagram 1-52
tolerances 8-4, 8-5
tolerances on tool members 5-25
tool materials 5-29
tooling costs 5-33
tools for sizing and coining 7-12
transformation diagrams of steel 1-52
transient liquid phase 6-15
transition points A3 and A1 1-44
trostite 1-44
TTT-diagram 1-43, 1-52
tumbling in dry lubricant 7-9
turn-over sizing 7-26
twinning 1-14
two-level part 5-10
I-22

INDEX

U
undercooling
step 0 (equilibrium) 1-44
step I 1-44
step II 1-44
step III 1-45
step IV 1-46
unit cell 1-7, 1-9, 1-10, 1-11, 1-12, 1-47
V
vacancy 1-14, 1-15
Van der Waals force 1-5
viscous or plastic flow 6-5
volume diffusion 1-20, 1-25, 6-5, 6-9
W
water-atomized iron and steel powders 2-6
water-atomized iron powders 3-6
water-atomizing process 2-10
withdrawal principle 5-8
withdrawal-type tool 5-11
world-wide usage of iron powder 2-3
Y
yield point 1-19
Z
zones of a continuos sintering furnace 6-26

I-23

You might also like